Skip to main content

Full text of "Naval hydrodynamics : boundary layer stability and transition, ship boundary layers and propeller hull interaction, cavitation, geophysical fluid dynamics : twelfth symposium"

See other formats


2 s2eec000 Toeo O 


AU A 


1OHM/18lN 


Twelfth Symposium 


NAVAL 
HYDRODYNAMICS 


Boundary Layer Stability and Transition 


Ship Boundary Layers and 
Propeller Hull Interaction 


Cavitation 
Geophysical Fluid Dynamics 


sponsored by the 

OFFICE OF NAVAL RESEARCH 
the 

DAVID W. TAYLOR NAVAL SHIP RESEARCH AND DEVELOPMENT CENTER 
and the 

NAVAL STUDIES BOARD 

of the 

NATIONAL RESEARCH COUNCIL 


MARINE 
BIOLOGICAL 
LABORATORY 


LIBRARY 


WOODS HGLE, MASS. 
Vie Tes (CEE 


NATIONAL ACADEMY OF SCIENCES 


Washington, DC 1979 


Partial support for the publication of these 
Proceedings was provided by the Office of Naval 
Research of the Department of the Navy. The 
content does not necessarily reflect the position 
or the policy of the Navy, the U.S. Government, or 
the National Academy of Sciences and no endorse- 
ment should be inferred. 


ISBN 0-309-02896-5 


Library of Congress Catalog Card No. 79-53803 


Available from: 


Office of Publications 
National Academy of Sciences 
2101 Constitution Avenue, N.W. 
Washington, D.C. 20418 


Printed in the United States of America 


PROGRAM COMMITTEE 


George F. Carrier, Chairman, Harvard University 


William E. Cummins, Vice Chairman, David W. Taylor Naval Shtp 
Research and Development Center 


Ralph D. Cooper, Office of Naval Research 
Stanley W. Doroff, Office of Naval Research 
Lee. M. Hunt, Nattonal Research Counctl 


Wen Chin Lin, David W. Taylor Naval Shtp Research and 
Development Center 


Justin H. McCarthy, Jr., David W. Taylor Naval Ship Research 
and Development Center 


Vincent J. Monacella, David W. Taylor Naval Ship Research 
and Development Center 


SYMPOSIUM AIDES 


Marguerite A. Bass 
Yetta S. Hassin 
Office of Naval Research 


Lavern Powell 
David W. Taylor Naval Ship Research and Development Center 


Bernice P. Hunt 
Joyce L. Wright 
Nattonal Academy of Sctences 


Grace Masuda 
Institute of Medicine 


Hope M. Bell 

Doris E. Bouadjemi 
Beatrice Bretzfield 

Mary G. Gordon 

Virginia A. Harrison 
Debra A. Tidwell 

Eva F. Tully 

Nattonal Research Counetl 


statal, 


= 
ted 


ety 
7h Leann 
“hin! iy H Terni 


Preface 


The Twelfth Symposium on Naval Hydrodynamics 
was held at Washington, D.C., during the period 
5-9 June 1978 under the joint sponsorship of the 
Office of Naval Research, the David W. Taylor Naval 
Ship Research and Development Center, and the 
National Academy of Sciences. 

The technical program of the Symposium con- 
sisted of eight sessions equally apportioned among 
the following four subjects of great current inter- 
est in the general field of naval hydrodynamics: 

(1) boundary layer stability and transition, (2) 
ship boundary layers and propeller hull interaction, 
(3) cavitation, and (4) geophysical fluid dynamics. 
Tours of the hydrodynamic research facilities of 
the David W. Taylor Naval Ship Research and Devel- 
opment Center and of Hydronautics, Inc., were also 
included in the technical program. 

It is interesting to recal that the National 
Academy of Sciences was a cosponsor of the First 
and Second Symposia in this series which were held 
respectively in 1956 and 1958. It is a great plea- 
sure to acknowledge once again the invaluable 
assistance of the Academy in launching these Sym- 
posia and in establishing the high standards of 
quality and style for them by which we are guided, 
even to this day. 

Similarly, the David W. Taylor Naval Ship 
Research and Development Center has played an 
important role in the series of Symposia on Naval 
Hydrodynamics from their very inception. Scien- 
tists and engineers from the Center have presented 
outstanding scientific papers at each of the Sym- 
posia and have, in addition, participated in an 
informal manner in the planning of many of the 
earlier ones. 

For these reasons the Office of Naval Research 
is especially pleased and honored at the opportu- 
nity presented by the cosponsorship of this Twelfth 


Symposium to renew and continue the fruitful col- 
laboration with its old scientific allies. We are 
deeply grateful for their generous assistance in 
the past and present, and look forward with confi- 
dence to their continued support in the future. 

Of the seemingly endless list of people who 
contributed in large and small ways to the planning 
and organizing of the Twelfth Symposium the follow- 
ing deserve special recognition: Professor George 
F. Carrier of Harvard University and the Naval 
Studies Board of the National Research Council, who 
served as chairman of the Program and Organizing 
Committee; Dr. William E. Cummins of the David W. 
Taylor Naval Ship Research and Development Center, 
who served as vice-chairman of the Committee, and 
his colleagues from the .:nter, Dr. Wen Chin Lin, 
Mr. Justin H. McCarthy, Jr. and Mr Vincent J. 
Monacella, who served on the Committee; Mr. Lee M. 
Hunt of the Naval Studies Board, who served on the 
Committee and who, with the able assistance of 
Miss Virginia A. Harrison, personally carried out 
the multitude of detailed arrangements required for 
the success of the Symposium; and Dr. Nelson T. 
Grisamore of the National Academy of Sciences, who 
edited these Proceedings. 

A special note of appreciation is extended to 
Mr. Phillip Eisenberg, President of Hydronautics, 
Inc., for his delightful after-dinner talk at the 
Symposium Banquet and for the tour of Hydronautics, 
Inc., which he graciously arranged for the partic- 
ipants of the Symposium. 

To all of these, and many more, the Office of 
Naval Research is forever indebted. 


Ralph D. Cooper 
Office of Naval Research 


: il ou , rad y 
i i i wana |. Vash i iy 
1 J 1 : : 1 i 
I _ iV OF y 
i it i igi il ‘ 
Wi we y 
1 y \ =! J 
I : i ra (ro see t|| 
i f Bein > Blu, 
[ fl i Di Tite 2 it 
i ho ee i Uy, ih Aaa seaieet 
i i i il ie i L 
i _ yee 
i ie : 
Uy ee fn vy 
u H THT Lt ik if 1 


hale 


ae) Le fant 
ipl! ire ie 


=i 


- yin, a gl tia ‘ 
11 A Te A ea es a 1 yaa Hatodbtiake 


i Ssaihie am 


iy 
5 be Gia 


: ok Lanes th 


wih Sa alae 


Contents 


Preface 
Ralph D. Cooper 
Introductory Address 
Courtland D. Perkins 
Introductory Address 
Robert K. Geiger 
Introductory Address 


Myron V. Ricketts 


SESSION I: BOUNDARY LAYER STABILITY AND TRANSITION 


Stability and Transition Investigations Using the 
Navier-Stokes Equations 


Hermann F. Fasel 

The Physical Processes Causing Breakdown to Turbulence 
M. Gaster 

The Instability of Oscillatory Boundary Layers 
Christian von Kerczek 

Heated Boundary Layers 
Eli Reshotko 


Discussion 


SESSION II: BOUNDARY LAYER STABILITY AND TRANSITION 


Nonparallel Stability of Two-Dimensional Heated Boundary 


Layer Flows 
N. M. El-Hady and A. H. Nayfeh 
Three-Dimensional Effects in Boundary Layer Stability 


Leslie M. Mack 


vii 


22 


25 


33 


48 


53 


63 


Experiments on Heat-Stabilized Laminar Boundary Layers 
in a Tube a, 


Steven J. Barker 


Some Effects of Several Freestream Factors on Cavitation 
Inception on Axisymmetric Bodies 86 


Edward M. Gates and Allan J. Acosta 


Discussion 109 


SESSION III: SHIP BOUNDARY LAYERS AND PROPELLER HULL INTERACTION 


Calculation of Thick Boundary Layer and Near Wake of Bodies 
of Revolution by a Differential Method ata} 


Virenda C. Patel and Yu-Tai Lee 
Stern Boundary-Layer Flow on Axisymmetric Bodies 127 
Thomas T. Huang, Nicholas Santelli, and Garnell Belt 


Theoretical Computation and Model and Full-Scale Correlation 
of the Flow at the Stern of a Submerged Body ‘158 


A. W. Moore and C. B. Wills 


Experimental and Theoretical Investigation of Ship Boundary 
Layer and Wake 169 


Shuji Hatano, Kazuhiro Mori, and Takio Hotta 


A General Method for Calculating Three-Dimensional Laminar 
and Turbulent Boundary Layers on Ship Hulls 188 


Tuncer Cebeci, K. C. Chang, and Kalle Kaups 


Study on the Structure of Ship Vortices Generated by 
Full Sterns 209 


Hiraku Tanaka and Takayasu Ueda 


SESSION IV: SHIP BOUNDARY LAYERS AND PROPELLER HULL INTERACTION 

Wake Scale Effects on a Twin-Screw Displacement Ship 225 
Arthur M. Reed and William G. Day, Jr. 

Influence of Propeller Action on Flow Field Around a Hull 248 
Shunichi Ishida 

Prediction Of the Velocity Field in Way of the ship Propeller 265 
I. A. Titov, A. F. Poostoshniy, and O. P. Orlov : 

Recent Theoretical and Experimental Developments in the 

Prediction of Propeller Induced Vibration Forces on 


Nearby Boundaries 278 


Bruce D. Cox, William S. Vorus, John P. Breslin, 
and Edwin P. Rood 


viii 


A Determination of the Free Air Content and Velocity in Front 
of the "Sydney-Express" Propeller in Connection with Pressure 
Fluctuation Measurements 


Andreas P. Keller and Ernst A. Weitendorf 


Discussion 


SESSION V: CAVITATION 

Pressure Fields and Cavitation in Turbulent Shear Flows 
Roger E. A. Arndt and William K. George 

Secondary Flow Generated Vortex Cavitation 
Michael L. Billet 

On the Linearized Theory of Hub Cavity with Swirl 
G. H. Schmidt and J. A. Sparenberg 

Unsteady Cavitation on an Oscillating Hydrofoil 
Young T. Shen and Frank B. Peterson 

Cavitation on Hydrofoils in Turbulent Shear Flow 
Hitoshi Murai, Akio Ihara, and Yasuyuki Tsurumi 

Scale Effects on Propeller Cavitation Inception 
G. Kuiper 


Discussion 


SESSION VI: CAVITATION 
A Holographic Study of the Influence of Boundary Layer and 
Surface Characteristics on Incipient and Developed 
Cavitation on Axisymmetric Bodies 
J. H. J. van der Meulen 
Mechanism and Scaling of Cavitation Erosion 
Hiroharu Kato, Toshio Maeda, and Atsushi Magaino 
Experimental Investigations of Cavitation Noise 
Goran Bark and Willem B. van Berlekom 
Cavitation Noise Modelling at Ship Hydrodynamic Laboratories 
G. A. Matveyev and A. S. Gorshkoff 
Fluid Jets and Fluid Sheets: A Direct Formulation 
P. M. Naghdi 


Discussion 


ix 


300 


319 


327 


340 


348 


362 


385 


400 


426 


433 


452 


470 


494 


500 


516 


SESSION VII: GEOPHYSICAL FLUID DYNAMICS 

The Boussinesq Regime for waves in a Gradually Varying Channel 
John W. Miles 

Study on Wind Waves as a Strongly Nonlinear Phenomenon 
Yoshiaki Toba 


An Interaction Mechanism betwee Large and Small Scales for 
Wind-Generated Water Waves 


Marten Landahl, Sheila Widnall, and Lennart Hultgren 


Preliminary Results of Some Stereophotographic Sorties Flown 
Over the Sea Surface 


L. H. Holtuijsen 


Gerstner Edge Waves in a Stratified Fluid Rotating about a 
Vertical Axis 


Eric Mollo-Christensen 
The Origin of the Oceanic Microstructure 


Grigoriy I. Barenblatt and Andrei S. Monin 


SESSION VIII: GEOPHYSICAL FLUID DYNAMICS 

The Rise of a Strong Inversion Caused by Heating at the Ground 
Robert R. Long and Lakshmi H. Kantha 

Laboratory Models of Double-Diffusive Processes in the Ocean 
J. Stewart Turner 

Buoyant Plumes in a Transverse Wind 
Chia-Shun Yih; Appendix by J. P. Benqué 

Internal Waves 
O. M. Phillips 

Breaking Internal Waves in Shear Flow 


S. A. Thorpe 


LIST OF PARTICIPANTS 


523 


529 


541 


555 


570 


574 


585 


596 


607 


618 


623 


629 


Introductory Address 


Dr. Courtland D. Perkins 
President, National Academy of Engineering 


On behalf of the National Academy of Engine- 
ering and the National Academy of Sciences it is 
my distinct pleasure and privilege to welcome you 
to our Nation's Capitol, to the home of both Acad- 
emies, and to the Twelfth Symposium on Naval Hydro- 
dynamics. 

We have welcomed the opportunity to join with 
the Office of Naval Research and the David W. Taylor 
Naval Ship Research and Development Center in organ- 
izing and hosting the Twelfth Symposium in this 
distinguished series of meetings. 

We have, as a matter of fact, a special inter- 
est in the continuing success of the series since 
we cosponsored the First and Second Symposia with 
the Office of Naval Research in 1956 and 1958. 
Therefore, it is as gratifying for us as it must 
be for the Office of Naval Research to find that 
the international community of fluid dynamics and 
related specialties continues to find these meetings 
a unique forum for the exchange of research results 
and the discussion of problem areas of concern to 
both military and commercial activities. 

The interest and the involvement of the Acad- 
emies in naval science and engineering, of course, 
has a much longer history. After a careful reading 
of the early history of the National Academy of 
Sciences, one is persuaded that the Academy would 
not have come into being in 1863 had it not been 
for the carefully laid plan and persuasive argu- 
ments of the Navy's Chief of Navigation, Commodore 
Charles Henry Davis. One is further impressed by 
the fact that perhaps a quarter of those who signed 
the Academy's Charter were affiliated with the Navy 
in one way or another. And it is significant that 
the first five studies conducted by the fledgling 


Academy were requested by the Navy. In case some 


of you may be interested, these were: 


On Protecting the Bottom of Iron Vessels 

On Magnette Deviation tn Iron Ships 

On Wind and Current Charts 

Sailing Directtons 

On the Exploston On the Untted States Steamer 
CHENANGO 


I don't want to leave you with the impression that 
the Academy worked only on naval problems during 
the 1863-65 period. We did another study entitled 
"On the Question of Tests for the Purity of Whiskey" 
--an investigation undoubtedly stimulated by 
President Lincoln's remark that he wished he could 
supply all his generals with whatever it was that 
General Ulysses S. Grant was drinking. 

I have taken this short detour through some 
early Academy history, not so much to demonstrate 
our own long and continuous interest in naval sci- 
ence and engineering but to recognize the important 
role played by the Navy in supporting science and 
engineering throughout its 200-year history. Over 
the past 32 years the Office of Naval Research has 
continued that tradition by serving as a model for 
enlightened government support of basic research. 

On a more personal note may I conclude by say- 
ing that as a former professor of aeronautical 
engineering at Princeton University your technical 
program is of special interest to me. Therefore, 

I wish you an interesting and productive meeting. 
We are pleased that you have chosen to meet at our 
institution, and the staff we have assembled to 
support you is available to assure that your stay 
is a pleasant one. 


tee pear Lib 
ate 


fe Nai 


ac 


vinenigll 


Dd ne te 


M ieite 4 


(pSaee 


a 
i! Ne ne 
m aa 


(a 


(ye 
ni 


Oh ' iw 


in 
ay ih alk ‘Ani 


TT ae viaje hy Nae Fen 


Vy, OY 
AE y) 
MU i 
f ] 
ee Nowe i 
cf Meas 
feet 
Hint = 
Hun 
yay ; 
i y ’ 
J nll A 
i 
f | 
‘ Th ’ | 
inn i ™ a A i el 
( i Lay, i an i , 


Introductory Address 


Rear Admiral Robert K. Geiger, USN 
Chief, Office of Naval Research 


On behalf of the Office of Naval Research I 
would like to extend a sincere welcome to all the 
participants of the Twelfth Symposium on Naval 
Hydrodynamics. 

I wish to express my thanks to the National 
Academy of sciences for its assistance and role as 
a host and cosponsor of the Symposium through its 
National Research Council. 

Thanks are also due to the third member of the 
triumvirate of cosponsors of this, the Twelfth Sym- 
posium on Naval Hydrodynamics, namely the David W. 
Taylor Naval Ship Research and Development Center, 
known more familiarly to most of us old-timers as 
the David Taylor Model Basin and often referred to 
affectionately as DTMB. This facility has been a 
major contributor to the scientific program of each 
of the Symposia in this series, as a glance at the 
proceedings of any of the Symposia will confirm. 

I am happy to say that the present meeting is no 
exception and that it is again well represented on 
the technical program. However, this is the first 
time that it has participated as a cosponsor and I 
am especially pleased to acknowledge the invaluable 
assistance that our old colleague and ally in the 
field of naval hydrodynamics research has rendered 
in the organization and management of the present 
Symposium. 

The first two Symposia of this series were 
held in 1956 and 1958 and were also sponsored by 
the Office of Naval Research and the National 
Academy of Sciences. Many of the guiding princi- 
ples that govern the organization of the Symposia 
in this series were established in these first 
meetings. For example, the selection of a limited 
number of central themes of timely naval hydro- 
dynamic interest upon which to focus the technical 
program of the meeting was introduced in the 
Second Symposium. 

From the very beginning, the international 
aspects of the Symposia were emphasized through the 


invitation of speakers from all over the world 
wherever outstanding research in naval hydrody- 
Namics was going on. Starting with the Third Sym- 
posium, the international aspects were strengthened 
by locating the meetings outside the United States 
and cosponsoring them with relevant organizations 
in host countries. 

The list of such meetings includes Symposia 
held in the Netherlands, Norway, Italy, France, and 
England, and we hope to continue this pattern into 
the future as long as the series of Symposia con- 
tinue to provide a useful forum for the exchange 
of valuable information on results of advanced re- 
search in the field of naval hydrodynamics. 

I am gratified to see so many representatives 
of several countries in addition to the United 
States, and the number of technical papers pre- 
sented by internationally known authorities in 
fluid dynamics and related fields. 

For the Navy, progress in hydrodynamics re- 
search has become increasingly urgent. The Navy 
must find ways to discover and correct the problems 
that a new design may run into before reaching the 
point of full-scale sea trials. 

Since the sea is the Navy's business and we 
have been involved in it a long time, we are ex- 
pected to know it well. Only investigators like 
yourselves are aware of how limited is our knowl- 
edge of the forces that impact on a buoyant body 
propelled through the water. As much as our under- 
standing has increased, we know we have much more 
to learn. This information can only be obtained 
through the arduous bit-by-bit process of basic re- 
search, such as you gentlemen pursue. 

Today our nation is faced with the dilemma 
that we must plan types of ships that are radically 
different in design from anything in the past. At 
the same time, these ships must be inexpensive to 
operate and maintain in addition to satisfying our 
traditional standards. 


The results of the research that will be re- 
ported at this Symposium should help us move toward 
that formidable goal. It is clear that all of you 
here today are dedicated scientists, so I do not 
need to urge you to keep pressing forward in your 
search for solutions to the frustrating problems 


in hydrodynamics. I would like to stress, however, 
that you maintain strong lines of communication so 
that as many people as possible can benefit when 
you inevitably succeed in your endeavors. 

Best wishes for a successful symposium. 


Introductory Address 


Captain Myron V. Ricketts, USN 
Commander, David W. Taylor 
Naval Ship Research and Development Center 


We at the David W. Taylor Naval Ship Research 
and Development Center are both pleased and proud 
to join with the Office of Naval Research and the 
National Academy of Sciences in sponsoring the 
Twelfth Symposiumon Naval Hydrodynamics. While 
not a sponsor of the four earlier symposia held in 
Washington, the Center was directly and indirectly 
involved with all of the previous meetings. Of the 
forty-one papers to be presented at the present 
Symposium, five are authored by Center researchers, 
roughly the same number of papers given by Center 
authors at earlier symposia. In addition, much of 
the other U.S. research to be presented in papers 
to this Symposium was supported by the U.S. Navy's 
General Hydrodynamics Research Program which the 
Center has administered for nearly thirty years. 

It is worthy to note that this year's confer- 
ence is directed mainly at the underlying physics 
of hydrodynamic processes. The papers are of quite 
a fundamental nature, perhaps more so than was true 
of many of the earlier symposia. The Symposium 
topics are of immense importance to both the mer- 
chant ship and naval communities: Boundary Layer 
Stabtltty and Transttton because of their relation- 
ship to vehicle drag, cavitation inception, and 
flow noise; Shtp Boundary Layers and Propeller/ 
Hull Interactton because a need to accurately pre- 
dict vehicle drag, propulsive efficiency, and 
vibration; Cavitation, a very major cause of ero- 


sion, vibration, and noise; and finally, Geo- 
phystcal Flutd Dynamtes which describes the envi- 
ronment in which ocean systems must operate. Each 
topic area is a subject of current and lively in- 
terest and has witnessed remarkable advances over 
the past few years. 

The very high quality of the research papers 
to be presented this week is typical of previous 
Naval Hydrodynamics Symposia and has earned for the 
series the reputation of being the preeminent inter- 
national conferences on ship hydrodynamics. Each 
symposium has constituted an exceedingly valuable 
open forum which promotes national and international 
ties and dialogues between researchers in the field 
of hydrodynamics. 

I would like to close by saying that my 
Center's namesake, Admiral David W. Taylor, the U.S. 
pioneer hydrodynamicist and foremost naval archi- 
tect, introducer to the U.S. of towing tanks, water 
tunnels, transformer of empiricism to scientific 
methods, would be very pleased to be associated 
with the Twelfth Symposium on Naval Hydrodynamics. 
On Wednesday we look forward to welcoming you on a 
tour of the hydrodynamic facilities at the Center. 
You will see work in progress at our rotating arm 
facility, seakeeping basin, towing tanks and turn- 
ing basin, and at our largest cavitation tunnel. 
Best wishes for a very successful conference. 


Pa) : ou i : Toy teat 
0 ly { iN i r Ud) had, Lar en 
n i 1 i : i : 1 en ee im 
on ‘ } enh i y h a i 1 en 
: j han Yo iT : , - a i i) Ne 
7 ; com, i i 7 TH Dai - un) a 
‘s ’ my we Y inet } iy as = 1 - \h ita eat iver 
u/ in fet ; ice i Xie ty thi ees Uy fis! | Delf 
; i - 7 i i} t ‘a, ae i i Ahh 9 ell \ 1) ee i} thi a, i Jaa fal yey] 
' i 10 Oe i at re i eee i 7 " iia, Whey ’ iH 


igs verter 
Lec! 


i Henly! 
RE ae 


ae eA Fi 


‘co 


: ieee | 
i Lae eee 


i 


yA Tee ath WER 
{ay ie! Be 


: Fy 
qe ae Ly i 


r ut fi fu i 
f iy it ' 
1 ete Th 
ee eth 
a ja ete uve 
i } : p 
‘ | Wee 
_ eer 
f I | 
f 
i = i 
| / l t " 1 ne i 
i 
j et 
! : . 3 ag 
i ae I ‘yh ee 
; ne 
i U 1 
1 1 : i i , 
uy 
ie! ty 
1 \ if 1 / 
, ere wu 
i f H) I = 
1 a 1 


Session I 


BOUNDARY LAYER STABILITY 
AND 
TRANSITION 


PHILLIP. S. KLEBANOFF 
Session Chairman 

National Bureau of Standards 
Washington, D.C. 


eed ih oN 
Ae Urea 
na 1 at ye 
= x x er 
‘sl ie aie el 

o 


ii : 
fe hi mn By 


qe idl 


i 
Ly 
faly I in 


Saimiri 


if eed AM 
Pathe en 


oie 
Payee 


ii 


10 ae Se cc 


5 
(ee 
PAS 


moe 


i ey My 
TIE SALMAN a 
Hie ata i 


Stability and Transition Investigations 
Using the Navier-Stokes Equations 


Hermann F. Fasel 
Universitat Stuttgart 


Stuttgart, Federal Republic of Germany 


SUMMARY 


With this paper an attempt is made to review the 
stability and transition simulations, performed at 
the University of Stuttgart, which are based on 
finite-difference solutions of the Navier-Stokes 
equations. Research in this area has demonstrated 
that implicit finite-difference methods for the 
solution of the complete Navier-Stokes equations 

for unsteady, two-dimensional, incompressible flows 
can be successfully applied to investigations of 
hydrodynamic stability and to certain aspects of 
transition. This approach of numerically solving 
the partial differential equations describing the 
underlying flow mechanisms promises to be a valuable 
aid in transition research. In particular, this 
concept may prove to be especially rewarding for 
investigations of aspects of stability and transition 
which as yet are not feasible with other theoretical 
models. 

There are two main reasons for the attractiveness 
of this approach: Firstly, no assumptions whatsoever 
have to be made concerning the basic flow field under 
investigation. Thus, for example, all possible 
effects resulting from the growth of a boundary layer 
in downstream direction can be included in such 
investigations. Even strongly converging or diverg- 
ing flows, or flows with separation and/or reattach- 
ment can be studied. Secondly, no restrictions 
have to be made concerning amplitude and form of 
the disturbances which are injected into the flow. 
Therefore, using larger disturbance amplitudes 
certain nonlinear effects of the amplification 
process can be readily investigated. 

The major aspects of this approach will be dis- 
cussed in this paper. Emphasis will be placed not 
only on conveying the advantages of such investi- 
gations but also on elaborating the difficulties 
and shortcomings of such numerical simulations. 
Finally, a conjecture concerning the course of 
future developments will be attempted. 


1. INTRODUCTION 


The phenomena occurring in transition from laminar 
to turbulent flow have been the subject of inten- 
sive research ever since the discovery that these 
two entirely different states of flow exist. 

From all the research efforts basically only one 
universally-accepted theoretical concept evolved, 
namely, linear stability theory, verified experi- 
mentally by the famous experiments of Schubauer 
and Skramstad (1943). 

However, experimental evidence has also shown 
that linear stability theory is only applicable 
for one 'special' transition process, namely, 
transition initiated by the presence of very small 
disturbances in the flow. In this case a substan- 
tial portion of the entire transition process is 
indeed well described by this theory, i.e. the 
amplification of two-dimensional disturbance waves 
(the so-called Tollmien-Schlichting waves) can be 
predicted adequately. But even for this special 
transition process, triggered by small disturbances, 
linear stability theory is inadequate in the 
description and investigation of the mechanisms 
that follow the growth of Tollmien-Schlichting 
waves, and which finally cause the breakdown to 
fully turbulent flow. Nevertheless, due to the 
relative success of the linear stability theory 
and its impressive experimental verification, the 
vast majority of theoretical transition investi- 
gations were, and still are, based on stability 
theory concepts, thus constantly improving and 
perfecting this theory. 

The inherent shortcomings of this concept 
nontheless (such as being applicable only when 
transition is initiated by small disturbances, or 
that certain assumptions concering the basic and 
disturbance flow have to be made to keep the 
resulting equations tractible) led to a search for 
other means to investigate transition. One of the 
more promising concepts that has emerged in recent 


10 


years is based on direct numerical solutions of the 
complete partial differential equations that 
describe the flow phenomena arising in the transi- 
tion process. This approach became feasible with 
the rapid progress in the development of large, 
high-speed digital computers. 

The main difficulties here arise from the fact 
that these flow phenomena can be adequately 
represented only when the complete Navier-Stokes 
equations (or certain modifications thereof) are 
used. Thus, this approach requires the solution 
of the Navier-Stokes equations for strongly time- 
varying flow fields, due to the highly unsteady 
nature of the transition processes. Additionally, 
complications increase because the numerical 
solutions have to yield reliable results for 
relatively high Reynolds numbers (higher than the 
critical Reynolds number) to allow onset of 
transition. For a numerical solution procedure it 
is therefore necessary to allow for adequate 
resolution of the large temporal and spatial 
gradients resulting from the occurrence of thin 
time-varying fluid layers with large gradients 
close to solid walls. 

The development of finite-difference methods, 
which are applicable for such complex, unsteady 
flow phenomena as thase occuring in laminar 
turbulent transition, is associated with numerous 
difficulties which will be elaborated upon in this 
Paper. Because of these difficulties relatively 
few previous attempts based on such an approach be- 
came known. Reasonably successful earlier investiga- 
tions of this kind (based also on finite-difference 
solutions) are reported, for example, for incom- 
pressible flows in a boundary layer [De Santo 
and Keller, (1962)], for Poiseuille and plane 
Poiseuille flow [Dixon and Hellums (1967), Crowder 
and Dalton (1971)] and for a compressible boundary- 
layer flow [Nagel (1967)]. These earlier attempts 
clearly demonstrated the usefulness and potential 
of such investigations. However, due either to 
insufficent resolution of the resulting gradients 
and/or assumptions made concerning the basic or 
disturbance flows, or to shortcomings of the differ- 
ence methods used, the results of these calculations 
were more of a qualitative nature. Therefore, 
relatively little information could be gained 
concerning the various phenomena arising in the 
laminar-turbulent transition process. 

Some years ago, a research effort was initiated 
at the University of Stuttgart aiming at the devel- 
opment of numerical methods for the solution of the 
Navier-Stokes equations which would be applicable 
for detailed investigations of various aspects of 
Stability and of phenomena occurring in transition. 
To date, an effective implicit finite-difference 
method has evolved for the calculation of unsteady, 
two-dimensional incompressible flows. The ap- 
plicability of the numerical method to investigate 
stability and two-dimensional transition phenomena 
has been demonstrated by realistic simulations of 
Tollmien-Schlichting waves. Detailed results of 
these calculations are discussed elsewhere [Fasel 
(1976)]. With calculations involving large ampli- 
tude disturbances [Fasel et al. (1977)] it was addi- 
tionally shown that numerical simulations using 
the implicit difference method yield results which 
enable insight into certain nonlinear mechanisms 
of the transition process. 

In this paper the major aspects of the numerical 
approach using finite-difference methods will be 


reviewed and the present state of the developments 
discussed. Emphasis will be placed on the advan- 


tages of the numerical approach in general and on 
directional options chosen for the present method. 
Special attention will also be focused on the 
difficulties and limitations of such simulations. 


2. SELECTION OF THE INTEGRATION DOMAIN 


For a numerical solution of the Navier-Stokes equa- 
tions using finite-difference techniques a finite 
domain in which the equations are being solved has 
to be specified. The selection of the integration 
domain determines the nature of a physical flow 
problem to be simulated, because the boundary con- 
ditions required along the boundaries of this domain 
determine to a large degree the solution within the 
domain. For reasons of simplicity, in the present 
studies only rectangular domains of the x,y plane 
were considered as depicted schematically in Figures 
1 and 2 with the direction of the basic, undisturbed 
flow being in the x-direction. Rectangular domains 
allow relatively easy application of difference 
methods by using simple rectangular meshes. For 
example the rectangular domain may be a section of 

a boundary-layer flow on a semi-infinite flat plate 
(Figure 1) or a section of a flow between two paral- 
lel plates (Figure 2). 

In selecting the integration domain one has to 
consider that boundary conditions must be found for 
the 'artificial' boundaries B-C in Figures 1 and 2 
and additionally for C-D in Figure 1. These con- 
ditions should allow physically meaningful solutions 
in the finite domain, i.e. solutions that would be 
obtained if the domain were not made finite by 
means of these artificial boundaries. Due to the 
spatially elliptic (in x,y) character of the Navier- 
Stokes equations application of finite-difference 
methods requires boundary conditions on all bound- 
aries of the x,y domain. Of course, in a mathemat- 
ical sense the equations are parabolic because of 
the time derivative (See section 3). Selection of 
boundary conditions for boundaries representing 
solid walls (such as A-B in Figures 1 and 2 and C-D 
in Figure 2) generally creates no additional diffi- 
culty although consistent implementation in the 
numerical scheme is frequently difficult to achieve. 
Also, free stream boundaries such as C-D in Figure 
1 for the boundary-layer flow can be handled in 
satisfactory fashion (see Section 4). 

However, the upstream (A-D) and to a larger ex- 
tent the downstream (B-C) boundary require special 
considerations because the specific treatment of 
these boundaries determines the approach to be 
taken in a prospective stability and transition 
simulation. In selecting the boundary conditions 


FIGURE 1. 


Integration domain for boundary 
layer on flat plate. 


FIGURE 2. 


Integration domain for plane 
Poiseuille flow. 


there are basically two different approaches which 
lead to entirely different conceptions of the trans- 
ition simulation: 


1) Use of periodicity conditions at the upstream 
(A-D) and downstream (B-C) boundary, i.e. 
corresponding disturbance quantities are 
equal at the two boundaries for all times. 


Here it is assumed that flow phenomena are spatially 
periodic in downstream direction where the integra- 
tion domain X contains integer multiples of the 
spatial wavelength. When the spatial development 

is forced to be periodic, the flow responds with a 
temporal development. Thus, with this arrangement 
the temporal reaction of the flow to an initial 
disturbance (at t=0) of the flow field can be 
studied. This case corresponds in linear stability 
theory to an eigenvalue problem with wave number 

a real and frequency 8 complex (6=6,+i6,), i.e. 
amplification in time. Figure 3, for example, shows 
a typical result of a finite-difference calculation 
based on such an approach for a plane Poiseuille 

low [Bestek and Fasel (1977) ]. Plotted here is a 
time signal for a case which is unstable according 
to linear stability theory. The flow is only dis- 
turbed once at t=0. After a certain time span, 
where considerable reorganization of the disturbance 
flow takes place, the disturbances assume a periodic 
character with a slight amplification in time- 
direction. 

The Navier-Stokes calculation for this approach 
may be conceived as a means of solving the eigen- 
value problem as in linear stability theory, with 
a and Reynolds number given and obtaining the fre- 
quency 8,, amplification rate 8;, and the amplitude 
distribution of the distrubance flow. Of course 
these answers could be obtained with considerably 
less effort from linear stability analysis. The 
advantage of this present approach is, however, 
that it can be easily extended to investigations 


- FIGURE 3. Temporal development of u'-disturbance at 
y/Ay = 3 for initially disturbed flow (small ampli- 
tude); spatially periodic case (plane Poiseuille flow). 


11 


of certain nonlinear effects by merely increasing 
the amplitude level of the initial disturbances 
[see, for example, George and Hellums (1972)]. An 
equivalent study of nonlinear effects formulated 

as an eigenvalue problem in a stability theory 
analysis would, on the other hand, become consider- 
ably more involved. 

A major drawback of this first approach is, how- 
ever, that it is pratically only applicable for 
basic flows that do not vary in downstream direction 
(parallel flows), because only then is the period- 
icity assumption for the disturbance flow a real- 
istic one. Thus, strictly speaking, boundary-layer 
flows could not be treated in this manner since they 
are basically (although very mildly) non-parallel. 
It has been shown that non-parallel effects can 
have a strong influence on the stability character- 
istics of this flow [Gaster (1974), Saric et al. 
(1977) J. 

A second, perhaps even more serious disadvantage 
of this model is that the disturbance development 
in downstream direction cannot be investigated. As 
observed in numerous laboratory experiements the 
phenomena of transition are not periodic in space 
but rather are inherently space dependent. The 
disturbance flow may vary rapidly in downstream 
direction. This space dependency of the transition 
process does not only occur for flows where the 
basic flow is already dependent on the downstream 
location. It also occurs when the basic flow does 
not vary in downstream direction, as was impress- 
ively demonstrated experimentally by Nishioka et 
al. (1975) for the parabolic profiles of plane 
Poiseuille flow between parallel plates. Thus, 
this model is not suitable for realistic studies 
of transition phenomena. 

However, finite difference simulations based on 
this approach become considerably less involved 
and are less costly in practical execution than for 
the second approach discussed subsequently. The 
former approach is therefore applicable for funda- 
mental investigations of various unresolved ques- 
tions in hydrodynamic stability (such as certain 
nonlinear effects) or for preliminary studies of 
flow simulations based on the approach discussed 
below. 


2) At the upstream boundary, time-dependent 
disturbances are introduced. Use of bound- 
ary conditions at the downstream boundary 
which allow downstream propagation of the 
spatial disturbance waves. 


This second approach differs entirely in concept 
from the first one. Here, the reaction of the 

flow field to the disturbances introduced at the 
upstream boundary is of interest, particularly the 
spatial developments of the ensuing disturbance 
waves. In contrast to the previous approach, this 
case corresponds in stability theory to an eigen- 
value problem with a complex (a=a,+ia;) and 6 real. 
A typical result for a boundary-layer flow of a 
calculation based on this concept is shown in Fig- 
ure 4. Plotted is the disturbance variable u' 
(velocity component in x-direction) versus the down- 
stream coordinate x. The downstream development 

of the disturbance (in this case amplification) may 
be clearly observed. Thus, this approach enables 
the calculation of the spatial reaction of the flow 
to upstream disturbances, and therefore realistic 
simulations of space-dependent transition phenomena 


FIGURE 4. Downstream development of u'-disturbance at 
y/Ay = 3 for boundary-layer flow disturbed periodically 
(small amplitude) at upstream boundary. 


as observed in laboratory experiments should be 
possible. 

For example, realistic numerical simulations of 
Tollmien-Schlichting waves (as observed in the 
Schubauer and Skramstad experiments) can be per- 
formed by using at the upstream boundary A-D per- 
iodic disturbances as produced by a vibrating ribbon 
in the physical experiments. If the location of 
A-D is considered to be somewhat downstream of the 
ribbon in the real experiments, eigenfunctions of 
linear stability theory may be conveniently used 
to disturb the flow in the numerical simulation. 
It was shown that the disturbance flow somewhat 
downstream of the ribbon is well described by 
linear stability theory when amplitudes are small. 

The disadvantage of the second approach is that 
the development of numerical methods to solve the 
resulting mathematical problem is considerably more 
difficult than in the first approach. Although in 
a strict mathematical sense both problems represent 
mixed initial-boundary-value problems, the main 
difference between the two concepts is that the 
first approach results in a predominantly initial 
value problem, where the temporal evolution of an 
initially disturbed flow field is calculated. 

The second concept leads to a predominantly 
boundary-value problem where the spatial reaction 
of the flow field (which is also time-dependent, 
of course) to disturbances introduced on the left 
boundary is to be calculated. In the latter case 
difficulties arise from the necessity of finding 
adequate downstream boundary conditions which 
allow unhindered passage of the disturbance waves 
propagating downstream, and properly implementing 
them into the numerical method. Since the aim of 
this research effort is directed toward realistic 
simulations of transition phenomena, emphasis in 
the development of finite-difference methods was 
placed on methods that were applicable to solving 
the mathematical problem resulting from the latter 
approach. The remainder of the discussions in this 
Paper are therefore also based on this concept. 


3. FORMULATIONS OF NAVIER-STOKES EQUATIONS FOR 
NUMERICAL METHODS 


The Navier-Stokes equations can be cast into various 
forms to be used as basis for a finite-difference 
method. Each formulation has its inherent advan- 
tages and disadvantages. The decision in favour of 
a particular formulation has to be governed by the 
physical flow problem to be investigated and by the 
difference scheme finally used. In most cases, and 
also particularly for the present investigations, 
such a decision is difficult to make beforehand. 
Extensive preliminary numerical experiments are 
necessary before a decision can be made in favour 
of a particular formulation. 


For two-dimensional, incompressible flows the 
stream-function-vorticity formulation is most 
widely used in numerical fluid dynamics. It is 
also a possible choice for the present investiga- 


tions. It consists of the vorticity-transport 
equation 
aw aw dU) 1 3 
et NY 1 
3 Sipe dy Re 2 (1) 


and a Poisson equation for the stream function 


Ay = w (2) 


~ 


where A is the Laplace operator, w is defined as 


oy = acl ' (3) 


TS gee a (4) 


With this definition of the stream function the 
continuity condition 


au ov 


ye ee (5) 
is satisfied for the continuum equations, however, 
not necessarily for the discretized equations. All 
variables in Eqs. (1) to (5) are dimensionless; 
they are related to their dimensional counterparts, 
denoted by bars, as follows 


— —_— = oe tu, 

alae = 22 = ee = = 

mR TREN AUS, Un eae 
(pe Oe ee ee | he SE 
wa ” 7 {7 UoL ’ 


where L is a characteristic length, Ug a reference 
velocity and Re a Reynolds number (v kinematic vis- 
cosity). Thus this formulation represents a system 
of two partial differential equations, each of 
second order, for the unknown variables w and yw 
because u and v in Eq. (1) can be eliminated using 
Eq. (4). 

A variation of this formulation is the so-called 
conservative form for which the vorticity-transport 
equation 


aw , d(uw) | Oa). Ine (6) 


ot ox oy Re 


is used instead of Eq. (1). With this formulation 
conservation of vorticity is guaranteed for the 
continuum equations. 

A second formulation of the governing equations 
also consists of a vorticity-transport equation (1) 
or (6). However, instead of the Poisson equation, 
(2), for ~, two Poisson equations for the velocity 
components u and v are used 


Au = : 
¥ (7) 


See Dat i'r 
Oe ax 


which can be derived from the definition of vortic- 
ity, (3), using the continuity equation, (5). This 
system of partial differential equations for the 
w,u,v formulation is of higher order than the w,\W) 
system. The higher order allows less restrictive 
boundary conditions which is advantageous in appli- 
cations to transition simulations as discussed in 
Section 4. 

A third form of the governing equations is the 
so-called primitive variable formulation with the 
two momentum equations 


du du du dp 
—— — Ss ob ——— ee —— 
cay ae oinaees: oy x” Re Au 
(8) 
av dv av dp al 
Bc RNs ay OM es 2 St 
t roa bod y By ee 


(where p = p/pus, with density 0) and a Poisson 
equation for the pressure 


ise Qe Og oy oe 


9 
ox oy ax One oY (9) 
which is derived from Eq. (8) using the continuity 
Eq. (5). 
There is also a conservative form of the primi- 
tive variable formulation (conserving momentum) 


jm . ee) ©) (om). G22 4b. 
yt oF qx + dy re ax Re Ae 
(10) 
ov, Sw , 0) 2 _ 2 , kL _y, 
are ox oy hy INS ; 


and a Poisson equation in a now different form 


32 (u*) a2(uv)  a2(v2) aD , Ll 
= - oo - 2 - = + A 
me ox2 2 ax dy ay2 at Re Dr(tt) 
with the so-called dilation term 
du Ch 
= — + — 2 
D x ay (12) 


The absence of the dilation terms in a Poisson equa- 
tion for the pressure may cause nonlinear numerical 
instability, which can be avoided when such terms 
are retained (Harlow and Welch, 1965). 


Conservative versus Nonconservative Formulation for 
Use in Transition Studies 


The evaluation of the relative merits of conserva- 
tive formulations over non-conservative ones is a 
widely investigated subject in numerical fluid 
dynamics [Roache (1976), Fasel (1978) ]. Neverthe- 
less, satisfactory answers have not yet been found 
except for compressible flows for which conserva- 
tive formulations are obviously advantageous. One 
argument in favour of conservative formulations is 
that better accuracy can be obtained. However, for 
- incompressible flow problems there are several ex- 
amples contradicting this claim. When evaluating 
possible advantages of a conservative formulation 


13 


one has to keep in mind that the respective quanti- 
ties (such as vorticity in the w, or w,u,v formu- 
lation or momentum for the u,v,p formulation) are 
initially only conserved for the continuum equations. 
The conservation property may be carried over to 
the discretized equations only if certain differ- 
ence approximations (in this case, central differ- 
ences) are used. For the implementation of the 
boundary conditions it is frequently very difficult 
or sometimes impossible to employ such difference 
approximations required to maintain the conserva- 
tion properties for the discretized equations. 

For the present investigations, comparison cal- 
culations during the early stage of the development 
of the numerical method have shown that, for the 
W,) or w,u,v systems, almost equivalent accuracy 
can be obtained with either formulation. Because 
the conservative formulation leads to a somewhat 
slower solution algorithm for the solution of the 
difference equations, preference was given there- 
fore to a non-conservative formulation. 


Vorticity Transport (Ww, or W,u,v) versus Primitive 
Variable (u,v,p) Formulation 


In reviewing literature on numerical simulations 
of viscous incompressible flows it is noticeable 
that formulations involving a vorticity-transport 
equation, rather than the primitive variable form- 
ulation, are preferred. The unpopularity of the 
u,v,p system is a result of numerable unsuccessful 
attempts in applying it to calculations of viscous 
incompressible flows. Although a few successful 
applications based on the u,v,p system are reported 
in more recent literature, there are still serious 
arguments against its use for stability and trans- 
ition simulations. Difficulties result from prob- 


lems associated with the use of a Poisson equation 
for the pressure. This equation is often a source 

of numerical instabilities, possibly due to difficul- 
ties of properly implementing the boundary conditions 
for pressure into the numerical scheme. Although 

the numerical instabilities could be brought under 
control, at least to a degree, (for example by intro- 
ducing the dilation terms in Eq. 11) so that solu- 
tions could be obtained for steady flow problems, 

the inherent inclination of this formulation to 
numerical instability still prohibits its use for 
transition simulations. Frequently numerical 
solutions based on this system are of a slightly 
oscillatory nature (although amplitudes are extremely 
small) and therefore interaction with oscillations 

of the physically meaningful disturbances as oc-— 
curring in transition studies cannot be avoided. 

For these reasons finite-difference methods de- 
vised for investigations of stability and transition 
are based on the equations in vorticity transport 
form, i.e. either on the w,l) system (Eqs. 1 and 2) 
or the w,u,v system (Eqs. 1 and 7). Nevertheless 
current efforts are also directed toward develop- 
ment of difference methods based on the equations 
in primitive-variable formulation. Emphasis is 
placed on extreme numerical stability in order to 
make this method also applicable for stability and 
transition studies. The continuing attraction of 
the equations in primitive-variable form results 
from the fact that, for the three-dimensional case, 
fewer fields of variables have to be stored than 
for a vorticity-transport formulation. For the 
three-dimensional case, storage requirements are an 


14 


order of magnitude even more critical than for the 
two-dimensional calculations. 


Use of Navier-Stokes Equations for the Disturbance 
Flow 


For stability and transition simulations, the depen- 
dent variables, which appear in the different form- 
ulations of the Navier-Stokes equations discussed 
previously, are those of the total flow, that is, 
including both the basic and the disturbance flow. 
There is an alternate approach, namely, to decompose 
the total flow into the basic flow and a disturbance 
flow such that 


u=U+tu!’ , v=Vt+v' , p=Ptp' , w=¥+p'. w=lt+w', (13) 


where the prime indicates the variables of the dis- 
turbance flow and the capital letters denote those 
of the basic flow. Substituting relationships (13) 
into various forms of the Navier-Stokes equations, 
it is possible to rewrite the equations with the 
disturbance variables as dependent variables. Sev- 
eral terms involving only the basic flow can be 
dropped, assuming the basic flow satisfies the 
Navier-Stokes equations. 

The aspect of directly solving the equations for 
the disturbance variables is an attractive one, 
since it is the disturbance conditions that are of 
interest when performing numerical stability and 
transition studies. For this reason this approach 
has probably been preferred in earlier attempts. 

It also allows for detailed investigations of the 
effects of the nonlinear (convective) terms. because, 
in a difference method based on this form, the 'lin- 
earization' can be conveniently switched on or off. 

A careful evaluation of this form of equations, 
however, reveals that it also has some major disad- 
vantages. The equations in disturbance form contain 
several additional terms (involving disturbance 
terms with terms of the basic flow) which are not 
present in a corresponding formulation for the total 
flow. Thus, in finite-difference solutions addi- 
tional numerical operations are required. A more 
serious disadvantage is that, because of the 
additional terms involving the basic flow, the 
basic flow quantities have to be kept in fast- 
access computer storage to be readily accessible 
for the numerical operations in order to avoid ex- 
cessive computation times. On the other hand, 
using the equations for the total flow the basic 
flow quantities are not directly involved in the 
solution algorithm. In this case they are only 
required for analysis and better respresentation 
of the results (for example to determine the dis- 
turbance quantities). For this purpose they can 
be stored in mass storage of lower speed accessi- 
bility. 

The availability of sufficient fast-access stor- 
age is, even with the latest computer generation, 
still a critical limitation for such numerical 
investigations of stability and transition. For 
large scale simulations involving large numbers of 
grid points, use of the disturbance formulation is 
prohibitive. For this reason, for the present re- 
search effort, use of the equations for the total 
flow variables was generally preferred instead of 
the disturbance formulation. However, the basic 
solution algorithm of the definite-difference method 


was developed such that it is applicable with only 
minor modifications for either formulation. 


4. BOUNDARY AND INITIAL CONDITIONS 


The selection of adequate boundary conditions and 
the practical implementation into a finite- 
difference scheme represents one of the major dif- 
ficulties in the development of a finite-difference 
model applicable for stability and transition stud- 
ies. Difficulties arise from the necessity that 
boundary conditions, selected and implemented along 
the artificial boundaries (see Section 2) for the 
finite integration domain, have to enable solutions 
that would be identical to solutions if the govern- 
ing equations were solved in the infinite domain. 
There is, of course, no way of checking this be- 
cause solutions for the infinite domain are not 
available. This indicates that, for selecting 
boundary conditions, it is necessary to rely on 
experience, intuition, and test calculations. 

For practical reasons the boundary conditions 
at these artificial boundaries have to be such that 
physically meaningful results can be obtained with 
a relatively small integration domain. The number 
of grid points, and therefore computer storage and 
amount of numerical operations required for a nu- 
merical solution, is directly dependent on the size 
of the integration domain. Thus, only with a rela- 
tively small domain may the computational costs of 
numerical simulations be kept within acceptable 
limits. This aspect is of particular importance 
during the testing phases of the numerical methods. 

There are also other difficulties resulting from 
the complicated nature of the governing equations. 
For the nonlinear systems of governing equations in 
the formulations of Section 3 it is not yet possible 
to decide if a given problem consisting of the 
governing equations and a set of boundary conditions 
is well-posed in the sense of Hadamard (1952). More- 
over, it is not obvious whether Hadamard's postulates 
for a well-posed problem are adequate to include 
physically meaningful solutions only. Additional 
difficulties may arise because finite-difference 
methods frequently require more boundary conditions 
than would be needed for the original differential 
formulation if exact solutions were possible 
[Richtmyer and Morton (1967) ]. From numerical ex- 
perimentation with model equations simpler than the 
full Navier-Stokes equations it is known that these 
additional 'numerical' boundary conditions are of- 
ten a source of numerical instabilities possibly © 
caused by certain inconsistencies. Therefore, one 
is confronted with the delicate task of selecting 
and implementing the extra conditons (where it is 
normally not known a priori which conditions are 
the extra ones) in such a way that the numerical 
stability of an otherwise stable method would not 
be adversely affected. 


Initial Conditions 

When the simulation of space dependent transition 
phenomena is of interest as in the present in- 
vestigation the reaction of the flow to disturbances 
introduced at the upstream boundary has to be cal- 
culated. In this case one may assume an undisturbed 
flow as initial condition at t=O enabling the dis- 
turbance waves introduced for t>0 to propagate down- 


stream into an undisturbed flow field. Denoting 

the undisturbed flow field with capital letters the 
initial conditions for the w,i) system can be written 
as 


wW(x,y,0) = 2(x,y) , 


14 
W (x,y ,0) = ¥ (x,y) ’ ( ) 
and for the w,u,v system 
w(x,y,0) = 2Q(x,y) , 
u(x,y,0) = U(x,y) , (15) 
v(x,y,0) = V(x,y) , 


The undisturbed flow field is obtained by solving 
the Navier-Stokes equations for the steady flow. 

Of course, for the flow between two parallel plates 
the Poiseuille profiles already represent exact 
solutions of the Navier-Stokes equations and can 
therefore be used directly. For the boundary-layer 
flow a solution has to be calculated numerically 
by solving the Navier-Stokes equations without the 
unsteady termdwft in Eq. (1). The argument could 
be raised that in this case Blasius profiles could 
be used instead. The differences between the 
Blasius solution and a numerical Navier-Stokes sol- 
ution are indeed very small. Nevertheless, for 
investigations with very small disturbance ampli- 
tudes, the differences can be of the same order of 
magnitude as the disturbances themselves and there- 
fore the transient character of the flow could 
become considerably distorted. The boundary condi- 
tions used for the calculation of the undisturbed, 
basic flow are discussed subsequently in connection 
with the conditions used for the calculation of the 
unsteady, disturbed flow. 


Boundary Conditions 
At solid walls (non-permeable, no-slip), such as 
boundary A-B of Figure 1 or A-B and C-D of Figure 
2, the velocity components vanish 

TO , WO - UO 7 WO oc (16) 
The vorticity-transport formulations (the u,v,p 
formulation will not be discussed further) require 


special treatment for the vorticity calculation at 


the walls. For the w,i) formulation vorticity can 
be calculated from the relationship 


2 
WO = = (17) 


derived from Eq. (2); for the w,u,v formulation 
either 


Qt) _ OA 
a Wisc ime) 
derived from Eq. (7b) or 
OS = (19) 


; resulting from Eq. (3) can be used. Equations (17)- 


15 


(19) are applicable for the calculation of both the 
steady, undisturbed and the unsteady, disturbed flow. 
At the upstream boundary A-D the disturbances are 
introduced by superimposing onto the profiles of a 
basic, undisturbed flow (denoted by subscript B; for 
example, Blasius profiles or Poiseuille profiles 
could be used for the cases considered in Figures 
1 and 2) so-called perturbation functions which are 
dependent on y and t only. Thus for the w, formu- 
lation we have 


w(O0,y,t) = Wply) + Pyly,t) , 


(20) 
DOr) = WG) se tera) 
and for the w,u,v formulation 
w(O,y,t) = SH) Pin Q¥pie) 
u(O,y,t) = ugly) + Pyly,t) , (21) 


v(O,y,t) = vply) + Pyly,t) 


For the calculation of the steady, undisturbed 
flow field the perturbation functions in Eqs. (20) 
and (21) of course vanish. For simulations of 
Tollmien-Schlichting waves, for example, the 
perturbation functions are periodic in time where 
amplitude distributions (or so-called perturbation 
profiles) as obtained from linear stability theory 
can be used. 

The freestream boundary C-D (Figure 1) for the 
boundary-layer flow is an artificial boundary and 
requires special considerations as discussed in 
Section 2. For both the calculation of the steady 
flow and the unsteady, disturbed flow, vorticity is 
assumed zero (w'=2=0). For boundary-layer type flows, 
vorticity for both basic and disturbance flow (when 
disturbances are introduced within the boundary lay- 
er) decays rapidly away from the wall and is practi- 
cally zero at a distance of two 6 (6 boundary layer 
thickness) from the wall. 

For the calculation of the steady flow using the 
W,u,Vv system suitable conditions for C-D are 


U = Ugg (x) (22) 


where the freestream velocity Ug (x) may be speci- 

fied according to the downstream pressure variation 
of the boundary layer flow. A condition for the v 

component can be derived from the continuity equa- 

tion, ((5))/,,,.using Eq. | (22) 


TORU ge LS 
dy dx 0 (23) 
For the w,i) system a condition equivalent to Eq. (22) 
can be used 


= = Ugg(x) - (24) 
The w',u',v' disturbances decay relatively slowly 
in direction normal to the wall. For example, for 
Tollmien-Schlichting waves the ' or v' amplitude 
at 66*, (for Re*=630, based on displacement thick- 
ness 6*) may still be close to 50% of the maximal 
amplitude. Therefore Dirichlet conditions (u'=v'= 
w'=0) could only be used if the freestream boundary 
were very far, for example 506*, from the wall. 


16 


This would be impractical due to the excessive 
amounts of grid points required. On the other hand, 
the conditions given below allow a relatively small 
integration domain in y-direction. They only postu- 
late that the disturbances decay asymtotically in 
y-direction. For the w,i formulation such a condi- 
tion is 

On ae ’ 

ae - ay : (25) 


and for the w,u,v formulation 


we att 
Onvaa - 
(26) 
dv! Plea! 
dy 


where a is the local wave number of the resulting 
disturbance waves. Test calculations have shown 
that with the conditions (25) or (26), together 
with the Dirichlet-type vorticity condition dis- 
cussed previously, physically meaningful results 
can be obtained when the integration domain in y- 
direction includes only two to three boundary-layer 
thicknesses. 

Selection and implementation of the boundary 
conditions at the downstream boundary B-C represents 
a very difficult task. These boundary conditions 
have to enable propagation of disturbances right 
through this boundary, where any effects causing 
even the slightest wave reflection have to be 
avoided. The conditions found most satisfactory 
in this respect are for the w, formulation 


Be ae 
ro apse 
(27) 
ay! Oe 
Spon ano ata 
and for the w,u,v formulation 
920! E- Dee ' 
Roe ans 
32u! 2 
Dac oteate BOO es te (28) 
av! aig Wie 
aT atv 


Numerical experiments with conditions (27) and (28) 
have shown that physically reasonable results are 
already possible when, for periodic upstream dis- 
turbance input, the length of the integration domain 
includes only three to four wavelengths. 

For the calculation of the steady flow (for the 
boundary-layer flow, for example) boundary. condi- 
tions which are compatible with those of the unsteady 
calculations are for the w, system 


7 = OA SSE SO 4 (29) 


and for the w,u,v system 


922 92u 92v 
ox2 Oke axe Oy og ox2 Y 


The boundary conditions Eqs. (27) or Eqs. (28) for 
the downstream boundary [also Eqs. (25) and (26) 
for the free stream boundary] can be derived assum- 
ing neutral, periodic behaviour of the disturbance 
flow. However, extensive test calculations have 
shown that use of such conditions does not enforce 
a strict periodic behaviour of the disturbance flow 
near these boundaries. Rather, these conditions 
allow damping or amplification of the disturbances 
even on these boundaries themselves. These con- 
ditions have also proven to be applicable for cal- 
culations with periodic disturbance input of large 
amplitudes as well as for non-periodic disturbance 
input (random disturbances, for example) [see Fasel 
et al. (1977) ]. 

For cases where a is not known a priori it can 
be determined interatively. Starting with an ini- 
tial guess ao (x) (a is generally a function of x, 
of course, although for the derivation of the 
boundary conditions it was assumed constant to 
arrive at simple relationships) an improved a(x) 
can be determined from the resulting disturbance 
waves developing in the integration domain. Even 
with relatively crude initial guesses ag(x) (for 
example ag=0) this interation loop converges 
rapidly, and for practical purposes two or three 
iterations are sufficient. 

There is no formal difference between the bound- 
ary conditions (27) and (28) used for the w,W and 
w,u,v formulation, respectively. Both sets of con- 
ditions specify relationships for the second deriva- 
tives in the disturbance variables. Nevertheless a 
subtle difference does exist. Condition (27) for 
w' implies that (due to the definition of w, Eq. 4b) 
for v' a relationship involving the first derivative 
is prescribed 


—=-a*p'. (31) 


This is obviously more restrictive than condition 
(28c) where for v' a second derivative is prescribed. 
For small periodic disturbances the two sets of 
conditions lead to practically the same results, 
although the results with the w,W system, together 
with conditions (27), exhibit subtle irregularities 
near the downstream boundary for the waves propa- 
gating through this boundary. The w,u,v system, 
together with conditions (28), however become su- 
perior to the w, system with conditions (27) when 
larger disturbance amplitudes are involved. In this 
case, reflection-type phenomena can be observed in 
increasing manner at the downstream boundary for the 
w,) system. For the investigation of the effects 

of a backward-facing step on transition [Fasel et 
al. (1977) ] the small vortices traveling downstream 
are caught at the downstream boundary when the w,\ 
system and conditions (27) are used, rendering the 
numerical results worthless. Using conditions (28) 
with the w,u,v system, on the other hand, allows 
smooth passage of these vortices through that bound- 
ary. . 

For these reasons conditions (28), in connection 
with the w,u,v system, have proven to be the best 
choice so far in properly treating the downstream 
boundary. The relatively small upstream influence 
of these conditions can be best demonstrated with 
typical results from test calculations. Figure 5 
for example, shows a comparison of the disturbance 
variable u' for calculations with small periodic 
disturbances where first in Eqs. (28) an adequate 
value for a (a=35.6, obtained from linear stability 


cee oe 40 x/Ax 
-0.0004 
— a in eq (28) from linear stability theory 'T 
—— a=0 in eq (28) y 
FIGURE 5. Downstream development of u'-disturbance at 


y/Ay = 3 for different boundary conditions at the 
downstream boundary (boundary layer on a flat plate). 


theory) was used while for the other calculation a 
was simply set zero. It is obvious that even with 
the poor value for a the upstream influence is re- 
stricted to a region of approximately one wavelength, 
while the disturbance further upstream is practi- 
cally unaffected. This relatively minor upstream 
influence can also be observed in Figure 6 where 
the amplification curves (for the maximum of u') 
are compared for the two cases. The disturbance 
amplification further than one wavelength upstream 
is practically unaffected by the value used for a 
in Eqs. (28). 


5. NUMERICAL METHOD 


A numerical method for transition studies has to 
generally allow for numerical solutions of a 
boundary-value problem for the calculation of the 
steady flow, i.e. solution of Eqs. (1) and (2) or 
Eqs. (1) and (7) (without dw/dt in Eq. 1) with ap- 
propriate boundary conditions discussed in Section 
4. Further the solution of a mixed initial-boundary- 
value problem for the calculation of the unsteady 
flow is required, i.e. solution of Eqs. (1) and (2) 
or Eqs. (1) and (7) with the boundary conditions for 
the unsteady, disturbed flow and initial conditions 
discussed in Section 4. The partial differential 
equations are of fourth order for the w,) formula-— 
tion and of even higher order for the w,u,v-system. 
For both formulations the governing equations are 
elliptic for the calculation of the steady flow and 
parabolic for the unsteady flow. In this paper the 
discussion is restricted to application of finite- 
difference methods for the solution of the mathe- 
matical problems posed. 

A difference method for investigations of hydro- 
dynamic stability and transition phenomena has to 
meet a number of requirements in order to ensure 


Lal 
Ao 
1.5 
a ineq.(28) from 
linear stability theory 
1.0 a=0 in eq (28) 
0 10 20 30 40 x/Ax 
"FIGURE 6. Amplification curves for maximum of u' for 


different boundary conditions at the downstream bound- 
ary (boundary layer on flat plate). 


17 


success. Some of the requirements deemed most 
important in this context are as follows: 


(i) Stability, convergence 


Rigorous mathematical proofs of (numerical) stabil- 
ity and convergence for nonlinear problems as dif- 
ficult as the one at hand have not been accomplished 
as yet. For the present investigation, however, 
stability of the numerical method is of fundamental 
importance. Numerical instability is frequently 
exhibited in form of oscillations which would be 
hardly discernible from the physically meaningful 
oscillations caused by introduced forced perturba- 
tions. Hence, a prospective difference method has 
to be highly stable, even for relatively large 
Reynolds numbers. 

In general, for transition studies of the kind 
considered in this paper convergence is also quite 
serious. Convergence is not necessarily guaranteed 
if for a properly posed problem the numerical scheme 
is stable and consistent as is the case for linear 
partial differential equations of second order 
[Lax's equivalence theorem, see Richtmyer and 
Morton (1967)]. However, experimenting first with 
small periodic disturbances one can at least empir- 
ically check the convergence behaviour of the nu- 
merical method by comparing calculations for various 
grid sizes with linear-stability-theory results and 
experimental measurements. Then for other dis- 
turbance inputs, such as large amplitude periodic 
disturbances, one hopes that the convergence char- 
acteristics do not change significantly. 


(11) Accuracy of second order 


For these investigations at least second-order ac- 
curacy of the numerical method (i.e. the truncation 
error of the difference analogue to the governing 
equations, initial and boundary conditions at least 
of second order) is required to exclude or minimize 
undesirable non-physical effects, such as artificial 
viscosity, when mesh intervals of practical sizes 
are used. 

(iii) Realistic resolution of the transient char- 
acter of unsteady flow fields 


Transition phenomena are of highly unsteady nature, 
with the time-dependent behaviour of the flow being 
of special interest. Thus, the difference method 
has to be such that realistic resolution of the 
transient character of such flow fields is possible. 
Therefore truly second-order accuracy is also de- 
sirable for the time derivative. 


(iv) Efficiency with respect to computational 
speed and required fast-access storage capacity 


Numerical solutions of the complete Navier-Stokes 
equations for unsteady flows at high Reynolds 
numbers require numerous time-consuming numerical 
operations. Therefore computers with large, fast- 
access computer storage capacity, reaching even the 
limits of modern computer systems, are necessary. 

A prospective difference method for transition 
simulations has to be extremely efficient, i.e. 
maximizing computational speed and minimizing re- 
quired computer storage capacity as much as possible, 


18 


in order to be capable at all of undertaking inves- 
tigations of this nature with the computers available 
today. 

Of the requirements discussed here, numerical 
stability is the most stringent one and hence has 
to be given most consideration. For this reason 
only implicit methods are suitable. Implicit meth- 
ods are generally much more stable than their im- 
plicit counterparts. For the adequate resolution 
of the large gradients, resulting from the strongly 
time-dependent flow fields to be investigated, rel- 
atively small spatial intervals Ax and Ay are re- 
quired. Using explicit methods this could lead to 
excessively small time-steps required to maintain 
numerical stability. For example, using an explicit 
counterpart to the present implicit method, the 
time-step, according to a linearized stability anal- 
ysis, would have to be more than 100 times smaller 
for a practical calculation than when using the 
corresponding implicit scheme. To satisfy require- 
ment (iv) attention has to be given to making the 
implicit difference method extremely efficient and 
also to meeting the other requirements discussed 
previously. 

Experimentation with various implicit difference 
schemes suggested that 'fully' implicit schemes are 
the most promising for transition studies. 'Fully' 
implicit means that all difference approximations 
and nodal values for the approximation of governing 
equations and boundary conditions are taken at the 
most recent time-level. For our fully implicit 
method three time-levels are employed to obtain a 
truncation error of second order for the time de- 
rivative dw/dt in Eq. (1). 

For all space derivatives, central difference 
approximations with second-order truncation error 
are employed. The implementation of the boundary 
conditions into the numerical scheme requires 
special care so that overall second-order accuracy 
can be maintained. 

This implicit scheme leads to two systems of 
equations for the w,i) formulation and to three 
systems of equations for the w,u,v formulation. 
These systems of equations can be solved by itera- 
tion. Because of the retention of full implicity 
the equation system resulting from the vorticity- 
transport equation is coupled with the Poisson 
equation systems via the nonlinear convection terms. 
It is additionally coupled with the systems result- 
ing from the Poisson equations via the calculation 
of the wall vorticity from Eq. (17) for the w,wW 
formulation and from either Eqs. (18) or (19) for 
the w,u,v formulation. 

A very effective solution algorithm based on 
line-iteration has been developed for our method 
for this coupled system. It is discussed elsewhere 
in more detail [Fasel (1978)]. This solution. algo- 
rithm has shown to be equally effective when the 
basic equations are transformed to allow for a vari- 
able mesh in the physical plane such as, for exam- 
ple, to concentrate grid points close to walls where 
high gradients are expected. Overrelaxation to 
accelerate convergence can be easily implemented as 
has been done for several calculations [Fasel et al. 
(1977) ]. Another advantage is that the solution 
algorithm is readily exchangeable to be applied for 
both the governing equations in w,W and w,u,v formu- 
lation. This has been successfully exploited in the 
investigations of the effects of a backward-facing 
step on transition. In this study both formulations 
were used in the integration domain; the w, formu- 


lation was used in the region containing the corners 
of the step which can be treated more conveniently 
with this formulation. For the domain bounded by 
the downstream boundary the w,u,v formulation was 
applied, because it allows use of less restrictive 
boundary conditions as discussed in Section 4. 

The effectiveness of this solution algorithm can 
be best judged by presenting a typical computation 
time for a practical calculation. For a periodi- 
cally disturbed flow with small disturbance ampli- 
tudes, using a 35 x 41 grid and calculating 260 
time-steps, the required CPU time on a CDC 6600 
is about five minutes, including the calculation 
of the steady flow. This is relatively little, 
considering that the flow is disturbed at every- 
time level and that full implicity is retained in 
the numerical method. 


6. NUMERICAL RESULTS 


The implicity difference method which we have devel- 
oped has been subjected to crucial test calcula- 
tions to verify its applicability to investigations 
of stability and transition. First, the reaction 
of the boundary-layer on a flat plate to periodic 
disturbances of small amplitudes was investigated 
in detail. It was demonstrated that the spatial 
propagation of Tollmien-Schlichting waves could be 
simulated where comparison of the numerical calcu- 
lations with results of linear stability theory and 
laboratory measurements showed good agreement. Re- 
sults of such calculations for the numerical method 
based on the w,u,v formulation are presented and 
discussed elsewhere [Fasel (1976) ]. 

The usefulness of the numerical simulations for 
the investigation of two-dimensional, nonlinear 
effects was demonstrated by calculating the reaction 
of a boundary-layer flow to periodic disturbances of 
larger amplitudes. Investigating the propagation 
of spatially growing or decaying disturbance waves 
in a plane Poiseuille flow (both in the linear and 
nonlinear regime) verified that the numerical method 
is not limited to boundary-layer flows but rather 
that it is equally applicable to other flows of 
importance. Finally, numerical investigations of 
of transition phenomena in the presence of a two- 
dimensional roughness element (backward-facing step) 
showed that simulations with this numerical model 
allow insight into processes which may possibly be 
important for understanding certain transition mech- 
anisms. Results of this investigation and of the 
investigations mentioned before are discussed in 
another paper [Fasel et al. (1977) ]. 

Because the purpose of this paper is to review 
Main aspects of numerical transition simulations, 
emphasis here is not on conveying new results or 
details of numerical calculations. Rather, results 
presented here are intended to be of exemplary 
nature and were selected in order to clearly demon- 
strate essential aspects of such simulations and 
to show what can be expected from such numerical 
calculations. 

The drawings:in Figures 7 and 8 should facilitate 
an evaluation of the potential of such numerical 
simulations, and, of course, also point out possible 
disadvantages and limitations. Figures 7 and 8 
show results for a boundary-layer flow on a flat 
plate, disturbed at the upstream boundary with small 
periodic disturbances. This case is particularly 
Suitable for demonstration purposes. The ensuing 


a) 


0.0005 


IN 44 
0.0000 ig 


Nj 


b) 


0.05 AN 
v La iY Cc 


i 
i ly co 
Ce ag 


‘Ni (C\ 
0.00 EX /, 


—= x/Ay 30 do 


FIGURE 7. 


b) v', c) w', d) w' (different view). 


Tollmien-Schlichting waves that can be studied from 
such calculations are thoroughly investigated, 
experimentally as well as theoretically, and the 
results of these calculations are therefore more 
intelligible than those of more complicated phe- 
nomena of transition. 

For these calculations, based on the w,u,v for- 
mulation, the Reynolds number at the upstream bound- 
ary is Re*=630. For the periodic disturbance input, 
for which perturbation profiles of linear stability 
theory gre u ged, the frequency parameter (defined 
as F=10 Bo/u, , with disturbance frequency 86) is 
We So | Abel Shis case the flow is unstable according 
to linear stability theory (the location of the left 
boundary corresponds to a point on the neutral curve) 
and therefore the disturbances should become ampli- 
fied in downstream direction. For the calculations 
an egqui-distant grid with 35 points in y-direction 
and 41 points in x-direction was used. 

In Figures 7 and 8 the function values of the 
disturbance flow (obtained by subtracting the 
quantities of the basic flow from those of the total 
flow) are plotted for all three fields of variables 
u',v',w', for which the total flow variables are 
directly obtained from the numerical calculations. 
To allow simultaneous representation of the func- 

- tion values at all grid points a perspective rep- 
resentation was chosen where the function values 
are plotted versus the downstream coordinate x/Ax 


Disturbance variables versus x/Ax and y/Ay (perspective representation) at t/At = 80; a) u', 


9) 


c) w 


i. 0.0015 


A 


and the coordinate normal to the wall y/Ay. These 
perspective representations allow the best possible 
qualitative survey of the large amount of data ob- 
tained from such calculations. 

In Figure 7 the disturbance variables u',v',w' 
are plotted for a time instance of t=80At, which 
corresponds to a time of two time periods after 
initiation of the disturbances at the upstream 
boundary. In Figures 7a, 7b, and 7c the view is in 
the direction away from the wall, looking slightly 
in upstream direction. In Figure 7d the view 
is also in the direction away from the wall, look- 
ing now, however, downstream. From these figures 
the propagation of the disturbance waves into the 
undisturbed flow field can be clearly observed. 

Figure 8 shows the corresponding drawings for 
the three variables u',v',w' at a time instance of 
t=250At, that is, more than two time periods after 
the disturbance wave reached the downstream bound- 
ary. These plots demonstrate that the downstream 
boundary conditions work properly. Obviously, the 
waves can smoothly pass through this boundary, 
causing no noticeable reflections. Even after hun- 
dreds of time-steps the flow at and near this 
boundary maintains its time-periodic character and 
therefore the state of the disturbance flow as rep- 
resented in Figure 8 would repeat itself periodi- 
cally if the calculations were continued for further 
time-steps. 


20 


a) 
0.0000 


b) 


WS 


Wi 


p Sr 


au 
XS 


\V7 \ 
Wey 


0.05 


FIGURE 8. Disturbance variables versus x/Ax and y/Ay 
b) v', c) w', da) w' (different view). 


In Figures 7 and 8 the large gradients normal 
to the wall of the u' and w' disturbances become 
clearly visible (for w' this can be best observed 
from Figures 7d and 8d) while v' changes more grad- 
ually. The large gradients observable in these re- 
sults indicate already the major difficulties and 
limitations in numerical simulations of transition 
phenomena. In a numerical solution method these 
large gradients have to be adequately resolved to 
obtain meaningful representation of essential physi- 
cal phenomena. For nonlinear disturbance waves re- 
sulting from disturbance input with larger amplitudes 
[Fasel et al. (1977)] or for other more complicated 
transition phenomena the gradients may become even 
considerably larger. Using finite-difference meth- 
ods of a given accuracy (for example, second order 
as for the present method) better resolution can 
only be achieved by using additional grid points. 
This, however, leads to ever larger equation sys- 
tems the sizes of which are limited by computer 
storage capacity and computation time. 

Some help can be expected from employing vari- 
able mesh systems allowing allocation of more grid 
points closer to walls, where the gradients are 
largest, and using fewer points further away where 
gradients are small. This can be best achieved 
using coordinate transformations for which test 
calculations have shown that sizable savings in 
the number of grid points, and also in computation 


- 0.0015 


c) 


Te BN 40 


d) 


0.0015 4 


0.0000 


0 


(perspective representation) at t/At = 250; a) u', 


time, are possible to achieve accuracy comparable 
with calculations in an equidistant grid. Addi- 
tional improvement may be expected from application 
of higher-order accurate difference schemes (higher 
than second order) which are presently in the state 
of development and about to be used in our numerical 
method. 

The results shown in Figures 7 and 8 also unveil 
the considerable potential and advantages of such 
numerical simulations. The finite-difference so- 
lutions produce a bulk of data, i.e. the values of 
the variables directly involved in the solution 
procedure are obtained for all grid points and for 
all time-levels that are calculated. The data can 
be conveniently stored on mass storage devices, 
such as magnetic tape (used for the present calcu- 
lations, for example). The data stored can be 
processed immediately or at any later data to ob- 
tain any specific information desired, or to produce 
additional data that might be deemed necessary for 
a more detailed evaluation of particular flow phe- 
nomena. For example, they can be used to obtain 
frequency spectra, Reynolds stresses, energy bal-— 
ances, amplitude distributions, or to produce con- 
tour plots (equivorticity lines, stream lines) etc. 
Another positive side of such numerical simulations 
is that if the data would be destroyed or lost, they 
could be reproduced identically, which would be 
hardly possible in comparable laboratory experiments. 


7. CONCLUDING REMARKS 


The objective of the present review was to discuss 
possible approaches to numerical simulations of sta- 
bility and transition based on numerical solutions of 
the Navier-Stokes equations using finite-difference 
methods. The approach, allowing investigations of 
spatially propagating disturbance waves, mainly 
elaborated upon in this paper, appears most promis-— 
ing for realistic numerical investigations of physi- 
cal phenomena occurring in transition. The immense 
amount of reproducible data obtained from such cal- 
culations allows detailed information of any part 
of the flow field which may be helpful to gain in- 
sight into essential mechanisms occurring in tran- 
sition. 

The restriction of the numerical model to two- 
dimensional flows has also a positive side. With 
this model truly two-dimensional numerical experi- 
ments can be performed while in laboratory experi- 
ments it is always difficult to completely exclude 
unwanted three-dimensional effects. Of course the 
later stages of transition are inherently three- 
dimensional in nature and therefore for a study of 
these later developments a three-dimensional model 
would be desirable. 

The main difficulties and limitations of such 
simulations result from the large gradients which 
occur in the transition process. For adequate 
resolution of the large gradients which become even 
larger for more complicated phenomena, increasing 
numbers of grid points are required which may lead 
to excessive requirements of computer storage and 
computation time. 

In spite of these difficulties the number of 
numerical simulations of transition, similar to 
the approach discussed in this paper, is likely to 
increase due to the enormous potential inherent in 
such investigations. Emphasis will probably be on 
the development of difference methods with higher 
accuracy which are applicable for such studies. 
Additionally, increasing use of numerical methods 
other than finite-difference methods is likely, 
such as spectral methods or finite-element methods. 
Finally, with continuing progress in the develop- 
ment to high-speed digital computers, detailed 
quantitative investigations of three-dimensional 
transition phenomena will probably become feasible 
in the near future. 

This research is supported by the Deutsche 
Forschungsgemeinschaft, Bonn-Bad Godesberg, con- 
tract Ep 5/7. 


REFERENCES 


Bestek, H., and H. Fasel (1977). Ein numerisches 
Verfahren zur Untersuchung angefachter, kleiner 
Storungen bei der ebenen Poiseuille-Strémung. 


21 


Paper presented at the GAMM-Meeting 1977 in 


Copenhagen/Denmark. To be published in ZAmMM 58 
(1978). 

Crowder, H. J., and C. Dalton (1971). On the sta- 
bility of Poiseuille flow in a pipe. J. Comp. 


Phys. 7, 12. 

De Santo, D. F., and H. B. Keller (1962). Numeri- 
cal studies of transition from laminar to tur- 
bulent flow over a flat plate. J. Soc. Ind. 
Appl. Math. 10, 569. 

Dixon, T. N., and J. D. Hellums (1967). A study on 
stability and incipient turbulence in Poiseuille 
and plane-Poiseuille flow by numerical finite- 
difference simulation. AIChE J. 13, 866. 

Fasel, H. (1976). Investigation of the stability 
of boundary layers by a finite-difference model 
of the Navier-Stokes equations. J. Fluid Mech. 
7a, S85. 

Fasel, H., H. Bestek, and R. Schefenacker (1977). 
Numerical simulation studies of transition 
phenomena in incompressible, two-dimensional 
flows. Proc. AGARD Conf. on Laminar-Turbulent 
Transition, Lyngby, Denmark 1977. AGARD-CP-224, 
Paper No. 14. 

Fasel, H. (1978). Recent developments in the 
numerical solution of the Navier-Stokes equa- 
tions and hydrodynamic stability problems. Proc. 
VKI-Lecture Series "Computational Fluid Dynamics", 
March 13-17, Brussels, Belgium. 

Gaster, M. (1974). On the effects of boundary- 
layer growth on flow stability. J. Fluid Mech. 
66, 465. 

George, W. D., and J. D. Hellums (1972). Hydro- 
dynamic stability in plane Poiseuille flow with 
finite amplitude disturbances. J. Fluid Mech. 
Bil (7) 5 

Hadamard, J. (1952). Lectures on Cauchy's Problem 
in Linear Partial Differential Equations. Dover. 

Harlow, F. H., and J. E. Welch (1965). Numerical 
calculation of time-dependent viscous imcom- 
pressible flow of fluid with free surface. 
Fluids 8, 2182. 

Nagel, A. L. (1967). Compressible boundary layer 
stability by time-integration of the Navier- 
Stokes equations. Boeing Scientific Research 
Laboratores,:'Flight Sciences Report No. 119. 

Nishioka, M., S. Iida, and Y. Ichikawa (1975). An 
experimental investigation of the stability of 
plane Poiseuille flow. J. Fluid Mech. 72, 731. 

Richtmyer, R. D., and K. W. Morton (1967). Differ- 
ence Methods for Initial Value Problems. Inter- 
science Publishers, Second Edition, New York. 

Roache, P. J. (1976). Computational Fluid Dynamics. 
Hermosa Publishers, Albuquerque. 

Saric, W. S., and A. H. Nayfeh (1977). Nonparallel 
stability of boundary layers with pressure 

gradients and suction. Proc. AGARD Conf. on 

Laminar-Turbulent Transition, Lyngby, Denmark 

1977. AGARD-CP-224, Paper No. 6. 


Phys. 


The Physical Processes Causing 
Breakdown to Turbulence 


M. Gaster 


National Maritime Institute 


Teddington, England 


I want to present some recent experimental observa- 
tions that provide further insight into the physical 
processes that occur in the transition from a lami- 
nar to a turbulent boundary layer. We know that 
external disturbances, such as free-stream turbu- 
lence and sound, excite small pertubations in the 
laminar flow, and that under certain conditions 
these may develop downstream in the form of growing 
wave trains. At low pertubation levels these un- 
stable travelling waves are adequately described 

by the linearised equations of motion. Measure- 
ments on weak artificially excited waves have, by 
and large, provided excellent confirmation of linear 
theory. Far downstream the amplitudes of the per- 
tubation velocities will, however, become too large 
for the neglect of the non-linear terms to be valid, 
and a non-linear description of the motion is nec- 
essary. Even in the relatively simple situation 

of the constrained parallel Poiseuille flow, which 
has been extensively studied, the non-linear the- 
ories so far developed can only weakly describe 
non-linear events, and even then the computations 
are very involved. These non-linear theoretical 
models are nevertheless very helpful in describing 
the various interactions between the fundamental, 
its harmonics, and the mean flow, but they cannot 
go far toward providing a model of the process of 
breakdown to turbulence, nor are they intended for 
that purposes. 

Non-linear analyses have been concerned mostly 
with the evolution of purely periodic wave trains. 
In the case of linear problems it is quite proper 
to consider any disturbance in terms of its Fourier 
elements. Knowledge of the behaviour of purely 
periodic wave trains enables more complex distur- 
bances to be described. Unfortunately this is not 
the case when the disturbance is non-linear, and 
the welcome simplification obtained by breaking down 
a problem into harmonics is no longer valid. When 
the initial disturbances arise from natural rather 
random stimuli the linear wave train will initially 
consist of a band of unstable waves. After some 


22 


amplification a slowly modulated almost sinusoidal 
oscillation will inevitably develop. When the 
selective amplification is very large, as is the 
case in many boundary layer flows, the modulations 
are slow, and it does not seem too much of an 
idealisation to treat the non-linear problems 
analytically as if it were a purely regular wave 
train. It turns out, however, that the degree of 
modulation does not have to be large for its in- 
fluence on the Reynolds stresses and thus the 'mean 
motion' to be very significant. In a typical ex- 
periment on a laminar boundary layer over a flat 
plate in a low turbulence wind tunnel one finds 
that the instability waves are modulated suffi- 
ciently to influence the transition process. It 

is found, for example, that breakdown to turbulence 
occurs violently and in a random manner quite un- 
like the type of breakdown that is observed in 
controlled periodic wave trains. Measurements on 
isolated wave packets also show the effect that 
modulation of the wave train has on transition, but 
in a more controlled way. 

Previously reported measurements [Gaster and 
Grand (1975)] on artifically excited wave packets 
showed consistent and quite well defined deviations 
from the structure predicted by linear theory. 
Since the maximum level of the velocity fluctua- 
tions measured lay below that for which significant 
non-linearity is exhibited by regular periodic wave 
trains, the reason for this behaviour was at that 
time unclear. In the experiments only one level 
of input excitation was used and so there was no 
direct way of assessing the importance of the non- 
linear terms. These experiments have been repeated 
at the National Maritime Institute with various 
levels of input excitation and it has now been con- 
clusively established that the previously observed 
warping of wave fronts and the non-Gaussian char- 
acter of some of the hot-wire signal envelopes arose 
from non-linearity. This behaviour can best be il- 
lustrated by showing a comparison of the hot-wire 
signals that arise: (a) with a sinusoidal input, 


and (b) a pulsed input. As in the previous series 
of experiments the boundary layer flow was excited 
by an acoustic device mounted in a recess on the 
reverse side of the flat plate. A small hole 
through the plate provided the necessary fluid 
dynamic coupling at a point on the boundary of the 
working face. Figure 1 shows a set of hot-wire 
anemometer records taken with the probe mounted 
just outside the boundary layer one metre down- 
stream of the leading edge. The exciter was driven 
sinusoidally at four different amplitude levels 
increasing from (i) to (iv). The velocity fluc- 
tuations appear to be regular and show no harmonic 
or other distortion until the level of turbulence 
intensity exceeded 1% peak-to-peak of the free- 
stream velocity (see iv). Exciting the flow with 
isolated pulses on other hand, produces a some- 
what different picture. Figure 2 again contains 
four hot-wire records obtained with different 
levels of drive applied impulsively. At the lowest 
level shown the signal consists of a smooth roughly 
Gaussian packet of ripples, but even a small in- 
crease in driving amplitude produces a clearly 
discernible distortion to this signal. These dis- 
tortions are similar to those obtained in the 
earlier experiments quoted. As the amplitude is 
further increased the signal becomes increasingly 
distorted until at some level a secondary burst 

of relatively high frequency oscillations appears. 
It should be remarked that the amplitude scaling 

on both Figures 1 and 2 are identical, showing that 
non-linear effects occur at much lower amplitudes 
for the impulsively applied disturbance than for 

a periodic one. In these particular experiments it 
appears that non-linearity becomes apparent in the 
hot-wire signal at a peak to peak amplitude of only 
1/sth that for a continuous wave train. 

The high frequency oscillation appears to be 
associated with a steep shear layer that forms 
within the velocity profile momentarily as the 
wave packets sweep past the measuring station. 
These shear layers initially appear on either side 
of the centre line, and not surprisingly therefore 
the peak levels of the high frequency secondary 
oscillation also arise off centre at roughly these 
locations. The high frequency waves grow rapidly 
with downstream distance, initially developing 
exponentially but later the growth levels off. At 
that stage the filtered secondary wave packets were 
observed to distort in a way reminiscent of the 


c— 0,05 Seeso = 


FIGURE 1. Hot-wire Signals from Sinusoidal excitation. 


23 


p————— (0) SESS) ——$_—$$_— 


FIGURE 2. Hot-wire Signals from pulsed excitation. 


primary wave packet. It was therefore conjectured 
that there might be yet a further level of insta- 
bility on the secondary wave oscillations when 
these became sufficiently large. Just two days 
before leaving for this meeting this idea was 
tested. Hot-wire signals from appropriate regions 
of the flow were filtered to see whether there was 
any signal above the frequency of the secondary 
oscillations. When the secondary wave amplitude 
was large, a burst of high frequency oscillations 
could be seen on the oscilloscope. Figure 3 shows 
the result of applying a high-pass filter, set to 
pass above 2 kHz, to such a hot-wire anemometer 
Signal. The time scale of this record is con- 
siderably expanded compared with that of Figures 1 
and 2, and shows that the oscillation frequency in 
the burst was around 5 kHz. The basic primary 
wave packet of roughly 150 Hz developed over 1m 
before breaking and supporting a secondary burst 
of 1 kHz. This secondary instability grew in am- 
plitude to levels large enough to indicate the in- 
fluence of non-linearities in a distance of roughly 
4 cm. The tertiary mode of 5 kHz detected at this 
stage seems likely to grow even more rapidly. One 
can only presume that further stages in this evo- 
lutionary process are inhibited by viscosity. 
These experiments on the non-linear wave packet 
and its breakdown to turbulence are as yet incom- 
plete and it is my purpose here to indicate only the 


r———— 04005 Seess, ——_——— 


FIGURE 3. High frequency burst. 


24 


most important features of the process. Firstly, 

a clear demonstration of the difference between 

a purely periodic wave train and a modulated train 
as far as the level at which non-linear effects 
occur is presented. The local breakdown observed 
in the wave packet case is similar to that observed 
in the breakdown of the modulated wave trains that 
arise from natural random excitations. Secondary 
breakdown does of course also occur on large enough 
periodic waves, but modulation seems to cause this 
phenomenon to take place at somewhat lower levels 
of primary disturbance and in a slightly different 
form. The artifically driven wave packet embodies 
some of the most important features found in natu- 
rally occurring waves, and since they can be gen- 
erated in a controlled manner the effects can be 
quantified. It is essential to understand this 
process if one is going to make estimates of where 
transition occurs on the basis of the amplitudes 
of instability waves calculated from linear theory. 
At present, most prediction methods rely solely on 
the intensity of the most unstable wave. This is 
clearly inadequate as breakdown is also dependent 


on the modulation of the wave train, and consequently 


the bandwidth of the amplified part of the spectrum 
must also be taken into account in some way yet to 
be established. 

Secondly, the transition from regular waves to 


turbulence appears to occur through a cascade pro- 
cess. The stresses induced by a modulated wave 
train cause steep shear layers to form in the bound- 
ary layer. These support instabilities of higher 
frequencies and shorter wavelengths than the waves 
that caused the distortions, and these grow to large 
amplitudes in appropriately shorter distances. This 
process must at some stage be tempered by viscosity, 
but in these experiments three levels of instability 
have been so far detected. The lowest frequency 
motion was artifically excited by the input pulse, 
while the two successviely higher frequencies were 
excited by random turbulence in the flow at the 
particular location in space and time where local 
instabilities existed. The development of a fine 
scale structure is thus a local, almost explosive, 
phenomenon. Such a cascade breakdown process pro- 
vides the necessary mechanism for the generation of 
fine scale motions that arise in a fully turbulent 
flow. 


REFERENCE 


Gaster, M., and I. Grant (1975). An Experimental 
Investigation of the Formation and Development 
of a Wave Packet in a Laminar Boundary Layer, 
Proc Ris) SOC Lond. A\ 34772 53—269% 


The Instability of 
Oscillatory Boundary Layers 


Christian von Kerczek 
David W. Taylor Naval Ship Research and Development 
Center, Bethesda, Maryland 


ABSTRACT 


The instability of the two-dimensional flat plate 
oscillatory boundary layer induced by a stream with 
velocity U, + U; cos wt is considered. The velocity 
amplitudes, U, and U;, are constants and U)/U, is 
assumed to be small. The instability of this oscil- 
latory boundary layer is analyzed by a time-dependent 
linear parallel flow instability theory. The change 
of the Tollmien-Schlichting growth rates due to the 
imposed oscillations are computed to second order in 
U|/U,- It is found that for imposed oscillation 
frequencies in the range of the Tollmien-Schlichting 
frequencies of the underlying Blasius flow, the 
boundary layer is stabilized by the oscillations of 
the external flow. 


1. INTRODUCTION 


In this paper, we study the instability of the two- 
dimensional oscillatory laminar boundary layer which 
forms on a flat plate that is exposed to a stream 
with a velocity, U, + U; cos wt, perpendicular to the 
plate's leading edge. The velocity amplitudes, U, 
and U;, are constants, w is the angular frequency of 
the oscillation, and t denotes time. The considera- 
tions of the instability of oscillatory flows has 
become an important field of research in recent years 
and has been reviewed by Davis (1976). The partic- 
ular class of problems concerned with the instability 
and laminar-turbulent transition of oscillatory bound- 
ary layers has been reviewed by Loehrke, Morkovin, 
and Fejer (1975). The latter review indicates that 
very few studies of instability and transition have 
focused directly on the subject of oscillatory bound- 
ary layers. Such studies that have concentrated on 
oscillatory boundary layers have been mainly experi- 
mental investigations which were restricted to low 
frequency oscillations compared to the oscillation 
frequency of unstable Tollmien-Schlichting waves. 

The only analytical work concerning the instability 


25 


of oscillatory boundary layers has been the quasi- 
steady analysis of Obremski and Morkovin (1969) which 
was aimed at these low frequency cases. 

The study of the instability of oscillatory 
boundary layers has technological as well as funda- 
mental importance. Examples of a fundamental nature 
for which the study of the instability of oscillatory 
flows may have relevance are the problems of how 
ambient disturbances affect the instability of the 
underlying steady boundary layer. Specific examples 
might be the effects of ambient acoustic waves or 
ambient turbulence on steady boundary layer insta- 
bility. The problem of the effects of ambient tur- 
bulence on the instability of a steady boundary layer 
probably is not completely accessible by the theory 
of the instability of oscillatory boundary layers. 
However, a sufficiently complex, but organized, am- 
bient oscillation may be adequate for duplicating 
some aspects of the effects of ambient turbulence 
on steady boundary layer instability. We are hope- 
ful that this may be the case because of similar 
phenomena in the field of nonlinear ordinary dif- 
ferential equations. The study of the instability 
of forced periodic solutions of nonlinear ordinary 
differential equations has furnished a much richer 
class of phenomena than the corresponding study of 
the instability of only the steady solutions of these 
equations [see, for example, Hayashi (1964); in 
particular, the results for the forced van der Pol 
equation, pp. 286-300]. 

In the present study, we focus on the very simple 
oscillatory boundary layer that was described ear- 
lier. The purely oscillatory part of this boundary 
layer is approximated by the oscillatory Stokes layer 
which has no spatial structure in the plane of the 
plate, i.e., it is an exactly parallel flow. Thus, 
this model problem may be too simple to reveal any 
particularly important features of realistic ambient 
disturbances. However, the model problem is a good 
starting point and serves as a basis on which to 
develop the appropriate methods of analysis for the 


26 


instability of oscillatory flows. We will be con- 
cerned mainly with moderate and high frequency os- 
cillations comparable to the oscillation frequencies 
of unstable Tollmien-Schlichting waves. Thus, a 
direct comparison of our results with the low- 
frequency experimental results cited by Loerke, 
Morkovin, and Fejer (1975) will not be possible. 

The method used here for analyzing the instability 
of the oscillatory boundary layer is a combination 
numerical and perturbation method [Yakubovich and 
Starzhinskii (1975)]. In this method, the changes 
in the amplification rates of the free disturbances 
of the underlying steady boundary layer are computed 
as perturbation series in the amplitude parameter, 
U,/Up, for any positive value of the frequency, w. 
Certain resonant and combination frequencies are of 
particular interest. The numerical method used here 
to evaluate the perturbation series allows the ef- 
ficient and easy generation of many terms of the 
series. 

The plan of this paper is as follows: In Sec- 
tion 2, we formulate the basic flow whose instability 
is to be examined along with the associated theory 
instability problem. Section 3 outlines the solu- 
tion method. Section 4 discussed the numerical 
results. Some concluding remarks concerning the 
instability of womewhat more complex oscillatory 
boundary layers are contained in Section 5. 


2. INSTABILITY THEORY 


The basic flow field whose instability is to be in- 
vestigated is the oscillatory boundary layer formed 
on a flat plate in a unidirectional stream with 
speed U, + U, cos wt perpendicular to the leading 
edge of the plate and parallel to its plane. Let 
the cartesian coordinate frame (x,y,z) be placed 
with its origin in the leading edge of the plate, 
the x-axis pointing downstream parallel to the plate, 
the y-axis perpendicular to the plane of the plate 
and the z-axis pointing in the spanwise direction. 
For values of the parameter, (ww/Uy) >> 1, the 
ratio, 8) = 6/65, of the boundary layer thickness, 
6 = Yxv/U,, to the oscillatory Stokes layer thick- 
ness, 5, = VY2v/w, is large and the oscillatory 
boundary layer resulting for small values of A = 
U)/U, can be approximated well [see Ackerberg and 
Phillips (1972)] by the sum of the Blasius profile 
Up (y) [see Rosenhead (1963), p. 225] and the Stokes 
layer profile Ugly, t) [Rosenhead (1963), p. 381]. 

Let us scale the x- and y-coordinates by the 
local value of the displacement thickness 


6x = 1.7208 Y¥xV/U (1) 


Then the transverse coordinate, n, is defined by 

nN = y/éx and x' = x/éx. The time scale is §*/U, 

so that dimensionless time is t' = tU,/5* and hence- 
forth the primes will be dropped. Then the basic 
oscillatory boundary layer profile is given ap- 
proximately by 


U(n,t) = fp(n) + pre [ite G7 ET 1] (2) 


where f£,(n) is the Blasius profile, B = 6x/5o, Q = 
wS*/Uo = 282/Rs,, and Rey = U,d*/v. 

We shall consider the instability of the basic 
flow (2) in a similar manner to the standard two- 


dimensional linear instability theory for steady 
boundary layers. In particular, the quasi-parallel 
temporal instability theory as outlined by Rosen- 
head (1963) is followed. The restriction to two- 
dimensional disturbances can be justified based on 
an extended version of Squires' theorem [see von 
Kerczek and Davis (1974)]. The perturbation ve- 
locities (u,v) are determined from the stream func— 
tion, W(x,n,t) = $(n,t)er* 


(3a,b) 


: ia 
v = Re ae Reiave e 
ax 


The disturbance equation for the perturbation ve- 
locities is then given by 


one Stay one oo _ 840 
ae EP = Ra 22, + ia (use ane eg (4) 
* 


where £ = 32/an2 = a2, i = V-1, Re(a) denotes the 
real part of a, and a is the wave number of the 
sinusoidally varying disturbance in the x-direction. 
The boundary conditions are 


= OF = (0, ens fy) = © (5a) 
an 
and 
gr = >0O asn>®. (5b) 


By analogy with Floquet theory for ordinary dif- 
ferential. equations with periodic coefficients 
[Coddington and Levinson (1958)], we seek solutions 
of (4) and (5) in the form 


O= Haney on (6) 


where g(n,t) is a periodic function of t with period 
27/Q. This is a reasonable choice of solution be- 
cause we are mainly interested in the oscillation 
induced changes of the principal disturbance mode 

of the Blasius flow. The principal disturbance mode 
of Blasius flow has multiplicity one. 

We shall adopt in this study an absolute defini- 
tion of instability which requires that some measure 
of the disturbance amplitude becomes infinite as 
t >, If the amplitude remains bounded as t >, 
then the flow is defined to be stable to infinites- 
imal disturbances. However, we must keep in mind 
that the local instantaneous amplitude may be im- 
portant in this linear theory because a disturbance 
may be transiently so large (but bounded) that the 
linear instability theory is no longer valid. Fur- 
thermore, the instantaneous magnitude as a multiple 
of the initial magnitude of the disturbance is an 
important quantity for assessing the likelihood of 
transition from laminar to turbulent flow. Thus we 
shall consider in detail the gross amplification 
rate G of a disturbance which we define by 
ann 
en dt 


G= (7) 


where em is the total energy of the disturbance de- 
fined by 


27 


efficients, a -a1+++/Ay, are determined by the 
boundary conditions once 4 is known. The matrices, 
Q,P,J, and V, are the respective representations of 
the operators ,£,£7, (fp£-9°fp/dy") and (f-3%df/ay2) , 
together with the boundary conditions (6) in the 
space, y, whose basis is the first N Chebyshev 
polynomials, T),..., Ty-1 lOrszag (1971)]. The 
function, of, is the Stokes layer profile. 


1/a 
(u2 + v2) axdn (8) 


sig 


Q = 
ae 


OSS 


co 
ae 
fe) 


Then the relative amplification ratio, ep, /en POs 
a disturbance as it grows during the time intérval 
from t, to t); can be shown to be 


a : = 
-(1+ § 

a. ao) 5, is ra oe CE esa 

aie = TICE) exp 2 A at (9) ) 

8 ¢ cS Note that the matrices, Q,P, and J, are real constant 
matrices and V is real and time periodic and of the 
form 

ag|2 (1) int (=1) -int 
if - =i 
u(e) = ff (= + a*|g]* } an (10) VS Ye FETS (15) 
° 
(1) (-1) ; 
where V and V are constant matrices. 


and rr = Re(X). 

Since the disturbance energy vropogates down the 
boundary layer at the group velocity, c. [see 
Gaster (1962)] one can compute the aanaelye ampli- 
fication ratio, en /ep , by calculating the integral 


The matrix, Q, is invertable so that we can 
multiply (13) by Ont to get 


3 ; ; da iL : Sy = 
in the exponential function of Eq. (9) over the as ae (P'+iaJ')atiaAv'a (16) 
spatial interval, X) to x], using the transformation, Sy 
dx = cgdt. a , 
g where P' = Q Ip etc.; henceforth we shall dispense 


with the primes in (16) 

The perturbation procedure is most easily and 
illuminatingly carried out by transforming (16) so 
that the matrix, (PtiaJ)/R , 148 in diagonal form. 
That is we will be working directly in the (approx- 
imate) eigenspace of the steady Orr-Sommerfeld 
equation for Blasius flow. Suppose that the in- 
e(Aota1At.--)t (11) vertible matrix, B, transforms (P+iad) /Re into 

diagonal form. Then let 


3. SOLUTION OF THE DISTRUBANCE EQUATION 


Solutions of Eq. (4) in the form (6) can be obtained 
as a series in A, 


g= (g +g, +- 5.6) 


where each term of (11) can be evaluated by solving 5S (17) 
appropriate perturbation equations obtained by sub- 
stituting (11) into (4) and (5). Such perturbation and substitute (17) into (16) and left-multiply by 
equations are basically inhomogeneous unsteady Orr- Bethe 
Sommerfeld equations and must be solved numerically. 
Our approach is equivalent to this except we reverse ab 2 2s 
the procedure by first executing a numerical pro- ae = Db + AEb (18) 
cedure which reduces the Eqs. (4) and (5) to a sys- 
tem of ordinary differential equations in time. where 
These are easily solved by perturbation theory to as 
high an order as desired. Salt F 
Let us first expand the function ¢ in the Cheby- DSB (pew) 8) 6 Aire +s Ayey | (19a) 
shev series o 
x 1 (GX) calighs (Si), Sale 
OW, = » a (e)0 (y) (12) i} SS gratis}! “Wyss = in) e +E e (19b) 
n mea 
n=1 and the notation, [an gosor ad, |: stands for a di- 
agonal matrix of order n. 
where the Day) = cos7!(n cos y), n = 0,1,...are the The problem is now to find solutions of (18) in 
Chebyshev polynomials of the first kind and where the form 
we have mapped the interval, neo, no), onto ye[-1,1] . = » Ne 
Then we use the t-method as described by Orszag b(t) = z(t)e (20) 
(1971) to obtain the system of ordinary differential 
equations where Z(t+27/2) = z(t). We are mainly interested 
in perturbations of magnitude A of the steady flat- 
dawn! 7 Spa = plate disturbance mode which becomes unstable far 
2 Ge ~ See ane CHLOE: cS) downstream of the leading edge. This mode is as- 
8 sociated with one of the eigenvalues of D, say A_, 
which for values of x between the two values, a < 
where Q,P,J and V are (N-4)x(N-4) term matrices and X], Satisfies ReA, > 0. It is known that Ap is a 
a = (a),-.-,a eae The dagger (+) superscript de- simple eignevalue [see Mack (1976)] so that a solu- 


notes the transpose of a vector or matrix. The co- tion of the form (20) can be expanded as 


28 
Z(t) Sze (e)) EVAZ (CE) INWIAN(S) 855, | (Amel) 
A= A, + Ao) + AZ Gio) hace (21b) 


Substituting these two expansions into (20) and 
(18) and equating the terms of equal order in A 
yields the set 


dz, 

Sas (DA oe = 0 (22) 
dz, “ * 

Sere a Dane = (E-o,1) 2, (23) 
dz, m v a 

Sra ea (D-ApI)z, = (Gg, We = Rae (24) 


ercr 


Note that the constant coefficient matrix of 
these equations is 


1D)) = att = ly gosoniio | 


where Y3 = A; - Ano) S pooap NY Sil o de 


The only 27/2 periodic solution that is possible 
for Eq. (22) when (A5--Ap) FEM roe SW OPI APS oocborel 
j *#oe is the solution 

B. (9 4) (25) 
where 6 is the Kronecker delta and c is an arbi- 
trary complex constant. This statement is merely 
a restatement of the fact that the eigenfunction 
corresponding to the eigenvalue, A,, is the p-th 
column of matrix B, i.e., the least stable eigen- 
mode of the underlying steady Blasius flow. 

Since the solution (20) requires that z(t) be 
periodic with period 27/2 in t, we shall need the 
inner product <f,g> defined by 


= - ) 
=> = * 
SE,GD Soe if ) £9, dt. (26) 


where the asterisk superscript denotes the complex 
conjugate. We shall also need the adjoint eigen- 
function of Eq. (22) that corresponds to the eigen- 
value, Yo = 0, and that is 21/2 periodic in t. This 
eigenfunction is 

+ 


7. = Cla) 6 
Me ( PE (27) 


For convenience we normalize Z,, and Vins that 


“a5 SS. 
o'%o 


by setting c=d=1. 

The solution of any one of the equations in the 
set (23), (24), etc. is obtained from the solution 
of the previous member, by the application of the 
Fredholm Alternative and the requirement that these 
solutions are unique, i.e., they do not contain 
multiples of the eigensolution of Eq. (22) and are 


27/2 periodic in time. All of the equations of the 
set (23), (24) etc. have the form 


— - (D-A_I)z. = h(t) (28) 
p J 


where h(t) is a periodic vector function which has a 
Fourier series representation of the form 


fA ( t) 2 ne 
kK==a0 


(29) 


where hy are constant vectors and the p-th component, 
hoo of hey, is zero. (This property is enforced by 
the solution procedures.) 

Then, application of the Fredholm Alternative 
for solving (28) yields the requirement that 


<h(t) -¥,> =0 (30) 


Assuming condition (30) to hold (this will be 
achieved by properly selecting the 55'S), the 
general solution, 24, can be written as 


z, = exp [(D-) 1) €] x 


t 


EB. 6 i exp[-(D-A 1) s]h(s)ds (31) 


(0) 


Equation (31) is easily evaluated because 


Vert 


Neonat 
exp (PSX t) el ae Waurere al (32) 


Equation (32) is the main reason for diagonalizing 
the matrix, (1/Rs,) (P+iad) . It makes evaluation 
of the exponential matrix and the integral of Eq. 
(31) trivial. Thus, by evaluating Eq. (31), and 
requiring that z(t) be unique and 21/2 periodic 
in t, values for the constant vector, Eq, are ob- 
tained which eliminate all the non-27/2 periodic 
functions from (31). The result of these calcula- 
tions is 


co 7 


h ; 
Z.(t) = oe ee (33) 
5} LkQ-Y 9 


k=-00 


The solution procedure then is to apply Eqs. (30) 
and (33) to each of the Eqs. (23), (24) etc. in 
sequence starting with (23). These calculations 
have been programmed and are quite easily performed. 
(Our program does these calculations to the 7th 
order term, but more terms can be easily incorpo- 
rated.) We are mainly interested in the first two 
perturbation terms which result in 


o, = 0; 2,(t) = eee Werte (38) 


where 


E. 
=(1 D (35a) 
z | ) 


+ 
= 
zi) = Ra ' (35b) 
and 
N! 
(il) (=) (5) en (GLP) 
= IDA Ga a ID a 0 36 
%2 , Gas ai 2 ae 
jaa 
Bo (t) = Fl e28Mt , lO) 4, g(-2)-2i8t 36) 
where 
N' ar 
(a). (1) 
Dn 1G 
2 £3 74 
=1(2) je 
nN 2i0-V, , (37a) 
N' 
(i) = (on) (2) (a) 
= = im eG pO aI ) 
(o) s ‘ vay 3) a) 3) 
j=l 
+ (37b) 
- O58 og /\-¥9) , 
N! + 
ak pl) ,C) 
=(=2) eo The? 
n = j=l . (37c) 
~ 212-5 


We note that the order A perturbation, 6), of 
the eigenvalue _ is zero so that the long-term 
effect of a flow? oscillation with amplitude A is 
only of order AZ. However, the short-term effect 
is still of order A because the eigenfunction, 
Z,(t), appears in the term I(t) in the relative 
amplification ratio, ep /en,# given by Eq. (9). In 
fact, the structure of the Matrix, E, is such that 
all values of 6. with odd indices are zero and A 
has an expansion in even powers of A about the 
simple eigenvalue, ro: This can be surmized easily 
from the fact that the phase of the imposed oscil- 
latory part of the boundary layer flow should not 
play a role in the modifications of the eigenvalue, 
X,. Furthermore, note that the solutions, 2) (t) 
and Zo(t), exhibit clearly the possible effects, at 
second order, of certain resonant couplings. None 
of the denominators in (35) and (37) are zero be- 
cause y. #£+ ik for any integer values of j or k; 
hence these solutions are uniformly valid for any 
positive value of the frequency, 2. It is possible, 
however, that at resonant frequencies such as at 
n=t8m (5), the value of o» will have a relative 
Maximum. Of particular importance is that in the 
low frequency limit, 2 > 0, the o4's may be singu- 
lar. The lower values of 2 will be an important 
consideration and will be discussed in detail in the 
next section. 


4. NUMERICAL RESULTS AND DISCUSSION 


‘Before describing the computational results that 
have been obtained, we emphasize that in this work 


29 


the instability of the oscillatory boundary layer 
as a whole is being compared to the instability of 
the underlying steady Blasius boundary layer. How- 
ever, it is easier to describe this comparison in 
the terminology of the oscillatory forcing of the 
Blasius boundary layer instability. For example, 
if the oscillatory boundary layer is less stable 
than the steady boundary layer by itself, then we 
describe this situation as one in which the imposed 
oscillations tend to destabilize the steady flow. 
The first set of calculations were made to test 
for resonant interactions at second order in A. By 
consulting the solutions (35) and (36), it can be 
seen that the mean effect of the imposed oscilla- 
tions on the eigenvalue, Aon is manifested by the 
term, 09. There are two types of resonances pos- 
sible. The first type is the "harmonic parametric 
resonance" which corresponds to values of 2 given 
by Wo/k = 1/2, 1, 2,... where w, is the response 
frequency of the disturbance, W, = Q M\p- The 
second type of resonance is the "combination reso- 
nance" corresponding to values of 2 given bydm 
(10ty 5) = 0 (note the denominators of solution 35). 
Figure 1 shows the computational results at certain 
frequencies 2 in the range, l<w./Q<3. It can be 
seen that the imposed oscillations stabilize the 
flow. Figure 1 shows that no resonance effects are 
predicted at either Wy /2 = 1,2),3, or at Wp /2 = 1417 
and 1.74, which correspond to the two possible com- 
bination resonances in the frequency range shown. 
This lack of resonance effect results mainly be- 
cause the external free stream oscillations induce 
a significant amount of oscillatory vorticity in- 
side the boundary layer only in a region very close 
to the wall. This can be seen by examination of 
the Stokes layer profile (14) where the exponential 
factor has a vertical decay constant, 8, which is 
equal to about 5 in the range of frequencies con- 
sidered. The main fluctuations of the disturbance 
velocity are concentrated at the mean critical layer, 
Nc * 0.5 [where no is given by cy = fp(nc) and c, 
is the mean phase velocity of the disturbance]. 
Thus, instead of the Stokes layer interacting 
directly with the disturbance of the underlying 
steady boundary layer at the level, nc, where most 
of the disturbance energy is being produced, it is 
confined mainly to the wall region where it cannot 
be very effective. Furthermore, the Stokes layer 
lacks a spatial structure in the x-direction that 
can match in some way the spatial structure of the 


9 SpaeatlieeeaGal ae MarR TP al ral) T 
-2 
w 
2 
% 
ie 
3 
ac 
-4 
-6 
1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6 2.8 3.0 
wp /2 


FIGURE 1. The growth rate perturbation Re 09 for a = 
= 1128 (on the neutral curve) Wp = Wo/Rex 
= 0,43 « n0-*. 


30 


disturbance mode. It is notable that the imposed 
oscillations have an increased stabilizing effect 
as 2 decreases (Wp /2 increases). This increased 
stabilizing effect can be expected for two reasons. 
The first reason is found in the solution (35) 
which shows that the terms, 


(aye 


Nast 
Oe (+12 alo 


may become unbounded if 


5!) ay ae 
pp pp 
remains bounded as 2 > 0, because y, = 0. Secondly, 


it can be seen in the Stokes layer profile (14) that 
the oscillations of the boundary layer become more 
effective in penetrating up to the critical layer 
when 2 decreases (i.e., B also decreases since 


B = YOR. Ne 
* 


However, we cannot use the present parallel flow- 
model at very low frequencies because in one period, 
27/2, of the imposed oscillation, a disturbance 
will propogate down.the boundary layer a distance, 
6x, that is too large for the parallel flow assump- 
tion to hold (i.e., constant boundary layer proper- 
ties in the x-direction). For example, the change, 
ORs, in the displacement thickness Reynolds number, 
R§,, Over the distance, 6x, (near the values of a = 
0.15 and R§, = 1200) is given approximately by 

SRe = Jeu) . (38) 
where N = Wo/Q. Thus, in the range of values of a 
and Rsy of our calculations 6Rgx = sO NPSonthakw tor 
N = 3, SRsy is nearly 20 percent of the value of 
Rgx- Under the circumstance, the parallel flow 
approximation is only roughly valid. Nevertheless, 
the values of 6Rs, as a fraction of Rg, decrease as 
one goes downstream of the neutral curve for fixed 
values of the frequency ratio, Wp /Q. Thus, the 
parallel flow approximation improves as one follows 
a constant frequency disturbance downstream of the 
neutral point. 

The second set of calculations that were per- 
formed was for the amplification of a fixed frequency 
disturbance propogating down the oscillatory bound- 
ary layer. Two values of Wo/2, equal to 2 and 3, 
were chosen for illustration. The disturbance ex- 
amined is an unstable Tollmien-Schlichting wave of 
constant absolute frequency Oe = Wo/Rg x = 0.43 x 107" 
along the constant frequency line a = 0.00133 Réy- 
This disturbance first begins to grow in the steady 
boundary layer at the values of a = 0.15, Rox = 
1128, and ceases to grow at about the values of a = 
0.3 and Rs, = 2255-, The disturbance trajectory a 
0.00133 Rs* passes nearly through the point in the 
a, Rs, plane of maximum rate of amplification. 

Figure 2 shows the values of Reo» obtained for 
the growing Tollmien-Schlichting wave along the 
trajectory, a = 0.00133 Rgx, at the two different 
values Wp /2 = 2 and 3. An interesting feature of 
the results in Figure 2 is that [Reo | increases 
with Rsgx although the quantity 8 also increases 
which would seem to indicate further decoupling of 
the oscillatory Stokes layer (14) from the distur- 
bance oscillations., Presumably, the values of 


-14 
1000 1200 1400 1600 1800 2000 2200 2400 
Rs, 
FIGURE 2. Growth rate perturbation Re o2 along a = 
0.00133 Rex- 


[Reap | decrease as Rox becomes sufficiently large 
for then 8 also becomes so large that the Stokes 
layer will almost completely disappear. It can be 
seen in Figure 2 that the stabilization of the 
boundary layer can be substantial for the value of 
Wp /2 = 3 and at the larger values of Rox- 

Figure 3 shows the values of Red = ReXp + A?Reo5 
for the value of A = 0.1 and the three values of 
Wp /2 —JO/2 mands se (W5/2 = 0 is equivalent to A = 
0). The total effect of the imposed oscillations 
with A = 0.1 is not very substantial at the value, 
Wp/Q = 2, but at the value of W/L = 3, the sta- 
bilization of the flow is significant. We note that 
an oscillation amplitude of A = 0.1 is a rather 
large value at the frequencies considered here and 
would require a large amount of power to achieve 
in an experimental test facility such as a wind 
tunnel unless the mean flow is very slow. 

The rates of amplification shown in Figure 3 can 
be summed according to formula (9) to obtain the 
relative amplification ratio, ep, /ep,- One can 
show that 


° 
2 
x 
ae 
-2 
1000 1200 1400 1600 1800 2000 2200 2400 
R5. 
FIGURE 3. The amplification rate A, along the trajec- 


tory 4 = 0.00133 Ray. 


q 
Wi 
to 


where Rg, and Roy) are the values of Rs, at the 
locations of the disturbance at the times, t, and 
t,, respectively. The value of cy, the group ve- 
locity, along the trajectory a = 0.00133 Rs,, was 
computed to be about 0.356. We neglected the 0(A2) 
modification of c_ due to the imposed oscillations. 
This modification of c, is 0(10 °) and thus does 
not affect the daiyareal , J, in a substantial way. 
From the results shown in Figure 3, one obtains (by 
a trapezoidal rule integration), the values of J =x 
11,8.7, and 6.3 for w,/Q = 0,2, and 3 respectively. 
The integral I(t) of (10) is evaluated by certain 
sums and products of the vector components of the 
solutions, (35) and (37). We omit the details. The 
resulting expression for I(t), to second order in A, 
has the following form 


I(t) = A] +A ,A2+A (By cosNt+Bysinxt) 
(40) 
+2 (Cy cos2Nt+Cysin2xNt) 


where A)>0,A>2,B,,Bo,C, and Cp are real numbers that 
depend on the Reynolds number, Rg,- These coeffi- 
cients have been computed along the disturbance 
trajectory, a = 0.00133 Rs, and are plotted in 
Figure 4. By using the values of Aj,B,, Bg, C, and 
Co from Figure 4 (A, = 1.0 by suitable normalization) 


in Eq. (40) for the value of A = 0.1 one finds that 
I(t) 
O.5 <=> & 
5 T(t.) 2 


at all the values of Reynolds number, Rye for which 
the disturbance grows. It is customary to assess 
the overall growth of a disturbance by considering 
the natural logarithm of the amplification ratio, 


ep/eq,- From (9) we have 
e 
Qn a Qn ze) cr ay 
ea oe (Gt) 


and one can see that although the term, &n I(t)/I(t,), 
contributes an oscillatory factor to &n e Jen, (re- 
call that, by following the disturbance down the 
plate, t = Rs,) this contribution is minor relative 
to the maximum value attained by J. Thus it can be 
seen that the major effect of the parallel free 
stream oscillations is to reduce the mean growth 
rate of the unstable disturbances. This effect is 
small for small values of A but can be significantly 
large at such large values of A as A > 0.1. We note 
that typical free stream turbulence rarely has a 
_velocity magnitude as large as 10 percent of the 
mean free stream speed. 

_ Experimental results on the effects of parallel 
free stream oscillation on the instability and 
transition of the flat plate boundary layer are re- 


31 


-10 =e el eee: 
1000 1200 1400 1600 1800 2000 2200 2400 
i, 
FIGURE 4. The coefficients of I(t) along a = 0.00133 


Rear W/2 = 3. 


viewed by Loehrke, Morkovin and Fejer (1975). How- 
ever, we shall not make any comparison with their 
experiments because these were for very low frequency 
oscillations (w_/Q * 10) for which our parallel 

flow instabilitY theory is of doubtful applicability. 
An appropriate analytical instability theory for 
comparison with these experiments is a quasi-steady 
and parallel flow theory [see Obremski and Morkovin 
(1969) ]. 


5. CONCLUDING REMARKS 


Our main result is that the parallel free stream 
oscillations, which manifest themselves in the 
Blasius boundary layer as a Stokes layer, lead to 

a mean stabilization of the flow. This stabiliza- 
tion is very weak except for oscillation amplitudes 
that are at least near 10 percent of the mean free 
stream speed. Precise experimental data on the 
effects of such oscillations on Blasius boundary 
layer instability is not available in the frequency 
range considered in this work. However, the results 
are in accord with transition data for oscillatory 
pipe flows. Sarpkaya (1966) has shown experimen- 
tally that transition is delayed substantially when 
harmonic axial oscillations are superimposed on 
steady pipe flow. Furthermore, von Kerczek and 
Davis (1975) have shown that the oscillatory Stokes 
layer by itself is very stable, probably at all 
Reynolds numbers, so that one might conjecture that 
if the Stokes layer begins to dominate the boundary 


32 


layer (which occurs for low frequencies 2 and large 
amplitudes A), then the Blasius boundary layer can 
be stabilized by these oscillations. However, the 
Stokes layer stability is very sensitive to extra- 
neous effects such as streamline curvature. For 
instance, experiments show that transition of plane 
Stokes layers occurs at Stokes layer Reynolds 
numbers, Rs, (where R = ARg,/8) on the order 

of 500 [see Li (1954)]. However, if a slight amount 
of streamline curvature exists, as would occur in 
Stokes layers induced on the bottom of a water chan- 
nel supporting free-surface gravity waves [see 
Collins (1963)], the transition Reynolds number is 
reduced to about 160. Thus, the effect on the in- 
stability of the Blasius boundary layer of free 
stream oscillations with a spatial structure such 

as Up + U, cos (kx-wt) can be expected to be different 


from the parallel flow oscillations considered above. 


It is well known that ambient turbulence tends 
to promote laminar to turbulent transition of the 
boundary layer. Thus, if some oscillatory boundary 
layer does in fact properly model certain features 
of the interaction of the ambient turbulence with 
the underlying steady boundary layer then it is to 
be expected that such a oscillatory boundary layer 
is less stable than the underlying steady boundary 
layer. Although the present numerical results show 
only a stabilizing effect for the type of oscilla- 
tion considered, as inferred above there is reason 
to believe that a more complex form of oscillation 
of the boundary layer can be destabilizing. The 
theory of the instability of forced oscillatory 
boundary layers provides an alternative point of 
view from that of Rogler and Reshotko (1974) and 
Mack (1975) on the role of the interaction of free- 
stream disturbances with Tollmien-Schlichting waves. 


ACKNOWLEDGEMENT 


This work was supported by the Naval Sea Systems 
Command. 


REFERENCES 


Ackerberg, R. C., and J. H. Phillips, (1972). The 
Unsteady Laminar Boundary Layer on a Semi- 
Infinite Flat Plate Due to Small Fluctuations in 
the Magnitude of the Free Stream Velocity. J. 
Fluid Mechanics, 51, 137-157. 


Coddington, E. A., and N. Levinson, (1955). The 
Theory of Ordinary Differential Equations, 
McGraw-Hill, New York. 

Collins, J. I. (1963). Inception of Turbulence at 
the Bed under Periodic Gravity Waves. J. Geo- 
physical Res., 68, 6007-6014. 

Craik, A. D. D., (1971). Nonlinear Resonant In- 
stability in Boundary Layers. J. Fluid Mechanics, 
50, PP SI8—413. 

Davis, S. H., (1976). The Stability of Time-Periodic 
Flows. Annual Review of Fluid Mechanics, 8, 
57-74. 

Gaster, M., (1963). A Note on the Relation Between 
Temporarily-Increasing and Spatially-Increasing 
Disturbances in Hydrodynamic Stability. J. Fluid 
Mechanics, 14, 222-224. 


Hayashi, C., (1964). Nonlinear Oscillations in 
Physical Systems, McGraw-Hill Book Co., New 
York. 


Kerczek, C. von, and S. H. Davis, (1974). Linear 
Stability Theory of Oscillatory Stokes Layers, 

J. Fluid Mechanics, 62, 753-773. 

Li, H., (1954). Tech. Mem. 47, Beach Erosion Board, 
U.S. Army Corps of Engineers, Washington, D.C. 

Loehrke, R. I., M. V. Morkovin, and A. A. Fejer 
(1975). Transition in Nonreversing Oscillating 
Boundary Layers, Transactions ASME, J. Fluid 
Engineering, 97, 534-549. 

Mack, L. M. (1975). Linear Stability Theory and 
the Problem of Supersonic Boundary-Layer Trans- 
ition, ATAA J., 13, 278-289. 

Mack, L. M., (1976). A Numerical Study of the 
Temporal Eigenvalue Spectrum of the Blasius 
Boundary Layer, J. Fluid Mechanics, 73, 497-520. 

Obremski, H. J. and M. V. Morkovin, (1969). Appli- 
cation of Quasi-Steady Stability Model to Periodic 
Boundary-Layer Flows, AIAA J., 7, 1298-1301. 

Orszag, S. A., (1971). Accurate Solution of the 
Orr-Sommerfeld Stability Equation, J. Fluid 
Mechanics, 50, 689-703. 

Rogler, H. L. and E. Reshotko, (1975). Disturbances 
in a Boundary Layer Introduced by a Low Intensity 
Array of Vortices, SIAM J. Appl. Math., 28, 431- 


462. 

Rosenhead, L., (1963). Laminar Boundary Layers, 
Oxford. 

Sarpkaya, T., (1966). Experimental Determination 


of the Critical Reynolds Number for Pulsating 
Poiseuille Flow, Trans. ASME, J. Basic Engineer- 
ing, 88, 589-598. 

Yakubovich, V. A. and V. M. Starzhinskii, (1975). 
Linear Differential Equations with Periodic Co- 
efficients, Translated from the Russian by D. 
Lauvish, John Wiley & Sons. 


Heated Boundary Layers 


Eli Reshotko 


Case Western Reserve University 


Cleveland, Ohio 


ABSTRACT 


Heating the walls on which laminar boundary layers 
develop in water can delay their transition to 
turbulent flow and lead to significant drag reduc-— 
tion. This paper describes the work done over the 
last several years at Case Western Reserve Univer- 
sity in examining the bases and consequences of the 
heating phenomenon. Included are theoretical and 
experimental studies of the stability of heated wa- 
ter boundary layers for both uniform and non-uniform 
wall temperature distributions, and experimental 
study of the effect of heating on laminar separa- 
tion and a quantitative assessment of the prospec-— 
tive drag reduction on underwater vehicles. 


1. INTRODUCTION 


It was noted many years ago in experiments at low 
subsonic speeds [Frick and McCullough (1942), 
Liepmann and Fila (1947)] that the transition lo- 
cation of the flat plate boundary layer in air is 
advanced as a result of plate heating. Based on 
this observation it had long been suspected that 
heating would have the opposite effect in water, 
namely that it would delay the onset of transition. 
This is because heating in water reduces the vis- 
cosity near the wall resulting in a fuller, more 
stable velocity profile for a flat plate than the 
Blasius profile. Cooling in water (and heating in 
air) on the other’ hand tends to give an inflected 
velocity profile which is less stable than the 
Blasius profile. 

These suspicions remained untested until con- 
firmed by the analysis of Wazzan, Okamura, and 
Smith (1968, 1970). These results triggered a 
significant activity in the United States to deter- 
mine whether wall heating could realistically be 


33 


used as a technique for drag reduction. A portion 
of this effort was undertaken at Case Western Re- 
serve University (CWRU) under the joint auspices 
of the Office of Naval Research and the General 
Hydrodynamics Research Program of the David W. 
Taylor Naval Ship Research and Development Center. 
The CWRU effort has been both analytical and 
experimental and is ongoing. This paper will re- 
view the results to date of the CWRU activity and 
indicate current and future directions. 


ANALYSIS OF THE STABILITY OF HEATED WATER 
BOUNDARY LAYERS 


The 
ers 


analysis of Wazzan et al. (1968, 1970) consid- 
the stability characteristics of the boundary 
layer to be governed by the disturbance vorticity 
equation including consideration of viscosity vari- 
ations in the basic flow but ignoring temperature 
fluctuations and the coupled viscosity fluctuations. 
The disturbance differential equation consists of 
the fourth-order Orr-Sommerfeld operator augmented 
by some lower order terms and is as follows: 


(Gc) (p"=026) - uN = - Slug?” - 2029" + alg) 


+ 2u'(6"" = 026) 
+ u"(o" + 029) ] (1) 


with boundary conditions 


(2) 


The analysis of Lowell and Reshotko (1974) on the 
other hand is based on the following coupled sixth- 
order system of vorticity and energy disturbance 
equations: 


34 


(U-c) [(p¢)" - a2 (6o)] - U"(pd) + ifr (U-c)2]' 


(od)! |" = 202) 26g)" ] 1 + abe 
p 


+ 2u! 


Nt 


+ 1S 

R 

N 

(f=) ee 

DIl[F  DI[D 

I oe 

nel] 

_ 

D 

eS 


(0d)' | ' + a2 


| 
e 


eae es se 

ahd ie E 5 

) i (3) 
= Wath AA Non |. = 1 WP \o aemtiye 
EE (Ge)) + (po)T] = GREE {tae + okT'] 


Die: 
- sae } (4) 
p 


ar | 
2 i 
N w 
LGR DN 
aa 
3 
\ 
! I 
real 
K 
S} 
DI 
a 
Cy SS CI|R 


with boundary conditions 


o(0) = ¢'(0) = 1 (0) (0) 


(5) 


$ (») o'(e) = t(@) 0) 

In equations (3) and (4) all properties of the 
basic flow are variable. The quantities r, m, and 
K are the density, viscosity, and thermal conduc- 
tivity fluctuation amplitudes respectively and 

the coupling comes about through the viscosity 
fluctuations that are directly related to the tem- 
perature fluctuations. 


Reg “min. crit. 
X 1073 
—O— Lowell & Reshotko 


(1974) 

12 —O-Wazzan, Okamura & 
Smith (1970) 

Ww To. = 60°F 
d 

10 el) @ 


dx 


MINIMUM CRITICAL REYNOLDS NUMBER 


100 200 300 


Ura HA) 


FIGURE 1. Effect of wall temperature on minimum cri- 


tical Reynolds number ,[from Lowell and Reshotko (1974)]. 


The results of these two analyses for the min- 
imum critical Reynolds number with wall heating are 
shown in Fig. 1. The curves are very much alike. 
Furthermore, the neutral stability characteristics 
and the growth rates as calculated in the aforemen- 
tioned analyses are sufficiently close so that there 
is no important quantitative difference between the 
two. The coupling of vorticity and temperature 
fluctuations through the viscosity seems therefore 
to be rather weak. 

As is seen in Figure 1, both sets of calculations 
predict significant boundary layer stabilization 
(increased minimum critical Reynolds number, de- 
creased disturbance amplification rates, etc.) with 
moderate heating, but display a maximum and sub- 
sequent decrease as the wall to free-stream tem- 
perature difference is further increased. The 
significant stabilization indicated for overheats 
of up to 40°C (70°F) prompted a study of the pos- 
sible drag reduction due to heating to see if this 
drag reduction technique was in fact worth pur- 
suing further. 


3. DRAG REDUCTION IN WATER BY HEATING 


It is shown in this section that significant reduc— 
tions of drag are available to water vehicles with 
on-board propulsion system is discharged through 
heating the laminar flow portion of the hull. The 
analysis is as follows [following Reshotko (1977) |: 
For a vehicle with an on-board propulsion system 


te E 
—_—_ 
Poy Ses 


Cm_) 


the friction drag is 
= dx + 
D q c wdx xJ Coy wdx (6) 


where q is the dynamic pressure, cre and Cry are 
respectively the laminar and turbulent friction 
coefficients, wdx is the area element at length x, 
L is the vehicle length, and Xty is the transition 
location. 

The total drag can be written 


D= D,,(D/D,,) (7) 


where D/Dp is the ratio of total to friction drag. 
For an axisymmetric body this ratio is a function 
of the fineness ratio of the configuration. 


Hoerner (1958) suggests that 


3/2 
D/D,, = il) sb 1.5(5) f wr Bao (8) 


The drag power can then be written 


Pr = Duan u,, (D/D,,) (CEA) (9) 


where [Cp A] is the quantity in brackets in equa- 
Editon (5) ie 


The power available for heating is related to 
the thermal efficiency of the power plant as 


follows: 
p s(22on |g. 2 2 (S=S a) q (10) 


where Neff is the effectiveness of transmitting 
the reject heat to the water in the desired manner. 

If one considers heating only the laminar por- 
tion of the hull then the power required to accom- 
plish such heating is 


x 
134 
= ar S 
Be pu ¢ 4 Che wdx (11) 


where c is the specific heat of water, cphe is the 
laminar Stanton number for the heated boundary 
May ermal ATE =) iti 

Applying the available heating power Pa to the 
laminar portion of the flow, (Py = Py) + after some 
simplification yields 


*tr 
oe Cha wax 

L c 

f ce cask f tr ~£Q wdx 

x ite ——— 

qs te eS at 2 2 (12) 
ff ee c wdx Ke || a0 ( = in ff 
° £2 F th e 


The left side is the ratio of overall friction drag 
to the laminar friction drag and is configuration 
dependent. The right side depends on the dimension- 
less ratio CAT/Uay and on the bracketed parameter 
in the denominator related to the amount of reject 
heat that can be transferred to the boundary layer. 
The bracketed parameter in the numerator is a 
Reynolds analogy factor which is configuration de- 
pendent. In order to close the calculation, a 
relation is needed between AT and transition 
Reynolds number Re, which is also dependent on 
configuration. ee 


Example - The Flat Plate 


In order to quantitatively evaluate the prospective 
drag reduction due to heating, it is necessary to 
choose a particular configuration. The flat plate 
is chosen because of its great simplicity and be- 
cause some information on transition with surface 
heating is available. The results should be repre- 
sentative of what can be obtained for slender shapes 
having pressure gradients that are not too large. 
For a flat plate (w = const) 


(13) 


ff x 
L eae 
x c dx = 0.074 ( ) 
fate 
a Reu/a Ree a 
L x 
tr 


and by Reynolds analogy 


35 


(e} 
a £2 5 OD (14) 


hy 2 


Thus for the case of the flat plate, Eq. (12) be- 
comes 


= x me Fe (15) 


1.328 1/2 oe (Dia i 
Re DEN ne) Wee 
x < th 


te 


The left side of equation (15) is the ratio of 
overall friction drag to laminar friction drag for 


a flat plate. 
The variation of transition Reynolds number 


Rex ~ with overheat AT depends on the choice of 
transition criterion. A criterion that has been 
shown to give plausible trends is the e? criterion 
of Smith and Gamberoni (1956) and Van Ingen (1956). 
For low speed flows, these authors correlated tran- 
sition Reynolds number over plates, wings, and 
bodies with the amplitude ratio using linear sta- 
bility theory of the most unstable frequency from 
its neutral point to the transition point. They 
found that the transition Reynolds number Rex,, as 
predicted by assuming an amplification factor of 
e? was seldom in error by more than 20%. Wazzan 
et al. (1970) have calculated and presented such a 
curve for heated flat plates in water a portion of 
which is shown in Figure 2. Although not quite 
shown on the figure, Rey i reaches a maximum value 
of about 260 x 10© at an overheat of about 43°C. 
The most recent data of Barker (1978) taken ina 
constant-diameter pipe are shown on this figure as 
well. Barker obtains a considerable increase of 
transition Reynolds number with heating in the 
entrance flow boundary layers and his data attests 


109 


@ Barker (1978) 


= 
[o) 
es} 


107 


TRANSITION REYNOLDS NUMBER Re x,, 


0 10 20 30 40 


WALL OVERHEAT, AT, °C 


FIGURE 2. Variation of transition Reynolds number for 
a flat plate with uniform wall overheat according to 


an "eo" transition criterion, T, = 60°F. 


36 


to the reasonability of the assumed transition 
schedule with overheat. 

Drag reduction calculations have been performed 
for plate speeds up to 24.4 m/sec (80 fps), for 
plate lengths of 3.05 m (10 ft), 15.24 m (50 ft). 
30.48 m (100 ft), 152.4 m (500 ft), and 304.8 m 
1) 


D) all 
(OOORBE)), endutonmvaluestof l= Gai Jof 2,5,and 


Cren 


Since the product DN o¢¢/ Op might be very close 


Vote 
9. 


to unity, one may view the aforementioned values of 


A ba 
the "efficiency factor" = 1) Nee! as approx- 
F th 


imately corresponding to n 
respectively. 

Results are presented in Figure 3 for the case 
of an efficiency factor of 5 (n, 0.17). Shown 
in Figure 3 are D/D the ENG of the drag with 
heating to that WER OuCanenG the reject heat for 
drag reduction purposes, the corresponding laminar 
fraction of the plate x;,,/L, the wall temperature 
rise of the laminar region, and finally the ratio 
of the computed drag with heating to that for fully 
laminar flow over the entire plate. 

Generally speaking the drag reduction becomes 
noticable as speeds exceed 10 m/sec (20 knots). 
Although the drag ratio is not a strong function 
of length, the overheat in the laminar region in- 


OSs, Osali7/7 ctarcl ()5ilf0) 


th 


1.0 L, m (ft) 


304.8 (1000) 


1.0 


creases quite significantly with vehicle length. 
For ne, 0.17 (Figure 3), drag reduction of about 
60% are atainable for vehicle speeds of 25 m/sec 
(v50 knots) but the vehicle is far from full lami- 
narization. The variation of drag ratio with Nth 
is shown in Figure 4 for selected cases. The lower 
the thermal efficiency, the larger the drag reduc- 
tion and vice-versa. The indication from the cal- 
culations is that full laminarization can be ob- 
tained in a number of cases (Figure 4) but only if 
Nth gets below about 0.03. Since the e” transition 
curve (Figure 2) has a maximum value of Rey,  be- 
low 3 x 108, vehicles with length Reynolds numbers 
above 3 x 108 cannot be completely laminarized. 

For a plate of given length at a prescribed 
speed, the fuel consumption (proportional to D/nty, 
the slope of a line through the origin in Figure 
4) increases as ntp is reduced. But it is far below 
that of the unheated plate. 


Real Configurations 


Real vehicle configurations involve additional fac- 
tors not considered in this flat-plate calculation. 
Favorable pressure gradient, for example, can be 
very effective in delaying transition while regions 
of adverse gradient are otherwise. Non-uniform 
longitudinal heating distributions can result in 
a more optimal use of the available heat. Effects 


Die 
= (= arent =5 
Dé th 


0.8 152.4 (500) a 08 ny ~ 0.17 L, m (ft) 
So oS 3.05 (10) 
Ww Pad 
O}+ 5 
4 = 0.6 
0.6 = 0 
a 
: 30.5 (100) 9 15.2 (50) 
SI © 
= 
< 15.2 (50) z 4 30.5 (100) 
0:4) a 0 
g 3.05 (10) 2 
fo = 
- s 
0.2 0.2 
50 kts 10 20 30 40 50 kts 
0 0 
0 5 10 15 20 25 0 5 10 15 20 25 
U_, m/sec U_, m/sec 
an, L, m (ft) 
(o) 
Bo 100 
oe L,m (ft) 
ES 
wa 50 
z z c 
Pac) <q 20 30.5 (100) 
co 2 
wi wu = 10 15.2 (50) 
a < 
re a 8 3.05 (10) 
35 
< 
34 
1 
0 5 10 15 20 25 0 5 10 15 20 25 
U_,, m/sec U__,, m/sec 
FIGURE 3. Drag reduction by use of reject heat of propulsion system for transition 
delay. D 
= t = 5, (n m WoIL7)) o 
[= Nth ) nese] U th 


1.0 Ue L 


m/sec (FT/SEC) mM (FT) 


15.2 (50) 15,2 (50) 
15.2 (50) 3.05 (10) 


24.4 (80) 15.2 (50) 
24.4 (80) 3.05 (10) 


st = 0 


D 
D 


Oo COMPLETE LAMINARIZATION 


Drac Ratio, 


0 cal 2 oe} 4 
PoweR PLanT THERMAL EFFICIENCY, 1. 


TH 


FIGURE 4. Effect of thermal efficiency of propulsive 
power plant on drag reduction. 


of surface roughness on transition are possibly 
more pronounced for heated surfaces than for un- 
heated. These factors are presently being studied 
both experimentally and analytically by a number 
of investigators for the purpose of obtaining an 
objective evaluation of the practical capabilities 
of this relatively simple and readily available 
means of drag reduction. The related experimental 
investigations done at CWRU will be described in 
the next two sections. 


4. STABILITY EXPERIMENTS IN WATER 


The first experimental study of flat plate boundary- 
layer stability in air was by Schubauer and Skram- 
stad (1948) who used hot wire anemometry to measure 
the growth characteristics of sinusoidal velocity 
disturbances introduced into the boundary layer by 
a vibrating ribbon. Ross et al.(1970) repeated the 
Schubauer and Skramstad experiment to obtain data 
for comparison with improved numerical solutions 
of the Orr-Sommerfeld equation. Similar stability 
experiments have been performed in water by Wortmann 
(1955) and Nice (1973). The results of these ex- 
periments are in agreement with the numerical solu- 
tions of the Orr-Sommerfeld equation except near 
the minimum critical Reynolds number, where the de- 
parture from parallel-flow theory seemingly results 
from the breakdown of the parallel flow assumption. 
Among the attempts to correct the parallel-flow 
formulation, those of Bouthier (1972, 1973) and 
Saric and Nayfeh (1975, 1977) using the method of 
multiple scales yield numerical results which dis- 
play the best agreement with experimental results. 
A natural extension of the above work is in the 
investigation of factors which can increase bound- 
ary layer stability. As indicated earlier, one of 
these factors is wall heating in water. The ob- 
jective of the experimental work done at CWRU was 
to see if the predicted increase in stability due 
to heating is in fact realized. To this end the 
stability of flat plate boundary layer was investi- 
gated on both a heated and unheated plate. For 
the heated plate, the case of uniform wall temper- 


37 


ature may be more interesting from an engineering 
viewpoint. For example, since the portion of the 
plate upstream of the minimum critical point of 
the unheated plate is stable without heating, why 
not begin heating at the minimum critical point 
and use more advantageously, the power that would 
have gone to heating the leading edge region? 

To systematize the approach to the problem, two 
types of nonuniform wall temperature distributions 
were studied: step changes in wall temperature 
of magnitude AT occuring at a location xg; and 
power law wall temperature distributions of the 
form T,,(x)-T,. = AxM for n both positive and nega- 
tive. The temperature T. is that of the external 
stream. In order to isolate the effect of the 
parameters, n and xg, on the boundary layer sta- 
bility, one of two quantities must be held fixed - 
either the total heating power put into the plate, 
Qtotal: or the local wall temperature difference 
at some reference location 1. (Xree) To: Since 
heat losses from the test plate used in this ex- 
periment could not be accurately measured, the 
total heating power put into the plate could not 
be related to the total convective heat transfer 
to the boundary layer. Therefore the wall tem- 
perature difference at Xyor¢, Ty (Xref) ~To, was held 
constant as n and x were varied, with Xref chosen 
in the region in which stability measurements were 
performed. 


Experiment 


The experiment was performed in a low turbulence 
water tunnel which has a test section 15.5 in. long, 
9 in. wide, and 6 in. high. The free stream tur- 
bulence intensity in the test section is 0.1 - 0.2% 
for free stream velocities ug Lellpte Sec 

The flat aluminum test plate, which is 13.6 in. 
long, 9 in. wide, and 0.625 in. thick is suspended 
from a frame which fits the top of the test section 
as shown in Figure 5. The origin of the coordinate 
system is located at the leading edge. The x- 
coordinate is the running length measured in the 
streamwise direction, y is measured normal to the 
surface, and z is the spanwise coordinate measured 
from the plate centerline. The rounded leading edge 
(1/32 inch radius) is located 0.425 in. below the 
top of the test section, thus forming a slot which 
spans the top of the test section. The turbulent 
wall boundary layer of the water tunnel is removed 
by suction through this slot. Suction is adjusted 
so as to locate the flow stagnation point at a 
stable position just downstream of the leading edge 
on the test side of the plate. A laminar boundary 
layer then develops along the plate starting from 
the stagnation point location. 

Plate heating is provided by eleven electric 
heating elements positioned as shown in Figure 5. 
Plate surface temperature is monitored by eleven 
thermistors imbedded in the surface of the plate 
along the centerline. However, because of the 
large temperature gradients which occur in the 
plate, the thermistors do not yield an accurate 
indication of the plate surface temperature. The 
surface temperature is determined from boundary 
layer temperature profiles measured with a hot- 
film anemometer operating as a resistance thermom- 
eter: 

The pressure distribution on the plate surface 
in both the spanwise and streamwise direction is 


38 


test section 


plexiglas plate 
mounting frame 


X (INCHES) 


monitored using static pressure taps in conjunction 
with a manometer board. Artificial velocity dis- 
turbances are introduced into the boundary layer 
with a phosphorbronze ribbon 0.001 in. thick and 
0.125 in. wide which is stretched across the plate 
surface 3.75 inches. behind the leading edge. Ribbon 
vibration is achieved by passing a sinusoidal cur- 
rent through the ribbon in the z-direction in the 
presence of a magnetic field maintained by horseshoe 
magnets located on top of the plate. 

A traversing mechanism located in the water 
tunnel diffuser downstream of the test section is 
used to position hot-film anemometer probes in the 
x and y direction for boundary layer profile 
measurements. The z-position of the probes is 
fixed at the plate centerline. 

Temperature measurements in the thermal boundary 
layer are made with a DISA 55D0O1 anemometer and a 
55F19 hot-film boundary layer probe operated in the 
constant current mode as a resistance thermometer. 
This unit is calibrated against the free stream 
temperature measured by thermistors extending in- 
to the free stream through the side walls of the 
test section. Boundary layer velocity measurements 
are made with a DISA hot-film system consisting of 
two 55F19 probes, a 11M01 constant temperature 
anemometer equipped with a 55M14 temperature com- 
pensated bridge, a linearizer, r.m.s. voltmeter, 
and d.c. voltmeter. The system is calibrated 
against the velocity measured by a pitot-static 
tube located in the center of the test section. A 
General Radio 1900-A wave analyzer is used to 
measure the r.m.s. amplitude of the anemometer 
signal resulting from ribbon-generated disturbances 
in the boundary layer. 

The mean velocity profile is measured at x = 5.5 
inches, which is the center of the region in which 
disturbance growth rates are measured. This posi- 
tion is also the value of x;,¢, the point at which 
the local wall temperature is held constant as the 
temperature distribution parameters n and x, are 
varied. The displacement thickness, 6*, is deter- 
mined by plotting the mean profile and using a polar 
planimeter to graphically perform the integration 


— o 
6* = Vx of (1 - 4) an, where n = yVu/vx 
Ue Ye 
Since the maximum wall temperature difference used 


in the present work,is T-T,.. = 8°F, the error in- 


upstream Flange 
aluminum suction 
transition piece ———\ \\ 
\ 
\\ 
ribbon-drive \\ 
) 
0. 


FIGURE 5. Test plate installation. 


curred by using the incompressible formulas given 
here to calculate 6* and n is only about 0.1%. All 
experimental results reported below are therefore 
based on the incompressibte forms of 6* and n. The 
Reynolds number, R,, = u 6*/v, is formed using the 
kinematic viscosity evaluated at the free stream 
temperature. 

For a fixed Reynolds number and ribbon frequency, 
the ribbon-generated disturbance amplitude is 
measured at five stations spaced 0.25 in. apart 
between x = 5 inches and x = 6 inches. In this 
region the pressure gradient is small (Falkner-Skan 
8 < 0.02) and there is no interaction between the 
ribbon-generated disturbance and the natural dis- 
turbances present in the boundary layer. The dis- 
turbance amplitude recorded at each station is the 
peak amplitude, defined as A(x) = [u'(n,x)/Velmax: 
found by searching through the boundary layer in 
the y-direction. The spatial disturbance growth 
rate is then calculated from the slope (aA/ax) | 
of a polynomial-curve fit of the A(x) data. By 
repeating the above process for several different 
frequencies the growth rate vs. frequency charac— 
teristics of the boundary layer are determined for 
a fixed Reynolds number and temperature distribution. 

All stability measurements reported here for 
non-uniform wall temperature distributions were 
performed near R,, = 800. At Reynolds numbers 
higher than 800 the ribbon-generated disturbances 
become more difficult to follow since background 
noise levels in the boundary layer increase with 
Reynolds number. At Reynolds numbers lower than 
800 the disturbance growth rates are already small 
for uniform wall temperature in the range 3°F Ss Hh 
(x)-To £ 8°F, and measurement of the decreased 
growth rates resulting from non-uniform wall tem- 
perature distribution is subject to large relative 
errors. 


xD 


Results and Discussion 
Uniform Wall Temperature Distributions 


The Mean Flow - A comparison between heated and 
unheated mean velocity profiles measured under 
identical flow conditions is shown in Figure 6 
together with the calculated unheated profile ob- 
tained using Lowell's (1974) program for 8B = -0.0036, 


which is the measured 8 for the case shown. For 

n > 6 the measured velocity is uniform to within 1%. 
The unheated boundary layer thickness for this case 
is 6 = 0.066 inches (n=6.3). Note that velocities 
measured in the region n< 0.75 are consistently 
higher than would be expected from the straight-line 
nature of the velocity profile in this region. 

These velocities may be subject to wall interference 
effects due to the size of the hot-film probe rel- 
ative to the boundary layer. At the last measured 
point, n=0.5, the prongs of the hot-film probe 
touch the wall. The probe prong diameter is 0.010 
inches (n=0.95 in the present case), while the 
sensing element diameter is 0.003 inches (n=0.29). 
The discrepancy shown in Figure 6 between measured 
and calculated profiles for Ty-To = 0 may be due to 
the integrated effect of the upstream pressure 
distribution on the measured profile. Note that 
the difference between the heated and unheated 
velocity profiles is within experimental error. 

The heated profile is slightly fuller than the 
unheated profile in agreement with Lowell's numer- 
ical solutions of the variable fluid property 
boundary layer equations. The calculated ratio of 
OP eattea nunheated for this case is 0.968 while 

the measured ratio is 0.967. 

Mean temperature profiles measured at varying 
values of T -T_ and R,, are compared to Lowell's 
(1974) solution of the boundary layer energy equa- 
tion in Figure 6. Note that the thermal boundary 
layer thickness is smaller than the velocity bound- 
ary layer thickness by approximately the ratio 
Sp/5y = pr-1/3 = 0.54, where the Prandtl number of 
water is taken as 6.3 at T, = 75°F. Further de- 
tails concerning the mean flow field may be found 
in Strazisar (1975). 

The Disturbance Flow Field - While the CWRU 
Water Tunnel has a relatively low turbulence level 
of 0.1% to 0.2%, this is still much higher than 
Ross et al. (1970) in air. It has nevertheless 
been ascertained by Strazisar (1975) that the pres- 
ent ribbon-generated disturbances do not interact 
with disturbances of other frequencies present in 
the tunnel turbulence and furthermore display the 
linearity required in order that the disturbances 
be considered "infinitesimal". 

The development of ribbon-generated disturbances 
just downstream of the ribbon is investigated to 
insure that the disturbances develop fully before 


—— Lowell's solution 


Tw-Ta =0°F 6 =-.0036 


—— Lowell's solution 
Tw-Tea =5F f=0 


39 


reaching the station where growth rates are first 
measured, namely x = 5 inches. Figure 7 shows 

the results for a decaying disturbance with 

Wo = We) SOTO, inj. = Gol Avse = SL dmenas, mie 
dimensionless frequency Wy is defined wy, = (27f) 
v/u,* where £ is the ribbon frequency. (The exper- 
imental lower branch neutral point at Rgx = 601 is 
at wy, = 150 x 1o-®.) points in the region n < 0.75 
are shown as broken symbols due to possible inter- 
ference effects because of probe proximity to the 
wall. The disturbance amplitude distribution 
through the boundary layer attains its final shape 
at x = 4.5 inches but the peak amplitude rises be- 
tween x = 4.0 and x = 4.5 inches. Downstream of 

x = 4.5 inches both the shape and peak amplitude 

of the disturbance display expected behavior as 
seen by comparison with the calculated eigenfunction 
for this frequency and Reynolds number obtained 
using Lowell's (1974) program. Since the measured 
wavelength of this disturbance is 0.66 inches the 
appropriate disturbance eigenfunction is seemingly 
established in less than 1-1/2 wave lengths. 

A measured disturbance temperature amplitude 
distribution is compared with the corresponding 
numerical solution in Figure 7. The calculated 
distribution is scaled by equalizing the area 
under the measured and calculated distributions 
in the region 0.75 <n <3. The shape of the 
disturbance temperature amplitude distribution is 
also found to be virtually independent of the 
disturbance frequency at a fixed R5x. 


Disturbance Growth Rates - Measured disturbance 
growth rates as a function of frequency for uni- 
form wall temperature are shown in Figure 8 for 
Rg*x = 800. The dimensionless spatial growth rate 


where A is the amplitude of the disturbance at the 
particular frequency under consideration. The 
solid lines in Figure 8 are curves faired through 
the measured points. The curve through the cir- 
cular symbols is for the unheated plate. It is 
evident that with increased heating of the plate, 
the growth rates progressively decrease and the 
range of disturbance frequencies receiving ampli- 
fication is diminished. Similar behavior is in- 
dicated at other Reynolds numbers as well. 


O Ty-Tao=O'F Reg» =940 DO Tyw—Teo =3.5°F R5+=863 
4 Ty-Ta =7.8°F Rs+=909 O Ty-Ta =5.4°F Rs«=910 
& Tyla =7.8°F Rs+=909 
1.0 
0.8 
0.6 
0.4 
0.2 
0) 
0 1 2 3 : 
FIGURE 6. Mean velocity and temperature pro- 
n 7 files for uniform wall heating. 


40 


(%) 


u 
Ue 


DISTURBANCE VELOCITY AMPLITUDE, 


Ribbon at x = 3.75 inches —— __ Lowell solution Rs + = 890 
Two Tee O°F A = 0.656 inches Tyy—Teo = SF w, X 108 = 81 
w, X 10° = 138 © Data for Ty—Ta. = 5.2°F 
© x=4.0in. w, X 108 = 83 Rs. = 890 
r 6 
O x=4.5in. 
a x=5.0in. 
© x=55in. 
——--—Solution by Lowell, Rs » = 601 
(x = 5.5 in.) = 
f 33.0 
+} I 
3 
he 
no 
w 
te 
< 
ao 
c 
E 2.0 
a ° 
w 
a 
=) 
= 
< 
oc 
Ww 
a 
ri 
F 1.0 
w 
S) 
Ww 
a 
=) 
= 
=) 
a 
= 
< 
0 
(0) 1 2, 3} 


180° PHASE SHIFT 


FIGURE 7. Velocity and temperature fluctua- 
tion amplitudes in the boundary layer. 


Neutral Stability - For reference, the neutral 


stability results for the unheated plate will be 
presented first. Neutral points obtained in the 
present experiment are plotted together with those 
from prior investigations in Figure 9. The solid 
line in Figure 9 is the non-parallel flow solution 
of Saric and Nayfeh (1975) while the dashed line 

is the corresponding parallel flow solution of 

the Orr-Sommerfeld equation. Lower branch neutral 
points in the region Rgx < 500, Wy < 210 x 107© are 


T.. = 75 F 

Ug = 4.4 ft/sec 
x = 5.5 inches 
Uy Veg 
OF 
3.14 
4.97 
8.87 


©Oavo 


DISTURBANCE 
FREQUENCY, 
w, X 108 


-4 


SPATIAL DISTURBANCE GROWTH RATE, - ~ X 10° 


-6 
FIGURE 8. Measured disturbance growth characteristics 


for uniform wall temperature distributions, Re xy = 800. 
( 
, 


denoted by bars in the present work because dis- 
tinct neutral points could not be identified. Ex- 
perimental results indicate that a neutral point 
lies somewhere in the barred region at each dis- 
turbance frequency considered. The present re- 
sults are in agreement with the experimental 
results of Ross et al. (1970), Schubauer and 


420 Theory 
@ Non-parallel flow solution (Ref. 7) 
p b --—--— Parallel flow solution (Ref. 12) 
360 Data 
o OC) Present investigation 


OQ Schubauer and Skramstad (Ref. 1) 
4 Ross etal. (Ref. 2) 
Wortmann (Ref. 3) 


300 


240 


180 


120 


DIMENSIONLESS FREQUENCY, w, X 10° 


60 


0 400 600 800 1000 1200 
DISPLACEMENT THICKNESS REYNOLDS NUMBER, Res « 


FIGURE 9. Neutral stability results for the unheated 
plate. (Solid symbols denote lower branch points, 
open symbols denote upper branch points). 


41 


4207 
360+ 
© 
cS) 
x 3004 
3 
© > 
(=) 
= (S) 
x. 2404 pa CI 
= 2 
> g 
9 = 180} 
2 1807 - 
3 3 
wy < 120} 
{120} = 
8 P 
2 ® 
< 60} fa) 60+ 
c 
= 
22) 0 0 + + + + 
iS) 0400 ~~ 600 800 0 400 600 800 #1000 


1000 1200 


DISTANCE DOWN THE PLATE, Rs+ = Vx 


a) Theory b) Experiment 


Skramstad (1948), and Wortmann (1955). and provide 
further verification of the departure from parallel- 
flow solutions in the region R,,< 500. The agree- 
ment obtained allows one to proceed to the case of 
the heated plate with some credibility. 

Experimentally determined neutral curves in the 
(Wy, Rg*) plane for nominal uniform wall tempera- 
ture differences of ky = 0,5,8°F are compared 
to the parallel flow results of Lowell (1974) in 
Figure 10. The experimental results are curves 
faired through the measured neutral points, which 
have not been shown for the sake of clarity. Com- 
parison between the calculated parallel-flow re- 
sults and experiment indicates that the departures 
between the two found near (Rg) min.crit for the 
unheated case persist in the heated cases. It is 
readily seen that with increases in Damo 
(RG t3) aden. Grete increases and also the range of 
frequencies receiving amplification decreases. 
Note that while the theoretical neutral curves 
according to Lowell's parallel flow calculation 
nest within each other, this does not happen ex- 
perimentally until Rsx exceeds 860. 

Predicted and measured values of (Rg*) min.crit 
are compared in Figure 11. The measured rate of 
increase in (R§*)min.crit Compares favorably with 
that predicted by Lowell (1974) and by Wazzan et al. 
(1970). Over the range of values of Ty-T, covered 
by the present work it is conjectured that the 


= 
fo) 
fo) 
oO 


0 2 4 6 8 
WALL TEMPERATURE DIFFERENCE, T\,-T.. (CF) 


S 


MINIMUM CRITICAL REYNOLDS NUMBER, 


(Re.+) MIN. CRIT. 


FIGURE 11. Effect of heating on minimum critical 
Reynolds number. 


DISTANCE DOWN THE PLATE, Rs «+ a/ xX 


1200 


FIGURE 10. 
for uniform wall temperature. 


Neutral stability characteristics 
ap. S 7525 


non-parallel flow nature of the boundary layer 
serves to reduce the value of (Rg*)min.crit by 
about 120 units from that predicted for parallel 
flow. This reduction seems independent of the 
level of wall heating. A more complete description 
of these results ig given in Strazisar, Reshotko, 
and Prahl (1977). 


Non-Uniform Wall Temperature Distributions 


As indicated earlier, the two types of non-uniform 
wall temperature distributions studied are a) the 
power-law type in which (Ty-T.) = Ax™ and b) step 
changes in wall temperature of magnitude AT = a 
T.. occuring at location x,. In the discussions 
that follow, n is the exponent of the power-law 
wall temperature distribution and s = x,/Xyef is 
the fraction of the distance to the measuring 
station (x;es = 5.5 inches) at which the step 
change in wall temperature is located. 

The Mean Flow - Mean velocity profiles for 
varying values of n, s and Tw(Xref)-To are compared 
to the Blasius profile in Figure 12. The discrep- 


x = 5.5 inches 
Ue = 4.65 ft/sec 
Te, = 75k 


U/U, 


Blasius 


FIGURE 12. Mean velocity profiles for varying wall 
temperature distributions. 


42 


ancy between the unheated profile and the Blasius due to equipment limitations. The thermal boundary 
solution may be due to small pressure gradient layer near the leading edge is too thin to make 
effects. temperature profile measurements with the hot film 


practical. The first indication of the wall tem— 
perature is thus provided by the thermistor imbedded 
in the plate surface at x = 1.2". The heater 


Mean temperature profiles and wall temperature 
distributions measured for values of Ty (Xye¢)-To 
=5°F are compared to relevant solutions of the 
boundary layer equations in Figure 19. These nearest the leading edge is located at 0.71" <x 
similar solutions were obtained by Runge-Kutta < 0.96". The actual wall temperature thus rises in 
integration of the coupled momentum and energy some unknown manner from Ty-To = 0 at the leading 
equations assuming variable viscosity and thermal edge to a value near the desired local wall tem- 
conductivity. Their development is not shown here. perature at x - 0.71". These limitations are more 
The error bars shown in Figure 13 represent the severe for increasingly negative values of n, which 
require large temperature differences near the 
leading edge, and may be the cause of the discrep- 
ancy between theory and experiment seen in Figure 
13 for the attempted n = -0.5 profile. 


maximum-measurement error. Agreement between the 
measured and predicted profiles is reasonable con- 
sidering the fact that the wall temperature cannot 
be monitored or maintained near the leading edge 


1.0 Ug = 4.7 ft/sec 
x (inches) 
e 2.0 
0.8 B30 
& 4.25 
©) > 
gle 0.6F 
Te 4 
BIE = 
i] i 
22 x! = 
4 
Ir 
\ 
ix 
0.2 eG 
- 2 4 6 
x (inches) 
0 
0 1 2 3 200 400 600 800 
nN 
Rs * 
1.0 Up = 4.53 ft/sec 
x (inches) 
©) Jes} 
0.8 O 4.25 
a 55 
° 7.0 _ 
0.6 a 
x ne 
' 
0.4 ix 
S 
IF 
0.2 
x (inches) 
0 : 
0 1 2 3 200 400 600 800 
n Rs 
1.0 Ug, = 4.79 ft/sec 
x (inches) 
0.8+ 
0.6 a 
Se 8 
IE 
i) 
0.4 is 
ze 
IF 
0.2 
2 4 
x (inches) 
FIGURE 13. Mean temperature profiles for 0 
power law wall temperature distributions, 0 1 2 3 200 400 600 800 


(36) Axn. n Rs« 


1.0 iy (Xe ep al an OnE 
= 07 5a 
US 4.4 ft/sec 


x = 5.5 inches 


0.8 


Theory Experiment: 
(eo) s=0.0 

ia) s= 0.35 
s= 0.68 


0.6 4 

x ee 
uw 

0.4 3 
= 
| 

x 

= 

0.2 E 

0 
0 1 2 3 200 
n 


Temperature profiles measured at x;of¢ = 5-5" for 
several yalues of S = x,/Xref, with AT = 5°F, are 
compared to analytic results in Figure 14. The 
actual wall temperature does not undergo a steep 
step change due to conduction of heat through the 
plate upstream of the first heater used in each 
case. As a result there is not a unique value of 
X,, the step change "location". For purposes of 
comparison solutions were obtained to the constant 
property energy equation assuming that the temper- 
ature profile developed entirely within the linear 
portion of the velocity profile. This is a reason- 
able assumption for the Prandtl number of water. 
Comparison of the measured profiles with these 
approximate step change solutions indicates that 
the best agreement between theory and experiment 
results when x, is taken as the x-location at 
which the wall temperature first begins to rise 
above the free stream temperature. The choice of 
X, is used in all of the results reported herein. 

The agreement between measured and predicted 
temperature profiles shown in Figures 13 and 14 


_—— Unheated 


—N=1.0 


1958\ w, x 108 


SPATIAL DISTURBANCE GROWTH RATE, - = Xx 10° 
° 


FIGURE 15. Measured disturbance growth characteristics 
for power law wall temperature distributions Ty(x) - To 
= Ax”, = 800. 


Ree 


43 


600 800 FIGURE 14. Mean temperature 
profiles at x = 5.5" for step 
Re* changes in wall temperature. 


for Ty(Xref)-To = 5°F is typical of that obtained 
at local wall temperature differences of 3°F and 
8°F as well. 

Disturbance Growth Rates - Disturbance growth 
rate characteristics for vaying values of n ata 
fixed Reynolds number near R§x = 800 with Ty (Xref) 
-T. = 5°, are shown in Figure 15. The unheated 
case is included for reference. The curves shown 
are faired through the measured (a;,W,) points, 
which are not shown for the sake of clarity. For 
n = +1.0 the maximum disturbance growth rate is 
greater than that for n=0 at a given value of Ty 
(Xye¢)-T,,, and the band of amplified disturbance 
frequencies moves to a higher frequency range. 
Similar results are obtained for Tw (Xref) -To = 3°F 
and 5°F,. 

Disturbance growth rates vs. frequency for 
various values of s, with AT = 5°F are shown in 
Figure 16 at a nominal Reynolds number of Rgx = 800. 
The unheated case is included for reference and 
measured points (a;,W,) are once again not plotted 
for the sake of clarity. The case s = 0 corresponds 


-68 


s 


Unheated —<S 


' 
nN 


SPATIAL DISTURBANCE GROWTH RATE, - « X 108 
& ° 


FIGURE 16. Measured growth characteristics for a step 
change increase in wall temperature, Rox = 800. 


44 


x 10° 


(- =) max 


(a) (b) 


FIGURE 17. Maximum growth rates for power law wall 
temperature distributions, Ty(x) - To = Axt, Rox = 800. 


to uniform wall heating beginning at the leading 
edge while the case s = 1 corresponds to a step 
change in temperature occurring at the measuring 
station x = 5.5 inches. The peak disturbance 
growth rate displays a minimum as s increases for 
each value of AT considered here. The band of 
amplified disturbance frequencies also moves toward 
a higher frequency range as s increases. 

Disturbance growth rate behavior as a function 
of wall temperature distribution is summarized in 
Figures 17 and 18, where (-01) max is defined as 
the maximum disturbance growth rate for a given 
value of Ty(Xye¢)-To at fixed values of n is shown 
in Figure 17b. We see that positive exponents can 
result in large disturbance growth rates at low 
wall heating levels. At higher levels of wall 
heating the relative reduction in (-0j)may, between 
any two temperature levels is greatest as n de- 
creases. 

The variation of (-;)max, with s at values of 
AT = 3°F and 5°F is shown in Figure 18. The min- 
imum in (-Gj)max at each wall heating level occurs 
near the minimum critical Reynolds number of the 
unheated boundary layer. The measured value of 
(Oe) annonce for AT = 0 is 400, which corresponds 
to s = 0.25 in Figure 18, while the predicted par- 
allel flow value of (Rg*) min.crit = 520 £0, AT) = 
O corresponds to s = 0.42. 

An attempt was made to use the program of Lowell 
and Reshotko (1974) to solve the parallel-flow 
spatial stability problem for power law, wall 
temperature distributions since the solution scheme 
allows the mean flow solution to be read directly 
into the coefficient matrix of the disturbance 
growth rate at a fixed frequency and Reynolds 


x 108 


(- “max 


0 2 4 6 8 


FIGURE 18. Maximum growth rates for step change in- 
creases in wall temperature. 


number is a minimum for n=O, and increases by 
maximum of 12% for values of n in the range 

- 1/2 £n‘£1. This behavior, which is not con- 
sistent with the experimental results, may be due 
to the fact that significant changes in wall tem- 
perature and therefore in the velocity and temper- 
ature distributions are taking place over one or 
two wave-lengths in violation of the parallel-flow 
assumptions. It is felt that a proper multiple 
scales formulation of the stability problem, which 
takes into account the rather rapid variation of 
wall temperature with x, is required to properly 
assess the present results for power-law and step 
function wall temperature variations. The results 
for non-uniform wall temperature distributions are 
given in more detail in the paper by Strazisar and 
Reshotko (1978). 


5. EFFECT OF WALL HEATING ON SEPARATION 


An underwater vehicle is basically a body of rev- 
olution having generally favorable pressure gra- 
dients forward of the maximum diameter and adverse 
pressure gradients downstream of the maximum 
diameter. If laminar flow can be maintained all 
the way to the adverse pressure gradient region 
then the boundary layer will be very easily 
separated unless measures are taken to delay such 
separation. 

An obvious way to delay separation is by suction. 
This however involves the complexities of suction 
slots, internal ducting and later discharge of 
the flow removed from the vehicle boundary layer. 

A "cleaner" possibility for separation delay if it 
in fact would work is heating. 

Wazzan et al. (1970) showed that heating can 
cause a separating profile to fill out significantly. 
Figure 19 indicates that for a Falkner-Skan B = 
-0.1988, an overheat of 90°F, converts a separating 
profile to one having the shape factor of a Blasius 
boundary layer. This motivated our proposal to 
investigate experimentally the potential effect of 
heating on delay of laminar separation. Subsequent 
calculations by Aroesty and Berger (1975) using an 


Y 


> 


FIGURE 19. Velocity profiles at various wall tempera- 
tures for 8 = -0.1988 [Wazzan et al. (1970)]. 


20 


-01 —04 -0.6 
Ger 


FIGURE 20. Effect of overheat on Falkner-Skan separa- 
tion parameter. 


approximate procedure showed that despite the 
large changes in profile shown in Figure 19, the 
value of 8 at separation did not change very much 
with heating (Figure 20). This was confirmed as 
also shown on Figure 20 by exact calculations of 
Strazisar (1975) using Lowell's (1974) program. 
The question of the length retardation of separa- 
tion on a real configuration nevertheless remained 
an open one. 


Experiment 


This experiment was also performed in the CWRU Low 
Turbulence Water Tunnel described in Section 4 
using a specially designed two-dimensional model 
having an NACA 635-015 profile. The model (Fig- 
ure 21) is designed as part of the upper wall of 
the test section of the water tunnel. The boundary 
layer developing on the upper wall of the nozzle 

is removed through a scoop with the bleed rate 
adjusted so that the stagnation streamline is 
straight and steady. Rod heaters (Figure 22) are 
provided over the length of boundary layer develop- 
ment. The tests were conducted at rather low unit 
Reynolds numbers so as to promote laminar flow 

in the separation region and to minimize the power 
needed for large temperature differences. The 
electric heaters distributed through the plate 
provide wall temperatures of the order of 60°F 


a Pressure tap 
(taps run down centerline of model) 


Suction duct 


Mounting frame 


Model 


FIGURE 21. Model as mounted in water tunnel. 


45 


Experimental model HEATER X_WNCHES 


NACA 639-015 profile 
Maximum thickness = 1.175" 
jh Oo ral 


Maximum output of each 


heater is 600 watts 


electric 
D heater 


1" electric 
4 ~ heater 


ayn 
8 D 


FIGURE 22. Location of rod heaters in ex perimental 
model. 


above the free-stream fluid temperature in the 
region of separation. Wall temperature distribu- 
tions are shown in Figure 23. 

The separation behavior was determined by com- 
binations of the following indicators: 1) indi- 
cation of separation by visual observations of a 
dye stream injected along the surface through static 
pressure holes, 2) location of separation indicated 
by the static pressure distribution along the plate, 
and 3) use of hot film anemometry to measure bound- 
ary layer velocity profiles. 

As with many water flow facilities, results are 
dependent on the state of cleanliness of the 
experimental equipment. In particular, the veloc- 
ity profiles were affected by the condition of 
the airfoil surface and the screens in the settling 
chamber. Even when the screens and airfoil surface 
were relatively clean, there was some scatter in 
the level of the boundary layer shape parameter 
as evidenced by the results for the unheated air- 
foil. The effects of heating on shape factor 
displayed consistent trends that were generally 
independent of facility condition. The experi- 
mental setup, procedures and measurement systems 
are described in detail by Timbo and Prahl (1977). 


HEATER VOLTAGE = 
5. 


8 
70, 
140. 


100 


mb 


FIGURE 23. Surface temperature distributions, To = 70°F. 


46 


Results 


When dye was injected into the boundary layer 
through the pressure taps along the centerline of 
the airfoil, it usually did not all move directly 
upstream on the centerline. Some of the dye moved 
initially spanwise and then upstream. Regardless 
of the path of the dye, its motion was never steady. 
When the airfoil surface was polished and the most 
accessible of the screens in the settling chamber 
cleaned, the most forward upstream position of the 
dye on the unheated airfoil was 4.1 inches (x/L = 
0.45). With heating, the patter of upstream dye 
flow remained indistinguishable from the unheated 
case. Thus for the wall overheats tested, up to 
80°F measured at x = 4.01" (x/L = 0.44 in Figure 
23), there was no separation delay discernable 
using the dye injection method. 

In looking at pressure distributions, separation 
is identified as the point where the experimental 
pressure distribution departs from the theoretical 
recompression distribution on the aft portion of 
the airfoil. The pressure taps will not indicate 
a separated boundary layer unless there is con- 
tinuous separation at the tap's position. Thus 
unless the upstream motion of the dye is very 
steady, which is usually not the case, the position 
of separation as determined by the most upstream 
penetration of dye is consistently farther upstream 
than indicated by pressure distributions. 

The separation point by examination of pressure 
distributions on unheated airfoil occurs at x*4.9" 
(x/L = 9.53). This is close to the location x = 5" 
predicted for separation using the Thwaites method. 
Heating, as reported by Timko and Prahl (1977), 
caused no significant alteration in the pressure 


4.0 


© -9/28 
QO -9/29 
3.8 4A -10/3 
: B = -0,1988 Vy —10/6 
4 Falkner-Skan 
Theoretical a Experimental 
ge Theoretical 


(Wazzan and Gazley, 1977) 


3.4 
3.2 
3.0 
2.8 
2.6 
2.4 


2.2 


2.0 
0 10 20 30 40 50 60 70 80 


Vergclies lz 


FIGURE 24. Variation,of shape factor H with wall heating. 


distributions and so again one cannot point to any 
delay of separation by heating from these data. 
Since the first two indicators showed negligible 
shift of separation with heating, the boundary layer 
velocity profiles were measured in some detail at 
a point upstream of separation with and without 
heating. Figure 24 shows the results for boundary 
layer shape factor at a station 3.88" downstream 
of the leading edge. Heating causes a reduction 
in shape factor from the unheated value. The un- 
heated profiles correspond to -0.17 < 8 < -0.15 
and the slight reduction in shape factor with 
heating is in accordance with expectation from the 
similar solutions of Wazzan and Gazley (1977). 
Despite these shape factor reductions the profiles 
are changing so rapidly with longitudinal distance 
(hence the scatter in Figure 24) that the separation 
location is hardly affected. 
Thus for the amounts of wall heating employed 
in this study the separation point does not move 
noticeably from its unheated position. This ina 
sense confirms the results of Aroesty and Berger 
(1975) .and of Strazisar (1975) (Figure 20) which 
show the theoretical insensitivity of the value 
of 8 at separation to heating. 


6. CONCLUDING REMARKS 


The studies to date reported herein together with 
those of Wazzan et al. (1970, 1977), Barker (1978) 
and others are such as to justify further investi- 
gation of the various elements of the heating 
phenomenon. Among the factors affecting the prac- 
tical application of heating is the combined effect 
of heating and roughness on stability and transition. 
The work of Kosecoff, Ko, and Merkle (1976) suggests 
that the roughness effect is due to the instability 
of the mean profile as distorted by the roughness. 
An alternative view being investigated at CWRU is 
that the roughness introduces disturbances into 

the boundary layer that may subsequently be ampli- 
fied by the Tollmien-Schlichting mechanism. In 

this view the wavelength of the roughness is im- 
portant as well as its height. An experiment has 
been planned that will map out the mean and dis- 
turbance flow-fields in the vicinity of roughness 
elements so that the relevant mechanism can be 
identified. This will provide a fluid mechanic 
characterization of roughness and help in further 
assessment of the effects of roughness on trans- 
ition of heated water boundary layers. With further 
attention given also to heat exchanger design pro- 
pulsion system, and fabrication techniques, there 
are promising prospects for the achievement of 

drag reduction by heating in water. 


ACKNOWLEDGEMENTS 


The author wishes to acknowledge the participation 
of the following colleagues in the effort reported 
in this paper: Dr. J. M. Prahl, Dr. M. Nice, 

Dr. R. L. Lowell, Dr. A. Strazisar, and Mr. M. 
Timko. All of us are grateful for the sponsorship 
of the work by the Fluid Dynamics Program of the 
Office of Naval Research and by the General Hydro- 
dynamics Research Program of the David W. Taylor 
Naval Ship Research and Development Center. 


REFERENCES 


Aroesty, J., and S. A. Berger, (1975). Controlling 
the Separation of Laminar Boundary Layers in 
Water: Heating and Suction. Report R-1789- 
ARPA, RAND Corp. 

Barker, S. J. (1978). Pipe Flow Experiments at 
Large Reynolds Numbers. Proc. 12th Symposium 
on Naval Hydrodynamics. 

Bouthier, M. (1972). Stabilite lineaire des 
ecoulements presques paralleles. J. de Mecani- 
que, 11, No. 4, 599-621. 

Bouthier, M. (1973). Stabilite lineaire des 
ecoulements presques paralleles. II. La couche 
limite de Blasium. J. de Mecanique, 12, 75-95. 

Frick, C. W., Jr., and G. B. McCullough, (1942). 
Tests of a Heated Low Drag Airfoil. NACA ARR. 

Hoerner, S. F. (1958). Fluid Dynamic Drag. 2nd. 
edition. 

Kosecoff, M. A., D. R. S. Ko, and C. L. Merkle, 
(1976). An Analytical Study of the Effect of 
Surface Roughness on the Stability of a Heated 
Water Boundary Layer. Physical Dynamics, Inc. 
Report PDT-76-131. 

Liepmann, H. W., and G. H. Fila, (1947). Investi- 
gations of Effects of Surface Temperature and 
Single Roughness Elements on Boundary Layer 
Transition. NACA Rept. 890. 

Lowell, R. L., and E. Reshotko, (1974). Numerical 
Study of the Stability of a Heated Water Bound- 
ary Layer. Report FTAS/TR-73-95, Case Western 
Reserve University. 

Nice, M. L.,(1973). Experimental Study of the 
Stability of a Heated Water Boundary Layer. 
Ph.D. Dissertation, Case Western Reserve 
University. (Also Nice, M. L., and J. M. Prahl. 
Report FTAS/TR-73-93, Case Western Reserve 
University. 

Reshotko, E. (1977). Drag Reduction in Water by 
Heating. Proceedings. 2nd International Con- 
ference on Drag Reduction. Cambridge, Sept. 
BHRA, Paper E2. 

Ross, J. A., F. H. Barnes, J. G. Burns, and M. A. 
S. Ross, (1970). The Flat Plate Boundary Layer, 
Part 3, Comparison of Theory with Experiments. 
J. Fluid Mech., 43, 819-832. 

Saric, W. S., and A. H. Nayfeh, (1975). Non- 
Parallel Stability of Boundary Layer Flows. 
Physics of Fluids, 18, 945-950. 

Saric, W. S., and A. H. Nayfeh, (1975). Non- 
Parallel Stability of Boundary Layers with Pre- 
sure Gradients and Suction. Paper No. 6, AGARD 
Conference on Laminar-Turbulent Transition, 
AGARD CP-224. 


47 


Schubauer, G. S., and H. K. Skramstad, (1948). 
Laminar Boundary Layer Oscillations and Tran- 
sition on a Flat Plate. NACA Report 909. 

Smith, A. M. O., and N. Gamberoni, (1956). ‘Tran- 
sition, Pressure Gradient and Stability Theory. 
Report EF 26388, Douglas Aircraft Co. 

Strazisar, A. J., (1975). Experimental study of the 
Stability of Heated Laminar Boundary Layers in 
Water. Ph.D. Dissertation, Case Western Reserve 
University. (Also, Strazisar, A. J., J. M. Prahl, 
and E. Reshotko. Report FTAS/TR-75-113, Case 
Western Reserve University, September 1975). 

Strazisar, A., (1975). Private communication. 

Strazisar, A. J., E. Reshotko, and J. M. Prahl, 
(1977). Experimental study of the stability 
of heated laminar boundary layers in water. 

J. Fluid Mech., 83, Pt. 2, 225-247. 

Strazisar, A. and E. Reshotko, (1977). Stability 
of Heated Laminar Boundary Layers in Water. 
AGARD Conference on Laminar-Turbulent Transition, 
AGARD-CP-224, Paper No. 10. 

Strazisar A., and E. Reshotko, (1978). Stability 
of Heated Laminar Boundary Layers in Water with 
Non-Uniform Surface Temperature. Physics of 
TECK, AIk 6 

Timko, M., and J. M. Prahl, (1977). Experimental 
Study of the Effect of Wall Heating on the 
Separation of a Laminar Boundary Layer in Water. 
Report FTAS/TR-77-135, Case Western Reserve 
University. 

Van Ingen, J. L., (1956). A Suggested Semi- 
Empirical Method for the Calculation of the 
Boundary Layer Transition Region. Report VTH-74, 
Dept. of Aero. Eng'g., University of Technology, 
Delft. 

Wazzan, A. R., T. T. Okamura, and A. M. O. Smith, 
(1968). Spatial and Temporal Stability Charts 
for the Falkner-Skan Boundary Layer Profiles. 
Douglas Aircraft Co, Report No. DAC-6708. 

Wazzan, A. R., T. T. Okamura, and A. M. O. Smith, 
(1970). The Stability and Transition of Heated 
and Cooled Incompressible Laminar Boundary Layers. 
Proc each pinteaeieatenrans her COM lat2 arn Gul, 
Elsevier, (ed. U. Grigull and E. Hahne). 

Wazzan, A. R., and C. Gazley, Jr., (1977). The 
Combined Effects of Pressure Gradient and Heat- 
ing on the Stability and Transition of Water 
Boundary Layers. Proceedings 2nd Int. Conf. on 
Drag Reduction, Cambridge, Sept. 1977, Paper E3. 

Wortmann, F. X., (1955). Untersuching instabiler 
Grenzschictschwingingen in einem Wasserkanan 
mit der Tellurmethode. 50 Jahre Grenzschicht- 
forschuug. Friedr. Vieweg and Sogn, Braunsch- 
weig, (ed. H. Gortler and W. Tollmien). 


Discussion 


CARL GAZLEY, Jr. 


Several of us* at Rand and UCLA have made a 
series of computations which serve to illuminate 
some of the experiments described by Professor 
Reshotko. His experiments with non-uniform wall 
temperature distribution indicate the sensitivity 
of the boundary-layer stability to the way the 
surface temperature changes with distance along the 
plate. For the power-law variation, AT = Ax , 
Reshotko's experiments for AT < 8°F appear to 
indicate decreased stability and increased ampli- 
fication rates as the experiment n decreases toward 
zero. Our computations indicate the same trend at 
low temperature differences, but also show a 
reversal at a temperature difference of about 20°F 
with an increasing stability with increasing n 
above this AT. In fact, very large increases occur 
for a AT above 30°. 

Our results were obtained both by exact numer-— 
ical techniques based on the Orr-Sommerfeld equa- 
tion [Wazzan and Gazley (1978)] and by a modifiac- 
tion of the Dunn-Lin approximation [Aroesty et al. 
(1978)]. The results for flat-plate flow in terms 
of the minimum critical Reynolds number based on 
displacement thickness are shown in Figures 1 and 
2 for values of n = 1 and 2 as a function of the 
local temperature difference. The modified Dunn- 
Lin approximation is seen to agree remarkably well 
with the exact computations. More extensive 
results of that approximation are shown in Figure 3 
for values of n ranging from zero to 2. For temp- 
erature differences above about 30°F, the advanta- 
geous effects of an increasing temperature differ- 
ence are seen to be very large. 


DL APPROX 


Res CRIT 


w= MODIFIED DUNN-LINN 
APPROXIMATION 


4 EXACT COMPUTATIONS 


0 10 20 30 40 50 60 
LOCAL aT=T. -T., °F 
Ww e 
FIGURE 1. Variation of critical Reynolds Number with 
local temperature difference. Flat plate with linear 
increase of temperature difference. 


*J. Aroesty, C. Gazley, Jr., G. M. Harpole, 
W. S. King, and A, R. Wazzan 


DL APPROX 


Ry cRIT 


eee MODIFIED DUNN-LIN 
APPROXIMATION 


0 EXACT COMPUTATIONS 


0 10 20 30 40 50 60 
LOCAL AT =T. -7., oF 
WwW e 


FIGURE 2. Variation of critical Reynolds Number with 
local temperature difference. Flat plate with tempera- 
ture difference increasing with the square of distance. 


Roocrit 


0 10 20 30 40 50 60 
LOCAL aT=T -T., °F 
WwW e 


FIGURE 3. Variation of Critical Reynolds Number with 
local temperature difference for several surface- 
temperature distributions. 


REFERENCES 


Wazzan, A. R., and C. Gazley, Jr. (1978). The Com- 
bined Effects of Pressure Gradient and Heating on 
the Stability and Transition of Water Boundary 
Layers. The Rand Corporatton, R-2175-ARPA. 
Aroesty, J., et al. (1978). Simple Relations for 
the Stability of Heated Laminar Boundary Layers in 
Water: Modified Dunn-Lin Method. 


49 


Author’s Reply 


ELI RESHOTKO 


Dr. Gazley and his colleagues have long been 
interested and active in the topic of heated 
boundary layers and his comments on the conse- 
quences of power-law temperature distributions 
are greatly appreciated. 


Let me first restate our experimental results. 


Referring to Figure 17 of the paper, our experi- 
ments for AT < 8°F appear to indicate decreased 
amplification rates as the exponent n decreases 
toward zero and in fact for some range of negative 
values of n, the disturbances become damped. In 
the temperature range AT < 8°F, neither our cal- 
culations (cited in the paper) nor Gazley's give 
any basis for this experimental result. 

Nayfeh and El-Hady (private communication) 
have recently pointed out that water boundary 
layers with non-isothermal walls cannot have 
similar boundary layer solutions because of the 
variable properties of water. They show that if 
one first calculates the non-similar boundary 
layer profiles expected at the measuring station 
of the Strazisar-Reshotko experiments and then 
analyzes the stability of these profiles, the 
resulting growth rates are qualitatively in accord 
with the Strazisar-Reshotko results as shown in 
the figure below supplied to me by Professor 
Nayfeh. Note in the figure that as n decreases, 
the growth rates also decrease, and although the 
calculated maximum growth rates are not negative 
for the non-parallel calculations with n = -0.5, 
they are very close to zero. This trend is oppo- 
site to what was obtained for the stability of 
similar boundary layer mean profiles. 

Nayfeh and El-Hady's calculations do not go 
beyond AT = 8°F. But I believe that they have 
made their point that when studying the stability 
of water boundary layers with power-law or other 
non-isothermal wall temperature distributions, 
one must analyze the stability of the appropriate 
non-similar boundary layer profiles in order to 
obtain even the correct qualitative trends. 
Therefore I believe that the results presented by 
Dr. Gazley in his comment must be reexamined. 


T T Ti T zi teaa 
lO Tw(Xpep) Te = 8 F 

| 2 = 

| RE, x 800 


—_— 


a SS / 
vA x 


We 
U, UNHEATED” \ 


y- n= +1.0 


AMPLIFICATION RATE oO/R*10° 


80 100 {20 140 |60 180 
FREQUENCY F <xI0° 


Effect of power law wall heating on stability of non- 
similar water boundary layer. ---- parallel, non- 
parallel. o = Im (a_ + €a,) where a is the quasi- 
parallel amplificatfon rate and ea) is the non-parallel 
contribution. 


x 
if 
i 
i 
i 
ii 
i 


Ne 


at oe 
i f 
(valle 


i ae 
' 2 
i Rtas) 
i va f 
i f 
1 if 
ovlichie 
i) 
ir i! 
j f 


eer iu 
ary . \ 
i i r 
5 Un 
t etn 
l 
be SE 
, \ 
yee : i f 
ay i 
th M i 
j Wty ' 
fy 
i \ 
i ; 
{ f 
} j 
ti 
x 
: 7 
' 
{ 
» 
in 
F t 
“ 


iy 
Mak Wie 


we bipeen ar 


Session IT 


BOUNDARY LAYER STABILITY 
AND 
TRANSITION 


KARL WIEGHARDT 

Session Chairman 

University of Hamburg 

Hamburg, Federal Republic of Germany 


s, 


Nonparallel Stability of Two-Dimensional 
Heated Boundary Layer Flows 


N. M. El-Hady and A. H. Nayfeh 
Virginia Polytechnic Institute and State 
University, Blacksburg, Virginia 


ABSTRACT 


The method of multiple scales is used to analyze the 
linear-nonparallel stability of two-dimensional 
heated liquid boundary layers. Included in the anal- 
ysis are disturbances due to velocity, pressure, 
temperature, density, and transport properties, as 
well as variations of the liquid properties with 
temperature. An equation is derived for the modula- 
tion of the wave amplitude with streamwise distance. 
Although the analysis is applicable to both uniform 
and nonuniform wall heating, numerical results are 
presented only for the uniform heating case. The 
numerical results are in good agreement with the 
experimental results of Strazisar, Reshotko, and 
Prahl. 


1. INTRODUCTION 


It is generally accepted that the instability of 
small amplitude disturbances in a laminar boundary 
layer is an integral part of the transition process. 
Significant changes in the boundary layer stability 
characteristics can be achieved by utilizing dif- 
ferent factors, such as pressure gradients, suction, 
injection, compliant boundaries, and heating or 
cooling of the boundary layer. 

Surface heating in a liquid boundary layer can 
be utilized to yield a mean velocity profile which 
is more stable than the Blasius profile. The rea- 
son is that heat transfer alters the shape of the 
boundary-layer temperature profile which in turn 
alters the velocity profile through the viscosity- 
temperature dependence. The effect of wall heating 
on the stability of boundary layers in water was 
investigated by Wazzan et al. (1968, 1970). They 
included the variation of the viscosity with tem- 
perature through the thermal boundary layer. They 
obtained a modified Orr-Summerfeld equation. How- 
ever, they did not include temperature fluctuations 
in the disturbance flowfield. Their results show 


53 


that while cooling the wall has a destabilizing ef- 
fect on the flow, moderate heating has a strong 
stabilizing effect. Lowell (1974) reformulated the 
problem by adding fluctuations for the temperature, 
density, and transport properties. The results of 
Lowell did not vary appreciably from those of Wazzan 
et al. (1970). 

The presently available analyses (Wazzan et al. 
and Lowell) for the stability of heated boundary 
layers in water are all parallel flow analyses. 
results of the parallel stability analyses do not 
agree with available experimental results. Strazisar 
et al. (1975, 1977) performed experiments on the 
stability of boundary layers on both unheated and 
uniformly heated flat plates. These experiments 
confirmed the increased stability resulting from 
wall heating in water. Strazisar and Reshotko (1977) 
extended their experiments to cases of nonuniform 
surface heating in the form of power-law temperature 
distributions; that is, Ty(x) - Te = Ax®. Their 
results are given only for a displacement thickness 
Reynolds number R* = 800 and indicate that, for a 
given level of wall heating, cases with n < O have 
the lowest growth rates. Strazisar and Reshotko 
(1977) found that applying Lowell's analysis (1974) 
to the case of power-law temperature distributions 
yielded results that did not agree with the experi- 
mental results. 

In this paper, we use the method of multiple 
scales (1973) to analyze the linear, nonparallel 
stability of two-dimensional boundary layers in 
water on a flat plate, taking into account uniform 
as well as nonuniform wall heating. We include 
disturbances in the temperature, density, and trans- 
port properties of the liquid in addition to dis- 
turbances in the velocities and pressure. However, 
we present numerical results only for the case of 
uniform wall heating and compare our results with 
the experimental data of Strazisar et al. (1975, 
1977). When the variation of the temperature, 
thermodynamic, and transport properties are ne- 
glected, the present solution reduces to those of 


The 


54 


Bouthier (1973), Nayfeh, Saris, and Mook (1974), 
Gaster (1974), and Saric and Nayfeh (1975, 1977). 
The formulation of the problem and method of 
solution is taken in the next section, the solution 
of the first-order problem is given in Section 3, 
the solution of the second-order problem is given 
in Section 4, the mean flow is discussed in Section 
5, and the numerical results and their comparison 


with the experimental data of Strazisar et al. (1975, 


1977) is given in Section 6. 


2. PROBLEM FORMULATION AND METHOD OF SOLUTION 


The present study is concerned with the two- 


dimensional, nonparallel stability of two-dimensional, 


viscous, heat conducting liquid boundary layers to 
small amplitude disturbances. The analysis takes 
into account variations in the fluid properties but 
neglects buoyancy, dissipation, and expansion ener- 
gies. All fluid properties are assumed to be known 
functions of the temperature alone. 

Dimensionless quantities are introduced by using 
a suitable reference length L* and the freestream 
values as reference quantities, where the star 
denotes dimensional quantities. 

To study the linear stability of a mean boundary- 
layer flow, we superpose a small time-dependent dis- 
turbance on each mean flow, thermodynamic, and 
trasport quantity. Thus, we let 


G(x,y,t) = Qo (x,y) + q(x,y,t) (1) 


where Qg(x,y) is a mean steady quantity and q(x,y,t) 
is an unsteady disturbance quantity. Here, q stands 
for the streamwise and transverse velocity compo- 
nents u and v, the temperature T, the pressure p, 
the density p, the specific heat c,, the viscosity 
u, and the thermal conductivity k. Substituting 

Eq. (1) into the Navier-Stokes and energy equations, 
subtracting the mean quantities and linearizing the 
resulting equations in the q's, we obtain the fol- 
lowing disturbance equation: 


or oe ( + ) + om ( + = 
ye v ope We 3s DW ay Pov + PVo) = 0 (2) 


Ci se My 
+ (32 + = @) 


eyo x Lh Ome 
oy R ox P 


avo , 5 2U0 
+ a(x ay FS )|} (4) 


or 0 oT 8T9 oT 
Pol se ap \b) 5 + Uo 9x oy ur Vo al 
c 
oT dTo 
+ (bog& + 9)/uy 22 + | 
@ C) C) 
an x 0 °Y 
femal O oT oO) 
RPr c { ax (xo ose ox ) 
0 
3 oT chy 
an, Go, ws 
qlee tee) ry 
Op ile (ep oo = functions (T) (6) 


Here, Cpg is the liquid specific heat at constant 
pressure, R = pAUaL*/ug is the Reynolds number and 
ies = cheba/KS is the freestream Prandtl number. 
Moreover, 


(Aer Aap GS 2 (e sdk) 


wir wiry 


(1 + 2e), 1 = 5 ule-1) (7) 


where e is the ratio of the second to the first 
viscosity coefficients (e = 0 is the Stokes assump- 
tion). 

The problem is completed by the specification of 
the boundary conditions; they are 


u=v=T=0O0=aty=0 (8) 


I Aue Se 0) AG “Ny ae (9) 

We restrict our analysis to mean flows which are 
slightly nonparallel; that is, the transverse ve- 
locity component is small compared with the stream- 
wise velocity component. This condition demands all 
mean flow variables to be weak functions of the 
streamwise position. 
mathematically by writing the mean flow variables in 
the form 


Up (x1,¥), Vo + EVo(xX1,Y) 


i] 


Ug 


Po = Pg(x1), To= To(x1-yY) 


Po = Po(x1+Y), © = 64 sna) 
0 ONS Day, Po Po 1rY 

Ho = vo(X1-¥), Ko = Ko(*1,Y) (10) 
where x; = ex with € being a small dimensionless 


parameter characterizing the nonparallelism of the 
mean flow. In what follows, we drop the caret from 
Vo- 


These assumptions are expressed 


To determine an approximate solution to Eqs. (2) 
-(10), we use the method of multiple scales [Nayfeh 
(1973)] and seek a first-order expansion for the 
eight dependent disturbance variables u, v, p, T, 
O, C., Hw and kK in the form of a traveling harmonic 
wave; that is, we expand each disturbance flow 
quantity in the form 


q(x ),y,t,) [qj] (x1,y) 
+ €q9(x,,y) + -.-Jexp(i0) (11) 
where 
Oo 5 oe aes 
a ae a9 (x1), re Tw (12) 


For the case of spatial stability, a is the complex 
wavenumber for the quasi-parallel flow problem and 
w is the disturbance frequency which is taken to be 
real. 

Substituting Eqs. (11) and (12) into Eqs. (2)- 
(10), transforming the time and the spatial deriv- 
atives from t and x to 8 and x), and equating the 
coefficients of ©9 and € on both sides, we obtain 
problems describing the q; and qo flow quantities. 
These problems are referred to as the first- and 
second-order problems and they are solved in the 
next two sections. 


3. THE FIRST-ORDER PROBLEM 
Substituting Eqs. (11) and (12) into Eqs. (2)-(10) 
and equating the coefficients of e€9 on both sides, 


we obtain the following 


: a) 
L} (U1, ,V,,P1,T1) = iagl(Pouy + (Ug - ag)! 


3 
+ oy (povi) = 0 (13) 


Lo (uj,,V],P1,T)) 


i duo : aod) dug 3U0 
aay a a 2h (eo. 
R oy ap Jv *O0P1 R dy \aTo dy 
_ 2 ot oun ia avi 
Roy oF 1 mw POO By 
_ 1 ao, Sin OE Ba 
R dT dy dy m O Re = © (2e) 
L3(4,,V),P)1,T}) 
9 Ww 1 2 
= = — + — 
[ teor0(vs =) R voad fr 
_ is ,. 20 i, duo 9U0 
R Y oy “1 in 2 aTg dy ql 


a7 ig PO Bye (15) 


55 


Ly (W,,V),P1,T)) 


=| ipgag( up - —) + ees 
PON 0 a9 Bie el 


0 2 dKQ OT] 
+ Vv, - 
Po 1 RPr co ay oy 
0 
1 a2T] _ 
RPr c 0 oy2 ae ile) 
Po 


wy = Ww = wh = Oo ae Wy = @) (17) 
Chie wala wy = O as y7o (18) 
Equations (13) - (18) constitute an eigenvalue 


problem, which is solved numerically. It is 
convenient to express it as a set of six first-order 
equations by introducing the new variables Zin de- 
fined by 


is _ uy = 
ai = ilo Bl see 0 Z213.= Vie 
oT] 
rah Sig, BG Sho B18. Fe (19) 


Then, Eqs. (13)-(18) can be rewritten in the compact 
form 


az 
Hat 
- = 0 = Np aeo pO 20 
y a 455215 for i ; (20) 
j= 
Bi = 4213 = Big =O at y =0 (21) 


Zllr 213, 215 10) ES 7 ae (22) 


where the a, are the elements of a 6 X 6 variable- 
coefficient matrix. The nineteen nonzero elements 
of this matrix are listed in Appendix I. 

We solve this eigenvalue problem by using SUPORT 
[Scott and Watts (1977)]. To set up the numerical 
problem, we first replace the boundary conditions 
(22) by a new set at y = y where y is a convenient 
location outside the boundary layer. Outside the 
boundary layer, the mean flow is independent of y 
and the coefficients aj; are constants. Hence, the 
general solution of Eqs. (20) can be expressed in 
the form 


6 
Ba S 2, A; ,cjexP Oy) fore A S Up Pros GO (23) 


where the \j are the eigenvalues of the matrix 
[a;;], the Ajj are the corresponding eigenvectors 
and the c+ are arbitrary constants. The real parts 
of three of the Aj are negative, while the real 
parts of the remaining \. are positive. Let us 
order these eignevalues so that the real parts of 
hy,A2, and 3 are negative. Then, the boundary 
condition (22) demands that cy,cs5 and cg are zero. 
To set up this condition for SUPORT, we first solve 
Eqs. (23) for the Cao Lo) and obtain 


6 
o5exp (1,y) = Waa, soe a) = Wpepceoc (22) 


Aa aba} ahat 


56 


where the matrix [b,;,] is the inverse of (A; -]- 
Setting cy = cs = cg = O in Eq. (24) leads to 


= 0 for j = 4,5, and 6 at y = y (25) 


where the bis are the elements of a 3 X 6 constant-— 
coefficient matrix. 

Using Eqs. (25) as the boundary condition at y 
= y and guessing a value for ag, we use SUPORT to 
integrate Eqs. (20) from y = y to y = O and attempt 
to satisfy the boundary conditions (21). If the 
guessed value for ag is the correct eigenvalue, the 
three boundary conditions will be satisfied. In 
general, the guessed value is not the correct value 
and the boundary conditions at the wall are not 
satisfied. A Newton-Raphson procedure is used to 
update the value of ag and the integration is re- 
peated until the wall-boundary conditions are satis- 
fied to within a prescribed accuracy. This leads 
to a value for ag and a further integration of 
Eqs. (20) leads to a solution that can be expressed 
in the form 


23 = A(x1) 6, (x) ,Y) nope SS DBAS oo FO (26) 


where A is still an undetermined function at this 
level of approximation. It is determined by im- 
posing a solvability condition at the next level of 
approximation. 


4. THE SECOND-ORDER PROBLEM 


With the solution of the first-order problem given 
in Eq. (26), the second-order problem becomes 


oa z 
5 = a, .Z,. 
y a aes 
=, 4 FA for i = 1,2,...,6 (27) 
al, (bre x] 
22] = 223 = 225 = 0 at y =0 (28) 


29112231 225 * O as yoy 2 (29) 


where the G. and F. are known functions of the Cie 
ag and the mean flow quantities. They are defined 
in Appendix II. 

Since the homogeneous parts of Eqs. (27)-(29) 
are the same as Eqs. (20)-(22) and since the latter 
have a nontrivial solution, the inhomogeneous Eqs. 
(27)-(29) have a solution if, and only if, a solva- 
bility condition is satisfied. In this case the 
solvability condition demands the inhomogeneities 
to be orthogonal to every solution of the adjoint 
homogeneous problem; that is, 


ic: Be ax nn dete va fay = 0 (30) 
i=1 


where the W;(x,,y) are the solutions of the adjoint 


homogeneous problem corresponding to the eigenvalue 
@g- Thus, they are the solutions of 


ow. 
i 
== + Ena =O) sore al = ilpArooe p@ 31) 
yy ayes j , , ( 


dy 
Joa 
Wo = Wy = We = O at y =0 (32) 


Wo, Wy, We > 0 as yy? (33) 

Substituting for the G; and Er from Appendix II 
into Eq. (30), we obtain the following equation for 
the evolution of the amplitude A: 


AAG) aes 
az ax} = ia, (x,) (34) 
where 
oo co 
6 6 
ia) = - » FW dy > G.W.dy (35) 

j=1 J j=1 a) 

0 0 


The solution of Eq. (34) can be written as 


A = Agexplie a (x1) dx] (36) 


where Ag is a constant of integration. 

To determine a1(x1), we need to evaluate dag/dx, 
and the 90; /0x)- To accomplish this, we differen- 
tiate Eqs. (20)-(22) with respect to x; and obtain 


; — )- : Ge ) 
zs a. 
dy \ dx} sai ij\dx, 


Sear Wl, sere at Se LHD paod® (37) 
1 al 


ae = ae = Fay =O at y=0 (38) 
Uist wis OES c 
dX] ¥ ax] z ax] “ao Ss er (Se) 


The initial conditions for the computational pro- 
cedures are chosen to exclude any multiple of the 
homogeneous solution. The H; are known functions 
of Ci, 4 and the mean flow quantities and their 
derivatives; they are given by 


6 OB 5 
= me) d 
Hy » c, oxy an 
aaah 9 
: 3 
CLs 
Ce Ds 3 8) oe AS Tp Proce (40) 
j=1 ax] 


Using the solvability condition of Eqs. (37)-(39), 
we find that 


6 
COM). = = : 
an »y HW. dy yy GW, dy (41) 


Therefore, to the first approximation 


2.) = BN eWET gs) Cesailes [va + e€a,)dx - iwt] + O(e) (42) 


Y 


where the z, are related to the disturbance variables 
by Eq. (19) and the constant Ap is determined from 
the initial conditions. It is clear from Eq. (42) 
that, in addition to the dependence of the eigen- 
solutions on x, the eigenvalue a is modified by 

€a,- The present solution reduces to those obtained 
by Nayfeh, Saric, and Mook (1974) and Saric and 
Nayfeh (1975) for the case of nonheat conducting 
flows. 


5. THE MEAN FLOW 


For flows whose thermodynamic and transport 
properties are functions of temperature, the 
two-dimensional boundary-layer equations for a 
zero-pressure gradient are 


a 
5 (p*u*) + ma (OA) = 6 (43) 
) du* 2) du* 
*y* + pxky* raves 44 
p*u = ON Se 7 Be Bp ) (44) 
oT oT a oT* 
*u*C* + o*v*ck = 
p*u*c ax* p*v SD dy* ay (K* ay*) (45) 


The temperature dependence of p and w) couples the 
momentum and the energy equations. Note that buoy- 
ancy and viscous dissipation effects are neglected. 
Although the stability analysis is applicable to 
any wall temperature variations, we present stability 
results for the case of constant wall temperature 
for which the flow is self similar. Thus, we intro- 
duce the transformation. 


= =f pdy* (46) 


where R, is the freestream x-Reynolds number defined 
by 


R _ p*U*x* /u* (47) 
x See S) 


Introducing this transformation in Eqs. (43)-(45) 

and solving the continuity equations for v, we trans- 
form the original set of partial-differential equa- 
tions into the following set of ordinary-differential 
equations: 


Q) du du 

an (pu an) a Arie () (48) 
a oT oT 
Ls wes = 49 
on (pk on) + we eS) on (0) (49) 


I) 7/ 


where 


lee 


i) 


nN 
g(n) = 5 Jf puan (50) 
0 


Note that all fluid properties are made dimension- 
less by using their freestream values. 

Equations (48)-(50) are supplemented by the fol- 
lowing boundary conditions: 


u=0, T= aes and g = 0 at n 0 (51) 


el Se Ls Cyayel ave = a AL as nto (52) 


where the subscript w denotes wall values. Equa- 
tions (48)-(52) are numerically integrated by using 
Runge-Kutta and Adams-Moulten integration techniques 
with the liquid thermodynamic and transport prop- 
erties computed at each integration step. All nu- 
merical results presented here are for water; the 
dependence of its thermodynamic and transport prop- 
erties on the temperature is given in Appendix III. 


6. ANALYTICAL RESULTS AND COMPARISON WITH 
EXPERIMENTS 


Although the analysis is applicable to both uniform 
and nonuniform wall heating, results are presented 
only for the case of uniform wall heating for which 
the mean flow is self similar. 

The only available experimental results for the 
stability of uniformly heated boundary-layer flows 
are those of Strazisar et al. (1975, 1977). Using 
a water tunnel, they introduced disturbances by 
vibrating a ribbon and measured the response by 
using a temperature compensated hot-film anemometry. 
They used the r.m.s. of the stream-wise component 
of the disturbance velocity, u, to calculate the 
growth rates. They determined the growth rate as a 
function of frequency at different Reynolds numbers. 

For a parallel mean flow, a, = 0, dp and A are 
constants, and the ¢_ are function of y only. Hence, 
one can unambiguously define the growth rate o of 
the distrubance as the imaginary part of Op; that 
is, 


o = - Im(a9) (53) 


This definition is equivalent to 


Oo = Re @ &nu) = Re (2 Qnv) = 
ax ox 


3 mene 
Rey (ine) eRe (ent) (54) 


On the other hand, for a nonparallel mean flow, aj, 
7 O, A and dg are functions of x, and the Gnyarce 
functions of both x and y. Thus, if one generalizes 
(53) to take into account €a), one obtains 


6 = - Im(d) + €a)) (55) 


which is not equivalent to (54). Moreover, the 
quantity a); and hence o depend on the normalization 
of the Cy because part of the Tn Can be absorbed in 


58 


A and a}. If one generalizes the definition (54) 
and uses (42), one obtains 
) 
o = - Im(a9 + €a)) + eRe(— Lnz_) (56) 
ox n 


Thus, the growth rate in (56) depends on the choice 
of S, because the axial and transverse variations 
of the Ty, are not the same. Since the Z, are func- 
tions of both y and x, one may term a stable flow 
unstable or vice versa. 

Since there are many possible definitions of the 
growth rate in a nonparallel flow, one should be 
careful in comparing analytical and experimental 
results. Saric and Nayfeh (1975, 1977) found that 
the best correlation between the nonparallel theory 
and available experimental data for the Blasius flow 
is obtained if one uses the definition (55). In 
this paper, we compare the definitions (55) and (56) 
evaluated at the value n where ¢; is a maximum. 

Figure 1 shows the variation of the calculated 
disturbance growth rates o/R with frequency FR=W/R 
for Twate = 0, 3,5, and 8°F and for the displacement 
thickness Reynolds number R* = 800. This range of 
Tw-Te is chosen for comparison with the existing 
experimental results. The growth rate is calculated 
by using the definition (55) and by normalizing 7, 
so that ¢)>exp(-agy) as ye~. This figure indicates 
that the disturbance growth rate decreases with in- 
creasing T,-T,. The maximum growth rate is reduced 
by approximately 56% by increasing the wall temper- 
ature by 5°F. The maximum growth rate is very small 
when the wall temperature is increased by 8°F at 
R* = 800. Figure 1 shows that the range of unstable 
frequencies decreases with increasing T,-Te. 


AMPLIFICATION RATE o/R« 10° 


80 100 120 140 160 180 
FREQUENCY FRxI0® 


FIGURE 1. The variation of the spatial growth rate 
with frequency for varying wall temperatures at R* = 
S00) =a = Nonparallel, ----- Parallel. 


MAXIMUM o/Rx10° 


Oo 400 600 800 1000 1200 1400 1600 


FIGURE 2. The variation of the maximum growth rate 
with streamwise position for varying wall temperatures. 
Nonparallel, ----- Parallel. 


Figure 2 shows the variation of the maximum 
growth rate obtained from our analysis with Ty-Te- 
It shows that the maximum growth rate decreases with 
increasing wall temperature at all Reynolds numbers. 

Figures 1 and 2 show a comparison between the 
growth rates based on the parallel, (53), and non- 
parallel, (55), stability theories. The nonparallel 
maximum growth rates are approximately 30% larger 
than the parallel ones. Moreover, the nonparallel 
critical Reynolds number is approximately 20% lower 
than the parallel one for all the values of T,-T 
considered as shown in Figure 2. 

Figures 3a-3d show comparisons of the experi- 
mental growth rates of Strazisar et al. and the 
nonparallel growth rates defined by (53), (55) and 
(56) for different values of T,-T, and different 
values of R*. These figures show good agreement 
between the growth rate defined by (55) and the ex- 
perimental results, in contrast with the parallel 
theory which underpredicts the experimental results 
by large amounts. Moreover, including the distor- 
tion of the eigenfunction with streamwise position 
in the definition of the growth rate yields a growth 
rate that is very close to the parallel one and 
hence underpredicts the experimental results by 
large amounts. 


© 


7. CONCLUSION 


The method of multiple scales is used to analyze 

the linear nonparallel stability of two-dimensional 
liquid boundary layers on a flat plate for the cases 
of uniform and nonuniform wall heatings. We include 
disturbances in the temperature, density, thermo- 
dynamic, and transport properties of the liquid in 
addition to disturbances in the velocities and 
pressure. The growth rates calculated from non- 
parallel results without including the distortion 

of the eigenfunction with streamwise position are 

in good agreement with the experimental results of 
Strazisar et al. (1975, 1977). The nonparallel 
results show that wall heating in water has a sta- 
bilizing effect on the flow; there is a decrease in 
the disturbance growth rates, a decrease in the 
range of unstable frequencies and an increase in 

the critical Reynolds number. 


a/R «10° 


80 100 120 140 


FR «I0° 


FIGURE 3a. Comparison of the analytical and the experi- 


mental spatial growth rates for various displacement 
thickness Reynolds numbers and wall temperatures. 
Experiments, Strazisar et al. (1975, 1977), 1) o = 


NCA pA) GS SINC, se Sei py SY) Kor Se —adin(ay ay Ger)) sp 
e€ 0} 1 
G1] 9x) 


o/Rx10° 


60 80 100 120 


FRx10° 


FIGURE 3b. Comparison of the analytical and the ex- 
perimental spatial growth rates for various displace- 


ment thickness Reynolds numbers and wall temperatures. 


Experiments, Strazisar et al. (1975, 1977), 1) o = 
Im(a49), 2) o = -Im(a9 + €0)), 3) O = -Im(ag + €0}) + 
€ a] 41 


269) ox) 


o/R*« 10° 


wl = 5.4 °F 
R*=910 


80 100 120 140 160 
FRx10° 


FIGURE 3c. Comparison of the analytical and the ex- 

perimental spatial growth rates for various displace- 

ment thickness Reynolds numbers and wall temperatures. 

Experiments, Strazisar et al. (1975, 1977), 1) 5 = 

-Im(ag), 2) Go = -Im(a9 + €01), 3) O = -Im(a9 + €4)) + 
a ailGa 


80 100 120 140 160 
FRx 10° 


FIGURE 3d. Comparison of the analytical and the ex- 
perimental spatial growth rates for various displace- 
ment thickness Reynolds numbers and wall temperatures. 
Experiments, Strazisar et al. (1975, 1977), 1) 5 = 
-Im(ag), 2) o = -Im(ag + €4)), 3) © = -Im(dp + €0]) + 


59 


60 


ACKNOWLEDGMENT 


The authors are indebted to Dr. W. S. Saric for 

many valuable comments. This work was supported 
by the NASA Langley Research Center Under Grant 

No. NSG 1255. 


REFERENCES 


Bouthier, M. (1973). Stabilité linéaire des 
écoulements presque paralléles. J. de Mécanique 
IA, WS 

Gaster, M. (1974). On the effects of boundary-layer 
growth on flow stability. J. Fluid Mech. 66, 
465. 

Lowell, R. S. (1974). Numerical study of the sta- 
bility of a heated water boundary layer, Ph.D. 
dissertation, Case Western Reserve University; 
also, Dept. Fluid, Thermal, and Aerospace Sci., 
Case Western Reserve Univ., Rep. FTAS/TR-73-93. 

Nayfeh, A. H. (1973). Perturbation Methods. Wiley, 
New York, Chap. 6. 

Nayfeh, A. H., W. S. Saric, and D. T. Mook (1974). 
Stability of nonparallel flows. Arch. Mech. 26, 
401. 

Saric, W. S., and A. H. Nayfeh (1975). Nonparallel 
stability of boundary layer flows, Phys. Fluids 
18, 945. 


APPENDIX I 


ai2 =] 
= 1Po0%0R _, Ox 2 
ari 7 (Ug aie! +0, 
= i duo 
a = = — —_ 
ze Uo dy 
= OoR dUo . 
a = 
23 Wo y joo ( 
— 140R 
a — 
24 7 
2 
= _ hao doo 
E29 Po To (Uo 
fog 2 & Eto ous: 
5 Ho dTo oy 
431 = = 10 
1 po 
Ag So =e 
2 Po Oy 


Saric, W. S., and A. H Nayfeh (1977). Nonparallel 
stability of boundary layers with pressure gra- 
dients and suction. AGARD Conference Proceedings 
No. 224, Laminar-Turbulent Transition, Paper 
No. 6. 

Scott, M. R., and H. A. Watts (1977). SUPORT-A 
computer code for two-point boundary value prob- 
lems via orthonormalization. SIAM J. Num. Anal. 
14, 40. 

Strazisar, A. J., and E. Reshotko (1977). Stability 
of heated laminar boundary layers in water with 
non-uniform surface temperature, AGARD Conference 
No. 224, Laminar-Turbulent Transition, Paper No. 
10. 

Strazisar, A. J., J. M. Prahl, and E. Reshotko (1975). 
Experimental study of stability of heated laminar 
boundary layers in water, Dept. Fluid, Thermal 
and Aerospace Sci., Case Western Reserve Univ., 
Rep. FTAS/TR-75-113. 

Strazisar, A. J., E. Reshotko, and J. M. Prahl (1977). 
Experimental study of stability of heated laminar 
boundary layers in water. J. Fluid Mech. 83, 225. 

Wazzan, A. R., T. T. Okamura, and A. M. O. Smith 
(1968). The stability of water flow over heated 
and cooled flat plates. J. Heat. Trans. 90, 109. 

Wazzan, A. R., DT. TT. Okamura, and! Aj Ma1Os eSmictch! 
(1970). The stability and transition of heated 
and cooled imcompressible boundary layers, Pro- 
ceedings, 4th International Heat Transfer Confer- 
ence, ed. Grigall U. and E. Hahne, Amsterdam. 


ai nsn= Iolo (2 duo 90) 
; R Ho Oy Po oy 


_ idovo 


\e dio 9Po 


if (2-p0 2 (200)24 


Ang = See Yc = > 
as R oH) QoHoPo dy dy Oo Po ay? Qo oy 


= 1QoU0 ] dio Uo if doo Oo (Uy _ Ww ) 


ne R uo dTo 8Y ~~ Woo ATo dy Lo 
eg eee 
Bing 2 2 Jaguar Ph (Us - aa 
ase = | 
salves RPr.c 3° ln. 
Ko dy 
ae ae 9090 ‘Un < wy rapa Le 
APPENDIX II 
ei i+ FiA = 0 
g Bot ros = - ms Ie s lk, 
Gs tee HESAV= Li, 
Binge tg een a ae 
Ge a FEA = 0 
Ge a Fell = = orn I, 
where 


ats doo GA _ J 20 
Ie P0G1 + Uo as “| dx, 3X4 en wv 


Xx 


day duo 


ar Ee (to dx, + Qo ep - Po 9x, 10} zt 


1, 2 (FF voce - PoUo)ti + R oy Sg 0 RH 


27 2) 
(SS wrosto re polar 


continued on page 62 


61 


62 


iy, abe (r An eg Soy) e ope (Uo th Vo atest} A 


= pogwelts = poVoge> + (Foc - poUo) + : [h sue = 
(2a) Ble 2 — (Shey + v Ste Shoe (y Blog 5 Boy Ics 
te es ea nae)ss} o> fe ee lapRwe 
e € Po 
(a eo cy Bee ng My - a Be Pou, 
+ Vo lls (RPE oko - 2oUo) a 2 paece 


APPENDIX III 


The variation of the water thermodynamic and transport properties 
with temperature is given by 


_ | (T* = 3,9863)2(T* + 288.9414) 374.3 
ex = | - “sog9n9.2 (T* + 68.12963) + 0.011445 exp(- =>) 


p* iin gm/m’,  T* in °C. 


1.002 ) _ 1.37023(i* = 20) + 8.36 x 1On(T = 20)2 
Tia UWS) SF We 


Log( 


Me Win Cio, Te Alin 2G 


K* = - 9.901090 + 0.1001982T* - 1.873892 x10 “T*? 
+ 1.039570 x 10° 7T*? 


k* in mwatts cm ?K?, T* in °K 


cS = 2c! = OsGsley 2 WO “Te? & 2G WOT Ae 
= 2.42139 x 105 °1** 


Gs in cal Gn K os Te? a Ol 


A discussion of the sources and accuracy of these formulas can be 
found in Lowell (1974). 


Three-Dimensional Effects in 
Boundary Layer Stability 


Leslie M. Mack 


California Institute of Technology 


Pasadena, 


SUMMARY 


Most work in linearized boundary-layer stability 
theory has been carried out either on the basis 

of two-dimensional mean flow and plane wave dis- 
turbances with the wavenumber in the flow direction, 
or, for a more general case, by a transformation 

of the equations to two-dimensional form. This 
procedure can obscure important physical aspects 

of wave propagation in two space dimensions. In 
this paper the stability equations are retained 

in three-dimensional form throughout. A method 

for treating spatially amplifying disturbances with 
a complex group velocity is adopted and applied 
first to oblique waves in a two-dimensional bound- 
ary layer, and then to the two-parameter yawed 
Falkner-Skan boundary layers. One parameter is 

the spanwise to chordwise velocity. For boundary 
layers with small crossflow, the maximum amplifi- 
cation rate with respect to frequency is calculated 
as a function of flow angle for waves whose normal 
is aligned with the flow. Next, the minimum crit- 
ical Reynolds number of zero-frequency crossflow 
instability is obtained for both large and small 
pressure gradients, and finally the instability 
properties of two particular boundary layers with 
crossflow instability are determined for all un- 
stable frequencies. 


1. INTRODUCTION 


Most work in linearized boundary-layer stability 
theory has been restricted to two-dimensional mean 
flows, and, for these flows, even further restricted 
to plane-wave disturbances with the wave normal in 
the flow direction.* The latter restriction is 
normally justified by reference to the theorem of 


*Such a wave is called two-dimensional because 

it has only two disturbance velocity components. 

All other plane waves have three velocity components 
in any coordinate system, and are called three 
dimensional. 


California 


63 


Squire (1933), which states that ina two-dimensional” 
incompressible boundary layer, the minimum critical 
Reynolds number is given by a two-dimensional wave. 
Even though the most unstable wave at a given 
Reynolds number is two dimensional in accordance 
with the theorem, the most unstable wave of a 
particular frequency can well be three dimensional. 
Furthermore, the unstable three-dimensional waves 
can have phase orientation angles (the angle between 
the local freestream direction and the wave normal) 
up to almost 80°. Any method for the estimation 

of transition that is based on stability theory 
must take this large range of unstable three di- 
mensional waves into account. For a supersonic 
two-dimensional boundary layer, even the most un- 
stable plane wave at a given Reynolds number is 
three dimensional. The two-dimensional waves be- 
come of little importance as the Mach number in- 
creases above one until the hypersonic regime is 
reached, where a two-dimensional second-mode wave 
is the most unstable. 

When we turn to three-dimensional boundary layers, 
there are no two-dimensional waves, but the trans- 
formation of Stuart [Gregory et al. (1955) ] reduces 
the three-dimensional temporal stability problem 
to a series of two-dimensional problems. That is, 
the temporal amplification rate can be obtained by 
solving a two-dimensional problem for the boundary- 
layer profile in the direction of the wave normal. 
This approach was carried through numerically by 
Brown (1961) for the rotating disk and a limited 
number of swept-wing boundary layers. When the 
same approach is applied to the spatial theory, 
it leads to complex velocity profiles and loses 
much of its utility except as a computational device. 

Instead of trying to make a two-dimensional 
world out of a three-dimensional world, it might 
as well be accepted that boundary-layer instability 
is inherently three dimensional, even with two- 
dimensional mean flow, and to formulate the insta- 
bility problem directly as three dimensional [Mack, 
(1977); this paper will be referred to as M77]. A 
transformation of the dependent variables reduces 
the order of the incompressible eigenvalue problem 


64 


from sixth to fourth order, but the velocity pro- 
files and wave parameters are not transformed. 
This approach is equally valid for the temporal 
and spatial theories, but for the latter a growth 
direction must be assigned before eigenvalues can 
be computed. In M77 this direction was taken 
equal to the direction of the real part of the 
group velocity and numerical results were obtained 
for two-dimensional incompressible and compressible 
flat-plate boundary layers and for the rotating 
disk boundary layer. 

In the present paper, a theoretical presentation 
is given in Section 2 to justify the use of a 
spatial mode whose direction of growth is determined 
by the complex group velocity. In Section 3, some 
results concerning three-dimensional spatial waves 
in the Blasius boundary layer are given as an 
example. In Section 4, we adapt the family of 
yawed-wedge three-dimensional boundary layers 
[Cooke (1950) ] for use in stability calculations. 
In Section 5, under Boundary Layers with Small 
Crossflow, we consider the effect of the flow angle 
(the angle between the local potential-flow direc- 
tion and the direction of the pressure gradient) on 
the maximum amplification rate for small pressure 
gradients. Next, in Section 5 we take up cross-— 
flow instability and determine the critical Rey- 
nolds number for several combinations of pressure 
gradient and flow angle. We then obtain the max- 
imum amplification rate and instability boundaries 
of all unstable frequencies as a function of the 
wavenumber vector for a favorable pressure-gradient 
boundary layer which is unstable at low Reynolds 
numbers only because of crossflow instability. 
Finally, in the last part of Section 5, we repeat 
the latter calculation for an adverse pressure- 
gradient boundary layer with crossflow instability 
at a Reynolds number where the boundary layer is 
unstable even without crossflow instability. In 
all of the examples, only the amplification rate 
is calculated, and on the basis of locally uniform 
flow. No results concerning wave amplitude are 
given, although in Section 2, we make use of a 
simple wave amplitude equation in order to properly 
define the spatial amplification rate. 


2. THREE DIMENSIONAL STABILITY THEORY 


Formulation and Transformations 


The linearized, incompressible, parallel-flow, 
dimensionless Navier-Stokes equations for the 
elementary modes 


u(x,y,z,t) f(y) 
v(x,y,z,t) a o (y) 
w(x,y,Z,t) h(y) 
p(x,y,zZ,t) tT (y) 


exp[i(ax + Bz - wt) ], (1) 


where u,v,w are the velocity fluctuations and p 
is the pressure fluctuation, can be reduced to 
(M77) 


' 
Z) = Zor 


zZi= [ao + 6. + OR(au Hew w)lZ) 


2 a2! 
+ (aU' + BW')RZ3 + i(a + B )RZy, (2) 


2 2 
d Cmts 
-i2, -|i(au + aw - 0) +2 */z,, 


N 
a= 
i] 


for the determination of the eigenvalues. The 
primes refer to differentiation with respect to 
y, and the dependent variables are 


Z, (y) af(y) + Bhly), Z3(y) = oly), 


Z,(y) = my). 


There are two additional uncoupled equations for 
h(y). In Eqs. (2), a and 8B are the complex wave- 
number components in the x and z directions, w is 
the complex frequency, U and W are the mean velo- 
city components in the x and z directions, and R 
is the Reynolds number UpL*/v*, where the velocity 
scale U* is the potential velocity and L* is a 
suitable length scale. Asterisks refer to dimen- 
sional quantities. The modes in Eq. (1) can be 
termed plane waves in the x,z plane because of the 
phase function, even though there is a modal struc— 
ture in the y direction. 

The boundary conditions are 


Z,(0) =O , 423(0) =0, (3) 


Z, (y) a> '@ 5 Z,(y) +0 as y +o. 
If we choose x to be the direction of the local 
potential flow, then z is the crossflow direction 
and 

WD) ao Ibe Wiy) > 0 asyreo. 
Thus U(y) is the mainflow velocity profile; W(y) is 
the crossflow profile. 

In the temporal stability theory, a and 8 are 
real, and Eqs. (2) can be reduced to two-dimension- 
al form in two different ways. The first transfor- 
mation is 


(4) 


When W = 0, this is the transformation of Squire 
(1933). It relates the eigenvalues of a three- 
dimensional wave of frequency w in a velocity pro- 
file (U,W) at Reynolds number R to the eigenvalues 
of a two-dimensional wave of frequency w/cosy in 

a velocity profile U + W tani at Reynolds number 

R cosy, where 


y = tan’ (8/a) 


is the phase orientation angle. 
The second transformation, 


2 =) 
2 +) 1B) )7 A au = ov + BW, 


MN 
2 


a 


RrSCReo; 


is that of Stuart [Gregory et al. (1955) ]. It 
relates the eigenvalues of a three-dimensional 
wave of frequency w in a velocity profile (U,W) at 
Reynolds number R to the eigenvalues of a two- 
dimensional wave of the same frequency in a veloc- 
ity profile U cosy + W sini at the same Reynolds 


number. The Squire transformation is most useful 
for a two-dimensional boundary layer because the 
velocity profile is unchanged. Thus all eigen- 
values of three-dimensional waves can be obtained 
from known eigenvalues of two-dimensional waves 
with no additional calculations. In a three- 
dimensional boundary layer, the velocity profile 
must change and the Stuart transformation is pre- 
ferred because the frequency can remain fixed at 
a given Reynolds number as the phase orientation 
angle ~ is varied. 


Spatial Stability Theory 


Statement of the Problem 


In the spatial stability theory, a and 8 are com- 
plex and w is real. Neither transformation is of 
much utility except when 


a;/B; = o,/8,- (6) 
When (6) is not satisfied,a is complex, and in the 
Squire transformation both R and w are also com- 
plex as well as U for a three-dimensional boundary 
layer. In the Stuart transformation, U is complex 
for all boundary layers. With complex quantities, 
we might as well deal directly with (2), as these 
equations have already been reduced to fourth order 


and nothing is to be gained from an additional trans- 


formation. There only remains the question, to be 
answered later in this Section, of whether any use 
can be made of the simplification offered by (6). 

It is convenient to define a real wavenumber 
vector 


eS 
k = (a,,B,) 4 
and a real spatial amplification rate vector 
> 
G = (-0,, —By), 
ae R 
in place of the complex vector K - io. The magni- 


tudes of the vectors are k and o, and their di- 
rections are given by the two angles 


v= tan (8, /a,), p= tan” "(B,/a,). 


Equation (6) is now seen to be a statement that 
k and 6 are parallel (J = ~). Plane waves with 
? A) have been termed inhomogeneous by Landau and 
Lifshitz (1960). 
The solution of the eigenvalue problem set up 
by (2) and (3) gives the complex dispersion relation 


o = O(K,0,x;2))- 


Even with w, x and z fixed, there remain four real 
wave parameters: k, W, o and ~. Only two of 
these can be determined in a single eigenvalue 
calculation, e.g., k and o with w and wt) specified. 
The angle i can be considered an independent 
variable on the same basis as the frequency. The 
problem is to choose . What we are looking for 
is a single spatial mode which serves the same 
purpose as a two-dimensional spatial mode in a 
two-dimensional boundary layer, where it represents 
the wave produced by a stationary harmonic source. 
“The amplification rate of this mode is used as a 


65 


measure of the relative instability of different 
velocity profiles, and its amplitude can be applied 
to the transition problem. 


Introduction of an Amplitude Equation 


In order to describe wave propagation in the non- 
uniform medium of the boundary layer, equations are 
needed for the wave amplitude and the change in the 
wavenumber vector in addition to the dispersion 
relation. Even though no amplitude calculations 
are included in this paper, a consideration of the 
amplitude equation will help us select jp. 

In a nonuniform medium the elementary modes (1) 
are not general enough and must be replaced by 


wGesy747e) = AGB) explo (})x]£ (y)expli (a,x 


+B 2 - wt) ]. (7) 


In this, the exponential amplitude factor has been 
written separately in terms of the spatial amplifi- 
cation rate o()). This amplification rate is the 
magnitude of G(K,P,w,X,zZ) considered as a function 
of k,W,w,x,zZ with a fixed value of J. Each j de- 
fines a coordinate 


x = cos x + siny Z 
along which the wave growth is directed. 

Nayfeh et al. (1978) have derived an equation 
for the amplitude factor A(x,z,t) on the basis of 
the multiple scales technique, with A considered 
to be a slowly varying function of x,z,t, as are 
a,8,w and f(y). In a uniform medium, and with A 
independent of time, their equation reduces to 


A 
@ BE a6 dB = ©, (8) 
Be (e) ote Z0z 
where C = (Cx,Cz) is the (complex) group velocity. 


We may note that (8) is also obtained from 


2 > 

a (V.c)A> = 0, (9) 
which is the energy conservation equation of Whit- 
ham's theory (1974). Davey (1972) has applied (9) 
to non-conservative wave motion in a two-dimensional 
mean flow, and refers to the amplitude function A 
as a pseudo amplitude, or the 'dispersive part' of 
the amplitude. 


Spatial Mode - Real Group Velocity 


ss 
We restrict ourselves first to the case of C real 
and define the orthogonal coordinates 


x = cos x + siny Za (10a) 

gr gx gr 

Zee = -siny,, x + cos, Zy (10b) 
where 

Ves tan (E/E) o 


The angle Ugr defines the direction of the charac— 
teristic coordinate xg,, which is identical to a 
group velocity trajectory, and A is constant along 
each characteristic according to (8). 


66 


The amplitude portion of (7) is now 
a(x,z) = A(z expla ()) x], (11) 


and (7) can be interpreted as a certain type of 
solution for a uniform medium when A is variable, 
provided only that A is constant along a character—- 
istic. A knowledge of A along some initial curve 
completely specifies a along the characteristics 
of A, and the characteristics of A are also the 
characteristics of a. Therefore we can write (11) 
as 2 is os 
A(Xgy) = ag (z exp (Og,Xgr) » (12) 


gr gr 


where (10) has been used to eliminate x and 

Sgr = 5 (p) cos (V-Pgy) « (13) 
Consequently, (7) becomes 
u(x,¥,Z,t) = ao (Zg,) exp (OgyXgy) f(y) exp 


[i (4px + Byz - wt)]. (14) 


If an is a constant everywhere, the spatial mode 
(14) represents a physical wave in the entire x,z 
plane that could be produced by a particular 
stationary harmonic line source in a uniform medium. 
ifag ag is constant only along a characteristic, we 
have a form of ray theory, and (14) in turn applies 
only along a characteristic (ray). In other words, 
x and z are constrained to follow the characteristic. 
The latter viewpoint is more useful for a general, 
nonuniform boundary layer, and also applies to a 
stationary harmonic line source in a uniform bound- 
ary layer when the locus and amplitude. distribution 
of the source are arbitrary. 

Equation (13) was derived in M77 from a general- 
ized Gaster relation between temporal and spatial 
amplification rates. Its meaning can best be seen 
from Figure 1, where the constant amplitude lines 
for the two growth directions Ugr and } are shown. 
These lines are normal to the direction of growth, 
just as the constant phase lines are normal to the 
direction of the wavenumber vector. A certain 
growth along Xgr in distance Ax xy requires the 
amplification rate along x to be 1/cos (b-bgr) 
larger than the amplification rate along Xgr to 
yield the same growth along x in the shorter dis- 
tance Ax = Axgr cos (¥-Pgr) - It is this relationship 
between o() and Sgr that is expressed by (13). For 
a fixed orientation of the constant amplitude lines 


N 


Use LINES OF 
#—— CONSTANT 
AMPLITUDE 


FIGURE 1. Wave growth in direction Xgr as described 
by constant amplitude lines normal to Xgr and to X. 


normal to Xgrr the growth in different directions 
follows the usual vector law with the amplification 
rate in direction V1 given by 
o(p,) = Go. cos(, - eae © 

We can_use (13) to (a) determine o,, from o(p) 
provided is known; (b) determine t if two 
neighboring values of o() are known; and (c) answer 
the question left open previously of whether we can 
make use of the simplification in the spatial theory 
afforded by (6). The latter is easily done. With 
(6), the transformation (5) applies to spatial waves 
and gives 


Se o(W) cosw. (15a) 


With v =w, (13) relates a() to ~O, by 


¥ COM ore 


= A 
= [o(v)cosw](1 + tany cane) cos eee (15b) 


It is evident from this expression that (6) is valid 
only for can O (or p = Wg,)- However, o() can 

be used to~calculate o_, if Jy, is known, on the 
same basis as any other o(i). This procedure is 
obviously to be avoided when the direction of k is 
perpendicular to that of o . 


Spatial Mode - Complex Group Velocity 


With a complex dispersion relation, the group veloc- 
ity, defined as 


oo dw aw 

é-G2, gS) 
is also complex. For pure temporal or spatial modes, 
@ is real only at points of maximum amplification 
rate. Consequently, it is important to know how 
the complex é affects the preceding analysis. With 
a and Ce complex, (8) is no longer hyperbolic, as 
pointed Out by Nayfeh et al. (1978). However, it 
is still possible to proceed by defining a real 
characteristic in the three-dimensional space (xy + 
ixj,z). Such a technique was used in a different 
context by Garabedian and Lieberstein (1958). 

The complex vector group velocity is conveniently 

described in terms of a complex magnitude and a 
complex angle by writing 


G@ S&€ cea, C= € sinh, (17a) 
x g Z g 
where 
2 2 
cai +o) (17b) 
x Zz 
is the complex magnitude, and 
VG = Vr + Wi (17c) 


is the complex angle. 
The complex counterparts of (10) are 


* 
Il 


cosW x + sin) 2z, 
g g 


N 
i] 


-sin) x + cos) 2, 
g g 


With x = x + ix,, and x, required to be real, 
x, = tanh . (tan Xi 1Z)) 
ab gi Cig, 46 
and 


aw = x % = a= = 
Xg = SOS 4 (an BEEN) tanh“ GA 2 


= 3e , 

With a real, the analysis for real C applies and 
gives for the now complex amplitude along the real 
characteristic, 


A(xg) = 4 (23,)explo() cos() - Vg)%g]. (18) 


This expression differs from (12) in that tg is 
complex, has been replaced by z, and z 


Xgr g gr 
(orthogonal to X'grr see below). 
We define 
Sab Picet Uae = a= wi) 
Soa Be Bey oe tanh veri Bore (19a) 


as the characteristic coordinate in the physical 
plane to replace Xgr - The angle between Xgr and 
Xgr is given by 


tan ()gy 2 Vgr) = -tanlgy tanh*}g;. (19b) 


We can now write the complex amplitude (18) as 
A(X) = Ay (Zgr) exp {ow [cos - Ygr) cosh*Pgi 
+i sin( = Ygr) coshi)g; sinhYg; ian} 
(20) 


The real part of the exponential factor defines the 
spatial amplification rate along ee to be 


= Sgt De 
o = o(v) cos(p - vee cosh Teas (21a) 


This expression differs from its real counterpart 
(13), aside from the factor cosh? wv gis in that ee 
Vgr is the real part of the canoes angle Vg 
not the angle formed by the real parts of ce ma 
Cz. When fp = Dgr' 

2 
a cosh Dag 0 (21b) 


Q 
ll 
Q 


and, unlike o_, o is not directly calculable as 
an eigenvalue?” The imaginary part of the exponen- 
tial factor of (20) gives the phase difference be- 
tween the elementary mode growing along x and the 
spatial mode (20) growing along x' . The phase 
difference can be written as JT 

= rae mr = Fr = P (22) 


Boe - a(W)= o(p) sin(y - Vespe) ENN) SHINY a0 


where a is the wavenumber component in the ore 


direction. We can now write the complex ¢ counter- 
part to the pure spatial mode (14) as 
) £(y) 
ore fi as ap B. Zz > [ee - a) |x - ut}. (23) 
With (23) we have arrived at the spatial mode 


that will be used for the numerical calculations 
to follow. The amplitude growth is along Xgr with 


WE 7,Bpe) = & 9 (2gr) exp (ox" ay 


67 


magnitude o given by (21b) The eigenvalues are 
preferably computed with p = Ygrr but as Ygr is 
generally not known in advance, or for computation- 

al convenience, they can be computed at a neighbor- 
ing ~ and o obtained from (2la). If W is sufficiently 
close to Wgr, the phase shift given by (22) is 
negligible and the orientation angle i is unaffected 
by the transformation. 

If Vg were independent of v, both (14) and (23) 
would also be expected to be independent of . How- 
ever, as v departs from Degen o(~) becomes large and 
the evaluation of the complex derivatives in (16) 
takes place in a region of the complex a and 6 
planes well removed from the points which give v 
The same difficulty exists in making comparisons 
between temporal and spatial amplification rates. 
Although the elementary modes with arbitrary |) are 
available for the solution of an initial value 
problem by superposition, we give physical signifi- 
cance _here only to the special spatial mode with 
yp =v All of the other spatial modes, as well 
as the2éombined temporal/spatial modes with a,f,w 
all complex, do not enter the present analysis 
except for computational purposes. 


OBLIQUE WAVES IN A TWO-DIMENSIONAL BOUNDARY LAYER 


Numerical Example of Transformation Formulas 


We shall first discuss the transformations from 
three- to two-dimensional form and then the trans- 
formation between a spatial mode with arbitrary 
growth direction and the mode with growth direction 
Vgr- A single numerical example for the Blasius 
boundary layer will suffice. We use the conventional 
dimensionless frequency parameter F = w*v*/U*T, 

and choose the length scale to be L* = (x*V*/U*) 2, 
With this choice, the Reynolds number appearing in 
(2) sig: Rs (Ut x*/v*)%. The subscript 1 refers to 
freestream conditions. 

For F = 0.2225 x 107+, R = 1600 and ) = 50°, a 
direct calculation of the eigenvalues with (6), i.e., 
~ = 50°, or the, completely equivalent two-dimensional 
calculations with either the Squire or Stuart 
transformations, gives 


ks OIG, a) = A119 x 10%. 


Application of the wavenumber transformation rule 
in (4) and (5) gives 


a = 0.1074, -a, = 2.648 x 10-3 
a ab 


for the complex wavenumber in the x direction. 

Ite Vgr is computed in the neighborhood of wp = 
50° from (13) by means of the assumption that o 
is independent of ) and with the frequency hela? 
constant, we find 


p= 9.39° 
Were 9 


to be an approximate value for the real part of the 
complex angle of the group velocity vector. (If 
the wavenumber is held constant, gy = 8-80°; a 
value closer to the angle formed by the real parts 
of Cy and C,.) The eigenvalues of the p - 50° wave 
with p = 9.39° are 


hk O10, G = Say & lor, 
gr 


68 
and in the x direction 
a = 0.1073,-a; = 3.085 x Oss). 


The eigenvalues computed with (6) differ from these 
values in the fourth decimal place, which means 
that as has an unacceptable error of 16.5%, an 
error which can also be calculated directly from 
(15b). Consequently, this example reiterates that 
(6), or the real Squire and Stuart transformations, 
can only be used if p = 0 (or p= bgr) - 

For the check of che transformation of an ele- 
mentary spatial mode with growth direction x to the 
'physical' mode with growth direction Xgr, we start 
by calculating the eigenvalues as a function of 
for 0 <p < 95° and the same F, R and yj as in the 
previous example. In addition, we calculate the 
complex group velocity by evaluating the complex 
derivatives of (2) from central differences for 
increments in dy, 8, of +0.001 about the calculated 
dy, By at each ~. The real and imaginary parts of 
the complex angle Ug are listed in columns 2 and 3 
of Table 1. The angle Ygr of the growth direction 
x!_, as obtained from (19b), is listed in Column 
4. Eigenvalues were computed as a function of jp 
with wy = 50° by integrating (2), starting at y/L = 
8.0, with a fourth-order Runge-Kutta integration 
and 80 equal integration steps. The results are 
listed in columns 5 and oF 

If (13) with Ugr = = tan” 
the o() given in column 6, a nearly constant 6 
is obtained out to about = 60°. For v SOO 5 eee 
decreases steadily, and at pw = 95° it is 21% lower 
than the Sgr for pp = v gr’ Columns 7 and 8 give the 
angle and Wace ciber : k for y= = v gr as calculated 
from the phase-shift formula (22) of the transform- 
ation for complex group velocity. The corresponding 
amplification rate, as calculated from (21a), is 
listed in column 9. Comparisons of directly com- 
puted eigenvalues with these k and o are provided 
in the last two columns. Column 10 lists the eigen- 
value k computed for the Vgr of column 2 and the wp 
of column 7. Column 11 lists the amplification 
rate 0 obtained from the eigenvalue Ogy accompany— 
ing k and from (21b). 

We see that the transformation formulas work 
quite well out to J = 60°, where the difference 


L(Cen/ Cue) is applied to 


between columns 9 and 11 is 0.13%. The change of 
the o in colum 9 with jj is only about half of the 
change given by the transformation with real group 
velocity and the correct jj x given by Cy, and Cz,. 
In this particular example, at least, the smallest 
change of o with | is found if (13) is used with 

v a also computed from (13) on the basis of two 
neighboring values of o(\) obtained with the fre- 
quency held constant and Sgr assumed to be indepen- 
dent of ~. The conclusion to be drawn is that in 
order to obtain the desired spatial amplification 
rate o as defined by (21b), o(W)may be computed at 
some convenient v which can differ from the correct 
Vor by as much as 40° or 50°, but should be as close 
as possible. Only later, after Vgr and 4 are 
know, is o(W) converted to o by the transformation 
formulas. Almost any of the methods discussed above 
for applying the transformations gives acceptable 
numerical accuracy. 


Effect of Obliqueness Angle on Instability 


The frequency F = 0.2225 x 107* used in the examples 
of the previous Section is the most unstable fre- 
quency at R = 1600, and the maximum amplification 
rate for this frequency occurs for ) = 0°. ‘The 
distribution of o with ~ is shown in Figure 2 for 
this frequency and F x lot = 0.280, 0.1490 and 
0.1008. The latter two frequencies are the most 
unstable for y = 60° and 75°, respectively. They 
have their peak amplification rates, not for ~ = 0°, 
but for y = 34.4° and 61.8°, respectively. These 
results demonstrate that although the maximum am- 
plification rate at a given Reynolds number with 
respect to both frequency and orientation occurs 

for a two-dimensional wave, the maximum amplification 
rate with respect to orientation of given frequency 
occurs for.a three-dimensional wave if the frequency 
is less than the most unstable frequency. 

The envelope curve formed by the individual 
frequency curves is also shown in Figure 2. This 
curve gives Om;y,, the maximum amplification rate 
with respect to frequency, as a function of i. The 
envelope curve emphasizes the wide range of unstable 
orientations in a two-dimensional boundary layer. 

It can be seen that oy,, is not reduced to one-half 


TABLE 1 Numerical check of spatial-mode transformation for complex group 


velocity. R = 1600, F = 0.2225 x 107", 


w= 50°. 


% ik @ & TO? 
‘ip k o(t) x 103 v k o x 103 

TF p vy. eke F oa Y of C7 
7) gr gi (9b)! etga, w= 50° (22) (21a) SX D = There 

1 2 3 4 5) 6 U 8 9 10 11 
0 9.23 =-4.03 9.18 (o)aleyexs} sical GIS) SS) @)qalleysye) 3.145 0.1669 3.146 
9.21 9.21 -—4.02 9.16 0.1669 3.127 50.00 0.1669 Salas} 0.1669 3.143 
30.0 Deals AIG, GS), alah 0.1670 3.340 50.02 0.1669 iS leSy/) 0.1669 3.138 
60.0 9.08 -4.00 9.04 0.1672 4.928 50.06 0.1670 2}, La OMG OMSL 26 
90.0 8.67 =-3.88 8.63 OO) akexsi7/ aS SO) S10) © Oaal7/7/ 2.977 0.1676 3.060 
95.0 83365) —SieSuskhiss. 0.1710 46.04 50.66 0.1688 2251/0) (0) WS 27 


ENVELOPE (o 


max) 


% 4 
x 
in} 

3 

2 

] 

0 

0 10 20 30 40 50 60 70 80 
WY (deg) 

FIGURE 2. Amplification rate as function of ~ for four 
frequencies. Blasius boundary layer, R = 1600. 


of its two-dimensional value until ) has increased 
to 60°. With unstable waves for -79° < p < 79°, a 
consideration of only the two-dimensional wave gives 
an incomplete picture of the instability of the 
boundary layer. 


THREE-DIMENSTONAL FALKNER-SKAN BOUNDARY LAYERS 


In order to study the influence of three dimension- 
ality in the mean flow on boundary-layer stability, 
it is necessary to have a family of boundary-layers 
where the magnitude of the crossflow can be varied 
in a systematic manner. The two-parameter yawed- 
wedge flows introduced by Cooke (1950) are suitable 
for this purpose. One parameter is the usual Falkner- 
Skan dimensionless pressure gradient; the other 
is the ratio of the spanwise and chordwise velocities. 
A combination of the two parameters makes it possible 
to simulate simple planar three-dimensional boundary 
layers. 
The inviscid velocity in the plane of the wedge 

and normal to the leading edge is 

U* = C#(x*) T) 

cy c 

where the wedge angle is (1/2) R and § = 2m/(mt1) . 
We shall refer to this velocity as the chordwise 
velocity. The velocity parallel to the leading 
edge, or spanwise velocity is 


W* = const. 
Sl] 


The subscript 1 refers to the local freestream. For 
this inviscid flow, the boundary-layer equations 

in the x_ direction, as shown by Cooke (1950), 
reduce to 


2 
£” + ££" +8 4) = |S O. 
h 2 


This equation is the usual Falkner-Skan equation 
for a two-dimensional boundary layer, and is inde- 


69 


pendent of the spanwise flow. The dependent vari- 
able f(n) is related to the dimensionless chordwise 
velocity by 


We S36 _f 2 2°) 5 
U m+1 


and the independent variable is the similarity 
variable 


iW) SPN oeers Il a 
r c 

where xo is measured normal to the leading edge. 

Once £(n) is known, the flow in the spanwise di- 


. * . 
rection Zs is obtained from 


where 


w* 
Si 


Both f£'(n) and g(n) are zero at n = O and approach 
unity as Tabulated values of g(n) fora 
few values of By may be found in Rosenhead (1963, 
p. 470). 

The final step is to use £'(n) and g(n) to con- 
struct the mainflow and crossflow velocity components 
needed for the stability equations. A flow geometry 
appropriate to a swept back wing is shown in Figure 
3. There is no undisturbed freestream for a Falkner- 
Skan flow, but such a direction is assumed and a 
yaw, or sweep, angle yw is defined with respect to 
it. The local freestream, or potential flow, is at 
an angle Pp with respect to the undisturbed free- 
stream. It is the potential flow that defines the 
x,Z coordinates of the stability equations. The 
angle of the potential flow with respect to the 
chord is 


n> ©. 


ws 
54, 


-1 
Cesta U* 4 
i 


and @ is related to We. and i by 


UNDISTURBED 
y FREESTREAM 


FIGURE 3. Diagram of coordinate systems used for 
Falkner-Skan-Cooke boundary layers. 


70 


With the local potential velocity, Up = (uxt + we?)’2, 
as the reference velocity, the dimensionless main- 
flow and crossflow velocity components are 
2 2 
f'(n) cos 6 + g(n) sin 6, 


U(n) (24a) 


W(n) [-£" on) + gin) | cos@ siné . (24b) 

These velocity profiles are defined by By, which 
fixes £'(n) and g(n), and the angle 6. We note 
from (24b) that for a given pressure gradient all 
crossflow profiles have the same shape; only the 
magnitude of the crossflow velocity changes with 
the flow direction. In contrast, according to 
(24a), the mainflow profiles change shape as §@ varies. 
For 0) = 0, U(n) = £.(n);) for 6 = 90°, UM) = g(n); 
for 6= 45°, the two functions make an equal con- 
tribution. 

When the velocity profiles (24) are used directly 
in the stability relations, (2), the velocity and 
length scales of the equations must be the same as 
in (24). This identifies the velocity scale as U*, 
the length scale as P 


V*X*/U* (x*) : ' 
{o cy Cc 


ie Ss 


and the Reynolds number Were as 
R=R/cos® , 
c 


where R_ = [vs oxy | 4 is the square root of 
c i) Ee re 


the Reynolds number along the chord. For positive 
pressure gradients (m > 0), 8 = 90° at x = O and 

8 > 0° as x > »; for adverse pressure gradients 

(m <0), 6) ="90° atix = 0) and! 0) 209 as x =); for 
adverse pressure gradients (m < 0), 8 = 0° at x = 
0 and 8 + 90° as x > ©. The Reynolds number R, is 
zero at x = 0 for all pressure gradients, as is 

R with one important exception. The exception is 
where m= 1 (8, = 1) is the stagnation-point solution; 
here it is the attachment-line solution. In the 
vicinity of x = 0, the chordwise velocity is 


U* = x* (aU* /dx*) x 
cy c cy c x=0 


The potential velocity along the attachment line is 
eno and the Reynolds number is 


R(x=0) = we / |v (a08, /8%2) x20] 


a non-zero value. 

For the purposes of this paper, 86 may be regarded 
as a free parameter, and the velocity profiles (24) 
used at any Reynolds number. However, for the flow 
over a given wedge, 8 can be set arbitrarily at only 
one Reynolds number. If 6,45 is 8 at Ro = (R)) 
the 6 at any other Ro is given by 


a | m/(m+1), 
tan® = tan® _. [a 7 | 


Cerehaic 


ref’ 


For m << 1, the dependence on R_ is so weak that 6 
is constant almost everywhere. “one way of choosing 
(Re)ref Within the context of Figure 3 is to make 
it the chord Reynolds number where p = 0; i-e., 
the local potential flow is in the direction of the 
undisturbed freestream. Then 6 is equal to the 
yaw angle j_. mee 
Figure 4 shows the crossflow velocity profiles 


FIGURE 4. Four crossflow velocity profiles, Falkner- 
Skan-Cooke boundary layers. INF, inflection point; 
MAX, maximum crossflow; SEP, separation pressure 
gradient (fy, = -0.1988377) . 


for 6 = 45° and four values of 8}. The inflection 
point and point of maximum crossflow velocity (Wmax) 
are also noted on the figure. In Figure 5, Wmax for 
@ = 45° is given as a function of 8} from near sep- 
aration to 8, = 1.0. The crossflow velocity for 
any other flow angle is obtained by multiplying the 
Wmax of the figure by cos8 sin@. The maximum cross-— 
flow velocity of 0.133 is generated by the separa- 
tion profiles rather than by the stagnation profiles, 
where W = 0.120. However, W varies rapidly 
with 8 men the neighborhood of separation, as do 
all Behen boundary-layer parameters, and for 8, = 
-0.190, W is only 0.102. 

The function g(n) is only weakly dependent on 


12 


ai 

: 

x 
6 

) 

= 

= 
4h 
| 
-0.2 0 0.2 0.4 0.6 0.8 1.0 

By, 
FIGURE 5. Effect of pressure gradient on maximum cross— 


flow, Falkner-Skan-Cooke boundary layers. 


al 


TABLE 2. Properties of three-dimensional Falkner-Skan-Cooke boundary layers. 
Bu 8 Ns ng" be Wiese W finf Ninf 
SEP 2.2 8.238 3.495 4.024 0.0102 0.00476 0.487 4.306 

5.0 8.236 3.489 4.010 0.0231 0.01077 1.100 

10.0 8.229 3.466 3.959 0.0455 0.02123 2.156 

40.0 8.095 3.075 3.280 0.1310 0.06214 5.709 

45.0 8.058 2.986 3.167 0.1330 0.06339 5.696 

50.0 8.017 2.897 3.064 0.1310 0.06274 5.516 
-0.10 45.0 6.522 1.985 2.698 0.0349 0.01619 1.498 3213 
-0.02 45.0 6.098 1.763 2.609 0.0058 0.00267 0.249 2.940 
0.02 45.0 BJoesul 1.682 2.578 -0.0054 -0.00248 =0)232 2.835 
0.04 45.0 5.854 1.646 2.564 -0.0104 -0.00480 -0.449 2.787 
0.10 45.0 5.646 Ae ‘SISAL 2ro2 9 O02 9) -0.01094 -1.029 2.659 
0.20 45.0 5.348 1.424 2.482 -0.0423 -0.01924 milo a3} 2.478 
1.0 2.4 3.143 0.6496 2.227 -0.0100 -0.00503 -0.406 1.524 

10.0 3.196 0.6603 2.226 -0.0410 -0.02021 -1.669 

40.0 3.574 0.8050 2.275 -0.1181 -0.05204 —)o Ie) 

45.0 3.621 Oss} Bo shoal —o)paliejal -0.05217 —B)E2Eil 

50.0 3.661 OF 87/06) (253325 5-0-8 -0.05081 5 Zeb) 

55.0 3.695 0.9024 2.366 -0.1127 -0.04804 —B), SS) 

80.0 So7/Sal 1.0153 2.524 -0.0410 -0.01704 AS NSH) 

87.6 Jo 72) -0.00416 -0.489 


1.0260 2.542 -0.0100 


Bre and, unlike f'(n), never has an inflection 

point even for an adverse pressure gradient. Indeed 
it remains close to the Blasius profile in shape, 

as underlined by a shape factor H (ratio of dis- 
placement to momentum thickness) that only changes 


have its wavenumber vector nearly aligned with the 
local potential flow, and we can restrict ourselves 
to waves with ~ = 0° for the purpose of determining 
the maximum amplification rate. With the temporal 
stability theory, this procedure is equivalent to 


from 2.703 to 2.539 as, goes from -0.1988377 (sep- 
aration) to 1.0 (stagnation). The weak dependence 
of g(n) on was first pointed out by Rott and 
Crabtree (1952), and made the basis of an approximate 
method for calculating boundary layers on yawed 
cylinders. For our purposes, it allows some of the 
results of the stability calculations to be antici- 
pated. For waves with the wavenumber vector aligned 
with the local potential flow, we can expect the 
amplification rate to vary smoothly from its value 
for a two-dimensional Falkner-Skan flow to a value 
not too far from Blasius as 6 goes from zero to 90°. 
The stability results in the next section will 
be presented in terms of the Reynolds number R and 
the similarity length scale L*. In order that the 
results may be converted to the length scales of 
the boundary-layer thickness, displacement thick- 
ness and momentum thickness, Table 2 lists the 
dimensionless quantities ng = 5/L*, ng* = 6*/L* and 
H = ng*/Ng of the mainflow profile for several com- 
binations of 6, and 6. Also listed are Wmax, the 
average crossflow velocity W = (wan) /ng, the 
deflection angle of the streamline at the inflection 
point, €jnf, and the location of the inflection 
point, Ninf. The quantity ng is defined as the 
point where U = 0.999. 


studying the two-dimensional instability of the 
mainflow profile, but is only approximately so in 
the spatial theory _unless Vgr = 0, 7S v 7 is 
usually small for y = 0°, even with large cross- 
flow, we may also view the = 0° spatial results 
as a measure of the instability of the mainflow 
profile. 

In order to place the three-dimensional effects 
in context, it is helpful to first consider a small 
deviation in the assumed pressure gradient on the 
maximum amplification rate of two-dimensional 
Falkner-Skan profiles. Figure 6 shows the maximum 
spatial amplification rate (with respect to frequency) 
as a function of Reynolds number for Blasius flow 
and for By = + 0.02. What is noteworthy about 
these results is the magnitude of the shift in Omax 
for what are quite small pressure gradients. It 
is evident that an experiment intended to measure 
amplification rates in a Blasius boundary layer to 
within an accuracy of 10% is required to maintain 
an exceptional uniformity in the flow. 

The effect of the flow angle 8 on the maximum 
spatial amplification rate of the waves with i = 0° 
is shown in Figure 7 for By, = + 0.02 and two Rey- 
nolds numbers. In these calculations, gr and bgi 
were both taken equal to zero. The amplification 
rate Omax is expressed as a ratio to the Blasius 
value (0,)max Shown in Figure 6. It will be re- 
called that with 6, = 0, g(n) = £(n), and the 
velocity profile remains the Blasius function for 
all flow angles. The effect of a non-zero flow 
angle withs, # 0 is destabilizing for a favorable 
pressure ergatthicrate , and stabilizing for an adverse 
pressure gradient. Consequently, it reduces the 
pressure-gradient effect shown in Figure 6. The 
reason for this result is easy to understand by 
reference to (24). We have already pointed out in 
Section 4 that the spanwise velocity profile g(n) 


STABILITY OF FALKNER-SKAN-COOKE BOUNDARY LAYERS 


Boundary Leyers with Small Crossflow 


In a two-dimensional boundary layer, the most un- 
stable wave is two dimensional. Therefore, we can 
expect that in three-dimensional boundary layers 
with small crossflow the most unstable wave will 


x10? 


? max 


0 0.5 1.0 1.5 2.0 2.5 3.0 
R x 109 


FIGURE 6. Effect of small pressure gradients on 
the maximum amplification rate with respect to fre- 
quency for two-dimensional Falkner-Skan boundary layers. 


is always close to the Blasius function. Thus as 
the flow angle increases from zero the amplification 
rate must change from the two-dimensional Falkner- 
Skan value at 6 = 0° to a value not far from Blasius 
at 6 = 90°. 

As discussed in Section 4, the only physically 
meaningful flow with 6 = 90° and a non-zero Reynolds 
number is the attachment-line flow (f, = 1.0). For 
all other values of 8}, R at this flow angle must be 
either zero (8, > 0) or infinite (fy < 0). With By 
= 1.0 and R = 1000 (R = 404.2, where Ro is the 
momentum-thickness Reynolds number), Omax/ (8b) max 
= 0.766. The minimum critical Reynolds number of 
this profile is (Rg) jy = 268 (the parallel-flow 
Blasius value is 201), yet turbulent bursts have 
been observed as low as Rg = 250 for small distur- 
bances by Poll (1977). 


1,20 


Omax/( uy Bynax 


9 (deg) 


FIGURE 7. Effect of flow angle on the maximum amplifi- 
cation rate with respect to frequency of = 0° waves 
for two boundary layers with small crossflow at two 
Reynolds numbers. 


We must still show that the waves with p = 0° 
properly represent the maximum instability of three- 
dimensional profiles with small crossflow. For this 
purpose a calculation was made of 0 as a function 
of p for Bh = -0.02, 9 = 45°, R = 1000 and F = 
0.4256 x 10-*, the most unstable frequency for ) = 
0° at this Reynolds number. It was found that the 
crossflow indeed introduces an asymmetry into the 
distribution of 0 with W, and the maximum of Oo is 
located at = -6.2° rather than at 0°. However, 
this maximum value differs from the Omay of Figure 
7 by only 0.7%. It was also determined that v r= 
-0.04° and Wgi = -0.3° (approximately) for ~ = 0°, 
which justifies taking both of these quantities 
zero in all of the ~ = O° calculations. 


Crossflow Instability 


Minimum Critical Reynolds Number of Steady 
Disturbances 


The instability that is unique to three-dimensional 
boundary layers is called crossflow instability. 

It was discovered experimentally by Gray (1952) and 
later given a detailed theoretical explanation by 
Stuart in Gregory et al. (1955). This instability 
arises from the inflection point of the crossflow 
velocity profile. As explained by Stuart, there 

is a particular direction close to the crossflow 
direction for which the mean velocity at the in- 
flection point of the resultant velocity profile 

is zero. Consequently, at sufficiently large 
Reynolds numbers unstable steady disturbances exist 
which have their constant phase lines nearly aligned 
with the potential flow. 

Although crossflow instability is by no means 
restricted to. steady disturbances, these disturbances 
do make a convenient starting point for our investi- 
gation. The reason is that a suitable initial 
guess for the angle ~, which must be known rather 
accurately for the eigenvalue search procedure to 
converge, is given by 


v= (By/[Bnl) (7/2 - leline)» 


where €jnf¢ is the streamline deflection angle listed 
in Table 2. It turns out that this value is with- 
in a fraction of a degree of the angle of the most 
unstable wavenumber. There is no such convenient 
rule for the wavenumber itself, but the inverse of 
Ning the location of the inflection point in the 
similarity coordinate, or better still 0.9/Ning is 
usually an adequate enough initial guess to ensure 
rapid convergence to an eigenvalue. 

As the crossflow is a maximum at 0 = 45° fora 
given By, we can expect the crossflow instability 
to also be a maximum near this angle. Figure 8 
shows the minimum critical Reynolds number Roy at 
§ = 45° for the zero-frequency disturbances as a 
function of f,. For comparison, Roy of the two- 
dimensional Falkner-Skan profiles, as computed by 
Wazzan et al. (1968), is also given. For adverse 
pressure gradients, the steady disturbances become 
unstable at Reynolds numbers well above the Roy of 
the two-dimensional profiles. On the contrary, for 
Bh > 0.07 the reverse is true, and for most pressure 
gradients in this range the steady disturbances 
become unstable at much lower Reynolds numbers than 
the two-dimensional Roy (for fp, = 1.0, the two- 


0.4 b . 1.0 
B, 


FIGURE 8. Minimum critical Reynolds number as function 
of pressure gradient: ——, steady disturbances, Falkner- 
Skan-Cooke boundary layers with 6 = 45°; ---, two- 
dimensional Falkner-Skan boundary layers [from Wazzan 

et al. (1968) ]. 


cr 


107} = 
BE SEP 
6E 
4 | | = | 
0 20 40 60 80 


9 (deg) 


FIGURE 9. Effect of flow angle on minimum critical 
Reynolds number of steady disturbances for fy, = 1.0 
- and separation boundary layers. 


73 


TABLE 3. Wave parameters at minimum critical 
Reynolds number of steady disturbances. 


8 i) R k y (p_) 
h cr cr (che! gunen 

SEP Ps? 535 0.213 -89.41 0.2 

5.0 237 0.213 -88.68 0.4 

10.0 iZaL 0.215 -87.44 0.9 

40.0 46.5 0.230 -83.54 3.0 

45.0 46.7 0.230 —{2}2}55)7/ 3.0 

50.0 48.4 OR2 351) —OSmons 3.0 

-0.10 45.0 276 0.295 -88.42 0.9 

-0.02 45.0 1885 0.310 -89.74 0.2 

0.02 45.0 2133 0.322 89.76 —{0)5 AL 

0.04 45.0 1129 0.327 89.53 —Ofy2 

0.10 45.0 527 0.339 88.93 0), 

0.20 45.0 328 0.358 88.12 ok, aL 

1.00 Qe 2755 OF553 89.60 -0.3 

10.0 671 0.547 88.33 ile ib 

40.0 219 0.545 84.88 -3.4 

45.0 2ale2) 0.540 84.70 =3}5'5) 

50.0 212 0.540 84.70 = 8\5'5) 

55.0 218 0.538 84.85 —3i3 

80.0 563 0.532 88.00 =1155) 

87.6 2325 ORS'S82 I) 5 5.1 O53} 


dimensional Rg; is 19,280 compared to Roy = 212 for 
zero-frequency crossflow instability). 

The distribution of Roy with 6 is shown in Figure 
9 for Bh = 1.0 over the complete range of 6, and 
for the separation profiles (fp, = -0.1988377) over 
the range 0° < 6 < 50°. Near 6 = 0° and 90°, Roy 
is very sensitive to 6; near, but not precisely at, 
@ = 45° Roy has a minimum. This minimum occurs 
close to the maximum of ell Are (cf. Table 2), which, 
unlike Wy3., is not symmetrical about 6 = 45°. Table 
3 lists the critical wave parameters for a few com- 
binations of 8, and 6. The extensive computations 
needed to fix these parameters precisely were not 
carried out in most cases, and so the values in the 
Table are not exact. The listed Ygr was obtained 
from (13); Ygi was not calculated. 


Boundary Layer with Crossflow Instability Only 


As an example of a boundary layer which is unstable 
at low Reynolds number only as a result of cross- 
flow instability, we select 8}, = 1.0 and @ = 45°, 
and present results for the complete range of un- 
stable frequencies. Although this pressure gradient 
can only occur at an attachment line, Figure 8 leads 
us to expect that all profiles with a strong favor- 
able pressure gradient will have similar results. 
For this type of profile, the minimum critical 
Reynolds number of the least stable frequency is 
very close to the R of Figure 7. We therefore 
choose a Reynolds number well above Rr where the 
instability is fully developed. 

Figure 10 provides a summary of the stability 
characteristics at R= 400. For a given frequency, 
the eignevalue o(i) can be computed as a function 
of either k or , with the other parameter given as 
the second eigenvalue. For strictly crossflow 
instability, k is the more suitable independent 
variable as i) can have an extremum in the unstable 
region. All unstable eigenvalues of a given fre- 
quency with a specified increment in k were calcu- 


74 


lated in a single computer run with Ugr = 0°, and 
then corrected to an approximate Ygr (k) obtained 
from (13) with constant wavenumber. A least-squares 
curve fit to o(k) provided Omay, to maximum spatial 
amplification rate with respect to the vector wave- 
number, and kmax and Wax, the magnitude and direc- 
tion of the wavenumber of Omax- 

Figure 10a gives Omax aS a function of the di- 
mensionless frequency F, and also shows the portion 
of the ~-F plane for which there is instability. 
The unstable region is enclosed between the curves 
marked and w_. These curves represent either 
neutral stability points or extrema of }. 

The corresponding wavenumber magnitudes are 
shown in Figure 10b. The negative frequencies 
signify that with taken to be continuous through 
F = 0, the phase velocity changes sign. If we 
choose ) so that the wavenumber and phase velocity 
are both positive, then it is | that changes sign 
at F = 0. Consequently, there are two groups of 
positive unstable frequencies with quite different 
phase orientations. The first group, which includes 
the peak amplification rate, is oriented anywhere 


from 5° to 31° (clockwise) from the direction opposite 


to the crossflow direction. The second group is 

oriented close to the crossflow direction itself. 
All of the unstable frequencies have in common 

that the direction of growth is within a few degrees 

of the potential~flow direction. The angle Ugr of 

Umax, aS computed from (13), is negative and has 

its largest magnitude of just under 6° near F = 

-0.60 x 10-+. Orientations other than Umax can 

have growth directions further removed from the 

flow direction. 


Boundary Layers with both Crossflow and Mainflow 
Instability 


As an example of a boundary layer which has both 
crossflow and mainflow instability at low Reynolds 
numbers, we select 8, = -0.10 and 6 = 45°. In con- 
trast to the previous case, the steady disturbances 
do not become unstable until a Reynolds number, R = 
276, where the peak amplification rate is already 
7.35 x 10-°. [For B} = -0.10 and @.= 0° omax = 
MMO Os! ate kes 22 Om according to Wazzan 
et al. (1968)]. The distribution of o with jp is 
shown in Figure 11 for F = 2.2 x 10-4, a frequency 
close to the most unstable frequency of F = 2.1 x 
10- . We see that with a maximum crossflow velocity 
of 0.0349 (cf. Table 2), the distribution of o about 
w = 0° is markedly asymmetric, and the maximum 
amplification rate of 7.31 x 1073 is located at jp = 
-29.4° rather than near zero. This asymmetry was 
barely perceptible for the small crossflow boundary 
layers of Figure 7 where the crossflow is only one- 
sixth as large. The o at pp = 0° of Figure 11 (5.82 
x 10-3) is close to Omax With respect to frequency 
of the » = 0° waves (5.91 x 10°3). Since this value 
is 20% below the peak amplification rate, the = 
O° waves are no longer adequate to represent the 
Maximum instability as with small crossflow boundary 
layers. Figure 11 also gives the distribution with 
of k and Wgr- The latter quantitiy was obtained 
from (13) with constant wavenumber, and we see that 
it remains within + 7.5° of the potential-flow 
direction throughout the unstable region. 

Because R = 276 is the minimum critical Reynolds 
number of the steady disturbances, the unstable 
region terminates in a neutral stability point at 


UNSTABLE 


0.2 ki STABLE 


I i 1 it 
O05) Os SNM ONNZESINSSO 
Fx104 


0 L er 
-2.0 -1.5 -1.0 -0.5 


FIGURE 10. Instability properties of 8, = 1.0, 8 = 45° 
Falkner-Skan-Cooke boundary layer at R = 400. (a) maxi- 
mum amplification rate with respect to Wavenumber and 
unstable ~ - F region; (b) unstable k-F region. 


F = 0. We are particularly interested here in Rey- 
nolds numbers where F = O is also unstable, and as 
an example, Figure 12 gives results for all unstable 
frequencies at R= 555. Figure 12a shows Omax as 

a function of F (here, as in Figure 10, Omax 1s the 
maximum with respect to k), as well as the unstable 
region of the k-F plane; the unstable region of the 
W-F plane appears in Figure 12b. These two unstable 
regions are quite different from those of Figure 10 
where there is only crossflow instability. The 
negative frequencies do resemble those of Figure 10 
in that the unstable range of ~ is small, of k is 
large, and with defined so that F > 0, the orien- 
tations are close to the crossflow direction. How- 
ever, for the higher frequencies, which are by far 


a x 10°, k x 10 


=70 -60 =50 =40 -30) -20) -10 0 10 20 30 40 
(deg) 


FIGURE 11. Effect of wavenumber angle on 9, k and Vgr 
for By, = -0.10, 8 = 45° Falkner-Skan-Cooke boundary 
layer at R = 276. F = 2.2 x 107". 


x103, k x 10 


STABLE 


UNSTABLE 


-60 
-80 
-100 
-0.2 0 0.2 0.4 0.6 0.8 1,0 1.2 1.4 
Fx10 
FIGURE 12. Instability properties of Bh = -0.10, 9 = 


45° Falkner-Skan-Cooke boundary layer at R = 555. 
(a) maximum amplification rate with respect to wave- 
number and unstable k-F region; (b) unstable -F region. 


the most unstable, the unstable regions of Figure 
12 bear more of a resemblance to those of a two- 
dimensional boundary layer than to Figure 10. The 
main differences from the two-dimensional case are 
the asymmetry about = 0° already noted in Figure 
11, the one-sidedness) Of Wnax, and, for F < 0.4 x 
10-4, the replacement of a lower cutoff frequency 
for instability by a rapid shift with decreasing 
frequency to waves oriented opposite to the cross- 
flow direction and which are unstable down to zero 
frequency. The instability shown in Figure 12 
represents primarily an evolution of the small cross- 
flow boundary layers of Figure 7 to larger cross- 
flow. Only the frequencies, say |F| < 0.2 x loser 
have to do with the pure crossflow instability of 
Figure 10. For frequencies near 0.4 x 1074 ,wy 
varies little with k in one part of the unstable 
region, as with crossflow instability; in the other 
part, as with mainflow instability, the opposite 
is true. This behavior becomes more pronounced at 
high Reynolds numbers. 


CONCLUDING REMARKS 


All of the numerical results that have been presented 
stem from the viewpoint adopted in Section 2 that 


75 


useful information concerning three-dimensional 
boundary-layer stability can be obtained from par- 
ticular pure spatial modes just as with two- 
dimensional boundary layers. Arguments were given 
to support using the modes whose growth direction 
is determined from (17) or, more exactly, from 
(19b). A transformation (2la), was derived to 
enable the use of waves with an arbitrary growth 
direction in calculating eigenvalues. The trans- 
formation used in the temporal theory to reduce 
the three-dimensional problem to a two-dimensional 
probelm in the direction of the wavenumber vector 
was shown to apply to spatial modes only when this 
direction is close to the correct growth direction, 
or the latter is the same as the potential-flow 
direction WVgr = 0°). 

The waves which have their wavenumber vector 
aligned with the local potential flow (i) = 0° when 
the x axis of the mean-flow coordinate sytem is 
also in the flow direction) always have their growth 
direction very close to the potential-flow direction. 
If the crossflow is small, the maximum amplification 
rate of the ~ = 0° waves is almost identical to the 
maximum amplification rate of the three-dimensional 
boundary layer. Consequently, if we are only in- 
terested in establishing the maximum amplification 
rate of a small crossflow boundary layer, it can be 
obtained from the mainflow profile alone. We used 
this approach to obtain the effect of the flow (yaw) 
angle on the instability of the Falkner-Skan-Cooke 
yawed-wedge boundary layers for small pressure 
gradients, and found that yaw reduces both the 
stabilizing effect of a favorable pressure gradient 
and the destabilizing effect of an adverse pressure 
gradient. 

With moderate or large crossflow, crossflow in- 
stability, which arises from the inflection point 
of the crossflow velocity profile, is present and 
can destabilize a boundary layer at low Reynolds 
numbers which would otherwise be stable. As befits 
the name, the unstable waves have their wavenumber 
vectors oriented near the crossflow (or opposite) 
direction. Also the instability covers a wide band 
of unstable frequencies (including zero) and wave- 
numbers. The growth direction of all unstable waves 
is still near the potential-flow direction. If the 
mainflow profile is also unstable, then the unstable 
frequencies near zero act as with pure crossflow 
instability and the higher frequencies as with pure 
mainflow instability. Intermediate frequencies 
have the latter behavior for small wavenumbers, and 
the former for large wavenumbers. 

The results demonstrate why crossflow is more of 
a problem for the maintenance of laminar flow with 
strong favorable pressure gradients than with ad- 
verse pressure gradients. In the former case, cross- 
flow provides a powerful instability mechanism 
even when the mainflow profile is stable; in the 
latter, the crossflow only increases the amplifi- 
cation rate over that of an already unstable main- 
flow profile. This increase is about 50% for the 
6 = 45° separation boundary layer. 


ACKNOWLEDGMENT 


This paper represents the results of one phase 

of research carried out at the Jet Propulsion Lab- 
oratory, California Institute of Technology under 
Contract No. NAS7-100 sponsored by the National 
Aeronautics and Space Administration. Financial 


76 


support is gratefully acknowledged from the Tactical 


Technology Office, Defense Advanced Research Projects 


Agency, and from Langley Research Center. 


REFERENCES 


Brown, W. B. (1961). A stability criterion for 
three-dimensional laminar boundary layers. 
Boundary Layer and Flow Control, Vol. 2, G. V. 
Lachmann, ed., Pergamon Press, New York, pp. 
913-923. 

Cooke, J. C. (1950). The boundary layer of a class 
of infinite yawed cylinders. Proc. Camb. Phil. 
Soc. 46, 645. 

Davey, A. (1972). The propagation of a weak non- 


linear wave. J. Fluid Mech. 53, 769. 

Grabedian, P. R., and H. M. Lieberstein (1958). 
On the numerical calculation of detached bow 
shock waves in hypersonic flow. J. Aero. Sci. 
25), LOSE 

Gray, W. E. (1952). The nature of the boundary 
layer at the nose of a swept back wing, Unpub- 
lished work Min. Aviation, London. 

Gregory, N., J. T. Stuart, and W. S. Walker (1955). 
On the stability of three-dimensional boundary 
layers with application to the flow due toa 
rotating disk. Phil. Trans. Roy. Soc. London 
A248, 155. 


Landau, L. D., and E. M. Lifshitz (1960). Electro- 


dynamics of Continuous Media, Pergamon Press, 
New York, p. 263. 

Mack, L. M. (1977). Transition prediction and 
linear stability theory. Laminar-Turbulent 
Transition, AGARD Conference Proceedings No. 224, 
Pp l= to di=22'5 

Nayfeh, A. H., A Padhye, and W. S. Saric (1978). 

The relation between temporal and spatial sta- 
bility in three-dimensional flows, AIAA Paper, to 
be presented. 

Poll, D. I. A. (1977). Leading edge transition on 
swept wings. Laminar-Turbulent Transition, 
AGARD Conference Proceedings No. 224, 21-1 to 
Aaloilal. 

Rosenhead, L. (1963). 
Oxford Univ. Press. 

Rott, N., and L. F. Crabtree (1952. Simplified 
laminar boundary layer calculations for bodies 
of revolution and for yawed wings. J. Aero. 
Seri do), 35s}q 

Squire, H. B. (1933). On the stability for three- 
dimensional disturbances of viscous fluid flow 
between parallel walls. Proc. Roy. Soc. London 
Al42, 621. 

Wazzan, A. R., T. T. Okamura, and A. M. O. Smith 
(1968). Spatial and temporal stability charts 
for the Falkner-Skan boundary-layer profiles. 
McDonnell Douglas Report No. DAC-67086, Long 
Beach, Calif. 

Whitham, G. B. (1974). Linear and Nonlinear Waves, 
Wiley-Interscience, New York. 


Laminar Boundary Layers, 


Experiments on Heat-Stabilized Laminar 
Boundary Layers in a Tube 


Steven J. Barker 


Poseidon Research* and University 
of California at Los Angeles 


ABSTRACT 


There has been considerable recent interest in the 
stabilization of water boundary layers by wall 
heating. Calculations based upon linear stability 
theory have predicted transition Reynolds numbers 
as high as 2 x 108 for a zero pressure gradient 
boundary layer over a heated wall. The flow tube 
experiment described in this paper was intended to 
investigate these predictions. The test boundary 
layer develops on the inside of a cylindrical tube, 
0.1 m in diameter and 6.1m in length. The dis- 
placement thickness is small relative to the tube 
radius under nearly all operating conditions. The 
tube is heated by electrical heaters on the outside 
wall. The location of transition can be determined 
by a heat flux measurement, by flush-mounted hot 
film probes, or by flow visualization at the tube 
exit. 

A transition Reynolds number of 107 can be ob- 
tained without heat, which shows that free stream 
turbulence and other perturbations are well con- 
trolled. At 7°C wall overheat, a transition 
Reynolds number of 42 x 10° has been obtained, 
which is at least as high as the prediction for 
that overheat. However, as temperature is further 
increased there have been no additional increases 
in transition Reynolds number, which is in contra- 
diction to the theory. 

Possible reasons for the differences between 
theory and experiment have also been investigated. 
New test section exits have been developed to 
determine the effects of downstream boundary con- 
ditions upon the flow. An instrumented section 
has been used to measure detailed velocity profiles 
in the boundary layer, and determine intermittency 
as a function of azimuthal angle. From these 
measurements we can evaluate the possibility of 


* 

This work was performed by the Marine Systems Division of 
Rockwell International, and Poseidon Research. It was 
sponsored by the Defense Advanced Research Projects Agency. 


77 


buoyancy-generated instabilities in the tube. 
Future tests will also investigate the influence 
of free stream turbulence, streamwise vorticity 
in the boundary layer, and wall temperature vari- 


ations. 


1. INTRODUCTION 


Numerical calculations such as those of Wazzan, 
Okamura, and Smith (1968, 1970) have predicted large 
increases in the transition Reynolds numbers of 
water boundary layers with the addition of wall 
heating. The stabilizing mechanism is the decrease 
in fluid viscosity near the wall resulting from the 
heating. This increases the negative curvature 

of the velocity profile, making the flow more stable 
to small disturbances. The present study is an 
experimental investigation of these predictions, 
using the boundary layer developing on the inside 
wall of a cylindrical tube. This boundary layer is 
thin relative to the tube diameter, so that it 
approximates a boundary layer over a flat plate. 

The numerical predictions of Wazzan et al. are 
based on two-dimensional, linear stability theory. 
The mean flow is assumed plane and parallel, and 
the superimposed small disturbance is described by 
a stream function, 


W(x,y,t) = o(y) exp ia(x-ct) (1) 


Here $(y) is the disturbance amplitude, a is the 
wavenumber and is assumed real, and c is the wave 
velocity which may be complex. The imaginary part 
of c determines whether the disturbance is tempo- 
rally amplified or damped. If we substitute this 
stream function into the Navier-Stokes equations 
and linearize, taking account of the variation of 
viscosity || with distance from the wall y, we find 


at nm 2 " 
GRe Lh (> Ze) Qa 


(WU = e) (OY = a4) = U"> = 


m 
a 6) + 2ut (on! - a6") + uM(o" + a6)] (2) 


78 


In this equation, U(x) is the external flow velocity 
and Re is the Reynolds number based upon free stream 
velocity U,, and boundary layer thickness§5. This 

is known as the "modified Orr-Sommerfeld equation," 
the variable viscosity terms. 

Wazzan et al. have solved Eq. (2) numerically 
for the boundary layer over a heated flat plate, 
using velocity profiles generated by the method of 
Kaups and Smith (1967). The solutions determine 
the critical Reynolds number, which is the lowest 
Reynolds number at which any disturbance has a 
positive amplification rate. The last step of the 
calculation is to relate the critical Reynolds 
number to the transition Reynolds number, using the 
"e to the ninth" criterion of A. M. O. Smith (1957). 
According to this empirical criterion, transition 
occurs when the most unstable disturbance has grown 
to e? (which is 8,103) times its original amplitude. 
The linear theory is used in calculating the growth 
of the disturbance to this amplitude. 

Strasizar, Prahl, and Reshotko (1975) have 
measured growth rates of disturbances generated by 
a vibrating ribbon in a heated boundary layer. They 
found neutral stability curves and were able to 
determine critical Reynolds numbers for wall over- 
heats of up to 5°F (2.8°C). They found that in this 
range of overheats the critical Reynolds numbers 
are in reasonable agreement with the theoretical 
predictions. These experiments were performed at 
moderate Reynolds numbers and did not yield data on 
transition or on stability at higher overheats. 

The results of the Wazzan et al. calculations 
predict that the transition Reynolds number of a 
zero pressure gradient boundary layer should increase 
with wall temperature up to about 70°F (39°C) of 
overheat if the free stream temperature is 60°F 
(16°C). At that overheat, the transition Reynolds 
number should be in excess of 2 x 10° (based upon 
distance from the leading edge). Thus the experi- 
ment designed to investigate these predictions must 
be able to generate a very high Reynolds number 
boundary layer while maintaining low free stream 
disturbance levels. The wall should be very smooth 
and its temperature must be precisely controlled. 
These are the chief considerations that led to the 
experimental geometry described below. 


2. EXPERIMENTAL APPARATUS 


Configuration 


A facility in which water is recirculated through 
the test section was not used for two reasons. (1) 
Heat is continuously added to the test section so 
that a recirculating experiment would require some 
sort of heat exchanger. (2) The free stream tur- 
bulence level in the test section must be less than 
0.05 percent, which has previously been difficult 
to achieve in a recirculating water facility. The 
experiment must then be of the "blow-down" type, 

in which water is removed from one reservoir and 
discharged into another. Run times of more than 
twenty minutes are desired, which requires large 
reservoirs. This led to the selection of the Colo- 
rado State University Engineering Research Center 
as the site of the experiment. Here the water 
supply is Horsetooth Reservoir, which provides 
water to the laboratory through a 0.6 m diameter 
pipe at a total pressure of 6.8 x 10° N/m? (100 1b/ 
in.*). The discharge runs into a smaller lake be- 


HORSETOOTH 
RESERVOIR 


FILTRATION 
TAN 


K 

(7-FT DIA) 24 IN. DIA. 
24 IN. DIA DISCHARGE 
UPSTREAM SETTLING LINE 

TUBE CHAMBER 
(24-1N. DIA 


FLOW TUBE 


BALL VALVE 


36:1 
VIBRATION ORIFICE PLATE 
ISOLATION DISCHARGE 
SECTION 


FIGURE 1. Experimental geometry. 


low the laboratory. At the maximum flow rate of 
this experiment (200 liters/sec), the run time is 
effectively unlimited. 

The flow tube apparatus consits of a settling 
chamber for turbulence management, a contraction 
section, a test section and various types of instru- 
mentation described below. A diagram of the experi- 
mental geometry is shown in Figure l. 


Settling Chamber 


The inside diameter of the settling chamber is 0.6 
m, the same as that of the supply line from the 
reservoir. The test section is 0.102 m in diameter, 
so that the contraction ratio is 35:1. The settling 
chamber is made up of four separable sections, as 
shown in Figure 2. The sections are made of fiber- 
glass to avoid heat transfer through the walls, and 
their total length is 3.35 m. Each end of each 
section is counter-bored to hold a 0.15 m long 
aluminum cylinder with a 1.3 cm wall thickness. 

Each cylinder will hold one or more turbulence 
manipulators, including screens, porous foam, or 
honeycomb material. This design allows the settling 
chamber to be assembled in different configurations, 
so that it can be optimized experimentally. 

The details of the design and optimization of 
the turbulence management system have been reported 
separately [Barker (1978)]. The configuration 
shown in Fibure 2 was arrived at after a great 
deal of testing. There is a considerable body 
of literature on the subject of turbulence 
Management, and this provided some guidelines 
for the optimization of the present system. 

The most detailed recent study is that of Loehrke 
and Nagib (1972), who measured mean velocity and 
turbulence level downstream of various turbulence 
Manipulators. Further recommendations for the 
construction of a turbulence management system 
have been given by Corrsin (1963), Bradshaw (1965), 
and Lumley and McMahon (1967). 

At the downstream end of the settling chamber is 
an additional 0.30 m long section containing porous 
wall boundary layer suction. Hot film anemometer 
surveys in the settling chamber have shown that 
at test section velocities above 9 m/sec (0.26 m/sec 
in the settling chamber) the boundary layer becomes 
turbulent before the flow enters the contraction. 

A thin turbulent boundary layer entering the strong 
favorable pressure gradient of the contraction 


79 


SECTION 


FIGURE 2. Schematic of turbulence manage- 


-— TURBULENCE MANAGEMENT SECTION ————~} 
VIBRATION. C 
}—-—4 FT cot meet IET) Lis Nar |e |_— -CONTRACTION 
pa u SECTION 
/10 PPI FOAM A Loy 
/ 
SUCTION / 
VoSECTIONS Seed 
UPSTREAM 34 MESH | 6 MESH 
BALL SCREEN \ 
VALVE \ 
‘1/8 CELL 1/8 CELL 34 MESH 
HONEY COMB HONEYCOMB. (2 SCREENS) 
3" THICK 3" THICK 


section will tend to "relaminarize," as described 
by Launder (1964) and Back et al. (1969). However, 
this would leave us with unknown initial conditons 
at the entrance to the test section. Therefore we 
have added the suction section to completely remove 
the turbulent boundary layer. This section has a 
0.1m length of porous wall surrounded by an annular 
plenum chamber. The suction flow from the plenum 
is controlled by a valve and a Venturi meter. At 
each test section velocity above 9 m/sec, the 
suction flow is adjusted to the minimum value 
necessary to remove the turbulent boundary layer 

at the contraction entrance. 


Contraction and Test Section 


The 35:1 contraction was designed by a potential 
flow calculation using the method of Chmielewski 
(1974). The length.to diameter ratio of the 
contraction was chosen by balancing the effect of 
relaminarization with that of the Goertler insta- 
bility in the concave-curved portion. A careful 
study of these two effects led to a length to 
diameter ratio of 2.25, which made the contraction 
1.37 m long. The contraction was constructed in 
two sections: a fiberglass upstream half and an 
aluminum downstream half. The joint between the 
two sections is in the region of greatest favorable 
pressure gradient, and has no measurable step across 
Estee 

Recent velocity measurements in the test section 
(discussed below) have led to the design and con- 
struction of a new contraction section to replace 
the original one. The new contraction will have 
an annular bleed flow surrounding an entrance 
section which is all convex. In this way the 
concave-curved wall, which can produce Goertler 
vortices, will be avoided entirely. Results using 
this new contraction will soon be available. 

The flow tube test section is 6.4 m in length 
and 0.102 m in diameter, with a 2.5 cm wall thickness. 
It is made of aluminum, and the inside wall has been 
polished to a surface roughness of less than 10-7 m 
RMS (4 micro-inches). Surface waviness has been 
measured as less than one part per thousand for 
wavelengths less than 2 cm. The tube has been 
optically aligned on site so that it is straight to 
within less than 0.018 cm over its entire length. 
The outside wall is covered with electrical band 
. heaters, which are connected together in groups 
covering about 0.30 m of length. Each heater group 


ment system. 


is servocontrolled by a system which maintains a 
preset temperature on a thermocouple located near 
the inside tube wall. In this way the inside wall 
temperature can be controlled independently of flow 
velocity, and different variations of temperature 
along the tube length can be studied. 

To avoid tripping the boundary layer, no pene- 
trations of the inside wall are allowed except at 
the downstream end. The only instrumentation in 
the test section is an array of thermocouples within 
the wall, spaced along the tube length. At each 
location, there is one thermocouple on the outside 
surface and one in a small hole drilled to within 
0.15 cm of the inside surface. The temperature 
difference between the two thermocouples determines 
the heat flux through the wall at a particular 
location. Since heat flux increases by a factor of 
about ten at the transition point, these temperature 
measurements should provide a good transition 
indicator. A total of 53 thermocouple voltages are 
digitized and recorded. 

During the earlier experiments, there was a single 
hot film anemometer probe at the downstream end of 
the test section. This probe was located within 
the boundary layer and was used to indicate inter- 
mittency only. In the more recent measurements, a 
new instrumented section has been developed and 
installed on the.downstream end of the test section. 
This section is 0.61 m long and its inside diameter 
matches that of the test section to within 2 x 107° 
m. Two types of measurement can be made in the 
instrumented section. Very small Pitot tubes can 
be used to traverse the boundary layer and measure 
mean velocity profiles, and flush mounted hot films 
can determine intermittency at various locations. 

Since the boundary layer is typically less than 
0.5 cm thick, the Pitot tubes must be very small. 
The one being used at present has a cross-section 
of 0.013 x 0.076 cm. The smaller dimension is 
oriented in the direction perpendicular to the wall. 
The tube is traversed from the wall to the free 
stream by a micrometer, which can position it with 
an uncertainty of +0.002 cm. In addition, the 
entire central portion of the tube can be rotated 
in the azimuthal direction so that the Pitot tube 
can be traversed about the circumference of the 
test section. The azimuthal rotation can be per- 
formed while the experiment is running. 

The hot film anemometers in the instrumented 
section are all mounted flush with the wall to avoid 
tripping the boundary layer. The Pitot tubes are 
removed from the section while hot film measurements 


80 


are being made. The films are used only to determine 
intermittency, hence they are not calibrated. There 
are eight hot film locations--two streamwise sep- 
arated stations each having four probes at different 
azimuthal angles. All eight outputs can be displayed 
simultaneously on oscilloscope traces or recorded 

on a photographic strip-chart recorder. 

A high static pressure must be maintained in the 
test section to avoid possible cavitation or out- 
gassing from heated walls. Therefore the pressure 
loss for controlling the flow velocity is located 
at the downstream end of the experiment. Originally, 
a set of sharp-edged orifice plates was used on the 
end of a 1 m long extension tube added to the test 
section. Concern over possible upstream influence 
of the disturbances generated at the orifice plate 
led to the development of a smooth contraction 
section for the downstream end. With the smooth 
contraction, it is possible to maintain laminar 
flow all the way to the exit of the experiment, and 
thus determine transition by flow visualization in 
the exit jet. In addition, a "plug nozzle" has 
been developed, which consists of a strut-supported 
central cone which can be moved in and out of the 
end of the test section. This adjustable exit 
valve permits us to vary the test section static 
pressure independently of flow velocity while main- 
taining laminar flow all the way to the exit. With 
any of these possible exit conditions, the test 
section velocity can be determined from the test 
section static pressure and the known discharge 
coefficient of the nozzle. 


3. RESULTS 


Free Stream Turbulence 


Mean and fluctuating velocities were measured in 
the settling chamber by a cylindrical hot film 
anemometer. The probe penetrated the settling 
chamber wall 0.1 m downstream of the boundary layer 
suction section, and could be traversed from the 
wall to the centerline. Mean velocities and tur- 
bulence levels were measured at many points, and 
turbulence spectra were measured at two or three 
points for each flow condition. In addition, a 
1.2 m long instrumented straight tube could be 
substituted for the 6.4 m test section. This short 
tube contained a Pitot tube, accelerometers, and 
hot film probes. The unheated transition Reynolds 
number was measured in the 1.2 m tube for each 
settling chamber configuration. This Reynolds 
number varied from 800,000 for the empty settling 
chamber with no turbulence manipulators to 5.0 x 
10® for the "best" configuration. This configura- 
tion (shown in Figure 2) includes one piece of porous 
foam, two sections of honeycomb, and four screens. 
The last screen is located 0.3 m upstream of the 
beginning of the contraction, and has a mesh of 24 
per cm. All screens in the settling chamber have 
more than 55 percent open area, in accordance with 
the findings of Bradshaw (1965). 

Detailed results of the velocity measurements 
in the settling chamber have been reported separately 
[Barker (1978)], and are only summarized here. At 
test section velocities less than 9 m/sec, the 
settling chamber boundary layer remains laminar and 
the only effect of the suction is to make it thinner. 
The turbulence level is about 0.07 percent at all 
distances from the wall for the configuration of 


Figure 2. At higher velocities the turbulence level 
near the wall reaches 3 or 4 percent with no suction, 
but remains 0.07 percent at distances from the wall 
greater than 2 cm. As the suction flow rate is 
increased, the mean velocity profile shows thinning 
of the boundary layer and the turbulence level near 
the wall drops rapidly. At the optimum suction 
rate, the highest turbulence level near the wall 

in the settling chamber is about 0.4 percent. The 
suction has no measurable effect upon the mean 
velocity profile or turbulence level more than 2 

cm from the wall. 

The settling chamber velocity measurements and 
the unheated transition Reynolds numbers indicate 
that the turbulence management system is performing 
well. If the turbulence level reduction through 
the contraction is proportional to the square root 
of the contraction ratio [Pankhurst and Holder 
(1952) ], then the turbulence level in the test 
section should be about 0.01 percent. This is 
lower than the turbulence level recorded in most 
wind tunnels, and certainly lower than any previ- 
ously reported water tunnel. 


Transition Reynolds Numbers 


Figure 3 shows measured transition Reynolds numbers 
as a function of wall overheat for the uniform wall 
temperature case. The results on the upper curve 
were obtained with the smooth, laminar flow nozzle 
at the downstream end of the test section, using 
flow visualization at the exit to determine tran- 
sition. The water temperature was approximately 
50°F (10°C) during these tests. Note that the 
transition Reynolds number rapidly increases with 
wall temperature up to 10°F (6°C) wall overheat, 

at which it has reached a value of 42 x 10°®. 

This represents a factor of four increase in tran- 
sition Reynolds number for a relatively small heat 
input. However, above 10°F there are no further 
increases in transition Reynolds number, while the 
theory predicts that it should increase up to about 
60°F (33°C) overheat. Previously published results 
[Barker and Jennings (1977) ] have shown that varying 
the wall temperature distribution does not change 


hh 


/ 
7 SMOOTH 
7 NOZZLE 


ORIFICE 
PLATE 


o 
1 
S 
x 
Ix 
= 7 WAZZAN, ET. AL. 
© B= 0.07 
ONE EXTENSION 
0 5 lo 15 20 25 30 
OVERHEAT, AT (°F) 
FIGURE 3. Transition Reynolds numbers measured at 
exit: one extension tube. 


81 


this result. In fact, uniform wall temperature has 
produced the largest transition Reynolds numbers to 
date. The primary difference between the results 
shown here and those published previously is that 
the present experimental curve reaches the limit 
Reynolds number of 42 x 10° at a lower overheat 
than before. This change is attributed to the 
improvement of the exit conditons with the develop- 
ment of the laminar flow nozzle. 

All of the data of Figure 3 were taken by main- 
taining laminar flow over the full length of the 
tube and observing transition at the exit. If the 
flow velocity is increased further, so that the 
transition region moves upstream in the test 
section, the measured transition Reynolds numbers 
are much lower. In addition, there is a hysteresis 
effect when transition is allowed to move more than 
about 1 m upstream from the exit. That is, to 
restore fully laminar flow over the full tube length 
the velocity must be reduced to a value lower than 
that which previously yielded fully laminar flow. 
This hysteresis may be a phenomenon which is accen- 
tuated by the flow tube geometry. The free stream 
in the flow tube is confined by the boundary layer, 
so that the boundary layer can influence the free 
stream once it becomes turbulent. This free stream 
influence could propagate upstream, which has led 
to conjecture about disturbances from the test 
section exit affecting the transition Reynolds 
number. 

To test this hypothesis of downstream disturbances 
affecting transition Reynolds number, a separate 
study has been conducted to determine the dependence 
of transition upon the tube exit geometry. As dis- FIGURE 4a. Exit jet from smooth nozzle: laminar 
cussed above, there are three types of exit nozzle boundary layer. 
available: orifice plates, the smcoth contraction, 
and the plug valve. In addition, the length of 
unheated tube between the heated test section and 
the exit can be varied from zero to 3.7 m in incre- 
ments of 1.22 m. For each configuration, transition 
can be determined either at the exit itself or at 
the end of the heated section. Transition at the 
exit is easily determined by flow visualization, as 
shown in Figure 4. This photograph of the smooth 
exit contraction shows laminar flow (4a) and turbu- 
lent flow (4b), both at a length Reynolds number 
of approximately 40 x 1O®. ma Figure 4a, note 
the glassy region very near the exit, which soon 
becomes milky in appearance as the air-water shear 
layer undergoes transition. The longitudinal streaks 
in Figure 4a are appraently due to Goertler vortices 
generated in the concave part of the smooth exit 
contraction. They are not seen with the plug valve 
exit, which has no concave region. 

The data of Figure 3 are for one 1.22 m extension 
section on the end of the heated section, followed 
by either the smooth contraction or the orifice 
plate. Transition is measured at the exit in either 
case. Note that the transition Reynolds numbers 
with the orifice plate exit are about 20 percent 
lower than with the smooth contraction, showing a 
definite effect of the exit condition. Figure 5 
shows the same comparison with 2.44 m of unheated 
extension tube between the test section and exit. 
Here we see a much larger difference between results 
with the orifice and with the smooth contraction. 
The smooth contraction transition Reynolds numbers 
are nearly the same as with 1.22 m of extension 
tube, while the orifice Reynolds numbers have 
dropped almost by a factor of two. Clearly the FIGURE 4b. Exit jet from smooth nozzle: turbulent 
effect of the exit condition upon transition Reynolds boundary layer. 


SMOOTH 
NOZZLE 


1076 


oseanea 


5 = 0.07 


Rex, 


ORIFICE 
PLATE 


TWO EXTENSIONS 


° 5 Te) 15 20 25 30 
OVERHEAT, AT (°F) 


FIGURE 5. Transition Reynolds numbers measured at exit: 
two extension tubes. 


number is far more pronounced here than for the 
shorter extension tube length. The most reasonable 
explanation of this lies in the fact that in the 
second case the boundary layer has passed over a 
much longer region of unheated wall, which should 
have a destabilizing effect. This less stable 
boundary layer is then more sensitive to external 
perturbations such as the disturbances created by 
the orifice plate exit. 

As the extension tube length is increased still 
further, the transition Reynolds numbers obtained 
with the smooth contraction begin to decrease. 
Apparently the destabilizing effect of the long 
unheated wall is felt even with the low disturbance 
exit condition. These results indicate that, under 
some conditions, a moderate length of unheated wall 
can be used downstream with no measurable reduction 
of transition Reynolds number. 

When transition is determined at a distance of 
1.4 m upstream of the exit rather than at the exit 
nozzle itself, the influence of the exit condition 
is greatly diminished. Taking the case of the 2.44 
m unheated extension as an example, there is a 
factor of 2.3 difference in the maximum transition 
Reynolds number obtained with the orifice and with 
the smooth contraction when transition is measured 
at the exit (Figure 5). However, when transition 
is measured 1.4 m upstream of the exit, the corre- 
sponding difference is only 15 percent in Reynolds 
number. Clearly the disturbances present at the 
exit nozzle can affect the transition process if 
it occurs near the nozzle, but this influence 
diminishes rapidly as transition moves upstream 
of the exit. Since the highest transition Reynolds 
numbers have consistently been obtained with laminar 
flow over the full length of the tube, most future 
measurements will be made using one of the two 
laminar flow exit conditions. 

Although it is difficult to assess uncertainties 
in transition Reynolds number in this experiment, 
some effort should be made. Results for the highest 
transition Reynolds numbers exhibit a large amount 
of scatter, but most of this can now be attributed 


to variations in the free stream particulate content. 


The purity of the water supply varies considerably 


with weather conditions at the site, and these 
changes in purity have been directly correlated 
with changes in transition Reynolds number. Under 
the most adverse conditions, this effect has reduced 
the maximum transition Reynolds number to less than 
15 x 108 (compared with 42 x 10© for "clean" water) . 
If we compare results that were obtained during 
periods of relatively high water purity, the stan- 
dard deviation in transition Reynolds number is 
about 10 percent of the mean. 

This extreme sensitivity of the results to water 
purity was quite unexpected, and an effort has been 
made to improve the water quality by filtering 
upstream of the settling chamber. Measurements of 
the particle concentration spectrum have been made 
using a Coulter Counter, and some of the results 
are shown in Figure 6. The bands on this figure 
indicate the typical ranges of concentration that 
are obtained in the present experiments, as well 
as in the NSRDC towing basin and the ocean. Note 
that the flow tube particle spectrum has a steeper 
slope than either the ocean or the tow basin, which 
implies that for particle sizes greater than 10 U, 
the flow tube water is much cleaner than the other 
two. The filtration system presently used in the 
flow tube effectively removes all particles larger 
than 100 iL. 

The reason for the strong sensitivity of results 
to relatively minor contamination of the water 
supply is not understood at present. The most 
likely mechanism seems to be a slight increase in 
wall roughness due to the adhesion of particles 
to the wall. Whatever the mechanism, this effect 
will clearly be of importance in hydrodynamic 
applications. 


Comparison with Theory 


Wazzan et al. (1970) have presented numerically 
predicted transition Reynolds numbers for heated 


PARTICLE CONCENTRATION (COUntS/,)) 


FLOW TUBE 
STATION | 


| 10 100 
PARTICLE MEAN DIAMETER (microns) 


FIGURE 6. Particle concentration spectra: flow tube, 
NSRDC towing basin, and open ocean. 


83 


10 ft/sec 
(3 m/sec) 


Ug * 


-0,10 
20 ft/sec 


- 0,08} 


40 ft/sec 


80 ft/sec 
-0,04 


-0,02 


(6.1 m/sec) 


(12.2 m/sec) 


(24.4 m/sec) 


X (FT) 


wall boundary layers with zero pressure gradient. 
More recently, similar calculations have been 
performed for boundary layers in favorable pressure 
gradient flows. Before comparing the flow tube 
results with such predictions we should estimate 
the favorable pressure gradient produced by the 
boundary layer displacement effect in the tube. 
The most common way to characterize streamwise 
pressure gradient in a boundary layer is by the 
similarity parameter 8 [Schlichting (1968)]. For 
the general class of wedge flows, the external 
velocity U is given by U Cx , and the parameter 
8 is then 2m/(m + 1). Both m and & are constants 
in any wedge flow, and are equal to zero for the 
zero pressure gradient boundary layer. We have 
calculated approximate local values of 8 in the 
flow tube, using the Blasius growth law for the 
boundary layer displacement thickness: 


6* = 1.72 (vx/Us) 4 (3) 
(The calculation can be iterated to include the 
effect of pressure gradient upon 6*, but the differ- 
ence is negligible.) The resulting values of 8 as 
a function of x at several values of U_ are shown 
in Figure 7. 8 is proportional to the square root 
of x, and thus has its largest value at the down- 
stream end of the tube. 

Figures 3 and 5, which show transition Reynolds 
numbers versus overheat for the flow tube, also 
include the theoretical predictions of Wazzan et al. 
(1970) for a 8 of 0.07. This represents an approx- 
imate average of 8 in the tube for the velocity 
range of interest. (Calculations using exact 8 
values from the tube will be done in the near future.) 
Note that the experimental results lie near or even 
above the 8 = 0.07 prediction for overheats from 
zero to 13°F (7°C). At this point the experimental 
curve quite suddenly levels out, while the predicted 
curve continues to rise at an increasing slope. 

The predicted curve reaches its maximum at a Reynolds 
number of about 250 x 10° (near 45°C overheat), 

while the experiment has never yielded more than 

AD 16° , 

There are several possible reasons for the 
disagreement between theory and experiment at the 
higher overheats. (1) The theory does not account 
for the destabilizing effects of density stratifi- 
cation, which will become increasingly important as 

“overheat is increased. Buoyancy effects may 


FIGURE 7. 
Uk50 


8 versus x for several values of 


destabilize the flow in three distinct ways: (a) 
the bottom of the tube wall is subject to thermal 
convection rolls, similar in form to the Goertler 
instability; (b) the side wall boundary layer will 
experience a cross-flow due to the rising fluid 
near the wall; and (c) the top wall boundary layer 
will grow in thickness faster than normal because 
of the fluid rising up from the sides. (2) The 
theory neglects the effects of temperature and 
viscosity fluctuations upon the growth of the 
velocity fluctuations. There is evidence that this 
is a reasonable approximation. (3) The theory relies 
upon the e? transition criterion, which may become 
increasingly incorrect at higher overheats. This 
criterion has never before been applied to boundary 
layers with inhomogeneous physical properties. 
There is a large distance between the minimum crit- 
ical point in the boundary layer and the predicted 
transition point using ey ab is questionable 
whether the region of linear growth can extend over 
such a large range of Reynolds numbers. (4) Wall 
roughness is not accounted for in the theory, and 
the importance of roughness will increase with wall 
heating (and with increased velocity) due to the 
thinning of the boundary layer. Roughnesses that 
are insignificant at zero or low overheat may become 
important as overheat increases. 


Velocity Profile Measurements 


In view of the differences between experimental 
results and computed transition Reynolds numbers, 
measurements have been made of boundary layer 
velocity profiles in the flow tube to try to 
establish the mechanism of transition. If the 
buoyancy effects described above are in fact 
significant, they should produce measurable devi- 
ations from axisymmetry in the mean velocity profiles. 
In addition, they might cause transition to occur 
earlier on the top, side, or bottom wall, depending 
upon which mechanism is predominant. We therefore, 
designed the instrumented section (described above) 
to be installed on the downstream end of the 6.1m 
test section. This contains Pitot tubes for mean 
velocity measurements and flush mounted hot film 
probes for intermittency measurements. The instru- 
mented section has been very successful in measuring 
mean velocity profiles in the flow tube. Figure 8 
shows a typical measured profile that has been 


84 


a 


= T T T >I 


@ = 90° (FROM TOC) 
QT= O°F 


VELOCITY RATIO, 
u/Up 


BLASIUS PROFILE 
0.5 


0.5 1.0 
DISTANCE FROM WALL, y/8 


FIGURE 8. Normalized velocity profile with zero over- 
heat, compared with Blasius profile. 


normalized and plotted with a curve representing 

the Blasius profile for a zero pressure gradient 
boundary layer. Actually, the agreement shown 

here is better than it should be due to the positive 
B of the flow tube boundary layer. 

The most surprising result that has been obtained 
with the instrumented section is the large deviation 
from axisymmetry in the profiles, even with no wall 
heat. Figure 9 shows a plot of 6* (displacement 
thickness), 8 (momentum thickness), and H (shape 
factor) versus azimuthal angle for no wall heat 
at a free stream velocity of 1.60 m/sec. The 
dashed lines indicate the calculated values for 
8 = 0 and § = 0.16, which is the value of 8 at the 
downstream end of the test section. The variations 
in 6* and @ are more than 50 percent, which was 
totally unexpected. Figure 10 shows an azimuthal 
velocity profile, that is, u versus ¢ at a fixed y. 
Here we see that the departure from axisymmetry is 
wave-like in nature, and that significant changes 
in velocity occur over a 15° change in $. 

This behavior suggests that the asymmetries may 
be caused by streamwise vortices within the boundary 
layer, which would have a cross-stream length scale 
on the order of the boundary layer thickness. 

Such vortices could be caused by the Goertler 
instability in the contraction section, as described 
above. To test this hypothesis, a new contraction 
section is presently being built which will avoid 


a : [ oe ——~ . = 4 


Oo 90 180 270 360 
AZIMUTHAL ANGLE 


FIGURE 9. Displacement thickness, momentum thickness, 
and shape factor vs. azimuthal angle for zero overheat, 
Us. = 155 cm/sec. 


the Goertler instability entirely. This new con- 
traction will have a fully convex inlet section 
surrounded by an annular bleed flow. All fluid 
from the settling chamber boundary layer will be 
removed by the bleed flow. 

Variations in mean velocity profiles due to 
heating have in fact been measured, but they are 
small relative to the changes with azimuthal angle 
shown in Figures 9 and 10. The shape factor H 
tends to decrease with increasing overheat as 
expected. However, no firm evidence of buoyancy- 
driven instabilities has yet been seen, even at 
low flow velocities and high overheats. 


150 . I ] | 
VELOCITY mt ln I ¢ ll! hy 


ly 
(cm/sec) nett gl J 


50 


: ° ——— eee! es en i =i 
vs. azimuthal angle, 0 30 60 90 '20 EO Tk dD ee ake 


AZIMUTHAL ANGLE 


FIGURE 10. Velocity, u 300 330 360 
( 


¢?, at y = 0.51 cm 


4. CONCLUSIONS 


The flow tube experiment has already demonstrated 
that wall heating can have a significant effect 

upon transition Reynolds numbers in water boundary 
layers. Although the maximum transition Reynolds 
number of 42 x 10° is well below the predicted 
maximum, this value has been obtained with only 

7°C wall overheat. The unheated transition Reynolds 
number of 10’ shows that disturbances are well 
controlled in the experiment. 

Possible causes for the differences between the 
predicted and realized transition Reynolds numbers 
at higher overheats are still under investigation. 
Preliminary results from the instrumented section 
indicate that buoyancy-driven instabilities are 
not a Significant factor. However, major deviations 
from boundary layer axisymmetry have been observed 
even with no wall heat. These perturbations of the 
unheated flow could themselves have an effect upon 
transition Reynolds numbers. This is particularly 
true if the actual disturbances are Goertler vortices, 
because these vortices would increase in strength 
with increasing flow velocity. Since the transition 
length is fixed at the end of the tube in this 
experiment, transition Reynolds number will be 
directly proportional to velocity. Thus the 
Goertler vortices could impose a limit in transition 
Reynolds number if they begin to dominate the 
transition process above some critical flow velocity. 
This hypothesis will be tested by the installation 
of the new contraction section, which eliminates 
the possibility of Goertler vortex formation. 


ACKNOWLEDGMENT 


The author wishes to acknowledge the participation 
and support of the Marine Systems Division of Rock- 
well International, and in particular Mr. Douglas 
Gile. In addition, the author acknowledges the 
Defense Advanced Research Projects Agency, who 
sponsored this research. 


85 


REFERENCES 

Back, L. H., R. F. Cuffel, and P. F. Massier (1969). 
AIAA Journal, 7, 4; 730. 

Barker, S. J., and C. Jennings (1977). The effect 


of wall heating on transition in water boundary 
layers. Proc. of NATO-AGARD Symposium on 
Laminar-Turbulent Transition, Copenhagen, 19-1. 
Barker, S. J. (1978). Turbulence Management in a 
High Speed Water Flow Facility. Submitted to 
ASME. 
Bradshaw, P. (1965). J. Fluid Mech., 22 pt. 4, 679. 
Chmielewski, G. E. (1974). J. Aircraft, 11, 8; 435. 
Corrsin, S. (1963). Turbulence: Experimental 
methods, in Handbuch der Physik, 8, pt. 2, 523. 
Kaups, K., and A. M. O. Smith (1967). The laminar 
boundary layer in water with variable properties. 
Proc. ASME-AIChE Heat Transfer Conf., Seattle, 
Wash. 
Launder, B. E. (1964). Laminarization of the 
Turbulent Boundary Layer by Acceleration, M. I. T. 
Gas Turbine Lab Report 77. Cambridge, Mass. 


Loehrke, R. I., and H. M. Nagib (1972). Experiments 
on Management of Free-Stream Turbulence. NATO- 
AGARD Report 598. 

Lumley, J. L., and J. F. McMahon (1967). Trans. 


ASME, D, 89, 764. 

Schlichting, H. (1968). 
McGraw-Hill, New York. 

Smith, A. M. O. (1957). Transition, pressure gra- 
dient, and stability theory, Proc. 9th Int. Con- 
gress on Appl. Mech., 4, 234, Brussels. 

Spangler, J. G., and C. S. Wells (1968) . 
WOUGNaAly Mola Shea a Sr 

Strasizar, A., J. M. Prahl, and E. Reshotko (1975). 
Experimental Study of Heated Laminar Boundary 
Layers in Water, Case Western Reserve Univ., 
Dept. of Fluid, Thermal, and Aerospace Science 
Report FT AS/TR-75-113. 

Wazzan, A. R., T. T. Okamura, and A. M. O. Smith 
(S68) aransi, ASME, eG, 90) jooR 

Wazzan, A. R., T. T. Okamura, and A. M. O. Smith 
(1970). The stability and transition of heated 
and cooled incompressible boundary layers. Proc. 
4th Int. Heat Transfer Conf., Paris. 


Boundary Layer Theory, 


AIAA 


Some Effects of Several 
Freestream Factors on Cavitation 


Inception of Axisymmetric Bodies 


Edward M. Gates 


University of Alberta 


Edmonton, Canada 


Allan J. Acosta 


California Institute of Technology 
Pasadena, California 


ABSTRACT 


Some of the effects of freestream turbulence and a 
dilute polymer solution on the fully wetted flow 
and the subsequent cavitation inception has been 
investigated for three different bodies. Two of 
these bodies possess a laminar separation and one 
does not. In the fully wetted investigation the 
flow on one of the bodies was found to be insensi- 
tive to the present disturbances whereas the other 
two were found by comparison to be very sensitive. 
Although there is a pronounced "Suppression" of 
inception by the polymer, it seems clear that the 
effects observed are due primarily to the change 
in the real fluid features of the flow past the 
bodies themselves and not to an intrinsic cavita- 
tion process. There appeared to be no special poly- 
mer effect, insofar as cavitation is concerned, on 
the body not having a laminar separation, confirm- 
ing the results of van der Meulen. Due to practical 
limitations the effects of turbulence per se on in- 
ception could not be separately evaluated. 

The inception index on all bodies was found to 
be greatly dependent on the distribution of nuclei 
within the water tunnel. For those cases in which 
a turbulent transition was established well upstream, 
travelling bubbles were a common form of cavitation 
observed on all test bodies. The number of these 
cavitation events were so few, however, that in one 
test facility having a resorber, it was just as 
likely for an attached cavity to form as it was to 
observe a travelling bubble. In both cases the 
inception index was far below the customary minimum 
pressure coefficient reference value. Nuclei counts 
made with the aid of holograms reveal significantly 
fewer microbubbles within the flow of this test 
facility than in those not having a resorber. 


1. INTRODUCTION 


Our understanding of the details of the process of 
cavitation inception (and thus our ability to scale 


86 


laboratory results to prototype conditions) is far 
from complete [e.g. Acosta and Parkin (1975), Morgan 
and Peterson (1977)]. This lack of understanding 

is well illustrated by our ability to do no more 
than indicate reasons which are believed to be 
responsible for the large variations in the results 
of the ITTC comparative test series [Lindgren and 
Johnsson (1966), Johnsson (1969)]. These results, 
some of which are presented in Figures 1 and 2, did, 
however, prompt a considerable amount of effort to 
investigate more systematically the factors influ- 
encing cavitation inception. In particular there 
are three areas in which there have been significant 
developments: (i) the influence of viscous effects 
on inception, (ii) the discovery that in some situ- 
ations the presence of drag-reducing polymers in the 
water cause a suppression of the inception index, 
and (iii) the development of equipment to accurately 
measure freestream nuclei populations. 


1.0 
oO ® . WITHOUT RESORBER FACILITY 
2 OS ** WITH RESORBER FACILITY 
< 
WwW 
a 
z 
0.8 
Zz 
9° TOKYO - (JAPAN 
a | 
Pe OL 
1S) 
Zz Com 
mo! PENN STATE(DTMB-BODY A) Ts _____ | 
roy SAFH -(U.S.A.) 
E “DTMB 36" BODY B 
(U.S.A.) 
FE 05 a | 
= 2 "*>TMB 36" BODY B -(U.S.A.) 
ps NPL 2- 
r3) ta! rah STATE (SAFH BODY) x aclen 
“10 20 30 40 50 60 
TUNNEL VELOCITY-U_ , ft/sec 
FIGURE 1. Results of the comparative inception test 


on a modified ellipsoidal headform sponsored by the 
International Towing Tank Conference, Lindgren and 
Johnsson (1966). 


5. Caltech 


8. SSPA 


Viscous Effects 


Parkin and Kermeen (1953) appear to be the first 
investigators to appreciate the influence of the 
boundary layer on the inception process. However, 
even though their interpretations of the experi- 
mental results were used in many subsequent incep- 
tion theories [e.g. van der Walle (1962), Holl and 
Kornhauser (1969) to name only two] further experi- 
mental investigations of these viscous effects 

were carried out only much later. 

Among these, Arakeri and Acosta (1963), by using 
the schlieren flow visualization technique, were 
able to observe cavitation inception within the 
structure of the flow. A primary feature of the flow 
observed by them was a laminar separation in which 
the cavitation was seen to occur first. There was 
further some suggestion by them that the laminar- 
to-turbulent transition itself may promote cavita- 
tion, perhaps through a mechanism similar to that 
for inception in turbulent pipe flow [Arndt and 
Daily (1969)]. In any case, it should be expected 
then, that any factor which could influence the 
presence of separation or even transition may also 
influence the inception of cavitation. One such 
well-known factor is freestream turbulence. [For 
recent accounts of these effects on transition see, 
e.g., Spangler and Wells (1968), Hall and Gibbings 
(1972), and Mack (1977)]. Unfortunately, the mea- 
surement of turbulence in water is more difficult 
than its aerodynamic counterpart and, until recent 


87 


FIGURE 2. Photographs of dif- 
ferent types of cavitation ob- 
served in the ITTC tests, 
Lindgren and Johnsson (1966). 


times, there has been no great demand for determin- 
ing the freestream turbulence in water tunnels. For 
reference we tabulate in Table I the turbulence 
levels for a few water tunnels for which this inform- 
ation is available (12th ITTC Cavitation Committee) . 


Polymer Effects 


It was inevitable that the much-heralded, drag- 
reducing polymer solutions would be the subject 

of cavitation experiments also. Very early in the 
course of this work Hoyt (1966) and Ellis et al. 
(1970) found that the inception index was reduced 
by as much as a factor of two for hemisphere-nosed 
bodies. There was, furthermore, a pronounced change 
in the physical appearance of the cavitation, once 
it was well developed, as subsequently illustrated 
by the beautiful photographs of Brennan (1970). Two 
possible explanations for the cavitation-suppression 
effect were then advanced: in the first, it was 
speculated that the dynamics of individual bubbles 
were changed by the presence of the polymer, and 

in the second, it was assumed that the basic viscous 
flow about the model was altered by the presence of 
the polymer. Ting and Ellis (1974) could find no 
difference in the collapse time of spark-generated 
bubbles in either water or polymer solutions weak- 
ening the idea that the bubble mechanics are impor- 
tant for this process. Later, however, Holl and 
co-workers (1974) in commenting on experiments 


88 


TABLE I 
Ottawa, Canada 0.75% 
Kriloff No. 2 

Leningrad, USSR 0.4% 
NPL No. 1 

Feltham, UK 0.5% 
MIT 

Massachusetts, USA 0.77% 
6'' Tunnel 

Minnesota, USA 0.8% 


carried out at the Garfield Thomas Water Tunnel 
(GTWT) noted that there appeared to be no laminar 
separation on a hemisphere nose body when polymer 
was added to the water, but no direct flow visual- 
ization was done. Later van der Meulen (1976) 
verified this speculation with the clever use of 
schlieren holography to observe simultaneously the 
viscous flow and cavitation inception on a 10 mm 
diameter hemisphere nose body. His results showed 
clearly that when polymer was injected into the 
boundary layer that the laminar separation was re- 
moved. van der Meulen suggested that the polymer 
removed the separation by causing an early transi- 
tion to a turbulent non-separating boundary layer. 
He then attributed the suppression effect to the 
removal of the large pressure fluctuations associ- 
ated with the transition zone of the free shear 
layer [Arakeri (1975)]. 


Freestream Nuclei 


It is generally accepted that inception begins at 
the nuclei in the liquid and that there are two 
sources for these nuclei--the test body surface 

and the incoming flow. At one time "surface nuclei" 
received considerable attention [e.g. Acosta and 
Hamaguchi (1967), Holl and Treaster (1966), Holl 
(1968), Peterson (1968) and van der Meulen (1972) ]. 
While on the one hand it was shown that under certain 
circumstances, but not in normal cavitation testing, 
surface nuclei could exert a controlling influence 
upon inception. It seemed evident on the other hand 
from the results of the ITTC tests that freestream 
nuclei were the more important. Further, the de- 
velopment of the concepts of cavitation event count- 
ing [Schiebe (1966) ] in conjunction with Johnson 

and Hsieh's (1966) trajectory calculations, the 

idea of “cavitation susceptibility" [Schiebe (1972) ] 
and the development of equipment to measure free- 
stream nuclei populations have led to more interest 
in the influence of freestream nuclei versus surface 
nuclei. In particular, the experiments of Keller 
(1972) have prompted considerable interest in mea- 
suring and relating freestream nuclei populations 

to inception. 

Morgan (1972) has reviewed the various types of 
instruments available for measuring freestream 
nuclei populations and Peterson et al. (1975) have 
made an experimental comparison of three of these, 
namely; light scattering, microscopy, and holog- 
raphy. At the moment holography seems the best 


Turbulence Levels in Some Water Tunnels 


ORL 


Pennsylvania State, USA 0.8% 


HSWT 


California Institute of 


Technology, USA 0.25% 
LIWT 

California Institute of 

Technology, USA 0.05% 

(present work) tor See % 


in that no "calibration" is required, a permanent 
record is obtained, a large volume is sampled, and, 
as Peterson observed, one can determine if the 
nuclei are solid particles or micro-bubbles. 

There is seen to be ample reason then to pursue 
these various freestream factors in inception re- 
search. Two are primarily fluid-dynamic in nature 
and of these the questions concerning freestream 
turbulence levels are of historic interest in fluid 
mechanics and naval architecture. The cavitation 
nuclei however are directly involved in the cavita- 
tion inception process and the recent experimental 
progress cited above make one hope for a more quan- 
titative predictive ability than in the past inso- 
far as inception is concerned. The present work is 
in the mainstream of these observations; briefly we 
report on observations made in two different flow 
facilities having widely different freestream nuclei 
distributions on identical bodies. In one of these, 
the freestream turbulence level is varied over nearly 
a factor of 100 (but not in a condition of cavita- 
tion then) and we confirm and extend the observa- 
tions of van der Meulen on the polymer effect. 
Schlieren photography is extensively used to visu- 
alize thermal boundary layers on the test bodies 
used and in-line holography is used to determine 
nuclei populations in the working section. 

Before discussing these effects we should com- 
ment briefly on the means used for the determination 
of the actual inception observation. A standard 
procedure has been to observe the test body under 
stroboscopic light and to say that inception occurs 
when macroscopic cavities or bubbles become visible 
on the model. However, this method is observer- 
dependent and the trend now is to use cavitation- 
event counters free of human judgment. Ellis et al. 
(1970) and Keller (1972) have developed optical 
techniques which count interruptions of light beams 
which are adjusted to graze the model surface where 
inception has been observed to occur. Peterson 
(1972), Brockett (1972), and Silberman et al. 

(1973) have also determined inception acoustically 
by locating a hydrophone inside the test model. 
There are problems of identifying the types and 
location of the cavitation phenomena occurring with 
these “events." Aside from the question of tech- 
nique, there is also the question of selecting 
appropriate threshold levels at which an event be- 
comes countable and also the event rate at which 
inception is defined to occur. At present there is 
no universal agreement of just what these values 
should be. 


ELLIPSE 


0/6 
0/2 
a 


HEMISPHERE NOSE NSRDC BODY SCHIEBE BODY 
Cemin = 9.75 
(BRASS) (COPPER) (CRES 


FIGURE 3. Definition of test models. 


20) EXPERIMENTAL EQUIPMENT AND METHODS 
Test Models 


Three axisymmetric test models were used in the 
present experiments: a brass hemisphere nose, a 
copper modified ellipsoidal (or NSRDC) body, anda 
stainless steel standard headform from the Schiebe 
series with a minimum pressure coefficient (Cpmin) 
of -0.75. The hemisphere nose and Schiebe bodies 
were fabricated specifically for these tests, 
whereas the NSRDC body is the same as that used 

by Brockett (1972) and Arakeri (1976). Each body 
is 5.08 cm in diameter and has a 0.423 cm diameter 


hole at the stagnation point for polymer injection. 


No quantitative measure of surface roughness was 
made, but each model was highly polished [a highly 
polished surface typically has a 0.1 x 10-7m rms 
finish, Beckwith and Buck (1961)]. The model ge- 
ometries are shown in Figure 3. 

The models were supported by a two-bladed sting 
in the LIWT and by a three-bladed sting in the 
HSWT with the nose of the model being about six 


SPARK-GAP LIGHT SOURCE (0.032"DIA.) 
SLIT. GATHERING LENS F.L.=3" 


REMOVABLE fe ——.——| +] STEADY LIGHT SOURCE 
MIRROR | 


F.L.=12" 2.5 


BIND OWS MAN OR WITH 
CORRECTOR LENS) 


TUNNEL TEST SECTION 
WITH HEATED BODY IN PLACE 


FOCUSING LENS 
F.L.=7", f25 


CUT - OFF PLATE 


FILM PLATE 


FIGURE 4. Schematic diagram of flow visualization 
system. 


89 


body diameters upstream of the sting in each case. 
Misalignment from the geometric tunnel center-line 
in both the LTWT and HSWT was measured to be about 
ORO Ee 


Flow Visualization 


Thermal boundary layers in the viscous flow past 
the test model were observed by schlieren photog- 
raphy. The particular schlieren configuation used 
is shown schematically in Figure 4 and is essen- 
tially the same as that used by Arakeri (1973). 
Also following Arakeri, the prerequisite density 
gradient was produced by heating the body with in- 
ternal cartridge type electric heaters. An example 
schlieren photograph obtained using this system is 
presented in Figure 5. 


Water Tunnel 


The two facilities used in the present experiments 
were the High Speed Water Tunnel (HSWT) and the 


FIGURE 5. A schlieren photograph of a 5 cm diameter 
hemisphere showing laminar separation and turbulent 
reattachment at a body Reynolds number of 2.6 * 10°. 
The maximum height of the separated region is about 
2 mm. 


90 


MUFFLER 
/- TURBULENCE GRID 


/ WORKING SECTION 
f [ ia" SQUARE x 100" LONG 


TUNNEL PRESSURE 
CONTROL VALVE— 


FIGURE 6. Diagram of the Low Turbulence Water Tunnel 
(LTWT) . 


Low Turbulence Water Tunnel (LTWT) both at the 
California Institute of Technology. Since the 
HSWT has been described in detail elsewhere [see 
Knapp et al. (1948) or Knapp, Daily, and Hammit 
(1970) ], it will only be noted here that one, it 
has a resorber and two, the freestream turbulence 
level has been measured to be about 0.2 percent by 
Professor S. Barker. The LTWT [Vanoni et al. 
(1950) ] is also a closed loop recirculating tunnel; 
but, as can be seen in Figure 6, it has no resorber. 
In this facility the maximum test section velocity 
and minimum cavitation number are approximately 8 
meters per second and 0.3 respectively. The unique 
feature of the LTIWT is that the freestream turbu- 
lence level in the test section can be varied from 
a very low value (for water tunnels) of 0.05 per- 
cent to a high value of 3.6 percent. The low tur- 
bulence level is obtained by use of small turning 
vanes in each elbow of the circuit, a yery gradual 
diffuser (included angle is 3°13'), a nozzle with a 
16:1 contraction ratio, and by turbulence. damping 
screens and honeycombs in the "stagnation" section 
of the tunnel just upstream of the nozzle. The 
configuration of screens and honeycombs which pro- 
duces the 0.05 percent turbulence level is shown 
schematically in Figure 7 (with the exception that 
no turbulence generating grid is installed) and is 
based upon the results of Loehrke and Nagib's (1972) 
report. 

By inserting different turbulence generating 


grids into the tunnel circuit the turbulent intensity 


can be gradually increased from 0.05 to approxi- 
mately 3.6 percent. The description of these grids 
is as follows: 


HONEYCOMB 
"7" TRIANGULAR CELLS 

TURBULENCE DAMPING SCREENS 
0.0075" DIA. WIRE, 22 meshes/lineal inch 


TURBULENCE 
GENERATING GRID 


SECOND HONEYCOMB 
1/8 x2 HEXAGONAL CELLS 


FIGURE 7. Sketch of LTWT contraction nozzle showing 
the turbulence manipulators. 


grid No. 1: 12.7mm diameter bars with 50.8mm 
mesn 

grid No. 2: 6.35mm diameter bars with 25.4mm 
mesh 

grid No. 3: three 25.4mm diameter horizontal 
bars on 76.2mm centers 

grid No. 4: 0.635mm diameter fishing line with 


19.05mm mesh 
Grids 1, 2, and 4 are located at the entrance to 
the test section as is shown in Figure 7 (the 
distance from these grids to the test model is 
approximately 1.2 meters). Grid No. 3 is located 
in the "stagnation" section immediately after the 
final turbulence damping screen. Grid No. 3 has 
this particular configuration because (after much 
trial and error) it was found to produce a turbu- 
lence level which is close to the levels measured 
in a number of other facilities--see Table 1. 

A DISA constant temperature anemometer was used 
to measure the turbulence levels in the test section. 
The probe was a wedge-shaped hot film type and was 
firmly mounted on the tunnel center-line at the 
model position (1.2 meters from the test section 
entrance). The results of these measurements have 
been summarized in Figure 8. 


Polymer Injection System 


The injection approach of introducing the polymer 
into the boundary layer versus filling the tunnel 
with a polymer solution (polymer ocean) was chosen. 
After considering a number of injection configura- 
tions [Wu (1971)] it was decided to follow van der 


5 
GRID #| 


e) Oo 


O D0 oo g GRID #2 


GRID #4 
A IS & & B& A 


GRID #3 
Ga BS 8 g@ Bp 


0.5 


TURBULENCE LEVEL - u‘/U, percent 


0.1 


FREESTREAM 
e 


0.05 
® @ 


O 5 10 i) 20 25 
TUNNEE VEEOCIinY SU titZselc 


FIGURE 8. Summary of turbulence intensity measurements 
in the LTWT. 


j-2:! CONTRACTION 


INJECTOR ~ foie BRASS 


SMOOTHING SECTION 
PACKED WITH POROUS FOAM 


Luoves FOR 
CARTRIDGE HEATERS 


FIGURE 9. Cross-section of injector used for these 
polymer experiments. The body diameter is 5 cm. 


Meulen's (1973) example and inject the polymer into 
the boundary layer through a hole at the stagnation 
point. To do this an injector was designed to in- 
troduce the polymer into the boundary layer without 
also introducing disturbances. The injector is 
shown schematically in Figure 9 assembled inside 
the hemisphere nose body and consists of first a 


settling chamber 12.7mm in diameter and 31.75mm long. 


This section was packed with porous plastic foam 
held in place by a sintered brass disc. The pur- 
pose of this section is to disperse the jet enter- 
ing the injector and provide a smooth flow into the 
9:1 contraction which follows. After the smooth 
contraction there is a tube with a length to diam- 
eter ratio of 22 and this tube ends at the surface 
of the model. 

To minimize polymer degradation, the polymer 
solutions were "pushed" through the injector from a 
reservoir by using compressed air instead of a pump. 
A check with a turbulent flow rheometer [the same 
one as used by Debrule (1972) ] showed degradation 
of the polymer after it passed through the injec- 
tion system to be minimal. Preliminary tests were 
carried out with water as the injectant to ensure 
that the injection process itself was not respon- 
sible for any observed changes in the flow. Results 
of these tests for the NSRDC body are presented in 
Figure 10 and show that even at an injection rate 
of three to ten times higher than actually used 
with polymer solutions no differences are detectable 
from the no-injection case. 


Nuclei Counter 


Nuclei distributions were deduced from holograms 

of the test fluid. The experimental apparatus and 
method is much the same as used by Peterson (1972), 
Feldberg and Shlemenson (1973) and is described in 
detail in Gates and Bacon (1978). Essentially it 
is a two-step image forming process. In the first 
step, a hologram of a sample volume of the water 

in the tunnel test section is recorded on a special 
high resolution film by a "holocamera." In the 
second step, the developed hologram is reconstructed 
producing a three dimensional image of the original 
volume which can be probed at the investigator's 
leisure. The holocamera and reconstruction system 
are shown schematically in Figure 11 and 12 respec- 
tively. 


91 


(b) 


FIGURE 10. Schlieren photographs showing the effect 

of injecting water on the NSRDC body at a body Reynolds 
number of 3.2 x 10°, (a) injection rate = O m&/sec, 

(b) 1.8 m&/sec, (c) 3.6 m&/sec, (d) 6.6 m&/sec, 

(e) 9.8 m&£/sec. No effect is observed. 


SAMPLE VOLUME 


TEST SECTION 
WINDOWS 


BC D 
1 


FIGURE 11. Diagram of the holocamera; (a) dielectric 

mirror, (b) iris, (c) dye-quench cell, (d) ruby-flash 

lamp assembly, (e) iris, (f£) dielectric mirror, 

(g) beam splitter, (h) neutral density filter, 

(1) beam expander lens, (j) 25y pinhole, (k) collimat- 
ing lens, (1) front surface mirror, (m) p.i-.n. diode, 

(n) film pack. 


ics) 
No 


[icc alee ee : and also to let the freestream bubbles go back into 
solution or rise to the high points in the tunnel 
TV CAMERA circuit. 


The same general test procedure was used in the 
HSWT except for small differences in pressure mea- 


MICROSCOPE surement. However, desinent cavitation observa- 
tions were also made in this facility. All holograms 
=== TV MONITOR made in the HSWT were done without the model in 
=>==2 RECONSTRUCTED place but at conditions of velocity and pressure at 
Ty 7t Wee which inception had been observed to occur. 


4. PRESENTATION AND DISCUSSION OF FULLY WETTED 


HOLOGRAM ON TRAVELING RESULTS 


CARRIAGE 


ig} b L aL 
— BEAM DIAMETER ~5cm reestream Turbulence Levels 


COLLIMATING LENS CSN ACIOAS 


The influence of gradually increasing freestream 
turbulence level upon the viscous flow about each 
test body is illustrated in the sequences of 


PIN HOLE 


|| ~~ microscore OBJECTIVE schlieren photographs presented in Figures 13 

‘| through 15. In each sequence of photographs the 
test body is seen in silhouette and the flow is 
from right to left. The magnification is such that 
the surface length shown in these photographs is 


He-Ne GAS LASER (5mw) 


FIGURE 12. Arrangements to reconstruct and read the 
holograms. 


3. GENERAL EXPERIMENTAL PROCEDURES 


Before any experiments were carried out, the water 
in each facility was de-aerated to reduce the 
number of freestream air bubbles produced in the 
tunnel circuit. This was of particular importance 
in the LIWT which has no resorber. During the 
present tests the air content in the LIWT was typ- 
ically between 7 - 8ppm whereas in the HSWT it was 
between 9 - 10ppm (air content levels were measured 
with a van Slyke blood gas analyzer). At these air 
contents there were very few macroscopic air bubbles 
visible in the flow approaching the model in the 
HSWT (as will be seen later). However, in the LTWT 
there were always many macroscopic air bubbles 
easily visible in the approaching flow. 

In a typical cavitation test in the LTWT, the 
tunnel velocity and polymer injection rate (if any) 
were first adjusted to the desired values. Incep- 
tion was then obtained by reducing as rapidly as 
possible the tunnel static pressure until the pres- 
ence of cavitation was visually observed on the 
model under stroboscopic illumination. At the 
point of inception, a schlieren photograph, a holo- 
gram, the tunnel velocity, and the tunnel static 
pressure were recorded simultaneously. Each test 
had to take less than forty seconds since by that 
time the abundant supply of cavitation bubbles gen- 
erated at the pump would reach the test section and (e) 
dramatically change the freestream conditions. Af- 
ter each test, the tunnel pressure was raised to on the flow past the NSRDC body (the flow is right to 
about one atmosphere and the tunnel allowed to cir- left) at a body Reynolds number of 1.6 x 105: (a) u'/v 
culate for five minutes. This recess between each = 0.05 percent, (b) 0.65, (c) 1.1, (a) 2.3, (e) 3.6 
test was required to let the ruby laser cool down percent. 


FIGURE 13. The effect of freestream turbulence level 


(d) 


FIGURE 14. The effect of freestream turbulence level 
on the flow past the hemisphere body at a body Reynolds 
number of 2.6 * 10°. (Same turbulence values as in 
Figure 13.) 


approximately 10mm. As can be seen in the first 
photograph of each of Figures 13 and 14, the NSRDC 
and the hemisphere nose bodies respectively have a 
laminar separation. Transition on these bodies oc- 
curs on the resulting free shear layer and the flow 
subsequently reattaches as a turbulent boundary 
layer. With increasing turbulence intensities the 
point of transition on the NSRDC body moved upstream 
on the free shear layer. As the position of tran- 
sition moved forward, the size of the separation 
bubble decreased until finally it disappeared when 
the position of transition and separation coincided. 
Once the point of transition moved upstream of the 
point of separation, no further observations of the 
thermal boundary layer could be made with the pres-— 
ent schlieren system. Unlike the NSRDC model, the 
increasing turbulence level seemed to have no ef- 
fect upon the viscous flow about the hemisphere 
nose body--as can readily be seen in Figure 14. 
This rather surprising result will be returned to 
later. 

As is shown in the first photograph of Figure 15, 
the Schiebe body has no laminar separation and tran- 
sition occurs on the model surface rather than on a 
free shear layer. With increasing freestream tur- 
bulence level two effects were noted; first, as can 
be seen in Figure 15, the position of transition 
moves substantially upstream and secondly, the ap- 
pearance of the disturbance appears to change. This 
change is not quite so evident in only a few pic-— 
tures, but we believe we observe more-or-less peri- 
odic and highly amplified boundary layer waves in 
Figure 15a and even b. However, for the higher 


93 


turbulence levels frequent “bursts" interspersed 
with a periodic phenomenon seemed to be more common. 
A random collection of schlieren photographs of the 
same body (Figure 16) at an intermediate turbulence 
level shows these various forms more clearly. 


Discussion 


To quantify the effects of turbulence level, the 
position, length, and maximum height of the separa- 
tion bubble were measured for the NSRDC and hemi- 
sphere nose bodies. For the Schiebe body, which 
has no separation, the position of transition was 
recorded--the position of transition being defined 
as that point at which the first noticeable dis- 
turbance occurs in the laminar boundary layer. These 
quantities are defined in Figure 17 and were mea- 
sured directly from the negatives of the schlieren 
photographs with the aid of a scale or reference 


(e) 


FIGURE 15. The effect of freestream turbulence level 
on the flow past the Schiebe body at a body Reynolds 

number of 2.5 x 10°. The turbulence levels are those 
in figure 13 and the regions shown are, at arc-length 

diameter ratios of (a) 0.82-1.07, (b) 0.76-1.01, 

(c) 0.60-0.85, (d) 0.61-0.86, (e) 0.47-0.63. 


94 


FIGURE 16. Random photographs of the flow past the 
Schiebe body at a Reynolds number of 3.4 x 10° with 
background turbulence level of 1.1 percent. The 
region shown covers the arc length diameter ratio of 
0.68 to 0.93. 


(a) POSITION OF LAMINAR SEPARATION OR 
TRANSITION WHICHEVER IS APPLICABLE 


U eet 


H 


(b) LENGTH AND HEIGHT OF SEPARATED REGION 


FIGURE 17. Definition sketch of separation location. 


079 
HEMISPHERE NOSE BODY 


[o) 
1 
x 


[o) 
NI 
a 


ESTIMATED SEPARATION 


082 LOCATION (THWAITES METHOD) 


NSROC BODY 


TURBULENCE LEVEL 
0.05% 
0.65% 

O78} 1.10% 


° 
8 


2.30% 
3.60% 


POSITION OF SEPARATION/ DIAMETER -(S/D), 


15x10° 25x10° 35x10° 45xi0> 
BODY REYNOLDS NUMBER- UD /y 

FIGURE 18. Observed separation locations as a func-— 

tion of turbulence level for two bodies. 


negative. Note that the position of transition on 
the free shear layer coincides with the definition 
of the end of the separated bubble. 

Each of these measured quantities was non- 
dimensionalized by dividing by the body diameter 
and are plotted versus the body Reynolds number 
with the freestream turbulence level as a parameter 
in Figures 18 through 21. For the NSRDC body, Fig- 
ure 19 shows that the size of the separation bubble 
decreases with increasing velocity and turbulence 
level--the critical Reynolds number being reduced 
from a value of greater than 4 x 10° at 0.05 joene= 
cent to near 2.5 x 10° at 3.6 percent. As was ex- 
pected from the schlieren photographs of the 
hemisphere nose (Figure 14), Figure 20 shows the 
length of the separation bubble is independent 
of turbulence level but decreases with increasing 
velocity. Finally, Figure 21 shows that as with 
the NSRDC body, the position of transition on the 
Schiebe body moves forward with increasing velocity 
and turbulence intensity. 

The most startling result of the above tests was 
the insensitivity of the boundary layer on the 
hemisphere nose to the present disturbances imposed 
by the freestream turbulence. Hall and Gibbings 


TURBULENCE LEVEL 
@ 005% 
0.65% 
1.10% 
2.30% 
3.60% 


o 
ie 
ce} 


0.08 


‘s 
(e) 
5 


SEPARATION LENGTH/BODY DIAMETER 
° 
n 


fo} 


1.5x10> 2.5x10° 3.5x105 4.5x10° 
BODY REYNOLDS NUMBER-UD/v 


FIGURE 19. The length of the separated region as a 
function of freestream turbulence level for the 
NSRDC body. 


TURBULENCE LEVEL 
0.05% 
0.65% 


0.09 1.10% 
2.30% 


3.60% 


0.08 


0.07 


0.06 


0.05 


SEPARATION LENGTH/ BODY DIAMETER 


15x10° 25x105 35x10 45x10° 
BODY REYNOLDS NUMBER - UD/v 


FIGURE 20. The length of the separated region as a 
function of freestream turbulence level for the 
hemisphere body. 


(1972) have summarized the available experimental 
data and semi-empirical correlations at that time 
for the combined effects of pressure gradient and 
freestream turbulence level upon transition. How- 
ever, this correlation does not predict an insensi- 
tivity to increasing turbulence levels. No doubt 
this discrepancy is related to the question of how 
the freestream disturbances are assumed to inter- 
act with the boundary layer. For example, van 
Driest and Blumer (1963) accounted for the effect 
of freestream turbulence by using Taylor's assump- 
tion that the unsteady perturbation induced in- 
stantaneous variations in the velocity gradient. 
But, as just noted, this type of correlation did 
not work. Later, Spangler and Wells (1968) demon- 
strated that not only the intensity, but also the 
energy spectrum and the nature of the disturbance 
must be taken into consideration. Reshotko (1976) 
and Mack (1977) have re-emphasized Spangler and 
Wells' conclusions and pointed out the lack of un- 
derstanding of the interaction mechanism between the 
freestream disturbance and the boundary layer* is 
one of the major obstacles in the consistent predic-— 
tion of transition. Thus, although the effect of 
freestream turbulence on these bodies cannot be pre- 
dicted with confidence, we at least may offer some 
speculation based on these ideas to explain the be- 
havior on the hemisphere nose body. 

It is readily possible using the approximate 
method of transition prediction suggested by Jaffe 
et al. (1970) in conjunction with the stability 
charts for the Falkner-Skan profiles computed by 
Wazzan et al. (1968b) to determine the critical 
frequency, or most unstable frequency for growth, 
for each body at a number of velocities. These 
estimates are presented in Table 2. We then esti- 
mate with the aid of measured energy spectra of 
grid generated turbulence, Tsuji (1956) that there 
is approximately sixty times as much energy avail- 
able in the freestream at the critical frequency of 
the NSRDC body than there is at the critical fre- 
quency of the hemisphere nose body. Furthermore, 
the distance from the position of neutral stability 
to the position of separation is only 0.07 diameters 
on the hemisphere nose model whereas on the NSRDC 


*This is the concept of boundary layer receptivity devel- 
oped by M. V. Morkovin [see the review of Reshotko (1976) ]. 


95 


T ar 
a TURBULENCE LEVEL 
= @ 0.05 % 
SS B® 0.65 % 
7p) ai.1% 1 
7 © 2.3% 
Zz 4 3.6% 
[e) o— 
= ie 
2p) = 
E Ss 
ra 6 | 
[= —~S 
Ww a ™: 
fo) nia 
Zz 
© 
= 
w 
Qa 


HE +t ee 
1x 10° 2x10° 3x105 4x0 ~ 
BODY REYNOLDS NUMBER-UD/v 


FIGURE 21. The location of transition on the Schiebe 
body as a function of turbulence level. 


body it is 0.40 diameters. Thus on the NSRDC body 
not only is there considerable more energy available 
at the critical frequency, but there is also more 
opportunity for disturbances to grow than for the 
hemisphere nose body. This same trend is also found 
for the Schiebe body at the low turbulence levels. 
The critical frequencies are even less than those 

of the NSRDC model (Table 2). There is, therefore, 
more energy available at those frequencies than even 
on the NSRDC model. Finally, the distance from the 
position of neutral stability to transition is be- 
tween 0.40 to 0.60 diameters--much the same as for 
the NSRDC model. 

We find it somewhat reasonable then, in retro- 
spect, for the hemisphere body to be found insensi- 
tive, in the present experiments, to the freestream 
disturbances. Regrettably, the present visual 
observations are not sufficiently quantitative to 
shed light on this basic problem of boundary layer 
receptivity to external disturbances and their sub- 
sequent growth into turbulence. 

By using an oil film technique, Brockett (1972) 
found the NSRDC model to have a critical velocity 
of 2.8 meters per second at 20°C and Peterson (1972) 
reports 4.2 meters per second at 10°C in the NSRDC 
12-inch water tunnel. The same body in the HSWT 
was found to have a critical velocity of about 9.2 
meters per second and it was observed to be above 
7.6 meters per second in the LTWT at 0.05 percent 
turbulence level. To reduce the value of the criti- 
cal velocity to 4 meters per second in the LIWT re- 
quired a 316 percent turbulence level, which is as 
can be seen from Table 1 a very high value for a 
water tunnel test section. (Initially it was thought 
unlikely that the disturbance level in the NSRDC 
facility is this high. However, after inspecting a 
drawing of the facility [Figure 2.3 pg. 26, Knapp 
et al. (1970)] such a high level does not seem so 
unlikely.) However, in this as well as in most 
water tunnel facilities the energy spectrum is not 
known, forestalling therefore a direct comparison 
of transition phenomena. 

The present observations of transition on the 
Schiebe body at the lowest turbulence level are 
compared with calculations of Wazzan* and experi- 


*Private communication. 


96 


TABLE II 
for Several Bodies 


Approximate Critical Boundary Layer Frequencies 


Rep Hemisphere Nose NSRDC Schie be 
(Hz) (Hz) (Hz) 
aEOWaex 10° 1070 670 -- 
Ze, 0) 2 10° 1800 1060 350 
3,33 % 19> 2140 1780 650 
Bear a10r 3350 2100 1200 


ments of van der Meulen (1976) in Figure 22. There 
is good agreement of the experimental results and 
also with Wazzan's e? calculations. 


Polymer Injection 
Observations 


The influence of gradually increasing the injection 
rate of the polymer solution upon the basic flow 
about each test body is illustrated in the sequences 
of schlieren photographs in Figures 23 through 25. 
In Figure 23(a) the maximum height of the separa-— 
tion bubble is 0.5mm and on the hemisphere nose 

body in 24(a) the maximum bubble height is 0.25mm. 
Unlike the freestream turbulence level, the presence 
of polymer in the boundary layer was found to 
influence the basic viscous flow on all the test 
models. As can be seen in the schlieren photographs, 
as the polymer injection rate was increased the 
position of transition moved upstream in each case. 
For the NSRDC and hemisphere nose models a critical 
injection rate was reached at which the positions of 
transition and separation coincided and the laminar 
separation was eliminated. At injection rates above 
this critical value the position of transition ap- 


Be © van der MEULEN (1976) 
= @ PRESENT OBSERVATIONS 
— (0.05 PERCENT 
gy 1.47 TURBULENCE LEVEL) al 
arid B® CALCULATIONS NO 
z HEATING 
oO L 4 CALCULATIONS 10°F | 
= 1.2 HEATING (WAZZAN & 
= GAZLEY, 1978) 
ep) 
za 
CO 

= 1.Or ° SA s 4 
e 2) i ~~ 
ve — 
© 0.8 2 
<a 
ie) 
F 06 
i) 
oO 
a 

— 4 4 ~~ 4 jt —d 

1 x 10° 2x 10° 3x10> 4x10° 
BODY REYNOLDS NUMBER-UD/y 

FIGURE 22. Comparison of transition observations on 
the Schiebe body. 


(d) 


FIGURE 23. Flow past the NSRDC body with injection 
of 500 wppm Polyox (WSR 301) at a Reynolds number of 
1.6 x 10°: (a) no injection; (b) 0.1 mg/sec, G = 
0.5 x 1055; (c)) 0.3) me/sec, G = 15) x) dogo; (a) ons 
mi/sec, G = 2.5 x 10-©. G is the dimensionless 
polymer injection rate. 


peared to move further upstream, but with the limited 
resolution of the present schlieren system, these 
poisitions could not be accurately determined. 


Discussion 


It would seem desirable to normalize somehow the 
injection rate of polymer fluid. We have chosen 
to do this by dividing the mass flux of polymer 


FIGURE 24. Flow past the hemisphere body with injec- 

tion of 100 wppm Polyox at a Reynolds number of 3.9 x 

10°. The dimensionless injection values are: (a) G = 
0, no injection, (b) 0.5 x 1076, (c) 1.1 x 1078, 

(G) 1.7 & 1078, (@) 2.9 2 10-9. 


material by the mass flux of the boundary layer 
displacement flow. Although this is an arbitrary 
normalization, in the present experiments the dis- 
placement effect of the injectant fluid was always 
much less than the boundary layer displacement 
thickness, 6,. Thus we define a quantity G 


za cQ 
SS npn ome 
wo ts 
where c is the polymer concentration (weight basis) 
in the injectant Q the volume flow rate of injectant 
(basically the same fluid as the test medium) with 
D and V. being the body diameter and tunnel ve- 
locity respectively. For the NSRDC and hemisphere 
models dtg5 was calculated at the position of the 
laminar separation whereas for the Schiebe body it 
was arbitrarily calculated at S/D = 1.00. The pres- 
ent results, so normalized, are presented in Fig- 
ures 26, 27, and 28. As with the freestream 
turbulence level, no change in the position of 
separation on the NSRDC and hemisphere nose models 
was observed when polymer was injected into the 
boundary layer. 
The results of the experiments show the presence 
. of very small quantities of Polyox to be destabiliz- 


97 


ing to the laminar boundary layers on the present 
test models. This destabilization effect has been 
observed before: in fully developed cavity flows 
past spheres and cylinders Brennen (1970) observed 
distortions in the cavity surface and separation 
line due to the presence of polymer. Brennan at- 
tributed the changes in cavity appearance to a 
polymer induced instability in the wetted surface 
flow on the headform. Sarpkaya (1973, 1974) in- 
vestigated the flow of dilute polymer solutions 
about cylinders and several airfoils and also ex- 
plained his observations by suggesting a polymer 
induced instability in the laminar boundary layer. 
Some later experiments by Tagori et al. (1974) 
support some of Sarpkaya's speculation for one of 
the airfoils. 

A destabilizing effect is rather contrary to the 
general impression obtained from the available lit- 
erature on the effects of drag-reducing polymers on 
fluid friction [see for example Hoyt (1972)]. We 
were unable however, to find in the available lit- 
erature any satisfactory explanation of the effect 
on transition of the polymer fluids. 


FIGURE 25. Flow past the Schiebe body at a Reynolds 
number of 4.2 x 10° with injection of 500 wppm Polyox. 
The dimensionless injection parameters are, (a) G = 0, 
(3) 23 8 1O-, (6) 1.5 = 16-5, @) 2.9 % 20°53, mach 
frame is 0.2 body diameters in length and they are 
centered at arc length ratios of 0.82, 0.75, 0.6, 
0.53, respectively. 


ite} 
iss) 


NO INJECTION 
G~0.25 x 107 
G~0.50 x 1o7& 
G~1.00x107& 


018 


al4 


Pad 


010 


—— (5) 


0.04 


SEPARATION LENGTH/BODY DIAMETER 


) D A 
ix105 2xi0> 3x10° 4xio> 
BODY REYNOLDS NUMBER - UD /v 


FIGURE 26. The length of the laminar separation as a 
function of polyox injection on the NSRDC body. 


Comparison of Present Results with those of van 
der Meulen 


van der Meulen (1976) has studied the influence of 
dilute polymer solutions (Polyox, WSR 301) upon the 
fully-wetted flow and cavitation inception for a 
hemisphere nose body and was the first, to our 
knowledge, to observe the Schiebe body (Comin = 
-0.75). He also was the first to inject the polymer 
solution at the stagnation point. To observe the 
flow on the test models, van der Meulen used pulsed 
ruby laser holography. However, to make the flow 
visible a salt was added to the polymer solution. 
In his case the injectant was a 2 percent salt-- 
500wppm Polyox solution. ‘ 

On the hemisphere nose body he observed that the 
injection of the salt-polymer solution eliminated 
the laminar separation and he further speculated 
that the polymer caused an early transition to a 
turbulent non-separating boundary layer. On the 
Schiebe body, which has no laminar separation, the 
laminar to turbulent transition point was found to 
move upstream of the no-injection position. The 
present results for this body are seen to agree 
qualitatively with those of van der Meulen (Figure 
28), although the deduced injection rates of the 


G = 0 (NO INJECTION) 
G~0.5x107& 
G~1.0x107& 
G~1.5x107& 


0.10 


008 


006 


004 


ace 


SEPARATION LENGTH/BODY DIAMETER 


\x105 3x105 5x10> 7x105 9x10° 
BODY REYNOLDS NUMBER - UD/v 


FIGURE 27. The length of the laminar separation as a 
function of polyox injection on the hemispehre nose 
body . 


@G=0, PRESENT RESULTS 

~2xlO& PRESENT RESULTS 
12 < 7xI10® ” ” 
~13x10°° 
~I5x10"© van der MEULEN (1976) 
~20xI0°° " " " 


9° 
2) 


ARC LENGTH LOCATION OF TRANSITION-(S/D), 
fe) c 
gS @ 


1 x 10% 2x10? 3x10° 4x10° 
BODY REYNOLDS NUMBER-UD/v 


FIGURE 28. The position of transition on the Schiebe 
body as a function of polymer injection. 


latter are rather larger. Even though freestream 
conditions of these two tests may not quite be the 
same, it is evident because of the nearly one order 
of magnitude change in Reynolds number that the 
polymer fluid is the chief agent of boundary layer 
instability. 


5. EFFECT OF FLOW VISUALIZATION ON TRANSITION 


It is now well documented that heating a laminar 
water boundary layer, tends to stabilize it [see 
for example Wazzan et al. (1968a, 1970)]. This 
point was further discussed with reference to the 
hemisphere and ITTC test bodies by Arakeri and 
Acosta (1973) who concluded that for the separating 
flows of these bodies, the effect of heating was 

on the order of only a few percent. Since the heat-— 
ing rate and velocity ranges are similar in the 
present experiments, it is expected that the in- 
fluence of heating on the hemisphere and NSRDC 
bodies is not Significant. However, there is some 
question as to the influence of heating on the non- 
separating flow on the Schiebe body. Shown in Fig- 
ure 22 are averaged observed values of the position 
of transition calculated by Wazzan with and without 
wall heating. First, it can be seen that there is 
good agreement between Wazzan's calculation for an 
unheated boundary layer with e? amplification and 
the observed position of transition. However, the 
point to be noted is that (with a wall temperature 
10°F above the ambient water temperature) these 
same calculations predict a 40 percent delay in 
transition at Rep = 2.5 x 10°. This would suggest 
that wall heating is important although not perhaps 
sufficient to alter major trends in the present ex- 
periments. There is however the qualification that 
the calculation assumes a constant wall temperature 
while this is not the actual case. 

An attempt to measure the actual wall tempera- 
ture was made by installing six thermocouples near 
the surface of the model at positions of S/D = 0.4, 
0.6, 0.8, 1.0, 1.2, 1.4. The position of neutral 
stability on this body is S/D = 0.37 and the average 
position of transition varied from S/D = 1.0 to 
S/D = 0.8. Since it is the heating in the boundary 
layer prior to transition that is of importance, 


the values of the wall temperatures at S/D = 0.4, 
0.6, 0.8, 1.0 were of the most interest. The total 
heat flux was set at 250 watts (about 3W/cm2) at 
which the schlieren effect was observable. The 
wall temperatures were then measured at increasing 
values of velocity. It was found that the maximum 
wall temperature between S/D = 0.4 and 1.0 varied 
from 3°C to 5°C above the ambient temperature. 
However, it must be emphasized that these are very 
conservative values since the thermocouples are 
actually somewhat below the surface in a region of 
a high temperature gradient. When this gradient 

is accounted for our estimate of the surface excess 
temperature is from 1-3°C, a smaller but not neg- 
ligible amount. van der Meulen avoided the tem- 
perature effect by injecting a two percent salt 
solution. On the whole this method and the present 
one agree quite favorably (Figure 22). There is, 
however, the possibility of instability via a de- 


(c) 


FIGURE 29. Schlieren photographs of the Schiebe body 
with and without salt water injection. The top photo- 
graph of each group is without injection; the bottom 
photograph shows the injection of MgSO, solution having 
a specific gravity of 1.02. The Reynolds number is 
Ue67) <9102in) (a) 26150) x) 102 in (((b)),, and) 3433) <1 dloe 

in (c). 


99 


0.65% TURBULENCE LEVEL 
fos. 


BAND TYPE 
BUBBLE TYPE 
LOWEST AIR CONTENT 


HIGHEST AIR CONTENT 
(BROCKETT 1972) 


(eo) 
@o 


fo) 
a 


fo) 
u 


CAVITATION INCEPTION NUMBER - Oj 
° fo) 
B o 


1x10° 2x10° 3x05 4xi05 
BODY REYNOLDS NUMBER - UD/V 


FIGURE 30. Cavitation inception on the NSRDC body. 


stabilizing density gradient. This point was ad- 
dressed experimentally and in Figure 29 matched 
pairs of schlieren photographs, without and with 
salt injection are presented. It was found that 
although the appearance of the transition changed 
markedly the location of transition did not change 
significantly. 


6. PRESENTATION OF CAVITATION INCEPTION RESULTS 
Freestream Turbulence Level 


The data on the influence of freestream turbulence 
level upon cavitation inception is limited because 
of the low maximum water speed in the LTWT of about 
8m/s but more importantly because the turbulence 
generating grids located at the entrance to the 
test section cavitated themselves before the test 
models did. Consequently, only the 0.05 and 0.65 
percent turbulence level configurations could be 
used. The NSRDC body was the only one to be so in- 
vestigated. Some of these inception data are sum-— 
marized in Figure 30 where they are compared with 
Brockett's (1972) data. Inception on the NSRDC 
body was always of the band type which occurred 
suddenly without any precursor bubble type cavita- 
tion. As can be seen in Figure 30, inception oc- 
curred at the same value of the inception index for 
both turbulence levels, but as illustrated in Fig- 
ure 31 the subsequent developed cavitation was much 
less steady at the higher turbulence intensity. 


The Effects of Polymer Solutions 
Hemisphere Nose Body 


The type of cavitation and the value of the incep- 
tion index were found to be strongly dependent on 
the amount of polymer present in the boundary layer. 
For a fixed polymer solution concentration and free- 
stream velocity the following changes in inception 
were observed to take place: at zero injection 
rate, incipient band type cavitation as illustrated 
in Figure 32(a) always occurred. At injection rates 
less than the critical value (the injection rate 

at which the separation would disappear), band type 
inception still occurred but as can be seen in Fig- 
ure 32(b) the surface of the developed cavitation 


100 


FIGURE 31. The physical appear- 
ance of cavitation on the NSRDC 
body at two turbulence levels in 
the LIWT. The Reynolds number 
is 3.4 x 10°. In (a) the turbu- 
lence level is 0.05 percent and 
the cavitation index is 0.44. 
The turbulence level in the re- 
Maining photographs is 0.65 per- 
cent and the cavitation index is 
about 0.35 for all cases. 


FIGURE 32. In these photographs 
500 wppm of polyox solution is 
injected at the nose of the hemi- 
sphere body. The cavitation 
index is 0.59, and the Reynolds 
number is 6.7 x 10° (HSWT). The 
dimensionless injection rate, G, 
is zero in (a) 1.9 x 107° in (b), 
4.4 x 10-© in (c), and 5.24 

1o-© in (d). In many instances 
the attached cavitation would 
disappear. 


has a definite wave structure and the separation 

line has become very irregular. Inspection of 
Schlieren photographs of the fully wetted flow at 
this injection rate showed that the position of 
transition on the free shear layer had moved upstream 
from the no injection case and that the separation 
region was smaller in size. With a further increase 
in the injection rate to near critical values, dif- 
ferent types of cavitation were observed depending 
upon the facility. In the HSWI, band type inception 
would occur intermittently in patches with irregular 
separation lines and surfaces as is shown in Figure 
32(c), (d). At injection rates above the critical 
value, the same type of behavior took place, but with 
the flow altering between fully wetted and patchy 
band type cavitation more rapidly. A decrease in 


the cavitation number at this injection rate would 
make the cavitation more "violent," but no steady 


attached cavitation could be obtained. At these 
near-and-above critical injection rates the fully 
wetted observations showed the laminar separation 
had been eliminated with only an occasional short 
reappearance. That is, the flow in the region of 
interest was almost always turbulent. If then the 
injection rate was suddenly reduced to zero, a large 
steady cavity would quickly form on the body. 

In the LTWT the same sequence of cavitation 
events with increasing injection rates would occur 
as in the HSWT. However, near and above critical 
injection rates, travelling bubble and band type 
cavitation would occur simultaneously, unlike the 
HSWT where no bubble type cavitation was observed. 


NO INJECTION 
G~0,35x 107° 
G~ 1.75 x 107 
G~ 2.3 xior& 
G~ 7.0 x 107 
G~9.0 x1lo7® 
G~ 14.6 x 107° 
G~ 23.6 x 1o7® 


fo) 
x 
o@pvpDegncno 


to} 
co) 


oO 
‘7 


fo) 
a 


CAVITATION INCEPTION NUMBER - 0; 
° 
ce) 


4 1 — 4 


Ixio> 2 3 4 5 6 ? 8 cS) Ix10® 
BODY REYNOLDS NUMBER - UD/V 
FIGURE 33. Cavitation inception with polymer 
injection on the hemisphere body. 
This difference will be discussed later. These 


inception data have been summarized in Figure 33. 


NSRDC Body 


The NSRDC body was tested only in the LTWT and it 
too was observed to go through a sequence of cavi- 
tation development similar to that of the hemisphere 
nose body in the LIWT; namely, that the injection 

of polymer at sub-critical rates changed the orig- 
inal band type inception to simultaneously occurring 
intermittent band and travelling bubble type incep- 
tion. At above critical injection rates the inter- 
mittency became more rapid but still no steady 
attached cavitation could be obtained. Examples 

of these types of cavitation are shown in Figure 

34. Notice in particular Figure 34(d) where only 
one cavitation bubble is visible at a cavitation 
number of 0.34. Values of the inception index 


101 


versus body Reynolds number are presented in Fig- 
ure 35. 


Schiebe Body 


The Schiebe body was tested in both the LIWT and the 
HSWT, but the influence of polymer was only studied 
in the LTWI. Again as for the hemisphere nose body, 
the type of cavitation depended upon the facility. 
In the LIWT, travelling bubble type inception always 
occurred and the presence of polymer was found to 
have no significant effect on either the type of 
cavitation or the inception index. Lowering of the 
tunnel pressure below the inception value produced 
a steady, attached cavity of the type normally as- 
sociated with the presence of a laminar separation. 
On the other hand, in the HSWT, travelling bubble 
type cavitation events were extremely rare. In- 
ception occurred with the sudden appearance of an 
unsteady attached cavity occasionally preceded by 
one or two travelling bubble events. Examples of 
these types of cavitation on the Schiebe body are 
given in Figure 36 and a summary of the inception 
data is given in Figure 37. A unique location of 
inception could not be accurately determined in 
either facility for this body. 


7. DISCUSSION 
Freestream Turbulence Level 


The main purpose of the investigation of freestream 
turbulence level upon cavitation inception was to 
determine if it could be a contributing factor to 
the differences in cavitation results on identical 
bodies tested in different facilities. In particu- 
lar, could the differences in cavitation inception 
on the same NSRDC test body between the CIT HSWT and 
the NSRDC 12-inch tunnel be explained by different 


FIGURE 34. The physical appear- 
ance of cavitation on the NSRDC 
body at a Reynolds number of 

3.4 x 10° in the LIWT with (a) 
no injection, (b) G = 3.4 x 10°77, 
cavitation index = 0.45 [same as 
in (a)I, (ce) 3.4 x 1077, cavita- 
tion index = 0.34, and (d) 7.1 x 
10-© at the same index! 


102 


—————— 
© NO INJECTION 
@ G~0.35x107° 

4 BROCKETT (1972) 
NO INJECTION 


o 
n 


oO 
on 


° ° 
a B 


CAVITATION INCEPTION NUMBER - O; 
° 
Nn 


ix10° 2 3 4 
BODY REYNOLDS NUMBER - UD/V 


FIGURE 35. Cavitation suppression by polymer injection 
on the NSRDC body. 


turbulence levels? From the proceeding discussion 
of the fully wetted results it appears that the 
differences in observed critical Reynolds numbers 
are probably due to a higher turbulence level in 
the NSRDC facility. It follows then that the dif- 
ferences in the type of inception for velocities 
less than 30 feet per second can be explained in 
terms of the different viscous flows. However, we 


FIGURE 36. 


: Photographs of cavitation on the (same) 
Schiebe body in the LIWT (upper picture) at a cavita- 
tion index of 0.52 and in the HSWT at an index of 0.41. 
The flow speeds are 7.3 and 14 m/s, respectively. 


@ HSWT PRESENT TESTS 
@ LTWT a i 
4 van der MEULEN (1976) 


fe) 
(2) 


S —:—~C, AT ESTIMATED POSITION 
OF TRANSITION 
x< 

AOS 

a 

z 

204 

) 

Looe 

00.3 

WO. 

z 


9° 
to 


Axio = 4 5 6 7 8 9 1x10® 
BODY REYNOLDS NUMBER-UD/v 


FIGURE 37. Cavitation inception on the Schiebe body 
in three different facilities. 


still need to account for the different freestream 
populations of nuclei, the subject of the next 
section. 


Polymer Injection 


Some inception data for the hemisphere nose body 
with polymer injections are given in Figure 38 as 
a function of the injection rate for two concentra- 
tions. The same data have been replotted in Figure 
39 against the non-dimensional injection parameter 
G. It can readily be seen that the two curves have 
collapsed onto one. A similar happy result was 
found when the dimensions of the laminar separation 
bubble on the hemisphere nose body were plotted 
versus the parameter G. These correlations of the 
inception index and separation bubble dimensions 
with G implies that the polymer "effectiveness" is 
proportional only to the amount present within the 
boundary layer, here taken to be the displacement 
thickness. 

For the NSRDC and hemisphere nose bodies it can 
be seen that increasing amounts of polymer in the 
boundary layer produce an increasing suppression 


° 
= 


© 50 WPPM 
@ 500 WPPM 


9 
oO 


03 


CAVITATION INCEPTION NUMBER - oj 


0 5 10 15 
POLYMER INJECTION RATE - ml/sec 


FIGURE 38. Cavitation index with polymer injection on 
the hemisphere body at a Reynolds number of 7.5 x 10°. 


Oo 50 WPPM 
@ 500 WPPM 


{e) 
(>) 


° fo) 
rs a 


CAVITATION INCEPTION NUMBER - Go; 
° 
ol 


(o) Ixl0"& 2 3 4 5 6 7 8x1o7& 


FIGURE 39. The data of figure 38 replotted against the 
dimensionless injection parameter G. 


of cavitation index. There is a limit, however, 
beyond which no further increase in cavitation 
suppression occurs. In the present experiments on 
the hemisphere nose body this limiting value of G 
is approximately 7 x 107© which also coincides with 
the removal of the laminar separation. These 
results and others are summarized in Figure 40 

where the maximum percent reduction in cavitation 
index has been plotted versus the Reynolds number. 
These include the "polymer ocean" results of Baker 
et al. (1973), Holl et al. (1974), and Ellis et al. 
(1970). However, the information from their re- 
ports is limited and all that can be said is that 
they give values approximately the same as those 
noted in the present case. The agreement is be- 
lieved to be reasonably good for experiments of this 
type insofar as the maximum effect goes. We presume 
that similar effects in "ocean" experiments could 
be achieved at much smaller concentrations if the 

G parameter has significance. 

During their cavitation tests Baker and Holl 
noted a change in the appearance of the developed 
cavitation. From photographic observations of these 
changes they speculated that the cavitation attenu- 
ation was due to a "flow reorientation in the region 
of the laminar separation bubble." They further 
suppose [Arndt et al. (1975)] that the amount of 
attenuation might depend on a Deborah number, 

TV /S51 where T is the molecular relaxation time 
of the molecule, V,, the freestream velocity, and 


von der MEULEN(1976)-500 WPPM 
PRESENT STUDY - SOOWPPM 

20 WPPM 
50 WPPM 
20 WPPM 
BOWPPM 
2OWPPM _ HOLL etal (53) 


} ELLIS etal (1970) | 


8 


| BAKER etal (1973) 


@eseorpcvreo 


Db 
fe) 


nN 
le) 


PERCENT REDUCTION IN INCEPTION 
CAVITATION NUMBER 
° 


Ixto® 2 3 4 5 6 7 8 9 Ixio® 
BODY REYNOLDS NUMBER - UD/V 


FIGURE 40. Maximum cavitation inception index suppres- 
. sion by polyox WSR 201 on the hemisphere nose. The 
Ellis and Baker results are for polymer "oceans." 


103 


6, is the boundary layer displacement thickness at 
separation. It now seems clearly established in 
our opinion, that the overall gross effect caused 
by the polymer in the flow about these bodies is a 
removal of the laminar separation by stimulation of 
transition and that this is indeed the origin of the 
flow "reorientation" noted by Baker and Holl. Pre- 
sumably, the molecular relaxation time has an im- 
portant role in boundary layer stability, but as 
yet this appears to be unknown; it may be that the 
parameter proposed by Arndt is important for some 
laminar flows with separation (as it is indeed for 
the flow about a circular cylinder), but we think 
not in the context of the present experiments. 

Since the suppression of cavitation upon these 
bodies is a result of the elimination of the laminar 
separation by the polymer it is worthwhile to com- 
pare the present results with those in which the 
separation is eliminated by another method. Arakeri 
and Acosta (1976) carried out a series of tests with 
a hemisphere nose body and an ITTC body using bound- 
ary layer trips to reduce the critical Reynolds num- 
ber in the HSWT. It was, briefly, found that with 
the trip present and at velocities above the new 
critical velocity, the occurrence of cavitation was 
significantly suppressed, and that at higher ve- 
locities the tunnel would choke from the model 
support before the body could be made to cavitate: 
The present polymer tests show a very similar large 
effect on inhibiting cavitation but not quite as 
dramatic as the tripped tests. 


8. FREESTREAM NUCLEI AND CAVITATION INCEPTION 
Some Observations in the LTIWT 


As will be recalled from the description of the 
LTIWT, this facility has no resorber which neces-— 
sitated cavitation data acquisition before pump- 
generated bubbles entered the test section. Ona 
number of occasions the cavitation on the NSRDC and 
hemisphere bodies was deliberately maintained and 
the pump-generated gas bubbles allowed to pass 
through the test section. As the number of free 
gas bubbles increased, the initially-occurring band 
type cavitation was gradually destroyed and replaced 
by travelling bubble type cavitation. An alterna- 
tive procedure was to lower tunnel static pressure 
so that the cavitation number had a value below 
-Cpmin but above the inception value and again 
allow the pump-generated bubbles to accumulate. 

The body would then eventually cavitate with in- 
ception then always being of the travelling bubble 
type. Schlieren observations of the basic viscous 
flow on the hemisphere nose were made at these 
gradually increasing freestream bubble populations 
and nuclei populations were measured when band type 
inception occurred and when this above deliberately- 
promoted bubble type inception occurred. The 
schlieren observations show (see Figure 41) that 

as the number of freestream nuclei increased, the 
laminar separation on the hemisphere nose became 
unsteady and was finally greatly diminished if not 
eliminated. Thus, in effect, the free-stream bub- 
bles serve to trip the boundary layer. 

Nuclei populations obtained when band type incep- 
tion occurred (0; = 0.44) are shown with distribu- 
tions obtained when deliberately promoted travelling 
bubble inception occurred (0; = 0.58, 0.73) in 
Figure 42. As can be seen in this figure, for 


104 


FIGURE 41. Flow past the hemi- 
sphere nose with many freestream 
bubbles showing boundary layer 
stimulation. 


nuclei with radii less than 100 microns all the dis- 
tributions are essentially the same whereas for 
nuclei greater than 100 microns radius the bubble- 
type inception distributions have many more nuclei 
than the band type inception distributions. Thus 
it seems possible that in facilities with many 
Macroscopic freestream gas bubbles, the normally 
occurring laminar separation on some bodies can be 
eliminated. The subsequent cavitation index and 
form of cavitation should then be controlled by the 
nuclei population. 4 

If so, the experiments on the NSRDC body at that 
facility and those tests on the same body in the 


HSWT (PARTICLES) 
(.24<o<.72) 


io! 


10'° 


le O =0.44 LTWT (BAND) 
© O-=0.44 
4 0-058 LTWT(BUBBLE) 
10° @ 0-073 

O O =066-PETERSON(I972) 
BUBBLE TYPE 


NUCLE! NUMBER DENSITY DISTRIBUTION FUNCTION (m-4) 


lo-& 1075 10-4 
NUCLEI RADIUS (m) 


PIGURE 42. Nuclei distributions measured by holography 
in the LIWT (all microbubbles) and in the HSWT (essen- 
tially only solid particles). 


LTWT, when bubble type inception was deliberately 


promoted, should be very similar. This is, in fact, 
the case as the inception numbers are more-or-less 
the same. Beyond that, nuclei distributions are 
known for the two tests [Peterson (1972) and Fig- 
ure 42] so that, following the philosophy of Silber- 
man et al. (1974), it is possible to estimate the 
number of "cavitable" nuclei per unit volume for 
each experimental point. A rough estimate of the 
number of travelling cavitation events can be easily 
made if we take Johnson and Hsieh's (1966) "capture" 
radius of 0.01 body radius to determine the flux 

of fluid through the cavitating region. These data, 
calculated and measured events are tabulated in 
Table 3. Peterson measured the event rate acous- 
tically and chose one event/sec as the threshold 
level because of the agreement with a “visual" in- 
ception estimate. (Only the visual estimate was 
made in the LTWT.) 


Observation in the HSWT 


On the whole, the agreement of observations and 
event rates is satisfactory and it seems clear in 
this circumstance that viscous effects are not of 
primary importance and that travelling bubble cavi- 
tation, the type studied by the St. Anthony Falls 
group, is the prevalent form. But, on all of the 
bodies studied we have seen different forms of 
cavitation occur, when separation was not present, 
if the number of freestream nuclei is very small, 

as it is presumably in the California Institute of 
Technology HSWT and other resorber facilities. Then, 
even on the Schiebe body we see attached forms of 
cavitation at inception (see Figure 36) at very low 
inception indices with only rare occurrences of 
travelling bubble cavitation [see also Arakeri et al. 
(1976) ]. In these circumstances the fluid and the 
nuclei that it contains pass through regions of 

some tension (up to about 1/2 at m in the HSWT). It 
is conceivable then that the substantial pressure 
fluctuations in transition regions [Huang and Hannon 
(1975) ] can initiate cavitation. This is the ra- 
tionale for Arakeri's (1975) inception-transition 
pressure coefficient correlation. Values of —Cptr 


105 


TABLE Ml CAVITATION HVENT RATES 
$5 Model oy ee R. Cavitatable Calculated Measured 
Facility Mat'l. Nuclei/cm Events/sec Events per sec 
(ft/sec) (microns) (est) 
NSRDC CU 0.62 29.86 We 0.5 Ong 1.0 
CU 0.66 29. 86 15 ie 8 Bie WO) 
AU-Plated 0.65 29.86 14 Ze Al 3.8 ee 
DELRIN 0.69 29. 86 18 OS 0.9 1.0 
DELRIN 0.71 29.86 21 2.4 4.3 0 
LTWT CU 0.58 20.15 21 Sc 9) 10.7 -- 
CU 0.64 20. 10 29 2.0 Z.4 -- 
CU 0.66 ZO). U5 32 0.9 We il -- 
CU O73 20.25 58 Ne 7 2.0 -- 


are also shown for the Schiebe body in Figure 37; 
again the correlation is suggestive but not con- 
clusive. 

Further evidence of the difference between a 
resorber facility and a recirculating tunnel is 
given by the nuclei distributions of the flow in the 
California Institute of Technology HSWT. These data 
are averaged in the graph of Figure 42. Following 
Peterson (1972) it is possible to distinguish par- 
ticulate matter from gaseous microbubbles down to 
about 10 micrometers. Thus we identify solid par- 
ticulates on the one hand and microbubbles on the 
other. All of the nuclei reported in Figure 42 
for the LTWT are microbubbles. It is significant 
that the HSWT shows a very similar distribution of 
solid particulates, but very few microbubbles. In 
about ten holograms made of the HSWT flow, within 
the various sample volumes that were counted, about 
100 particles/cm? on the average were found. How- 
ever, of these, less than one on the average was a 
microbubble, too few even to hazard a guess as to 
the distribution. This finding certainly tends to 
explain the experimental trends in this facility if 
it is assumed (as appears evident) that the solid 
particulates do not act as nucleating sources. 

In closing this section we have perhaps come 
full circle in inception research to re-emphasize 
the important role of the cavitation nuclei. The 
influence of laminar inception on cavitation is now 
much clearer as are the effects of the processes 
that cause stimulation of the boundary layer. If 
there are many nuclei present (so that a large ten- 
sion on the body does not exist prior to cavitation) 
it is likely that travelling bubble cavitation will 
predominate, then the notion of a "standard body" 
to deduce cavitation susceptibility appears to be 
useful. However, with only a few nuclei other more 
_complex forms of cavitation are seen at inception. 


Comparison of Nuclei Distributions 


Data from several other investigations, reduced to 
the number density distribution function, N(R), by 
the following approximation 


number of nuclei per unit 
Ry + R, with radii between RQ, and R, 
ee ee 


(R,- R)) 


are shown in Figure 43. A tabulation of the mea- 
suring techniques and test conditions for each in- 
vestigation is given in Table 4. All the data have 
approximately the same slope, but the values of the 
distribution function can differ by several orders 
of magnitude, i.e., although the nuclei population 
changed by several orders of magnitude, the dis- 
tribution of the nuclei sizes remains constant. 
The large differences in populations is undoubtedly 
a consequence of the large variation in conditions 
which existed in the water when the data was col- 
lected and is no doubt one of the contributing fac- 
tors to the lack of repeatability seen in cavitation 
esitese 

A goal of cavitation research is to be able to 
predict the inception of cavitation and thus be able 
to scale laboratory results to prototype conditions. 
It is interesting then to compare nuclei populations 
in water tunnels to those in the ocean. Medwin 
acoustically measured bubble populations in the 
ocean near Monterey, California, and in Figure 43 
two of his measured distributions are presented. 
The summer distribution agrees reasonably well with 
the distributions obtained under cavitating con- 
ditions in strongly deaerated water. However, in 
the LTWT there are considerably more bubbles than 
found in the ocean for radii greater than about 30 


106 


TABLE IV COMPARISON OF NUCLEI MEASUREMENTS 


Investigator 


Gavrilov (1970) 


Peterson et al (1975) 


Arndt & Keller (1976) 


Keller & Weitendorf (1976) 


Medwin (1977) 


Peterson (1974) 


U.S. Navy (Naval Ocean 
System Center, San Diego, 
California. Courtesy 

Dr. T. Lang) 


Present Tests 


1975) (o=0.49, 
SCATTERING 


Measuring 
Technique 


Acoustic 


Light Scattering 
Holography 
Microscopy 


Light Scattering 


Light Scattering 


Acoustic 


Coulter Counter 


Coulter Counter 


Holography 


Facility 


Water Tunnel 
ene ID) 


Water Tunnel 
at NSMB 


Water Tunnel 
at Hamburg 
Model Basin 


Monterey Bay, 
California 


Santa Catalina 
Channel 


California 


San Diego Bay 
and offshore 


LTIWT 


Conditions at time 
of Measurement 


Standing tap water 


At inception on 50 mm diameter 
NSRDC body o=0.49 


Cavitation tests on a sharp edged 
disc. Air contents: 6.3 and 
12.5 ppm 


Propeller test, gassed water, 
Air content: ~30 ppm 


Various depths and seasons 


Various depths 


Various depths 


Air content ~ 7ppm, o=0.44 


@ PRESENT TESTS LTWT 
AIR CONTENT ~ 7 ppm 
a =0.44 


PETERSON etal 
(1975)(0=0.49, © 
HOLOGRAPHY) ' 


10"! 


\- 


GAVRILOV (1970) \ 


1o9 AFTER STANDING 
5 HOURS) 


NUMBER DENSITY DISTRIBUTION FUNCTION N(R),m“ 


MEDWIN (1977) 
(OCEAN, FEBRUARY) 


RNDT & KELLER 
976) 


AIR CONTENT~ | 
12.5 ppm 


ot 


1 10 
RADIUS R (micrometers) 


KELLER & 
WE!ITENDORF 
(1976) 

GASSED WATER 
AIR CONTENT~ 
30 ppm 


ARNOT & 
KELLER (1976) 


\ 
V\meowin (1977) 
(OCEAN, AUGUST) 


FIGURE 43. Nuclei distributions from various sources. 


micrometers. 


Further, in the winter the measured 


bubble population in the ocean is one order of 
magnitude less than in the summer. We see then it 
is actually possible for laboratory facilities to 
have much higher nuclei populations than actually 
occur in the ocean. Medwin concludes interestingly 
that the microbubbles had a biological as well as 
physical origin because the concentration of bubbles 
increased with depth. This observation is perhaps 
of importance for the Coulter Counter measurements 
of Peterson (1974) and Lang (1977). The particulates 
measured there, although thought to be of organic 
material, may actually also contain some gas. 
Finally, it is amazing to observe the wide range 

of applicability of fairly simple power laws for 
particulate and microbubble populations. 


9. CONCLUSIONS 


It is clear that the onset of cavitation and its 
physical appearance at this onset can be greatly 
affected by freestream turbulence and the presence 
of minute amounts of long chain polymer solutes. 
The present results support the conclusion that 
these effects are indirect insofar as cavitation 
goes and that the primary effect is on the viscous 
flow past the test body. The polymer solutions in 
particular promote an early boundary layer transi- 
tion which forestalls the presence of laminar separa- 
tion much as does boundary layer stimulation by 
freestream turbulence or trips. It follows that 
cavitation on bodies not having laminar separation 


should not be much affected by freestream turbu- 
lence or polymer solutions. This appears to be 
the case if the test medium has "many" freestream 
nuclei so that travelling bubble cavitation is 
predominant. However, if only a few nuclei are 
present, attached forms of cavitation occur at 
inception even on nonseparating bodies. From 
recent nuclei measurements in the ocean it appears 
that some test facilities may have too many nuclei 
and others possibly too few. 


ACKNOWLEDGMENTS 


This work was supported by the Department of the 
Navy, Office of Naval Research under Contract 
NOO014-76-C-0156 (in part) and by the Naval Sea 
Systems Command, General Hydromechanics Research 
Program, administered by the David W. Taylor Naval 
Ship Research and Development Center under Contract 
NOO014-75-C-0378. This assistance is gratefully 
acknowledged. Special thanks are due to Mrs. 
Barbara Hawk for manuscript preparation and to 
Joseph Katz and David Faulkner for their painstaking 
efforts in hologram analysis. Finally we thank 
Professor M. Morkovin for his careful and helpful 
review of the manuscript. 


REFERENCES 


Acosta, A. J., and H. Hamaguchi (1967). Cavitation 
inception on the ITTC standard headform. Hydro- 
dynamics Laboratory, Report No. E-149.1, Cali- 
fornia Institute of Technology. 

Acosta, A. J., and B. R. Parkin (1975). Cavitation 
inception--A selective review, J. Ship Res., 19, 
4; 193-205. 

Arakeri, V. H., and A. J. Acosta (1973). Viscous 
effects in inception of cavitation on axi- 
symmetric bodies. J. Fluids Eng. 95, Ser. 1, 

4; 519-528. 

Arakeri, V. H. (1975). A note on the transition 
observations on an axisymmetric body and some 
related fluctuating wall pressure measurements. 
Je Hluids Eng. , 97, ‘Sex. 1, 1; 82=s8i7-: 

Arakeri, V. H., and A. J. Acosta (1976). Cavita- 
tion inception observations on axisymmetric 
bodies at supercritical Reynolds numbers. J. 
Ship Res., 20, 1; 40-50. 

Arndt ReEe Acie .n las Basllet siden Wie sHOUsy) and 
C. B. Baker (1975). A note on the inhibition 
of cavitation in dilute polymer solutions. App. 
Res. Lab. Tech. Memo. 75-285, The Penn. State 
Univ. 

Arndt, R. E. A. and J. W. Daily (1969). Cavitation 
in turbulent boundary layers. Cavitation State 
of Knowledge, ASME; 64-86. 

Arndt, R. E. A. and A. P. Keller (1976). Free gas 
content effects on cavitation inception and noise 
in a free shear flow; Two phase flow and cavi- 
tation in power generation systems. IAHR Symp. 
Grenoble, March 16, 1976. 

Baker, C. B., R. E. A. Arndt, and J. W. Holl (1973). 
Effect of various concentrations of WSR-301 
polyethylene oxide in water upon the cavitation 
performance of 1/4-inch and 2-inch hemispherical 
nosed bodies. App. Res. Lab. Tech. Memo. 73-257, 
The Penn. State Univ. 

Beckwith, T. G., and N. L. Buck (1961). Mechanical 
Measurements. Addison-Wesley, Reading, Mass. 


107 


Brennen, C. (1970). Some cavitation experiments 
with dilute polymer solutions. J. of Fluid 
Mech., 44, part 1; 51-63. 

Brockett, T. (1972). Some environmental effects 
on headform cavitation inception. NSDRC Rep. 
3974. 

Cavitation Committee (1969). Summary of replies to 
questionnaire on water tunnel methods. 12th ITTC. 

Debrule, P. M. (1972). Friction and heat transfer 
coefficients in smooth and rough pipes with 
dilute polymer solutions. Ph.D. Dissertation, 
California Institute of Technology. 

Ellis, A. T., J. G. Waugh, and R. Y. Ting (1970). 
Cavitation suppression and stress effects in 
high-speed flows of water with dilute macro- 
molecular additives. J. of Basic Eng., Sept.; 
459-466. 

Feldberg, L. A., and K. T. Shlemenson (1973). The 
holographic study of cavitation nuclei. Dis- 
cussion to Proc. IUTAM Symp. on Nonsteady Flow 
of Water at High Speeds. Leningrad, USSR 
(English version; Moscow 1973); 106-111. 

Gates, E. M., and J. Bacon (1978). Determination 
of cavitation nuclei distributions by holography. 
Wo Glasto RAG, AA, ip Zs, 

Gavrilov, L. R. (1970). Free gas content of a 
liquid and acoustical techniques for its measure- 
ment. Sov. Phys-Acoustics, 15, 3; 285-295. 

Hall, D. J., and J. C. Gibbings (1972). Influence 
of stream turbulence and pressure gradient upon 
boundary layer transition. J. of Mech. Eng. 
Sethe da Bp alsvaS iss. 

Holl, J. W., and A. L. Treaster (1966). Cavitation 
hysteresis. J. of Basic Eng., Trans. ASME, 

Ser, ID, GH jp I@SoAily. 

Holl, J. W. (1968). Sources of cavitation nuclei. 
Pres. at the 15th Amer. Towing Tank Conf., 
Ottawa, Canada. 

Holl, J. W., and A. L. Kornhauser (1969). Thermo- 
dynamic effects on desinent cavitation on hemi- 
spherical nosed bodies in water of temperatures 
from 80 degrees F to 260 degrees F. Pres. at 
the Joint Conf. of App. Mech. and Fluids Eng. 
Divisions, ASME, Northwestern Univ. ASME paper 
No. 69-—PE=1- 

Holl, J. W., R. E. A. Arndt, M. L. Billet, and 
C. B. Baker (1974). Cavitation research at the 
Garfield Thomas Water Tunnel. App. Res. Lab. 
Tech. Memo. 74-62, The Penn. State Univ. 

Hoyt, J. W. (1966). Effects of High-polymer solu- 
tions on a cavitating body. Proc. llth ITTC, 
Tokyo. 

Hoyt, J. W. (1972). The effects of additives on 
fluid friction. J. of Basic Eng., June; 258-285. 

Huang, T. T., and D. E. Hannan (1975). Pressure 
fluctuations in the regions of flow transitions. 
DTNSRDC Rep. 4723. 

Jaffe, N. A., T. T. Okamura, and A. M. O. Smith 
(1970). Determination of spatial amplification 
factors and their application to predicting 
‘transaid onl vArAA mie, 1S),) 0217) SOL —30/8r 

Johnson, Jr., V. E., and T. Hsieh (1966). The 
influence of the trajactories of gas nuclei on 
cavitation inception. 6th Symp. on Nav. Hydro. 


Johnsson, C. A. (1969). Cavitation inception on 
headforms, further tests. 12th ITTC, Rome; 
BSI 392% 


Keller, A. P. (1972). The influence of the cavita- 
tion nucleus spectrum on cavitation inception, 
investigated with a scattered light counting 
method. J. of Basic Eng., Dec.; 917-925. 


108 


Knapp, R. T., J. Levy, J. P. O'Neill, and F. B. 
Brown (1948). The Hydrodynamics Laboratory 
of the California Institute of Technology. 

Trans. ASME, 70; 437-457. 

Knapp, R. T., J. W. Daily, and F. G. Hammitt (1970). 
Cavitation. McGraw-Hill, New York. 

Lang, T. G., (1977). Naval Ocean Systems, Private 
Comm. 

Lindgren, H., and C. A. Johnsson (1966). Cavitation 
inception on head forms ITTC comparative experi- 
ments. Pub. of the Swedish State Shipbuilding 
Exper. Tank, No. 58. (Also presented at the 
llth ITTC Tokyo, 1966). 

Loehrke, R. I., and H. M. Nagib (1972). Experiments 
on management of freestream turbulence. AGARD 
Rep. No. 598. 

Mack, L. M. (1977). Transition and laminar in- 
Stabwlattye die ee ae PUD 7715). 

Medwin, H. (1970). In situ acoustic measurements 
of bubble populations in coastal ocean waters. 

J. Geophys. Res. 75, 599-611. 

Medwin, H. (1977). In situ acoustic measurements 
of microbubbles at sea. J. Geophys. Res. 82, 

Gp SPARS 

Morgan, W. B. (1972). Air content and nuclei mea- 
surements. 13th ITTC, Rep. of Cavitation Comm. 

Morgan, W. B., and F. B. Peterson (1977). Cavita- 
tion inception. 18th Amer. Tow. Tank Conf., 

U.S. Naval Acad., Annapolis, Md. 

Parkin, B. R., and R. W. Kermeen (1953). Incipient 
cavitation and boundary layer interaction on a 
streamlined body. Hydro. Lab. Rep. E-35.2, 
California Institute of Technology. 

Peterson, F. B. (1968). Cavitation originating at 
liquid-solid interfaces. NSDRC Rep. 2799. 

Peterson, F. B. (1972). Hydrodynamic cavitation 
and some considerations of the infltience of free 
gas content. 9th Symp. on Naval Hydro. 

Peterson, F. B., F. Danel, A. Keller, and Y. Lecoffe 
(1975). Determination of bubble and particulate 
spectra and number density in a water tunnel 
with three optical techniques. 14th ITTC. 

Peterson, L. L., (1974). The propagation of sun- 
light and the size distribution of suspended 
particles in a municipally polluted ocean water. 
Ph.D. Thesis, California Institute of Technology. 

Reshotko, E. (1976). Boundary-layer stability and 


transition. Ann. Rev. of Fluid Mech., 8; 311-349. 


Sarpkaya, T., P. G. Rainey, and R. E. Kell (1973). 
Flow of dilute polymer (Polyox WSR-301) solu- 
tions about circular cylinders. J. of Fluid 
Mech., 57; 177-208. 

Sarpkaya, T. (1974). On the performance of hydro- 
foils in dilute Polyox solutions. Int'l. Conf. 
on Drag-Reduction, Churchill Coll., Cambridge, 
England, 4-6 Sept. 

Schiebe, F. R. (1966). Cavitation occurrence 
counting--a new technique in inception research. 
Cavitation Forum, ASME, N.Y., 8-9. 

Schiebe, F. R. (1972). Measurement of the cavita- 
tion susceptibility of water using standard 
bodies. St. Anthony Falls Hyd. Lab. Project 
Rep. No. 118, Univ. of Minn. 

Silberman, E., F. Schiebe, and E. Mrosla (1973). 
The use of standard bodies to measure the 
cavitation strength of water. St. Anthony 
Falls Hyd. Lab., Rep. No. 141, Univ. of Minn. 


Spangler, J. G., and C. S. Wells, Jr. (1968). Ef- 
fects of freestream disturbances on boundary 
layer transition. AIAA J., 6; 543-545. 

Tagori, T., K. Masunaga, H. Okamoto, and M. Suzuki 
(1974). Visualization of flow of dilute polymer 
solutions around two-dimensional hydrofoils. 
2nd Symp. of Flow Visualization, Univ. of 
Tokyo, July; 83-88. 

Ting, R. Y., and A. T. Ellis (1974). Bubble growth 
in dilute polymer solutions. Physics of Fluids, 
ili VAG VA GDS 

Tsuji, H. (1956). Experimental studies on the 
spectrum of isotropic turbulence behind two 
grids. J. of Phys. Soc. of Japan, 11, 10; 
1096-1104. 

van der Meulen, J. H. J. (1972). Cavitation on 
hemispherical nosed teflon bodies. Int'l. Ship- 
building Prog., 19, 218; 333-341. 

van der Meulen, J. H. J. (1973). Cavitation sup- 
pression by polymer injection. ASME Cavitation 
and Polyphase Flow Forum, N.Y.; 48. 

van der Meulen, J. H. J. (1976). A holographic 
study of cavitation on axisymmetric bodies and 
the influence of polymer additives. Doctoral 
Thesis, Netherlands Ship Model Basin. 

van der Walle, F. (1962). On the growth of nuclei 
and the related scaling factors in cavitation 
inception. 4th Symp. on Naval Hydro. 

van Driest, E. R., and C. B. Blumer (1963). Bound- 
ary layer transition: freestream turbulence 
and pressure gradient effects. AIAA J., 1, 6; 
1303-1306. 

Vanoni, V. A., E. Hsu, and R. W. Davies (1950). 
Dynamics of particulate matter in fluid sus- 
pensions. Hydro. Lab. Rep. No. 7l.la, California 
Institute of Technology. 

Wazzan, A. R., T. T. Okamura, and A. M. O. Smith 
(1968a). The stability of water flow over 
heated and cooled flat plates. J. of Heat 
Transfer, Feb.; 109-114. 

Wazzan, A. R., T. T. Okamura, and A. M. O. Smith 
(1968b). Spatial and temporal stability charts 
for the Falkner-Skan boundary layer profiles. 
D.A.C. rep. 67086, McDonnell Douglas Corp. 

Wazzan, A. R., G. Keltner, T. T. Okamura, and A. 

M. O. Smith (1972). Spatial stability of stag- 
nation water boundary layer with heat transfer. 
Physics of Fluids, 15, 12; 2114-2118. 

Wazzan, A. R., C. Gazley, Jr. (1978). The combined 
effects of pressure gradient and heating on the 
stability and transition of water boundary 
layers. Rand Rept. R-2175-ARPA. 

Wu, J. (1971). Some techniques of ejecting additive 
solutions for drag reduction. Hydronautics 
Technical Rept. 7101-1. 


Additional Reference* 
van Ingen, J. L. (1977). "Transition, pressure 


gradient, suction, separation and stability 
theory," paper 20 AGARD proc. CP-224. 


* 
Suggested by Professor M. V. Morkovin 


Discussion 


M. A. WEISSMAN 


My question was "What is your definition of 
growth rate?" This is quite a crucial point, for 
in comparing theory to experiment, we must make 
sure that we are comparing like to like. 

The meaning of growth rate for nonparallel 
flow is not obvious. Let us consider El—Hady and 
Nayfeh's lowest order solution (Eq. 42): 


A = iN feGe ry)explif (a + ca )dx - iwt] (1) 
1 0 1 0 1 


The downstream growth of the magnitude of this 
function is not purely contained in the expotential 
factor. The change in the eigenfunction, Tt, with 
x, also contributes to "growth." In fact, a com- 
miete definition of growth would be 


1 a|z,| 


er el ex 


which reduces to 


G= @) + eG, + Tel Oe, (2) 


109 


using (1), where it is understood that Gy and G4 
are the negative and imaginary parts of o, and a,. 
[Bouthier (1972), Gaster (1974), and Eagles and 
Weissman (1975)]. 

Equation 2 shows that the growth rate is 
actually a function of y. (It is also a function 
of the flow quantity under consideration, see the 
above mentioned references.) However, if we agree 
to measure the growth rate at a particular 
y-position and if the eigenfunction is normalized 
at that position (so that 3|c|/ax = 0 at that 
position), then the influence of the changing 
eigenfunction on growth rate will disappear (for 
this particular definition of growth rate). The 
poit is that a, is not uniquely defined; it depends 
on the normalization used for Z. [This can also 
be seen from examination of the equation defining 
Oye Eq. 35]. The authors have neglected to explain 
what their normalization was. 


REFERENCES 


Eagles, P. M., and M. A. Weissman (1975). On the 
Stability of Slowly Varying Flow. J. Flutd. Mech. 
69, 241-262. 


110 


Author’s Reply 


ALI H. NAYFEH 


The growth rate in a parallel flow can be 
unambiguously defined, but it cannot be unambig- 
uously defined in a nonparallel flow. Because the 
eigenfunctions are functions of y as well as x, 
Saric and Nayfeh (1977) note that stable flows may 
be termed unstable and vice versa. Saric and 
Nayfeh (1977) discussed in great detail the differ- 
ent possible definitions of the growth rate and 
compared these definitions with all available exper- 
imental data for the Blasius flow. They found that 


all the experimental data (neutral curves or growth 
rates) obtained at the values of n for which |u| 
has a maxima can be correlated with the nonparallel 
results if the growth rate is defined as in (55). 
For the heated liquid problem, we arrived at the 
same conclusion. Including the distortion of the 
eigenfunction with the streamwise position, the 
definition, (56), underpredicts the growth by 

large amounts. 


Discussion 


G. CHAHINE and D. H. FRUMAN 


The question of whether polymer solutions 
affect cavitation inception through changes of the 
flow structure or through the inhibition of bubble 
growth has been the subject of much controversy. 

In this excellent paper the authors seem to adhere 
to the first school of thought and disregard the 
second. We think that there is ample evidence of 
the profound flow changes introduced by the ejected 
polymers to support, at least partially, their 
contention. However, evidence also exists showing 
that the onset of acoustically generated cavitation 
is delayed by the presence of minute amounts of 
polymers and asbestos fibers [Hoyt (1977)]. also, 
in investigating the behavior of spark-generated 
bubbles in the vicinity of a solid wall, the dis- 
cussers have observed significant changes being 
promoted by the presence of the polymers. 

Figure 1 shows the geometric dimensions that 
have been considered in the analysis of the bubble 
behavior. The displacement of point A, where the 
re-entering jet originates, divided by the maximum 
lateral dimension of the bubble, Rumax, is plotted 
in Figure 2 as a function of the dimensionless time 
parameter, t/tp, and the parameter, n, which is the 
ratio between Rumax and £, the distance between the 
center of the spherical initial bubble and the wall. 
As shown, the polymer solution has a retarding 


A 


FIGURE 1 


lo} 
o oO 
Y—e1 1 
Se 
Le 
oA 
{eo 
lee 
a 
1 4 1 4 Hono n 


effect on the re-entering jet. This effect in- 
creases with increasing n [Chahine and Fruman 
(1979)]. Together with results shown in Hoyt, 

our data further confirm that, in the absence of 
flow, bubble behavior is affected by the intrinsic 
properties of dilute polymer solutions. 


4 
4 
4 
4 


Ra ] 
Bs A= A 1 
1s rae ‘A 
er, 
Le mR 1 
Ma fo) © SS 
~ Wi 


RUN | ) | LIQUID 
6 PY59|0.50| DISTILLED WATER eo \a 4 
; 4 PY 71/0.56) POLYOX 250 PPMW Ge Wye 
3-4 @ IBY, GBH 139) ies eee ae: aah 4 
r © PY641/1. 25] DISTILLED WATER 7 1 
IP t/Te. 
AO nn Fi CA ere eer rer ell eee a ; SES | 
ie) 0,5 1 1,5 2 2,5 
FIGURE 2 
REFERENCES 


Hoyt, J. W., (1977). 
tions and Fiber Suspensions. 
phase Flow Forum, ASME. 
Chahine, G. L., and D. H. Fruman (1979). Dilute 
Polymer Solution Effects on Bubble Growth and Col- 
lapse. Physics of Flutds. 


Cavitation in Polymer Solu- 
Cavitatton and Poly- 


112 


Authors’ Reply 


EDWARD M. GATES and ALLAN J. ACOSTA 


Messrs. Chahine and Fruman have raised the 
question of the relative importance of polymer- 
induced changes in bubble growth versus induced 
changes in the flow structure with regard to the 
suppression of cavitation. Although both experi- 
mental [Ellis and Ting (1970); Chahine and Fruman! 
(1979)] and theoretical [Street (1968); Fogler and 
Goddard (1970)] work demonstrate that in "no-flow" 
situations the growth and collapse rates in polymer 
solutions are different than those in pure water, 
the magnitude and sense (Street predicts an in- 
crease in bubble growth rate) of the changes are 
open to question. On the other hand, the results 
of Hoyt (1976), Brennen (1970), van der Meulen 
(1976), and the present work show drag-reducing 
polymers have a very dramatic effect upon the flow 
structure in jets and axisymmetric bodies. The 
authors believe that in the present work the influ- 
ence of these profound flow alterations predominate 
over any influence of modified bubble dynamics as 
nicely shown by them as evidenced by the following 
observations: 

First, it was observed in the LIWT that cavi- 
tation inception on the non-separating Schiebe body 
was not influenced by viscous considerations and 
was of the travelling bubble type. In this situa- 
tion we would expect that if the polymer effect 
upon bubble dynamics was significant, it should be 
well illustrated under these circumstances. How- 
ever, we (like van der Meulen) observed no change 
in either the cavitation index or the appearance 


of the cavitation at inception. Second, on the 
hemisphere nose and NSRDC bodies a similarly large 
suppression of the inception index was obtained by 
Arakeri and Acosta (1976) through the elimination 
of the laminar separation by a mechanical boundary 
layer trip - a situation for which there is no 
change of bubble dynamics. 

From these observations we infer that the in- 
fluence of the polymer on cavitation inception is 
dominated by changes in the flow structure rather 
than modified bubble dynamics. However, in "non- 
flow" sitations it must be assumed that modified 
bubble dynamics are responsible for the observed 
changes and the work of Messrs. Chahine and Fruman 
is a useful addition to this area of study. 


REFERENCES 


1 See Reference from Chahine and Fruman discussion. 
Street, J. R., (1968). The Rheology of Phase 
Growth in Elastic Liquids. Trans. Soc. Rheol., 

12, ps LOS. 

Fogler, H. S., and J. D. Goddard (1970). Collapse 
of Spherical Cavities in Viscoelastic Fluids. 
Phystes of Flutds, 13, (5), pp. 1135-1141. 

Hoyt, J. W., (1976). Effect of Polymer Additives 
on Jet Cavitation. Journal of Fluids Engineering, 
Trans. ASME, March, pp. 106-112 


Session IIT 


SHIP BOUNDARY LAYERS 
AND 
PROPELLER HULL INTERACTION 


PETER N. JOUBERT 
Session Chairman 
University of Melbourne 
Melbourne, Australia 


wh 


A raae 


Calculation of Thick Boundary Layer and 
Near Wake of Bodies of 


Revolution by a Differential Method 


Wa Go 


Iowa City, Iowa 


ABSTRACT 


The differential equations of the thick axisymmetric 
turbulent boundary layer and wake are solved using 

a finite-difference method. The equations include 
longitudinal and transverse surface curvature terms 
as well as the static-pressure variation across the 
boundary layer and wake. Closure of the mean-flow 
equations is affected by a rate equation for the 
Reynolds stress deduced from the turbulent kinetic- 
energy equation. The results of the method are 
compared with the two sets of data obtained at the 
Towa Institute of Hydraulic Research from experi- 
ments in the tail region of a modified spheroid 

and low-drag body of revolution, and also with the 
predictions of a simple integral approach proposed 
earlier. It is shown that the differential approach 
is superior, provided due account is taken of the 
normal pressure variation and the direct influence 
of the extra rates of strain, associated with the 
longitudinal and transverse surface curvatures, on 
the length scale of the turbulence. 


1. INTRODUCTION 


In the absence of flow separation, the boundary 
layer on a pointed-tailed body of revolution con- 
tinues to grow in thickness up to the tail. Over 
the rear quarter of the length of a typical body, 
the boundary layer thickness becomes large enough 
to invalidate the assumptions of conventional thin 
boundary-layer theory. The measurements of Patel, 
Nakayama, and Damian (1974) on a modified spheroid 
as well as those of Patel and Lee (1977) on a low- 
drag body indicate that the breakdown of thin bound- 
ary layer approximations is manifested by several 
concurrent flow features, namely (a) the boundary 
layer thickness is no longer small compared with 
the local transverse and longitudinal radii of sur- 
- face curvature, (b) the velocity component normal 
to the wall is not small, (c) the pressure is not 


115 


Patel and Y. T. Lee 
The University of Iowa 


constant across the boundary layer, and (d) the 
pressure distribution on the body surface does not 
conform with that predicted by potential flow theory, 
as a consequence of the interaction between the 
thick boundary layer and the external inviscid flow. 
These features have been recognized in the develop- 
ment of the simple integral method of Patel (1974) 
for the calculation of a thick axisymmetric bound- 
ary layer, and later on, in the formulation of the 
interaction scheme of Nakayama, Patel, and Landweber 
(1976a,b) which attempted to couple the boundary 
layer, the near wake and the external inviscid flow 
by means of successive iterations. Although the 
overall iteration scheme proved to be quite success-~ 
ful, the treatment of the boundary layer using the 
integral method, and particularly its extension to 
calculate the near wake, required many assumptions 
which remain untested. The purpose of the present 
work was therefore to develop a more rational pro- 
cedure in which the differential equations of the 
thick boundary layer and the near wake are solved 
by means of a numerical method, since it appeared 
that such a procedure would provide not only a 

more reliable vehicle for the extension of the 
boundary layer solution into the wake, but also 
yield the detailed information on the velocity 
profiles required for the interaction calculations. 
This paper describes the new differential method 
and evaluates its performance relative to the inte- 
gral method as well as the available experimental 
information. 


2. DIFFERENTIAL EQUATIONS AND TURBULENCE MODEL 


In the (x,y,¢) coordinate system shown in Figure 1, 
x and y are distances measured along and normal to 
the body surface, respectively, and ¢ is the azi- 
muthal angle. As shown by Patel (1973) and Nakayama, 
Patel, and Landweber (1976b), the momentum equa- 
tions of a thick axisymmetric turbulent boundary 
layer may be written 


116 


FIGURE 1. Coordinate system and notation. 


hyrt 
We Oe US pee i) Se Or (ee 
hy ox oy hy phy ox rhy oy p 
(1) 
U_ ov OV Kees} 1 op 
— —+V — - — UL +=—- 2 =0 (2) 
hy, ox oy hy po oy 
and the continuity equation is 
a a ¥ 
53 (Ur) + ay (xh,V) = 0 (3) 


U and V are the components of mean velocity in the 
x and y directions, respectively; h, = 1 + ky, Kk 
being the longitudinal surface curvature; T = —puv 
+ u dU/dy, where p is density, wu is viscosity and 
-puv is the Reynolds stress; r = r, + y cos 8 is the 
radial distance measured from the body axis, 8 
being the angle between the tangent to the surface 
and the axis of the body; and p is the static pres- 
sure. These equations resulted from order of mag- 
nitude considerations and an examination of the 
data from the modified spheroid experiments of 
Patel, Nakayama, and Damian (1974). Specifically, 
from Eq. (2) we note that the static pressure varies 
across the boundary layer and that the gradient of 
the pressure in the direction normal to the surface 
is associated primarily with the curvature of the 
mean streamlines. 

Equations (1), (2), and (3) also apply to the 
wake, with k = 0 and 0 = O (i-.e., r = y)- In place 
of the no-slip boundary conditions on the body sur- 
face, however, the conditions on the wake center- 
line are dU/dy = O and t = O. 

If the Reynolds stress is determined by a one- 
equation model using the turbulent kinetic-energy 
equation, as proposed by Bradshaw, Ferriss, and 
Atwell (1967), then the appropriate closure equa- 
tion for the flow outside the viscous sublayer and 
the blending zone is 


1 U oT chs ou 
Daa) Naan ox a= ; {2 7 “) 
- T 3/2 
Hid T max AT, 
+ = — = = 
Pighiet, Aaa/ow mp he Tw 2 e 


aj p 


where a, is a constant (=0.15), G is a diffusion 
function and 2 is a length-scale function identified 


with the usual mixing length. G and % are assumed 
to be universal functions of y/é, where 6 is the 
boundary layer thickness. The particular forms of 
these functions proposed by Bradshaw et al. (1967) 
for a thin boundary layer have gained wide accep- 
tance and have proved adequate for the prediction 
of a variety of boundary layers developing under 
the influence of different pressure gradients and 
upstream history. In the adoption of this closure 
model for the treatment of thick boundary layers 
and wakes, however, it is necessary to consider the 
influence of transverse and longitudinal surface 
curvatures on the turbulence. 

Figure 2 shows the conventional transverse and 
longitudinal curvature parameters for the modified 
spheroid and low-drag body [Patel and Lee (1977) ]. 
The ratio of the boundary-layer thickness to the 
transverse radius of curvature, 6/ro, is seen to be 
more than twice as large in the latter case as in 
the former. In both cases, however, 6/ro is less 
than 0.4 up to X/L = 0.75, so that the boundary 
layers may be regarded as thin up to that station. 
Over the rear one-quarter of the body length, the 
influence of transverse curvature would prevail 
not only through the geometrical terms in the mo- 
mentum and continuity equations but also through 
any direct effect on the turbulence. The precise 
nature of the latter is not known at the present 
time since the turbulence is also affected by the 
longitudinal curvature of the streamlines associated 
with the curvature of the surface as well as the 
curvature induced by the rapid thickening of the 
boundary layer over the tail. 

The longitudinal surface curvature parameter ké 
is seen to be quite different for the two bodies. 
In the case of the modified spheroid, the curvature 
is convex up to X/L = 0.933 and zero thereafter, 
while that of the low-drag body is initially convex 
and becomes concave for X/L > 0.772. Several 
recent studies with nominally two-dimensional thin 


© Modified Spheroid 


& Low-Orag Body 


Ly 
0.4 Os 0.6 0.7 0.8 0.9 10 
X/L 
FIGURE 2. Ratios of boundary-layer thickness to the 


longitudinal and transverse radii of surface curvature. 


turbulent boundary layers [Bradshaw (1969, 1973), 
So and Mellor (1972, 1973, 1975), Meroney and 
Bradshaw (1975); Ramaprian and Shivaprasad (1977); 
Shivaprasad and Ramaprian (1977)] have indicated 
that even mild (ké~0.01) longitudinal surface 
curvature exerts a dramatic influence on the turbu- 
lence structure. In particular, it is noted that 
quantities such as the mixing length 2, the struc- 
ture parameter a, = -uw/q2 and the shear-stress 
correlation coefficient uv/(Vva- vv“) are influenced 
markedly, and experiments indicate that convex 
streamline curvature leads to a reduction in these, 
whereas concave curvature has an opposite effect. 
The turbulence measurements on the modified spheroid 
and the low-drag body appear to confirm these ob- 
servatons although the relative influence of longi- 
tudinal streamline curvature and transverse surface 
curvature could not be separated readily. 

Bradshaw (1973) has argued that whenever a thin 
turbulent shear layer experiences an extra rate of 
strain, i.e., in addition to the usual dU/dy, the 
response of the turbulence parameters is an order 
of magnitude greater than one would expect from 
an observation of the appropriate extra terms in 
the mean-flow equations of momentum and continuity. 
For THIN shear layers and SMALL extra rates of 
strain he proposed a simple linear correction for 
the length scale of the turbulence, viz. 


Be a 
= U/oy 


(5) 
where £, is the length scale with the usual rate 
of strain, dU/dy, & is the length scale with the 
extra rate of strain, e, and a is a constant of 
the order of 10. For the axisymmetric boundary 
layer being considered here, there are two extra 
rates of strain: 


Pcs! (6) 
ns TP icy 


due to the longitudinal curvature, and 
fy oe es ot a) 
ies 


due to the convergence or divergence of the stream- 


lines (in planes parallel to the surface) associated 


with the changes in the transverse curvature. The 
former is a shearing strain while the latter is a 
plain strain, and it is not certain whether the 
two effects can be added simply in using Eq. (5) 
as recommended by Bradshaw (1973). If this is the 
case, however, we would expect a greater reduction 
in £ in the tail region of the modified spheroid, 
where kK is positive and dr,/dx is negative, than 
on the low-drag body, where kK becomes negative and 
would therefore tend to offset the influence of 
the negative dr,/dx. Although the available data 
appear to bear this out to some extent, a direct 
comparison between Eqs. (5), (6), and (7) and the 
data was not attempted, especially in view of 
Bradshaw's [Bradshaw and Unsworth (1976) ] assertion 
that Eq. (5) should be used in conjunction with a 
simple rate equation which accounts for the up- 
stream extra rate-of-strain history. He proposes 


ae 
Me eft : 
eS & SS S= 
Lo dU/dy So 


-and 


117 


ie yh Sa ese (9) 

dx eff 106 
where e is the actual rate of strain, eef¢ is its 
effective value and 106 represents the "lag length" 
over which the boundary layer responds to a change 
ine. In order to determine the merit of this 
proposal, it is of course necessary to incorporate 
it in an actual calculation and make a comparison 
between the predictons and measurement. Such an 
attempt has been made here. 

The functions £2, and G used in the present study 
are shown in Figure 3. For the wake calculation, 
the linear variation of 2, in the wall region is 
replaced by the constant value of 0.09, as shown 
by the dotted line in the figure. The local dis- 
tribution of the length scale, %, is thus given by 
Eqs. (6) through (9) while the diffusion function, 
G, and the structure parameter, a ,, retain their 
thin-boundary-layer values. 


3. SOLUTION OF THE DIFFERENTIAL EQUATIONS 


A numerical method available for the solution of 
equations corresponding to (1), (3), and (4) for 

a thin two-dimensional boundary layer was modified 
to introduce the longitudinal- and transverse- 
curvature terms. Instead of incorporating the y- 
momentum, Eq. (2), into the solution procedure, 
however, changes were made such that a prescribed 
variation, across the boundary layer, of the pres- 
sure gradient dp/dx could be used. This implies 
that the pressure field is known a priori. The 
solution of Eqs. (1), (3), and (4) together with 
Eqs. (6), (7), (8), and (9) can then be obtained 
through step-by-step integration by marching down- 
stream from some initial station where the velocity 
and shear-stress profiles are prescribed. A 
staggered mesh, explicit numerical scheme, similar 
to that used by Nash (1969), was used to integrate 
the equations in the domain between the first mesh 
point away from the surface (or the wake center- 
line) to some distance, typically 1.25 6, outside 
the boundary layer and the wake. The fifteen mesh 
points across the boundary layer are distributed 


Lap 1 
Fi LOF aT 
75 f, (Wake) 
C0) Sg RS ee 
6 08 5| 
(G)y=8 


06 


04 


0.2 


FIGURE 3. Distributions of empirical functions, 29 
and G. 


118 


non-uniformly to provide a greater concentration 
near the wall and the wake centerline. Instead of 
carrying out the integration of the equations up 
to the wall, i.e., through the viscous sublayer 
and the blending zone, the numerical solution at 
the first mesh point, located in the fully turbulent 
part of the boundary layer, is matched to the wall 
using the law of the wall. In the extension of the 
method to the wake, the matching between the first 
mesh point and the wake centerline is accomplished 
by using the conditions 3U/dy = O and t = O on the 
centerline. The main differences between the 
boundary layer and wake calculation procedures are 
therefore the treatment of the flow between the 
first mesh point and the wall or the wake center- 
line, and the change in 2, at the tail. Note that 
the local value of £2 in the boundary layer as well 
as the wake is different from 2, due to the lag, 
Eq. (8). The length scale recovers the reference 
distribution % 4 asymptotically in the far wake. 
Since the near wake data from the low-drag body 
indicated that most of the adjustment from the 
boundary layer to the far wake is accomplished over 
roughly five initial wake thicknesses, the lag 
length for the wake calculation was taken to be 

5 6, rather than 10 6 used for the boundary-layer 
calculation on the basis of Bradshaw's (1973) sug- 
gestion. Since the extra rates of strain vanish 
at the tail (k = 0, dr j/dx = 0), the length scale 
approaches the £2, distribution at about five wake 
radii downstream of the tail. 

Preliminary calculations performed with the dif- 
ferential method described above quickly indicated 
that the extra rates of strain in both experiments 
were much larger than those examined by Bradshaw 
(1973) in support of the linear length-scale 
correction formula of Eq. (8). In fact, the use 
of the linear formula led to a rapid decrease in 2 
and indicated almost total destruction of the 
Reynolds stress across the boundary layer in the 
tail region and the near wake. In view of this, 
recourse was made to a non-linear correction formula 
in the form 


ae 

L eff,- 1 
— = 1- — 

Qo { DU/ay? (8a) 
which reduces to the linear one, Eq. (8), for 
small extra rates of strain. Equations (1), (3), 
and (4), together with (6), (7), (8a), and (9), 
were then solved with the following inputs: 


A: the measured wall pressure distribution Cow 
(i-e., no normal pressure variation) and 
L(y/8) = &o(y/8) 

B: the measured Cpy with %£(y/s) corrected for 
only the longitudinal curvature (e = eg) 

C: the measured Cpy with 2(y/6) corrected for 
only the streamline convergence (e = e;) 

D: as above, but with e = Cn sr Gs 

E: using e = e€, + ey in Eqs. (8a) and (9), and 
a variable dp/dx across the boundary layer 
evaluated by assuming a linear variation in 
p from y = 0 to y = 6 and using the measured 
values of Cpw, Cpg and 6. 


Thus, case A corresponds to an axisymmetric bound- 
ary layer with thin, two-dimensional boundary-layer 
physics. The other cases enable the evaluation of 
the relative influence of the extra rates of strain 
as well as the static pressure variation through 


the boundary layer. The calculations were started 
with the velocity and shear-stress profiles mea- 
sured at X/L = 0.662 on the modified spheroid and 
at X/L = 0.601 on the low-drag body. 


4. COMPARISONS WITH EXPERIMENT 


The major results of the calculations are summarized 
in Figure 4(a-k) for the low-drag body and in 

Figure 5(a-h) for the modified spheroid. However, 
in the latter case the calculations are restricted 
to the boundary layer since detailed measurements 
were not made in the wake. Both figures contain 
comparisons between the experimental and calculated 
velocity, shear-stress, and mixing-length profiles 
at a few representative axial stations as well as 
the development of the integral parameters, 62, Ad, 
H, H, and Cg, with axial distance. These parameters 
are defined by 


Seis “en sae PE, BeBe GO) 
0 U OU 
6 6 
A, = fi (ay = =) ray, Ao = ie cz (il > =) xash7, 
U 
6 6 6 
H = Aj/Ad2 (11) 
and 
tw 
Ce = (12) 
4pU> 
6 


Where Us is the velocity component at the edge of 
the boundary layer and wake (y = 6), tangent to the 
body surface for the boundary layer and parallel 


fe) %, 


a ~ UV y2 


e Yu, 
~--= (A) Cg¢¢ =O. 
—— a ((E) e-e,+ ey, 


. op 
Linear Vax 


Vy 
io) 0.0005 0.0010 0.0015 0.0020 
-uv 
/y2 
FIGURE 4(a). Comparison of measurements with the solu- 


tion of the differential equations, low-drag body. Ve- 
locity and shear stress profiles at X/L = 0.920. 


[ U a, U u— T T T T T aT I 
fo) 7Uo fo) %, 
0.1000} a la 4 ~< Yo 
72 f 4 TUN y2 
V, 0.0750 
O'My. Ot | 
sae A | 
0.0875 + (A)@ eff = 0 7777 (A) eee =O 
—-—(B)e=e, 0.0625 4 —— (=) GOO, 4 
— = 0) 6 8 ; ap 
0.0750+ t 4 Linear 7 9x 
—-—(D) e=e,+e; 
0.0500 a 
=e,+ op, 
Me (E)e ey 4, Linear 9P/,, 
ie 4 Wy 
0.0375 4 
0.0250 a} 
0.0125 4 
4 
1 1 
(0) . 1.0 1.2 
di Q V, 
Yu5» “Up 
L 4 n 4 4 
fo) 0.0005 0.0010 0.0015 0.0020 
TUNZ 2 
t tl fe it i FIGURE 4(c). Velocity and shear stress profiles at 
(0) 0.0005 0.0010 0.0015 0.0020 X/L = 1.000. 
- Uy, 
u 72 
FIGURE 4(b). Velocity and shear stress profiles at 
X/L = 0.960. 
T T maT T T T 
T T T T T Q 
° 
Oo Vy Us 
= 0.0750 a uv y2 | 
0.0750 4 T UV 2 a ° 
° 
nj 
e Wy e 70, 
° 
—--— (A) Cg¢¢ = 0 fe) 
=-~~ (A) egg = 0 ° 0.0625 ct ° | 
0.06254, ° . (E) e=eyt &, 0 
Trimm Say ESO 2 Linear 9P, 
tines® aP/ax g inear 7 ax 
eg 0.0500 
(o) 4 
° yy 
yy FS | 7 
J 0.0375 7 
0.0250 | 
| 
0.0125 4 
(e) JJ] 
“ 12 
a 10 1.2 
Uc" /u n 4 4 4 
< : 1 : 4 0.0005 00010 00015 00020 
ie) 0.0005 0.0010 0.0015 0.0020 -uy, - 
oy, U 
7u3 ‘ 
FIGURE 4(d). Velocity and shear stress profiles at FIGURE 4(e). Velocity and shear stress profiles at 
X7/L = 1.06. X/L, = 1.20. 


ELS) 


120 


= T eal T T Sa 
Q 
° /U, 
00750 4 TONE ] 
Vv 
O 7Uo 
———— (A) Cpg¢= 0 ° 
0.0625 Gi ; 
—— (E) e= e+ e,, 
Linear 9P/9, 
0.0500 4 
MA 
0.0375 4 
0.0250 4 
0.0125 4 
fo} 1 di Sl 
a OR, 08 1.0 12 
uy 0 
L 4 ——i 4 4 
{e) 0.0005 0.0010 0.0015 0.0020 
—uy, 
7u2 


FIGURE 4(f) . 


Velocity and shear stress profiles at 


X/L = 2.472. 
=r T T T T = T T te 
0.125+ 4 
Asymp. Value 
I © Experimental 8 2 
o.100- = \ === (A) e44=0 2 | 
. toy, —-— (D) e=ey +e, - 
X\ —— (e)e=ey +e, Linear 9M, ' 
, 0.075+ J 
0 
Ke 
8 fe} 
i x 
0.050; al 
0,025+ 4 
| 
| 
° 
Co} 
05 are 
FIGURE 4(h). Boundary layer and wake thickness. 
| ——4 : : ro i 
es Asymp. Value 
2i2 hasan ye 4 
} Experimental 
aroma 
| 
| SIM Gy 3O x 
eff 
ech —-— (dD) e=ep+e, - 1 
—— (E)e-e, +e, | 
18+ Linear 9P/, | 
HLH 
16+ 4 
14+ | 
+ 
ler tq 
8 
10+ J = —) n 1 n 1 sf. 
OS NOG IO; 7s OG NLO!S 10 i 12 3 147 25 


FIGURE 4(3) 


Shape parameters. 


X/L 
ol © 0.601 | 


vy 0.920 
a 0.960 
x 1.000 
Colculation, e-ey+e,, Lineor my, 


0.10 


0,08 


0.06 


0.04 F 


0.02 


FIGURE 4(g) . 


[ T =a T ——T, To 
0.16 Le 5 
IV 2 106 
o 120 | 
O.14F x 247 
Calculation, e = ey +e. 
oz} * Linear 9%, J 


Mixing length profiles. 


T T T T T T T T fH 
Asymp. Value 
L a 8, 
eS 2 } Experimental 7 
QO! Bs 
== (A) €g¢¢=0 x 
—-— (D) e=-ey +e - 
20F a © ( Nee 4 
3, o —— (Eye=ey+e,, \ 
TS Linear P73, 
10 
USE 
es 
Lz 
G10) 
lo 
if ° 
o.5- 
G2 
005 


FIGURE 4(1i). 


Planar and axisymmetric momentum deficits. 


Ty T T T 
L i x,Q Preston Tubes | 
OES (1.651 mm, 0.711 mm) 
6 fe} Clauser Plot 
g —=-—= (A) Gg4¢= 0 
0.004 + Tp Gee Sy >I 
(E)e= ept+e,, 
Linear OM 
0.003 + DE al] 
Cr 
0.002 + 4 
Oooltf- =| 
0 N avs a 1 i 
0.5 0.6 0.7 0.8 0.9 ie) 
“i 
FIGURE 4(k). Wall shear stress. 


y 
A 
0.0250 + 
0.0125 
10) 0.2 0.4 0.6 0.8 1.0 1.2 
Q V, 
Aso Ms 
\ 1 fl L j 
ie) 0.0005 0.0010 0.0015 0.0020 
- UV, 
40 
FIGURE 5. Comparison of measurements with the solution 


° My, 
a —UuV 2 
“U2 
e Yul, 
a UN Gas = ©) 
° ° 
—— (£) e=eytey, 
Linear 9P/9, i 
e ° 
e ° 


of the differential equations, modified spheroid. 


(a) Velocity and shear stress profiles at X/L = 0.930. 


0.0750- 


0.0625+ 


0.0500 


T T >is T aa T 
fo) Hy, | 
a -UVs 2 ° 
Vv ° 
e 70, ° 
--- (A) eef¢ = 0 ° 4 
— (E) e=e, +e, 6 


Linear 9P/5, 


| 
' 
1 
I 


0.0375 | J 
y 
AL 
0.0250 + 4 
0.0125 4 
4 
{e) 1.0 1.2 
—aee | 
(0) 0.0005 0.0010 0.0015 0.0020 
TON 
FIGURE 5(c). Velocity and shear stress profiles at 
X/L = 0.990. 


T T Q “1 T = T 3 
0.0750+ oO, | 
oO 
a -UV, 2 
° 
Vv 
e My, 
OCGIs | | ae (Ale eff = O 4 
Eee Ee 
4 A (B) ey 
—--—(C)e=e, 
ie) 


X/L = 0.960. 


a 
0.0250 e 
0.0125 4 
at 
ie) 1.2 
lo) 0.0005 0.0010 0.0015 0.0020 
- Uy, 
/u2 

FIGURE 5(b). Velocity and shear stress profiles at 


0.12 


0.10 


0.08 


0,06 


0.04 


0.02 


XP 
0.622 


0.960 
0.990 
Calculation, e=@¢+@,, Linear om, 


xogqgoao 


0.930 4 


FIGURE 5(d). 


Mixing length profiles. 


121 


122 


= T p= T ao 
0100+ ——== (A) Cee = 0 
—-— (D)e=ep +e, 
(E)e=e 9 +ey, 
=== : ap 
— Linear “'/ 
0.075 oe Ox 
TosL 
8 
7 
t  o.0s0b 
0.025- 
i lL te 
0.5 06 
FIGURE 5(e). Boundary layer thickness. 
A H 
H 
Bal © 5 
Ses WW) Gage O 
<a> (0) e= ep +e, 
2.07 (eye =Op tes, Linear OM, 7 
Or 4 
AH is 
16) =| 
Lae 4 
ea = 
1.0 l | n i a 
0.5 0.6 0.7 08 0.9 10 
X71. 
FIGURE 5(g). Shape parameter. 


to the axis for the wake. In the interest of clar- 
ity, the results of all the calculations (cases A 
through E) are shown only at one axial station 
(Figure 4b and 5b), those at other stations being 
qualitatively similar. 

Considering the most detailed figures, 4b and 
5b, first, it is clear that the predictions are 
rather poor when the length scale, &, is assumed 
to be the same as that in a thin boundary layer 
(case A). This is particularly evident in the pre- 
diction of the shear-stress profiles across the 
boundary layer and the near wake. Incorporation 
of the correction to & to account for the extra 
rate of strain due to longitudinal curvature (case 
B) leads to a marginal improvement in the case of 
the low-drag body and a dramatic improvement for 
the modified spheroid. This is to be expected in 
view of the grossly different surface curvature 
histories of the two bodies as noted earlier 


(se = Sa ina T To 
a 8) 
(le AN 
720) )—, 2 =| 
SA) Cort =O 
et am ORC CIE fe) 
(E)e=ep+e,, 
= 1S) = £ 4 
8 Linear Op 
L 
(x10) 
Qo lO 
ue 
(x10) 
OS) |r 
(o) rt 
O5 0.6 
FIGURE 5(f). Planar and axisymmetric momentum deficits. 
T T lin erarlies 
0.005 a Preston Tube 
| O° Clouser Plot 
Sa == WN) Gaggs © 
—-— (D)e=e, +e, 
O00) 
(Eve=e) +e), 
Linear oy, 
0.003 
Cr 
HNO | 
0.001 
(e) 
o5 06 0.7 08 0.9 1.0 
ATi 
FIGURE 5(h). Wall shear stress. 


(Figure 2). Nevertheless, it is clear that this 
correction by itself is not sufficient to account 
for the differences between the data and the calcu- 
lations with thin boundary-layer turbulence models 
(case A). The application of the correction for 
the extra rate of strain due to the transverse 
curvature (case C) appears to account for a major 
portion of these differences for both bodies. The 
influence of transverse curvature is in fact seen 

to be somewhat larger for the low-drag body as 
would be expected from the fact that 6/ro is greater 
in that case (Figure 2). The simple addition of 

the effects of the two rates of strain (case D) 
leads to a significant improvement in the prediction 
of both the velocity profiles and the shear stress 
profiles. The incorporation of a variable pressure 
gradient across the boundary layer (case E), which 
is an attempt to account for the normal pressure 
gradients, appears to make a significant improve- 


ment in the prediction of the velocity profile in 
the case of the modified spheroid, but its influence 
is small, and confined to the outer part of the 
boundary layer, in the case of the low-drag body 

Examination of the velocity and shear-stress 
profiles at several axial stations shown in Figures 
4a-f and 5a-c suggests that the incorporation of 
the non-linear length-scale correction of Eq. (8a), 
the associated rate Eq. (9) and the static-pressure 
variation in the equations of the thick boundary 
layer, which already include the direct longitudinal 
and transverse curvature terms, leads to satis- 
factory overall agreement with the data for both 
bodies. It is particularly noteworthy that the 
velocity and shear stress distributions in the 
far wake (X/L = 2.472) of the low-drag body are 
predicted with good accuracy. The level of 
agreement can obviously be improved further by 
appropriate modifications in the empirical functions 
in the turbulent kinetic-energy equation and changes 
in the lag-length used in the length-scale equation. 
The predictions of the shear stress profiles are 
consistent with those of the mixing-length distri- 
butions shown in Figures 4g and 5e insofar as lower 
shear stresses correspond to an over correction in 
the mixing length. These comparisons provide 
further insight into the manner in which the length 
scale must be modified to improve the correlation 
between the calculation method and experiment. It 
is apparent that the consistent discrepancy between 
the calculated and measured velocity and shear- 
stress profiles near the outer edge of the boundary 
layer and wake stems from a poor representation of 
the length scale distribution. 

It is interesting to note that, for both bodies 
the calculation precedure predicts normal components 
of mean velocity which are of the same order of 
Magnitude as those measured. The relatively close 
agreement between the predictions and experiment 
for both components of velocity is perhaps a good 
indication of the axial symmetry achieved in the 
experiments. The large values of the normal veloc- 
ity and the influence of static pressure variation 
noted above would appear to indicate that incorpora- 
tion of the y-momentum equation in the calculation 
procedure would be worthwhile. Note that this has 
been avoided in the present calculations by using 
the measured pressure distributions at the surface 
and the outer edge of the boundary layer. 

Finally, the comparisons made in Figures 4 (i-k) 
and 5 (e-h) with respect to the integral parameters 
show several interesting and consistent features. 

It is observed that the prediction of the physical 
thickness of the boundary layer and the wake is 
insensitive to the changes in 2 as well as the in- 
clusion of static pressure variation. The under 
estimation of the thickness is associated with the 
discrepancy, noted earlier, in the velocity profile 
near the outer edge of the boundary layer and wake. 
The planar momentum thickness 69 and the momentum- 
deficit area A» are also insensitive to changes in 
2. The variation of static pressure across the 
boundary layer appears to make a small but notice- 
able contribution to the development of A» in both 
cases. However, it is not large enough to account 
for the differences between the calculations and 
experiment. The predictions of the shape parameters, 
H and H, presented in Figures 4j and 5g, appear-to 
be satisfactory, especially in view of the rather 
_large scale of the plots. Nevertheless, there is 

a systematic difference between the data and the 


123 


calculation in the tail region and wake of the low- 
drag body. As indicated earlier, this can be im- 
proved by modifications in the empirical functions 
and the lag length. The predictions of the wall 
shear stress, shown in Figures 4k and 5h, indicate 
that the present method gives acceptable results 
for both bodies. 


5. COMPARISONS WITH THE INTEGRAL APPROACH 


An integral method for the calculation of a thick 
axisymmetric boundary layer was described by Patel 
(1974) and its extension to the wake was proposed 
by Nakayama, Patel, and Landweber (1976b). A few 
possible improvements in this method were examined 
recently relative to the description of the velocity 
profiles in the near wake and these are discussed 
by Patel and Lee (1977). The most recent version 
of this method has been used here to calculate the 
development of the boundary layer and the wake of 
the low-drag body in order to assess its performance 
relative to the experimental data (which were not 
available at the time the method and its extension 
were proposed) and the more elaborate differential 
method. 

The results of the calculations are shown in 
Figure 6. It is seen that the performance of the 
integral method is comparable with that of the 
differential method (compare Figures 4h-k with 
6a-d) with respect to the prediction of the bound- 
ary layer up to the tail. The prediction of the 
near wake is, however, distinctly inferior to that 
of the differential method, particular with respect 
to the physical thickness 6 and momentum deficit 
area Ay. The main conclusion to emerge from these 
calculations is that the integral method is capable 
of giving a good overall description of the flow 
features with considerably less computing effort. 
The differential approach is to be preferred, how- 
ever, Since it affords the opportunity for further 
refinement and gives greater details which may be 
necessary for many applications. A more thorough 
discussion of the integral method and its short- 
comings is given in Patel and Lee (1977). 


6. CONCLUSIONS 


From the present solutions of the differential 
equations, using the (one-equation) turbulent 
kinetic-energy model of Bradshaw, Ferriss, and 
Atwell (1967), it is clear that methods developed 
for thin shear layers cannot be relied upon to pre- 
dict the behavior of the thick boundary layer and 
wake of a body of revolution. Although these cal- 
culations have demonstrated that the boundary-layer 
calculation can be readily extended to the wake 

and that a fairly satisfactory prediction procedure 
can be developed by incorporating ad hoc corrections 
to the model for the extra rates of strain, along 
the lines recommended by Bradshaw (1973), it is 
indeed surprising that such modifications, proposed 
originally for small extra rates of strain and thin 
shear layers, work so well for the two bodies which 
are substantially different in shape. In keeping 
with recent trends in the formulation of turbulence 
models, one inquires whether thick axisymmetric 
boundary layers and near wakes ought to be treated 
by the so-called two-equation models. From the 
yapid changes in the mixing-length indicated by 


124 


0.125 4 


ze ——_ Integral Method x 
0.100 IN ° Experiment 8 7 


0.075 + \ x | 
oy, ; 
ty 
AL 
ol 
0.025 + 
FIGURE 6. Comparison of ex- ©) 
periments with the solution of 6 | 
the integral equations, low- 05 
drag body. (a) Boundary layer 
and wake thickness- 
ere T aa T T =I T Saloon eal 
Asymp. Value 
Integral Method x 
2.5 A Experiment 8. 
° Experiment Qo 
FIGURE 6(b). Planar and axisymmetric momen- 


tum deficits. 


22+ ———_ Integral Method x 
4 Experiment H 
fo} Experiment H 
2.0F 
18- 
H,H 
16 
ise oy 6) 
x 
Lab ° = 
a 
° 
Lo 1 i 1 1 1 1 th 1 ~~ +} 
05 06 0.7 0.8 0.9 1.0 il 12 13 1.4 2.5 


FIGURE 6(c). Shape parameter. ae 


Ir Tigra pian bz ll ir 
0.005 | x4 Preston Tubes =| 
(1.651 mm, 0.711 mm) 
6 ° Clauser Plot 
Q a Integral 
0.004 |- s mm Method =| 
° 
g 
0,003 - 4 
0.002 - 4 
0.00] | =| 
(@) Jt ! 1 i 1 
0.5 06 0.7 08 09 1.0 
x 
ae 
FIGURE 6(d). Wall snear stress. 


the data, this would appear to be desirable since 
it would provide an extra equation for the length- 
scale of the turbulence in addition to that for 

its intensity. This would also enable the incorpo- 
ration of the variations in the structure parameter, 
a,, observed in the experiments. However, the 
recent work of Launder, Priddin and Sharma (1977) 
and Chambers and Wilcox (1977) indicates that even 
two-equation models, at least of the type available 
at the present time, require further modifications 
to account for the extra rates of strain stemming 
from such effects as streamline curvature, stream- 
line convergence, and rotation, two of which are 
present in the case examined here. 

In addition to the problem of turbulence models, 
the thick boundary layer and the near wake contain 
the complication of normal pressure gradients. The 
available data show that there exist substantial 
variations of static pressure across the boundary 
layer. The calculations presented here as well 
as those performed with the integral method by Patel 
and Lee (1977), suggest that the influence of the 
normal pressure gradients on the development of 
the boundary layer and the near wake is not negli- 
gible although it is masked by the rather major 
effects of the transverse and longitudinal surface 
curvatures on the turbulence. If normal pressure 
variations are to be taken into account in a method 
based on the differential equations, it is neces- 
sary to include the y-momentum equation in the 
solution procedure and regard the pressure as an 
additional unknown. This is perhaps best accom- 
plished by means of an iterative scheme such as 
that proposed by Nakayama, Patel, and Landweber 
(1976a,b), although other possibilities can be ex- 
plored. In view of the success of the present dif- 
ferential method, it is proposed to incorporate 
the present method in this iterative scheme, in 
place of the integral method, to study the viscous- 
inviscid interaction in the tail region in greater 
detail. é 

The representative calculations presented in 
Section 5 demonstrate the overall reliability of 

~the simple integral method of Patel (1974) for the 


125 


prediction of the thick boundary layer. Its ex- 
tension to the wake is not altogether satisfactory 
and this is attributed largely to the lack of a 
systematic procedure for the description of the 
velocity profiles in the near wake. This method 

is ideally suited, however, for rapid calculations 
to determine the state of the boundary layer in the 
tail region for certain applications. 


ACKNOWLEDGMENTS 


This research was carried out under the sponsor- 
ship of the Naval Sea Systems Command, General 
Hydro-Mechanics Research Program, Sub-project 

SRO23 O01 O01, administered by the David W. Taylor 
Naval Ship Research and Development Center, Contract 
NO0014-75-C-0273. The authors acknowledge the 
assistance of Professor B. R. Ramaprian through 
several stimulating discussions on the influence 

of longitudinal surface curvature on turbulent 
boundary layers. 


REFERENCES 


Bradshaw, P. (1969). The Analogy Between Stream- 
line Curvature and Buoyancy in Turbulent Shear 
WAV Va | Wie saul) We=yelalg Ska, 7/76 

Bradshaw, P. (1973). Effects of Streamline Curva- 
ture on Turbulent Flow. AGARDograph No. 169. 

Bradshaw, P., and K. Unsworth (1976). Computation 
of Complex Turbulent Flows, in Reviews of Viscous 
Flows. Proceedings of the Lockheed-Georgia 
Company Symposium, 448. 

Bradshaw, P., D. H. Ferriss, and N. P. Atwell 
(1967). Calculation of Boundary Layer Develop- 
ment Using the Turbulent Energy Equation. J. 
Fluid Mech. 28, 593. 

Chambers), Ls 5Le and) DriGe Witlicox (S77) — Grittical 
Examination of Two-Equation Turbulence Closure 
Models for Boundary Layers. AIAA Journal 15, 

6; 821. 

Launder, B. E., C. H. Priddin, and B. I. Sharma 
(1977). The Calculation of Turbulent Boundary 
Layers on Spinning and Curved Surfaces. ASME, 
J. Fluids Engineering 99, No. 1, 231. See also 
discussions by P. Bradshaw and G. Mellor, (1977), 
ASME, J. Fluids Engineering 99, 2; 435. 

Meroney, R. N., and P. Bradshaw (1975). Turbulent 
Boundary Layer Growth Over Longitudinally Curved 
Surfaces. AIAA Journal 13, 1448. 

Nakayama, A., V. C. Patel, and L. Landweber (1976a). 
Flow Interaction Near the Tail of a Body of 
Revolution, Part I: Flow Exterior to Boundary 
Layer and Wake. ASME, J. Fluids Engineering 98, 
Sie 

Nakayama, A., V. C. Patel, and L. Landweber (1976b) . 
Flow Interaction Near the Tail of a Body of 
Revolution, Part II: Iterative Solution for 
Flow Within and Exterior to Boundary Layer and 
Wake. ASME, J. Fluids Engineering 98, 538. 

Nash, J. F. (1969). The Calculation of Three- 
Dimensional Turbulent Boundary Layers in Incom- 
pressible Flow. J. Fluid Mech. 37, 625. 

Patel, V. C. (1973). On the Equations of a Thick 
Axisymmetric Turbulent Boundary Layer. Towa 
Institute of Hydraulic Research, ITHR Report 
No. 143. 

Patel, V. C. (1974). A Simple Integral Method 
for the Calculation of Thick Axisymmetric Tur- 


126 


bulent Boundary Layers. The Aeronautical 
Quarterly 25, 47. 

Patel, V. C., and Y. T. Lee (1977). Thick Axisym- 
metric Turbulent Boundary Layer and Near Wake of 
a Low Drag Body of Revolution. Iowa Institute 
of Hydraulic Research, IIHR Report No. 210. 

Patel, V. C., A. Nakayama, and R. Damian (1974). 
Measurements in the Thick Axisymmetric Turbulent 
Boundary Layer Near the Tail of a Body of 
Revolution. J. Fluid Mech. 63, 345. 

Ramaprian, B. R., and B. G. Shivaprasad (1977). 
Mean Flow Measurements in Turbulent Boundary 
Layers Along Mildly-Curved Surfaces. AIAA 
Journal 15, 189. 


Shivaprasad, B. G., and B. R. Ramaprian (1977). 
Some Effects of Longitudinal Wall-Curvature 
on Turbulent Boundary Layers. Proceedings of 
Symposium on Turbulent Shear Flows, Penn State 
University, 9.21. 

So, R. M. C., and G. L. Mellor (1972). An Experi- 
mental Investigation of Turbulent Boundary Layers 
Along Curved Surfaces. NASA-CR-1940. 

So, R- M. C., and G. L. Mellor (1974). Experiment 
on Convex Curvature Effects in Turbulent Bound- 
ary Layers. J. Fluid Mech. 60, 43. 

So, R. M. C., and G. L. Mellor (1975). Experiment 
on Turbulent Boundary Layers on Concave Wall. 
The Aeronautical Quarterly 26, 35. 


Stern Boundary-Layer Flow on 
Axisymmetric Bodies 


TA DavHuang, N. Santelilkl, wand! Go Belt 
David W. Taylor Naval Ship Research and Development 
Center, Bethesda, Maryland 


ABSTRACT 


Measurements of static pressure distributions, mean 
velocity profiles, and distributions of turbulence 
intensities and Reynolds stress were made across the 
stern boundary-layers on two axisymmetric bodies. 

In order to avoid tunnel blockage, the entire after- 
body was placed in the open-jet test section of the 
DTNSRDC Anechoic Wind Tunnel. The numerical itera- 
tion scheme which uses the boundary layer and open 
wake displacement body is found to model satisfac- 
torily the interaction between the thick stern bound- 
ary layer and the external potential flow. The 
measured static pressure distributions across the 
entire stern boundary layer and the near wake are 
predicted well by potential flow computations for 
the displacement bodies. The measured distributions 
of mean velocity and eddy viscosity over the stern, 
except in the tail region (X/L > 0.90), are also 
well-predicted when the Douglas CS differential 
boundary-layer method is used in conjunction with 
the inviscid pressure distribution on the displace- 
ment body. However, the measured distributions of 
turbulence intensity, eddy viscosity, and mixing- 
length parameters in the tail region are found to 

be much smaller than those of a thin boundary layer. 
An approximate similarity characteristic for the 
thick axisymmetric stern boundary layer is obtained 
when the mixing-length parameters in the tail region 
are normalized by the square-root of the boundary- 
layer cross-sectional area instead of the boundary- 
layer thickness. 


1. INTRODUCTION 


Many single-screw ship propellers operate inside of 
thick stern boundary layers. An accurate prediction 
of velocity inflow to the propeller is essential to 
meet the ever-increasing demand for improving pro- 
peller performance. Huang et al. (1976) used a 

- Laser Doppler Velocimeter (LDV) to measure the ve- 


127 


locity profiles on axisymmetric models with and 
without a propeller in operation. The measured 
difference between these velocity profiles has 
provided the necessary clues to formulate an inviscid 
interaction theory for propellers and thick boundary 
layers. An iterative scheme was employed to compute 
the velocity profiles of the thick axisymmetric 
boundary layer. In this approach, the initial 
boundary-layer computation proceeds making use of 
the potential-flow pressure distribution on the body 
[Hess and Smith (1966)]. The flow calculations are 
then repeated for a modified body and wake geometry, 
by adding the computed local displacement thickness 
as suggested by Preston (1945) and Lighthill (1958). 
Potential-flow methods are then used to compute the 
pressure distribution around the modified body and 
the boundary-layer calculations are repeated using 
the new pressure distribution. The basic iterative 
scheme is continued until the pressure distributions 
on the body from two successive approximations agree 
to within a given error criterion (1 percent). 

The Douglas CS differential boundary-layer method 
[Cebeci and Smith (1974)], modified to properly ac- 
acount for the effects of transverse curvature, was 
used to calculate the boundary-layer over the axi- 
symmetric body. The integral wake relations given 
by Granville (1958) were used to calculate the dis- 
placement thickness in the wake. In the stern/ 
near-wake region (0.95 £ x/L £ 1.05), where X is the 
axial distance from the nose and L is the total 
length, a fifth-degree polynomial was used, with 
the constants determined by the condition that the 
thickness, slope, and curvature be equal to those 
calculated by the boundary-layer method at X/L = 
0.95 and by the integral wake relations at X/L = 
1.05. Comparison with experimental results of Huang 
et al. (1976) show that the potential-flow/boundary- 
layer interaction computer program predicts accurate 
values of pressure, shear stress, and velocity pro- 
files over the forward 90 percent of the bodies, 
where the boundary layers are thin compared with 
the radii of the bodies. Over the last 10 percent 


128 


of body length, the measured shear stress and ve- 
locity profiles became smaller than those predicted 
by the theory. These differences are more notice- 
able over the last 5 percent of the body length 
where the boundary-layer thicknesses are greater 
than the radii of the bodies, especially for fuller 
sterns. 

In order to examine the thick stern boundary- 
layer properties in detail, it is necessary to 
measure the distributions of static pressure, tur- 
bulence intensities and Reynolds stress across the 
thick stern boundary layer. The magnitudes of the 
eddy viscosity and the mixing-length parameter were 
determined and compared with those obtained for 
thin boundary layers. It is found that the eddy 
viscosity and the mixing length for thick boundary 
layers are smaller than those of thin boundary 
layers. An improvement to the Douglas CS differ- 
ential method can be made by modifying the mixing- 
length model in the tail region. The distributions 
of measured static pressure, which were found to be 
nonuniform across the thick stern boundary layers 
and near wake, can be approximated very well by 
potential flow computations for the displacement 
bodies. The gross curvature effects of the mean 
streamlines on the static pressure distributions 
outside the displacement surface are represented 
very well by those of the potential-flow stream- 


TABLE 1 - Offsets for Model 1 


K/L Y/t Y/R X/L 
0.0000 9.0000 0.0000 - 2684 
0050 -0100 2193 28S) 
o0se -0142 3118 - 2883 
0149 -0175 3835 - 2982 
0199 0202 4441 3082 
+0249 -0227 4975 .3181 
0208 .0248 5454 ~ 3280 
0348 0268 5891 - 3380 
0398 -C287 6291 ~3479 
9447 -0303 6659 “3979 
0497 OSS) 7000 . 3678 
05417 0333 7315 SON AMaT| 
6596 -0347 7607 . 3877 
0646 -0359 7877 .3976 
0596 0379 8126 .4076 
0746 -0381 8355 -4175 
0795 -0390 8567 +4274 
0845 -0399 8760 - 4374 
0895 -0407 8936 -4473 
og44 -0414 9097 -4573 
oo94 0421 9241 »4672 
-1044 .0427 9371 -4771 
1093 0432 9466 -4871 
1143 -0437 9587 -4970 
1H}93 -0441 9676 -5070 
1243 .0444 9752 -5169 
1292 .0447 9816 -5268 
1342 -0450 9869 “9268 
11392 -0452 9912 ~5467 
144) -0453 9445 ~5567 
1491 -0454 9969 -5666 
1541 -90455 9986 .9765 
1590 -0448 9836 ~ 5865 
1640 .0456 1.0000 -5964 
1690 .0456 1.0000 .6064 
1740 .0456 1.0000 -6188 
1789 .0456 1.0000 .6264 
1839 .0456 1.0000 .6378 
1889 .0456 1.0000 -6454 
1938 .0456 1.0000 - 6567 
1988 -0456 1.0000 .6681 
2087 -0456 1.0000 ~6757 
2187 -0456 1.00C0 -G871 
2286 .0456 1.0000 -6984 
2356 -0456 1.0000 7060 
2485 .0456 1.0000 ~7174 
2584 -0456 1.0000 -7250 


lines of the fictitious displacement body. Thus, 
the nonuniform static pressure distributions across 
the thick stern boundary layer can be interpreted 
mainly as an inviscid phenomenon and can be assumed 
to have little effect on the stern boundary-layer 
development. 

Two axisymmetric bodies without flow separation, 
Afterbodies 1 and 2 of Huang et al. (1976), were 
chosen for this investigation. Their geometric 
simplicity offers considerable experimental and 
computational convenience in treating fundamental 
aspects of thick stern boundary layers. Afterbody 
1 is a fine convex stern while Afterbody 2 is a 
full convex stern. 

In the following discussion, the experimental 
techniques and geometries of the model are given 
in detail. The measurements of mean velocities, 
turbulence intensities, and Reynolds stresses were 
analyzed to obtain eddy viscosity and mixing length. 
The application of the present results to improve- 
ment of the accuracy of boundary-layer computations 
over the entire stern is outlined. 


2. WIND TUNNEL AND MODELS 


The experimental investigation was conducted in the 
wind tunnel of the DTNSRDC anechoic flow facility. 


NA Y/R X/L Y/L Y/R 
-0456 1.0000 - 7363 -0427 -9382 
-0456 1.0000 -7477 +0421 -9235 
0456 1.0000 .7553 -0416 SAY 
-0456 1.0000 - 7666 -0408 -8951 
0455 1.0000 - 7780 -0399 -8755 
0456 1.0000 - 7856 -0392 -8615 
0956 1.0000 ESTO: -0382 - 8387 
0456 1.0000 -8045 -0375 -8225 
0456 1.0000 -8159 -0363 - 7963 
0456 1.0000 -8273 0350 SASS) 
0456 1.0000 -8349 -0341 - 7477 
-0456 1.0000 -B8462 -0326 si NSS) 
0456 1.0000 -8576 -0310 - 6807 
0456 1.0000 - 8652 -0299 - 6560 
0456 1.0000 -8765 -0281 - 6167 
0456 1.0000 8841 .0268 -5889 
0456 1.0000 -8955 -0248 -5445 
9455 1.6000 +9069 +0226 -4970 
0456 1.0000 -9144 -0211 - 4633 
0456 1.0000 -9245 -0189 ~4147 
0456 1.0000 +9344 -0166 3636 
0456 1.0000 -9443 -0140 . 3078 
0456 1.0000 ~9513 -0122 - 2673 
0456 1.0000 9563 -0108 - 2380 
0456 1.0000 -9612 -0095 - 2080 
0456 1.0000 -9642 .0087 - 1900 
0456 1.0000 -9662 -0081 -1778 
0456 1.0000 - 9682 -0076 - 1669 
0456 1.0000 9692 -0074 -1613 
0456 1.0000 -9702 -0072 6 US 
0456 1.0000 -9722 -0068 1489 
0456 1.0000 -9732 -0066 ~1453 
0456 1.0000 OO -0063 - 1393 
0456 1.0000 +9771 -0062 1364 
0456 1.0000 OPA -0059 +1293 
0456 9999) -9811 -0056 -1222 
0455 SON, -9831 -0053 +1165 
0455 9988 -9851 -0050 1107 
0455 9977 -9871 -C048 -1051 
0453 9952 -9881 -C046 -1018 
0452 g915 «9901 -0043 -0951 
0450 9883 +9920 -0040 - 0880 
0448 9R23 -9940 -0036 -0782 
0444 9748 -9960 -0028 -0625 
0441 9690 -9980 -0019 -0413 
0437 9588 1.0000 0.0000 9.0000 
0433 SSH 


The wind tunnel has a 2.44 m by 2.44 m closed-jet 
test section, followed by a 7.16 m by 7.16 m open- 
jet test section. The length of the open-jet sec- 
tion is 6.40 m. The maximum air speed which can 

be achieved is 61 m/sec; in the present experiments, 
the velocity of the wind tunnel was held constant 
at 30.48 m/sec. The measured ambient turbulence 
level in the open-jet test section without the model 
in place was 0.1 percent. Integration of the mea- 
sured noise spectrum levels in the open-jet test 
section, over the frequency range of 0 to 10,000 Hz, 
indicated that the typical background acoustic 
noise at 30.48 m/sec was around 93 db re 0.0002 
dyn/cm2. These levels of ambient turbulence and 
acoustic noise were considered low enough so as not 
to unfavorably affect the measurements of boundary- 
layer characteristics. 

Two axisymmetric convex afterbodies without 
stern separation were used for the present experi- 
mental investigation. Their afterbody length/ 
diameter ratios (La/D) were 4.308 and 2.247. The 
detailed offsets for Models 1 and 2 are given in 
Tables 1 and 2. Each afterbody was connected to a 
parallel middle body of length Ly and an existing 
streamlined forebody with a bow-entrance length 
diameter ratio (L,/D) of 1.82. The total length 
of each model (L) is fixed at a constant value of 
3.066 m. The diameter of the parallel middle body 
(emp) is 27.94 cm. The common forebody and a 
portion of the parallel were constructed of wood. 


TABLE 2 - Offsets for Model 2 


X/L Y/L Y/R x/L Y/L 
0.0000 0.0090 0.0000 . 2684 -0456 
.9050 -0100 .2193 . 2783 .0456 
.0099 .0142 .3118 .2883 .0456 
.0149 0175 . 3836 . 2982 -0456 
0199 .0202 4443 . 3082 0456 
.0249 .0227 -4975 23181 .0456 
-0298 .0249 5455 .3280 -0456 
-0348 .0268 -5891 . 3380 .0456 
-0398 -0287 -6291 .3479 -0456 
.0447 .0303 -6659 .3579 .0456 
.0497 -0319 .7000 .3678 .0456 
.0547 .0333 .7316 SCUY -0456 
.0595 0347 .7606 .3877 .0456 
-9646 0359 7877 .3976 -0456 
.0696 0370 -8126 .4076 0456 
0746 -0381 8357 .4175 -0456 
.0795 .0390 .8566 .4274 -0456 
.0845 0399 .8761 .4374 -0456 
.0895 -0407 .8937 .4473 - 0456 
.0944 .0414 .9097 4573 -0456 
.0994 -0421 .9241 4672 0456 
.1044 .0427 .9372 4774 -0456 
-1093 .0432 -9487 .4871 -0456 
.1143 .0437 -9588 .4970 -0456 
-1193 .0441 -9677 .5C070 .0456 
.1243 .0444 -9751 .5169 0456 
.1292 -0447 9817 5268 -0456 
.1342 -0450 . 9869 .5368 -0456 
.1392 .0452 .9913 .5467 0456 
1441 0453 .9945 5567 -0456 
-1491 .0454 -9969 .5666 -0456 
21541 .0455 - 9987 5765 .0456 
.1590 .0455 -9996 .5865 -0456 
-1640 -0456 1.0000 5964 .0456 
.1690 .0456 1.00069 6064 .0456 
.1740 -0456 1.0000 6188 -0456 
.1789 .0456 1.0000 .6264 -0456 
.1839 .0456 1.0000 .6378 -0456 
.1889 .0456 1.0000 6454 -0456 
-1938 0456 1.0000 6567 -0456 
-1988 .0456 1.0000 .6681 0456 
. 2087 .0456 1.0000 6757 -0456 
.2187 -0456 1.0000 .6871 .0456 
.2286 .0456 1.0000 -6984 .0456 
. 2386 -0456 1.0000 .7060 -0456 
.2485 9456 7%.0000 7174 .0456 
- 2584 -045€ 1.0000 .7250 -0456 


AAAS) 


The afterbody and the remaining portions of the 
parallel middle body were constructed of molded 
fiberglass; specified profile tolerances were held 
to less than +0.4 mm, all imperfections were re- 
moved, meridians were faired, and the fiberglass 
was polished to a 0.64-micron rms surface finish. 
The tail ends of the afterbody were shaped to ac- 
commodate the hub of an existing propeller. This 
modification caused a considerable change of body 
curvature in the region of X/L 2 0.96. However, 
as will be seen later, the thicknesses of the 
boundary layer in this region are much larger than 
the local radii of the body. This deficiency does 
not cause serious degradation of boundary-layer 
flow at that point. 

The model was supported by two streamlined struts 
separated by roughly one-third of the model length. 
The upstream strut had a 15 cm chord and the down- 
stream strut a 3 cm chord. The disturbances gener- 
ated by the supporting struts were within the region 
below the horizontal centerplane. Prior to the 
experiments, pressure taps and Preston tubes were 
used to check the axisymmetric characteristics of 
the stern flow at X/L 0.90, 0.95, and 0.98. The 
circumferential variations of pressure and surface 
shear stress on the upper half of the two after- 
bodies at these three locations were within two 
percent. All the final measurements were made in 
each body's vertical centerplane along the upper 
meridian where there was little extraneous effect 


Y/R X/L Y/L Y/R 
1.0000 .7363 -0456 1.0000 
1.0000 .7477 -0456 1.0000 
1.0000 -7553 -0456 1.0000 
1.0000 . 7666 -0456 1.0000 
1.0000 .7780 0456 1.0000 
1.0000 .7856 -0456 1.0000 
1.0000 .7952 0456 1.0000 
1.0000 - 8050 .0455 .9996 
1.0000 -8147 .0454 .9959 
14.0000 .8245 -0450 -9871 
1.0000 .8342 -0443 .9723 
1.0000 .8459 .0431 .9452 
1.0000 . 8556 -0417 £9153 
1.0000 .8654 .0400 .8789 
1.0000 -8751 .0381 . 8364 
1.0000 .8849 .0359 .7881 
1.0000 -8946 .0335 -7349 
1.0000 .9044 .0309 .6775 
1.0000 -9141 .0281 .6162 
1.0000 .9239 .0251 .5514 
1.0000 .9336 .0220 . 4840 
1.0000 .9453 .0182 .3993 
1.0000 .9512 sO1G2" a Ses 
1.0000 .9570 .0142 Sit 
1.0000 .9609 .0128 . 2808 
1.0000 - 9648 .0114 .2501 
1.0000 .9662 .0109 . 2383 
1.0000 . 9682 .0101 .2221 
1.0000 .9692 .0098 .2145 
1.0000 .9702 -0094 .2055 
1.0000 .9722 .0087 .1901 
1.0000 .9732 .0083 .1818 
1.0000 .9751 .0075 .1654 
1.0000 £9771 .0068 .1490 
1.0000 .9791 .0060 -1309 
1.0000 9811 .0056 i222 
1.0000 .9831 -0053 -1165 
1.0000 .9351 -0050 -1108 
1.0000 -9871 -0048 .1052 
1.0000 .9881 .0046 .1019 
1.0000 -9901 .0043 .0951 
1.0000 +9920 .0040 .0879 
1.0000 .9940 .0036 .0781 
1.0000 -9960 .0029 . 0626 
1.0000 .9980 .0019 .0412 
1.0000 1.0000 0.0000 0.00060 
1.0000 


130 


from the supporting strut. One half of the model 
length protruded from the closed-jet working sec- 
tion of the wind tunnel into the open-jet test 
section. The ambient static pressure coefficients 
across and along the entire open-jet chamber (7.16 
x 7.16 x 6.4 m) were found to vary less than 0.3 
percent of dynamic pressure. The tunnel blockage 
and the longitudinal pressure gradient along the 
tunnel length were almost completely removed by 
testing the afterbody in the open-jet test section. 
The location of boundary-layer transition from 
laminar to turbulent flow was artifically induced 
by a 0.61 mm diameter trip wire located at X/L = 
0.05. When the flow was probed with a hot-wire, 
the trip wire was found to effectively stimulate 
the flow at a location 1 cm downstream from the 
wire. As a result of the parasitic drag of the 
wire, the boundary layer can be theoretically con- 
sidered to become turbulent at a virtual origin 
upstream of the trip wire. This virtual origin for 
the turbulent flow is defined such that the sum of 
the laminar frictional drag from the body nose to 
the trip wire, the parasitic drag of the trip wire, 
and the turbulent frictional drag after the trip 
wire equals the sum of the laminar frictional drag 
from the nose to the virtual origin and the turbu- 
lent frictional drag from the virtual origin to 
the after end of the model [McCarthy et al. (1976)]. 
The location of the virtual origin on the forebody 
with a 0.61 mm trip wire at X/L = 0.05 was found 
to be at X/L = 0.015 for a length Reynolds number 
of 5.9 x 10©. The location of transition in the 
mathematical model for the present boundary-layer 
calculation is specified at this virtual origin. 
The length Reynolds number based on the distance 
from the trip wire to the end of the parallel middle 
body is larger than for 4 x 10© for the two after- 
bodies. It can be assumed that a fully established 
axisymmetric turbulent boundary layer exists at the 
beginning of the afterbody and that the trip wire 
has no peculiar effect on the boundary-layer char- 
acteristics of the stern. 


3. INSTRUMENTATION 


A 1.83-cm Preston tube was taped to the stern at 
successively further aft locations in order to 
measure the shear stress distribution along the 
upper meridian of each stern. The Preston tube 
used was calibrated in a 2.54-cm water pipe flow 
facility described by Huang and von Kerczek (1972). 
Pressure taps (0.8 mm diameter) were used to mea- 
sure steady pressures at the same locations as the 
Preston tubes. The taps were connected by "Tygon" 
plastic tubes to a scanning valve located inside 
the model. The output tube from the scanning valve 
was run from the model through the supporting strut 
to a precision pressure transducer located on the 
quiescent floor of the open-jet chamber. The pres- 
sure transducer was a Validyne Model DP 15-560 de- 
signed for measuring low pressure up to + 1.4 x 10" 
dyn/cm? (0.2 psi). The zero-drift, linearity, and 
hysteresis of this transducer system were carefully 
checked and the overall accuracy was found to be 
within 0.5 percent of the dynamic pressure. 

A Prandtl type pitot-static pressure probe of 
3.125-mm diameter with four equally spaced holes 
located at three diameters aft of the nose was used 
to measure static pressure across the boundary 
layer. The yaw sensitivity of the static pressure 


probe was examined by yawing the probe in the free- 
stream. It was found that the measured static pres- 
sure was insensitive to the probe angle up to 5° 
yaw. The response of measured static pressure to 
probe angle was nearly a cosine function of yaw 
angle for yaw angles less than 15°. The static 
pressure probe was aligned parallel to the model 
axis for all of the static pressure measurements. 
The local angles between the resultant velocity of 
the boundary-layer flow and probe axis were found 
to be less than 15° (5° for most cases). The maxi- 
mum static pressure coefficient in the boundary 
layer was less than 0.2. Thus, the error in the 
measured static pressure caused by not aligning the 
probe with the local flow was less than 0.8 percent 
of the dynamic pressure. 

The mean axial and radial velocity components 
and the Reynolds stress were measured by a TSI, Inc. 
Model 1241 "X" wire. The probe elements were 0.05 
mm in diameter with a sensing length of 1.0 mn. 

The spacing between the two cross elements is 1.0 
mm. A two-channel TSI Model 1050-1 hot-wire ane- 
mometer and linearizer were used. The "X" wire, 
together with temperature compensated probes, were 
calibrated at the factory and supplied with their 
individual linearization polynomial coefficients. 
This eliminated the time-consuming linearization 
process. The frequency response of the anemometer 
system claimed by the manufacturer is dc to 200 kHz. 
Calibration of the "X" wire was made before and after 
each set of measurements. It was found that this 
hot-wire anemometer system had a 40.5 percent ac- 
curacy (40.15 m/s accuracy at the free stream ve- 
locity of 30.5 m/sec) during the entire experiment. 
The accuracy of cross-flow velocity measurements 

by the cross wire was estimated by yawing the cross-— 
wire in the free stream. It was found that the ac- 
curacy of the measured cross-flow velocities was 
about one percent of the free stream velocity. 

The linearized signals were fed into a Time/Data 
Model 1923-C Real-Time analyzer. Both channels of 
analog signal were digitized at a rate of 80 points 
per second for ten seconds. These data were imme- 
diametely analyzed by a computer code to obtain the 
individual components of mean velocity, turbulence 
fluctuation, and Reynolds stress on a real time 
basis. 

A traversing system enclosed in a 15 cm chord, 
streamlined strut was used to support both the 
static pressure probe and the cross-wire probe. The 
traversing system was mounted either on an I-beam 
along the axis of the lower floor of the open-jet 
chamber or on the ceiling of the closed-jet section. 
The combination of these two mounting arrangements 
allowed the measurements to be made at any axial 
location along the stern and up to 50 percent of 
the body length downstream from the aft end of the 
body. Positioning of the traversing system was 
achieved by manual adjustment in the axial direction 
and by remote control in the radial direction. The 
total radial traverse of the probe was 25 cm. The 
radial position of the probe was monitored by a 
potentiometer to with a +0.01 mm accuracy. 


4. COMPARISON OF EXPERIMENTAL AND THEORETICAL 
RESULTS 


In the following, the experimental results for the 
thick stern boundary layers are presented and com- 
pared with theoretical results. The theories used 


in the comparison are the Douglas CS differential 
boundary-layer method in conjunction with the dis- 
placement body concept. The iteration procedures 
for numerical computation are given by Huang et al. 
(1976). In this investigation, the displacement 
body concept for solving the interaction between 
the thick stern boundary layer and potential flow 
will be examined and an eddy-viscosity model will 
be evaluated. 


Measured and Computed Pressure and Shear Stress 
Distributions 


Significant improvement in the accuracy of measur- 
ing surface pressure and shear stress have been made 
by using a precision pressure transducer. The 
present results are more reliable than the earlier 
results of Huang et al. (1976), although the dif- 
ferences are small. 

The measured and computed values of the pressure 
coefficient, Cp = 2(p - Pa) /OUSs. are compared in 
Figure 1 for Afterbody 1 and in Figure 2, for After- 
body 2; p is the local static pressure, p is the 
mass density of the fluid, U, is the free-stream 
velocity and po is the ambient pressure (the qui- 
escent chamber static pressure of the open-jet sec- 
tion). The pressure coefficients computed on the 
displacement body were carried radially back to the 
hull surface and the radial distribution of pres- 
sure at a given axial station was assumed to be a 
constant between the hull surface and the fictitious 
displacement surface. The maximum error in the 
static pressure associated with this assumption is 
less than two percent of the dynamic pressure (next 
section). The agreement between theory and measure- 
ment is excellent for both afterbodies. The results 


0.15 


Ry = 6.6 x 10° 


THEORY 


MEASUREMENT 
0.10 ° 


0.65 0.70 0.75 - 0.80 


131 


suggest that the displacement body concept as used 
by Huang et al. (1976) permits accurate computation 
of the pressure distribution on the stern. 

The measured and computed distributions of local 
shear stress, C;, are compared in Figure 3. The 
agreement between theory and measurement is 
also very good for both afterbodies except for 
x/L > 0.95 where the measured values of C, are lower 
than the computed values. 


Measured and Computed Static Pressure Distribution 


The measured and computed static pressure coeffi- 
cients for Afterbody 1 are compared in Figure 4 at 
various locations across the stern boundary layer 
and in Figure 5 for the near wake. Figures 6 and 7 
show the comparisons for Afterbody 2. The off-body 
option of the Douglas potential-flow computer code 
was used to compute the static pressure distribu- 
tions off the displacement body. As can be seen in 
Figures 4 through 7, the computed static pressure 
distributions across the entire stern boundary layer 
and near wake mostly agree well with the measured 
static pressure distributions. The discrepancy 
between the measured and computed values of C, is 
in general less than 0.01 which is about the accuracy 
of the measurement. 

As will be seen later, both displacement bodies 
are convex from the parallel middle body up to X/L 
= 0.91 and become concave downstream from X/L > 
0.91. However, the actual afterbodies are convex 
all the way up to X/L = 0.96. The measured values 
of C, shown in Figures 4 through 7 increase with 
radial distance for X/L < 0.91, indicating that 
the mean streamlines are convex; and measured values 
of Cp decrease with radial distance for X/L > 0.91, 


0.85 0.90 0.95 1.0 


X/L 


FIGURE 1. Computed and measured stern pressure distribution on afterbody 1. 


0.20 


= 68 x 10° 


THEORY 
0.10 MEASUREMENT 


p 
0.0 
—0.10 
—0.20 
0.65 0.70 0.75 0.80 0.85 0.90 0.95 1.0 
X/L 
FIGURE 2. Computed and measured stern pressure distribution on afterbody 2. 
0.003 
MEASUREMENT 
AFTERBODY 1 
Ry = 6.6 x 10° 
0.002 THEORY 


MEASUREMENT 


AFTERBODY 2 
R, = 6.8 x 10° 
THEORY 


0.6 0.7 0.8 0.9 1.0 


mputed and measured shear stress distribution on afterbodies 1 and 2. 


FIGURE 4. 


X/L = 0.755 


THEORY 


X/L = 0.914 DISPLACEMENT 
SURFACE 


MEASUREMENT 
ON THE BODY 


Toa ed: 


MEASUREMENTS THEORY 


v 


Computed and measured static pressure distributions across stern boundary layer of afterbody 1. 


Tages aes ae | ea 


MEASUREMENT THEORY 


O 


X/L = 1.000 


MEASUREMENT 


THEORY 
DISPLACEMENT SURFACE 


FIGURE 5. Computed and measured 
static pressure distributions 
across near wake of afterbody 1. 


134 


3.0 


° | X/L_ MEASUREMENT THEORY 
by 0.846 ° —___ 
. 0.934 © —_-— 
L, | 0.970 4a 
in 0.977 Do soa 
O 1.000 O 
° TENS 
y\ 
2.0 oe) 
fo) ANS) 
\\ 
‘ ro 
X/L = 0.846 9° sa 
r ° \ 
"max O OY AN 
\O 
oO WO 
ON 
QD 
ato KS MEASUREMENT RIN AGO 
THEORY Ww. ‘Od 
DISPLACEMENT XY rat H 
X/L = 0.934 
SURFACE QUT AN WY 
OY 
MEASUREMENT 3 
CNiecopy oo as eNE 
emo 
Q, MT? 
SON 
i] 
—0.20 —0.10 0.10 0.20 
c 
p 
FIGURE 6. Computed and measured static pressure distributions across stern boundary layer of afterbody 2. 


indicating that the mean streamlines are concave. 
Thus, the curvatures of the mean streamlines are 
more closely related to the curvatures of the dis- 
placement body than the actual body. The close 
agreement between the computed and measured static 
pressure distributions again supports the displace- 
ment body concept for computing the potential flow 
outside of the displacement surface. 


Yo is the body radius; x is the axial distance; u 
and vy, are the mean velocity components respectively 
parallel to and normal to the meridan of the body 

(s and n directions); v is the kinematic viscosity 

of the fluid; Tul is the Reynolds stress; and u! 
and vy, are the velocity fluctuations in the s and n 
directions respectively. The Douglas CS method as- 
sumes that the Reynolds stress depends upon the local 
flow parameters only, e.g., 


5. MEASURED AND COMPUTED MEAN VELOCITY PROFILE Sal gpa for! oS nis ne 
=ulv' = 5 = (3) 
The incompressible steady continuity and momentum Se 6 2 HON iy SS © 
equations for thin axisymmetric turbulent boundary i. 
layers are . du. 
where ¢, = 22 = gn (eddy viscosity in the inner 
d(ru,)/ds + d(xrv,)/dn = 0 (1) fo) region) 
and co 
€. = 0.0168 y (Ul=u))idn = 00168 UR TOA, 
u_du_/3s + v du /dn ° G35 if @ SQ er tere p 
Sis Ties, 
(2) : é . . 
= -dp/pds + d[r(vdu /on)=u'v']/ron (eddy viscosity in the outer region), 
s sn r 
2 = 0-4 xine) {2 = exp = = in (yy 
u,(s,0) = v(s,0) = 0 at n = 0 OS a e 
where (mixing-length parameter in the inner 
region) , 
r(s,n) = xr _(s,n) + n cosa 
fe} Tw —5 : 
A = 26 Lira , (Van Driest's damping 


a= tan”* (dr, /ax) 


factor), 


135 


6 Us where u,/U, and V1,/U, are computed by the CS method. 

oF =f (1 - —)dn, [displacement thickness The potential-flow pressure is assumed to be con- 
fo} e (planar definition) ] stant between the body surface and the displacement 
; surface and is equal to the pressure Pq computed on 
y 2 Ant & BLE (2) Tees (Ghisetisteney the displacement body. The value of U, used in Eqs. 

E18 6 Recto) (4) and (5) is equal to vl - Pq and U, is assumed 

, 
to be parallel to the body surface. 
& = 8995 , (boundary-layer thickness), The displacement-body concept can be used to 


improve the computed values of u, and We outside 


ae (wall sheamistress)), of the displacement Surface of thick boundary layers, 


e.g., 
Ue is the potential-flow velocity used in the 
boundary-layer calculations, and at y,, ej is equal u(r) u, (n) U_cos(8-a) 
to €,. A computer code to solve for the values U ~ U_cos(@-a) U Cese! 
us/Ue and vp/Ue has been developed by Cebeci and © P ° 
Smith (1975) using Keller's numerical box scheme. : (6) 
The velocity components measured in the present v,,(n) Smee) ; 
investigation are u, and v,, the components in the ~ U sin(6—a) U Sei 
axial and the radial directions of the axisymmetric P ©) 
body. The computed values of uy and v; are given 
5 v_(n) u_(n) U_cos (6-a) 
Vi 1g s p 
SSS 58 ee a SS “Sting 
a, (x) us (n) U, v,,(n) U UR t eeS(O=@) Us 
U = U v cosa - yp sina, (4) 
fe) e fe} e fe} vy, (™) UL oaeWre) 
+ 7 cosa (7) 
v(x) a, (n) UL v (n) U umn (=e) U, , 
7 = a — sina + i cosa, (5) 
fo) e fo) e [o) where the variation of the inviscid static pressure, 


MEASUREMENT THEORY 


X/L = 1.000 


MEASUREMENT 


THEORY 


DISPLACEMENT 
SURFACE 


0 0.05 c 0.10 0.15 0.20 
p 


PIGURE 7. Computed and measured static pressure distributions across near wake of afterbody 2. 


MEASUREMENT 


MEASUREMENT 


| 


X/L 


(A) 0.755 


(B) 0.846 


(C) 0.914 


(D) 0.934 


X/L 


(E) 0.964 


(F) 0.977 


(G) 1.000 


THEORY 


(A) X/L = 0.755 


me ee )=—DIRECT 


IMPROVED 


ea 


THEORY 
O 


ed mean 


axial 


a olig 


(E) X/L = 0.964 


(F) X/L = 0.977 


pas = 0 0.2 
U U 
o oO 
easier OLRECT 
— — — — —— So mproveod 
and radial velocity 


(D) X/L = 0.934 


(G) X/L = 1.00 


0.4 0.6 0.8 1.0 


distributions across stern boundary 


C,,(r), across the thick boundary layer is expressed 
in terms of the inviscid resultant velocity U 

Name) S valve Cp(r)Jand @ is the angle between the 
inviscid resultant velocity U, and body axis (0 is 
positive when Up is directed away from the axis). 

In the first improvement the values u,g/(U,,cos(0-a) ) 
are taken as the computed values of f' = us/U, in 
the CS method with Ue equal to the inviscid resul- 
tant velocity on the displacement body. At the edge 
boundary layer, the value of ug(U,cos(8-a)) is equal 
to 1.0. The value of v,/(U.sin(68-a))is also equal 
to 1.0 since the boundary-layer-induced normal ve- 
locity is assumed to be equal to the inviscid normal 
velocity of the displacement body at that point 
(Lighthill (1958)]. The theoretical proof for an 
axisymmetric body has not been worked out in the 
literature and will not be given here. However, 

the validity of the assumption will be borne out 

by the experimental measurements of Vy: Therefore, 
Eqs. (6) and (7) reduce to the proper limit at the 
edge of the boundary layer, e.g., 


u(r=6)) U 
TE Kane: yg cost: (8) 
fo) ° 
v.(r=5)) U 
uU = T sin6é, (9) 
fe} 


which are the inviscid axial and radial velocity 
components of the displacement body, where Us = 


1.4 


1.2 


1.0 


0.8 


0.4 


0.2 


AL ey7) 
Vlas Coe (CTO Vo Outside of the boundary layer, 


Eqs. (B) and X9) are also valid so long as the 

local inviscid values of U, and 6 for the displace- 
ment body are used. The improved values in Eqs. 

(6) and (7) account for the variation of the in- 
viscid static pressure and potential-flow vector 
across the thick boundary layer and make appropriate 
use of the results of the CS method. As already 
noted, the variation of static pressure computed 
across the boundary layer outside of the displace- 
ment surface agrees quite well with the experimental 
results. 

Figure 8 shows the comparison of the mean axial 
and radial velocity profiles at several axial sta- 
tions on Afterbody 1, and Figure 9 shows the mea- 
sured axial velocity profiles across the near wake 
of Afterbody 1. The theoretical results at X/L = 
1.00 were calculated at X/L = 0.998. Figures 10 
and 11 show comparisons of the measured and computed 
velocity profiles for Afterbody 2. The mean axial 
and radial velocity components uy, and Vy were mea- 
sured by a cross-wire probe and the experimental 
accuracy of measurements of u,,/Ug and Vr/U5 were 
respectively about 0.5 percent and 1.0 percent. 

As shown in Figures 8 and 10, the theoretically 
computed velocities, which account for the variation 
of static pressure distribution across the thick 
boundary layer, agree better with the measured axial 
and radial profiles outside of the displacement sur- 
face. These results suggest that a simple improve- 
ment of the existing boundary-layer computation 
method can be made for the thick stern boundary 


FIGURE 9. Measured mean axial velocity distributions 
across near wake of afterbody 1. 


138 


0.8 


0.6 


I~)— —9- — — 0 


A 


(C) X/L = 0.970 


(B) X/L = 0.934 


(D) X/L = 0.977 


> 


0.2 
t) 
uy Mp o x 
a ee — (LDV) 
X/L Uy U, ra uy 
(A) 0.840 © QO 0.962 THEORY 
(B) 0.934 Q 0.484 oO uy v 
(C0970 A JA 0206 uU ~ 
° ° 
(D) 0.977 O  o.149 (0) — = — - — DIRECT 
oem ocean Sao IMPROVED 
FIGURE 10. Measured and computed mean axial and radial velocity distributions across stern boundary 
layer of afterbody 2. 3 


layer by means of the displacement body concept. 
However, it is important to point out that the 
measured axial velocity profiles in the inner region 
are in general smaller than the theoretical values. 
The eddy viscosity model plays an important role in 
this region; therefore, it is essential to examine 
the eddy viscosity model used for computing the 
thick stern boundary layer. Figures 8 and 10 also 
show the comparison of the axial velocities mea- 
sured by the cross-wire and by LDV (Huang et al., 
1976). The agreement is very good inside the bound- 
ary layer. However, due to the artifical seeding 

of oil mist required for the LDV, the axial veloc-— 
ities near the edge of the boundary layer measured 
by LDV are smaller than that by the cross-wire. 


6. COMPARISON OF MEASURED AND COMPUTED INTEGRAL 


PARAMETERS 


The integral parameters are derived from the mea- 
sured velocity distribution. The two-dimensional 
displacement thickness is defined as 


= u (x) 
x (10) 


where 6, is the boundary thickness measured radially 
normal to the body axis and Ux(r) is the value of 
the axial component of inviscid flow velocity com- 


puted about the displacement body. The value of 


, 


U,,(r) is computed by the potential-flow method ex- 
cept inside of the displacement surface where it is 
assumed that UL (x) = UL, (XQ) with re) being the radius 
of the displacement surface. The boundary-layer 
thickness 6, is defined at the radial position where 
the measured value of uy (r) equals) (01995 3U57\(@>) eae 
is difficult to obtain 6, precisely since the ac- 
curacy of the Uy/U, measurement is only about 0.005. 
Nevertheless, the overall accuracy of the values of 
6, estimated in the present investigation is about 
10 percent. 

A measure of the mass-flux deficit in the thick 
axisymmetric boundary layer is defined by 


r_ +6 ry +6% 
(0) "32 u(r) OMG 
= i oS rdr = rdr abil 
i U_(z) ff are 
xr x ie 
fo) 
where r_ is the local body radius and 6* is the 
axisymmetric displacement thickness. Thus, the 
axisymmetric displacement thickness becomes 
Ge a 1 O% Yr = 
2 z= (2) +2 (12) 
12 r 2 
max max max 


where Ypax is the maximum radius of the body. 

The displacement body in the present investiga- 
tion is defined by rq = 5# + xr, rather than the 
planar definition, rg 6* + ro. Similarily, a 
measure of the momenta lax deficit is defined by 


FIGURE 11. 
across near wake of afterbody 2. 


Measured mean axial velocity distributions 
5 


r +6 
fo) 


x u(r) 7 u(r) 
Do A saeray| cee ee 
x x 


The measured and computed values of 6* and 6, are 
shown in Figure 12 for Afterbody 1 and in Figure 
13 for Afterbody 2. The measured values of d¢ and 
6, for X/L > 0.90 are slightly larger than the 
computed values for both bodies. 

The transverse curvatures of the boundary-layer 
flow with respect to the body radius, (ro + Oe) Pes 
and (ro + 5,)/ro, are also shown in Figures 12 and 
13. A drastic increase of the values of (ro + Ss) AZ) 
and (rg + 6;)/Yg occurs at X/L = 0.9, indicating the 
important effect of transverse curvature on the 
stern. The longitudinal curvature of the body is 
denoted by K, = (d°r,/dx*) [1 + (dro/dx*)]~3/% and 
the longitudinal curvature of the displacement body 
is denoted by Kg = (d2rg/dx2) [1 + (drg/dx) 2173/2. 

A positive sign for K, or Kg indicates concave sur- 
pacer The values of nae and Ka' max are shown 

in Figures 12 and 13. There is a significant dif- 
ference between Ky and Ka in the thick boundary 
layer region. In each case, the curvature of the 
displacement body is convex up to X/L = 0.92, then 
changes to concave and remains concave throughout 
the entire thick boundary-layer and near-wake region. 
The curvature of the body surface is convex up to 
X/L = 0.96. As already shown in Figures 4 and 6, 
the measured distributions of static pressure and 


139 


hence the curvatures of the mean streamlines are 
much more closely related to the displacement body 
than to the actual body. The magnitudes of the 
maximum concave and convex radii of curvature of 

the displacement bodies are estimated to be 8 max 
and 30 Yrmax for Afterbody 1 (Figure 12) and 7 rmax 
and 8 Yrmax for Afterbody 2 (Figure 13), respectively. 
The magnitudes of the radii of curvature of the 

mean streamlines outside of the displacement body 
are expected to be larger than 10 rmax.- 


7. MEASURED TURBULENCE CHARACTERISTICS 


The cross-wire probe was used to measure the tur- 
bulence characteristics in the thick boundary layer. 
The measured Reynolds stresses and the measured 
mean velocity profiles were used to obtain eddy 
viscosity and mixing length. 


Measured Reynolds Stresses 


The turbulence characteristics in the thick boundary 
layer can be represented by the distributions of 
Reynolds stresses, namely, -u'v', u'2, v'2, and 
w'2, where u', v', and w' are the turbulence fluc- 
tuations in the axial, radial, and azimuthal direc-— 
tions, respectively. Figures 14 and 15 show the 
measured distribution of Reynolds stress =v’ /U,2 
and three components of turbulence intensity at 
several axial locations along the two afterbodies. 
In general, for a given location, the intensity of 
the axial turbulence-velocity component has the 
highest value and the intensity of the radial com- 
ponent has the smallest value. The degree of 
anisotropy decreases as the stern boundary layer be- 
comes thicker. Furthermore, the increased boundary- 
layer thickness is accompanied by a reduction of 
turbulence intensities and a more uniform distribu- 
tion of turbulence intensities in the inner region. 
The variation along the body of the radial location 
of the maximum values of the measured Reynolds stress 
=u 9/057 layer is small. The spatial resolution of 
the cross-wire probe is not fine enough to measure 
the Reynolds stress distributions in the inner re- 
gion when the boundary layer is thin. As the stern 
boundary layer increases in thickness, the location 
of maximum Reynolds stress moves away from the wall 
(Figures 14 and 15). The values of Reynolds stress 
=u'v" decrease quickly from the maximum value to 
zero at the edge of the boundary layer. As shown 
in Figures 14 and 15, the shape of the Reynolds 
stress distribution curves in the outer region is 
quite similar for all the thick boundary layers. 
It is interesting to note that the shapes of the 
Reynolds stress distributions in the inner regions 
are different from those measured in the wake at 
X/L = 1.057 and 1.182 (Figures 14 and 15); this is 
a typical characteristic of a developing wake 
([Chevray (1968)]. The Reynolds stresses experience 
a drastic reduction in magnitude near the edge of 
the boundary layer. 

A turbulence structure parameter defined by aj 
= =u'v'/aq2, where q?2 = uj," + v,'2 + w'2, is of 
interest. The measured distributions of a, are 
shown in Figure 16. Most thin boundary layer data 
show that a, is almost constant (a, = 0.15) between 
0.05 and 0.86. The present thick stern axisymmetric 
data shown in Figure 16 indicate that a, is almost 
constant up to 0.6 6,, and the magnitudes of aj 


140 


x 
3 
aE 
eS 
= 
RE 
ae 
x 
3 
RE 
~o 
0.75 
FIGURE 12. 


0.75 


0.80 


AFTERBODY 1 


0.80 


Transverse 


AFTERBODY 2 


0.80 


0.85 


and longitudinal curvature parameters for a 


Qj TRANSVERSE CURVATURE 


0.85 


1a] irvature parameter: 


~~ 


anes: 
—— — 
———_ LONGITUDINAL CURVATURE 


r 
o max 


0.90 0.95 1.00 1.05 1.10 


EDGE OF BOUNDARY LAYER AND NEAR 


THEORY 
MEASUREMENT 


An De mS 


DISPLACEMENT SURFACE 


Se Se SS ON 


THEORY MEASUREMENT 


0.90 0.95 1.00 1.05 1.10 


xX/L 


(r+ 5 )/r 
oO r oO Pa 
> 


EDGE OF BOUNDARY LAYER AND NEAR WAKE 
THEORY MEASUREMENT 


Foi Ses 


THEORY 


DISPLACEMENT SURFACE 


ed, er 
MEASUREMENT 


0.90 0.95 1.00 1.05 


1.20 


1.20 


FIGURE 14. 


(A) X/L = 0.755 


(C) X/L = 0.934 


(E) X/L = 1.0076 (F) X/L = 1.057 


0.02 0.04 
V w7/u, > v v7/u, . IV w'7/U, 3 


Measured distributions of Reynolds stresses for afterbody 1. 


(B) X/L = 0.846 


0.06 


<u'v/U 
o 


(G) X/L = 1.182 


141 


(B) = 0.934 


T ] Useerorrl Lapp leita amin cel firstly 


(A) X/L = 0.840 


(D) X/L = 0.977 


(C) X/L = 0.970 


0.10 


e) 
0.04 0.06 


Vo?u,. Voru,. Vw?su,, -uvsu,? 


molds stresses for afterbody 2 


0.08 


it) 0.02 


Measurea distributions of Rey 


a 


decrease toward the edge of the boundary layer. The 
values of a; also decrease in the inner region of 
wake at X/L = 1.057 and 1.182 of Afterbody 1. It 
should be pointed out that the measured values of 

q? contain the free-stream turbulence fluctuation, 
no attempt having been made to remove the free- 
stream turbulence fluctuation from the measured 
values of q*. The measured reduction of a) near 

the outer edge of the boundary layer is in part 
caused by the larger contribution of the free-stream 
turbulence to q2 than to -u'v". Nevertheless, the 
measured values of the turbulence structure param- 
eter a, are quite constant across the inner portion 
of the boundary layer where the effect of free- 
stream turbulence is small. 


Eddy Viscosity and Mixing Length 


The measured distributions of shear stress =u'Tv™ 
and mean velocity gradient, du,/dr, were used to 
calculate the variations of eddy viscosity and 
mixing length across the thick stern boundary layers 
according to the following definitions 


Tg ou (14) 
SU = ea 
or 
and 
ou du, 
ee 2. =e mS 
u'v Q les | ay (15) 


The experimentally-determined distributions of 
eddy viscosity, €/Ugd5p*1 are shown in Figure 17 for 
Afterbody 1 and in Figure 18 for Afterbody 2, where 
Us is the potential-flow velocity at the edge of 
the boundary layer and 6,* is the displacement 
thickness (based on the planar definition, Eq. 10). 
Figures 19 and 20 show the experimentally-determined 
distributions of mixing length 2/6, for the after- 
bedies, where 6, is the boundary-layer thickness 
measured normal to the body axis. As shown in 
Figures 19 and 20, the measured distributions of 


0.018 
0.016 


0.014 


THIN 
0.012 BOUNDARY 


143 


eddy viscosity agree resonably well with the eddy- 
viscosity model of Cebeci and Smith (1974, Eq. 3) 
when the boundary layers are thin. However, as the 
stern boundary layer thickens, the measured values 
of e/Us Sp* in the thick stern boundary layers are 
only about 1/6 of the values for thin boundary 
layers given by the Cebeci and Smith model (1974). 
The measured distributions of mixing length shown 
in Figures 19 and 20 also agree quite well with the 
thin boundary layer results of Bradshaw, Ferriss, 
and Atwell (1967). Again as the boundary thickens, 
the measured values of 2/6, reduce drastically. 
The values of 2/5, in the thick stern boundary 
layers are only about 1/3 of those of the thin 
boundary layers. Similar reductions of eddy vis- 
cosity and mixing length in thick stern boundary 
layers were also measured by Patel et al. (1974, 
USN) 3 

As the axisymmetric boundary layer thickens in 
the stern region, the boundary layer thickness Sy 
and the displacement thickness Sp* increase dras- 
tically. However, the values of eddy viscosity and 
mixing length do not have enough time to respond to 
this change. Therefore, neither the eddy viscosity 
model of Cebeci and Smith (1974), nor the mixing 
length results of Bradshaw, Ferriss, and Atwell 
(1967) can be applied to the thick stern boundary 
layer. 


8. TURBULENCE MODELS 


In most works, the basic assumption made in the 
differential methods for calculating turbulent 
boundary layers is that the mixing length or eddy 
viscosity is uniquely related to the mean velocity 
gradient and the boundary-layer thickness parameter 
at a given location. So long as the boundary layer 
is thin and the change in boundary-layer properties 
due to the pressure gradient is gradual, this simple 
assumption is know to be satisfactory [see e.g., 
Cebeci and Smith (1974)]. When the past history of 
boundary layer characteristics is important, Brad- 


FIGURE 17. Measured distributions 
of eddy viscosity for afterbody 1. 


144 


shaw et al. (1967) argue that the turbulence energy 
equation can be used to model the memory effect. 
In order to determine the rate of change of tur- 
bulent intensity along a mean streamline, three 
assumptions have to be made: namely, that turbu- 
lence intensity is directly proportional to the 
local Reynolds stress, aj =u /q2 0.15; that 


= =u'v'/ 
the dissipation rate is determined by the local 
Reynolds stress and a length scale depending on 

n/é; and the energy diffusion is directly pro- 
portional to the local Reynolds stress with a fac- 
tor depending on the mixing value of Reynolds stress. 
On the basis of thin boundary-layer data two em- 
pirical functions for the last two assumptions were 
proposed by Bradshaw et al. (1967). The first as- 
sumption, 2/6 £,; (n/é), was found not to be 


0.018 


0.016 


0.014 


0.012 


0.010 


U;5 
6 0.008 
0.006 


0.004 


0.002 


Measured distribu- 


FIGURE 18. 


0.2 


0.10 


0.08 


0.06 


0.04 


0.02 


Lstribu- 


applicable to the present thick axisymmetric stern 
boundary layers. The deviation of the apparent 
mixing length along the curved boundary from that 

of a thin flat boundary was also noted and dis- 
cussed by Bradshaw (1969). A simple linear cor- 
rection to the length scale of the turbulence by 

the extra rate of strain was made by Bradshaw (1973). 
The extension of this concept has just been made for 
the thick axisymmetric boundary layer by Patel et 

Gilly (ALS) 7S) 

It is important to note that the boundary-layer 
thickness of a typical axisymmetric body increases 
drastically at the stern. Most of the rapid change 
takes place within a streamwise distance of a few 
boundary-layer thicknesses. Most of the empirical 
functions for solving the turbulence energy equa- 


Ole! OSS ae ee 
= 0.0168 
THIN Usd, 
BOUNDARY 
LAYER 
O € 0.0168 
fe} = 6 
Usdp 1 +55 ee) 
5, 
X/L 


0.4 


THIN BOUNDARY 


LAYER 2 


© 0.755 
O 0846 
©} 0.934 
A 0.964 
QO 1.0076 


© 1.057 
QO 1.182 


0.10 


0.08 


0.06 


0.04 


tion will undergo rapid changes in basic forms. The 
one known for certain is the empirical function for 
mixing length. Therefore, it may be difficult to 
compute the rate of change of the turbulence energy 
or the extra rate of strain in the region. 

Fortunately the present measured distributions 
of Reynolds stresses shown in Figures 14 and 15 are 
quite similar in the outer region and differences 
appear in the inner region where the turbulence is 
reduced in intensity and more homogeneous. In such 
an axisymmetric flow configuration, the character- 
istic length scale is more closely related to the 
entire turbulence annulus between the body surface 
and the edge of the boundary layer rather than the 
radial distance between the two. Therefore, we 
propose that the mixing length of an axisymmetric 
turbu-length boundary layer is proportional to the 
square root of this area when the thickness in- 
creases drastically at the stern: 


Qo We, + S.) 2 - ie 


In order to examine this simple hypothesis, the 
present measured values of 2/V(rg + 6,) 2 = ro? 
together with the data of Patel et al. (1974, 1977) 
are shown in Figure 21. The solid line is the best 
fit of the present data. The present values of 2 
are slightly greater than those for Patel's modified 
spheroid (1974) and are slightly lower than those 
for Patel's low-drag body (1977). The data in 
Figure 21 support this simple hypothesis although 
the data are quite scattered due to large varia- 
tions of stern configurations and Reynolds number, 
and probable measuring errors. 

The existing thin turbulent boundary-layer dif- 
ferential methods can be applied to the forward 
portion of the axisymmetric body up to the station 
where the boundary layer thickness increases to 
about 20 percent of the body radius. Further down- 


THIN BOUNDARY 
LAYER 


145 


FIGURE 20. Measured distribu- 
tions of mixing length for 
afterbody 2. 


stream, the apparent mixing length of the thick 
axisymmetric stern boundary layer (2) can be roughly 
approximated by the mixing length for a thin flat 
boundary layer (2,) by 


Ghee 3.336 (to) 
ag 


which is the solid line of Figure 21. At the aft 
end of the stern Yo is zero and the value of 2/2, 
is 1/3.33. This simple approximation of the mixing 
length for thick axisymmetric stern turbulent bound- 
ary layers can be incorporated into most existing 
differential methods. As noted earlier, the mea- 
sured axial velocities inside the thick boundary 
layer (especially in the inner region) are smaller 
than the computed values (Figures 8 and 10). The 
present CS method overestimates the magnitude of 
eddy viscosity (Eq. 3) for the thick stern boundary 
layer. While the mixing length approximations ob- 
tained in the present investigation can be incorpo- 
rated into the CS method to predict more accurately 
the thick stern boundary-layer velocities, further 
refinement of the theoretical methods is desirable. 


9. CONCLUSIONS 


In this paper, we have described recent experimental 
investigations of the thick turbulent boundary lay- 
ers on two axisymmetric sterns without shoulder flow 
separation. A comprehensive set of boundary layer 
measurements, including mean and turbulence veloc-— 
ity profiles and static pressure distributions, are 
presented. Two major conclusions can be drawn: 

The Lighthill/Preston displacement body concept 
has been proven experimentally to be an efficient 
and accurate tool for treating the viscid and in- 


146 


(A) PRESENT AFTERBODIES 2 
—— 


© 0840 
0 084 Cj 0.934 
© 0934 [0970 
Q 0964 © 0977 


QO 1.0076 


0.05 


(B) MODIFIED SPHEROID 
PATEL ET AL (1974) 


(C) LOW-DRAG BODY 
PATEL ET AL (1977) 


y/5 


viscid stern flow interation on axisymmetric bodies. 
The measured static pressure distributions on the 
body and across the entire thick boundary layer and 
wake were predicted by the displacement body method 
to an accuracy within one percent of dynamic pres- 
sure. Theoretical predictions of the measured 
axial and radial velocity profiles outside the dis- 
placement surface were improved significantly when 
the variations of the static pressure and radial 
velocity of the displacement body were incorporated 
into the computation. 

Neither the measured values of eddy viscosity 
nor mixing length were found to be proportional to 
the local displacement thickness or the local 
boundary-layer thickness of the thick axisymmetric 
boundary layer. As the boundary layer thickens 
rapidly at the stern, the turbulence characteristics 
in the outer region remain quite similar but the 
turbulence reduces its intensity and becomes more 
uniformly distributed in the inner region. The 
measured mixing length of the thick axisymmetric 
stern boundary layer was found to be proportional 
to the square root of the area of the turbulent 
annulus between the body surface and the edge of 
boundary layer. This simple similarity hypothesis 
can be incorporated into existing differential 
boundary-layer computation methods. 


ACKNOWLEDGMENT 


The work reported herein was funded under the 
David W. Taylor Naval Ship R&D Center's Independent 
Research Program, Program Element Number 61152N, 
Project Number ZR 000 O1. 


REFERENCES 


Bradshaw, P., D. H. Ferriss, and N. P. Atwell 
(1967). Calculation of boundary layer develop- 
ment using the turbulent energy equation. J. 
Fluid Mech., 28, 593-616. 

Bradshaw, P. (1969). The analogy between streamline 
curvature and buoyancy in turbulent shear flow. 
J. Fluid Mech., 36, 177-191. 

Bradshaw, P. (1973). Effects of streamline curvature 
on turbulent flow, AGARDograph No. 169. 

Cebeci, T., and A. M. O. Smith (1974). Analysis of 
turbulent boundary layers. Academic Press, New 
York. 

Chevray, R. (1968). The turbulent wake of a body 
of revolution, ASME. J. of Basic Engineering, 
90, 275-284. 


Granville, P.S. (1953). The calculation of the 


viscous drag of bodies of revolution. David 
Taylor Model Basin Report 849. 
Hess, J. L., and A. M. O. Smith (1966). Calculation 


of potential flow about arbitrary bodies. 
in Aeronautical Sciences, Vol. 8, Pergamon Press, 
New York, Chapter 1. 


Progress 


147 


Huang, T. T., and C. H. von Kerczek (1972). Shear 
stress and pressure distribution on a surface 
ship model: theory and experiment. Ninth ONR 
Symposium on Naval Hydrodynamics, Paris; avail- 
able in U.S. Government Printing Office, ACR-203- 
Viole loos —20Or 

Huang, T. T. et al. (1976). Propeller/stern/ 
boundary-layer interaction on axisymmetric bodies: 
theory and experiment. David Taylor Naval Ship 
Research and Development Center Report 76-0113. 

Lighthill, M. J. (1958). On displacement thickness. 
J. Fluid Mech., 4, 383-392. 

McCarthy, J. H., J. L. Power, and T. T. Huang (1976). 
The roles of transition, laminar separation, and 
turbulence stimulation in the analysis of axi- 
symmetric body drag. Eleventh ONR Symposium on 
Naval Hydrodynamics, London; Published by 
Mechanical Engineering Publications Ltd., 

London and New York. 

Patel, V. C., A. Nakayama, and R. Damian (1974). 
Measurements in the thick axisymmetric turbu- 
lent boundary layer near the tail of a body of 
revolution. J. Fluid Mech., 63, 345-362. 

Patel, V. C., Y. T. Lee, and O. Guven (1977). Mea- 
surements in the thick axisymmetric turbulent 
boundary layer and the near wake of a low-drag 
body of revolution. Symposium on Turbulent Shear 
Stress, Pennsylvania State University, University 
PEWS IDs Do A998) 5 B3S\0 

Patel, V. C. and Y. T. Lee (1978). Calculation 
of thick boundary layer and near wake of bodies 
of revolution by a differential method. ONR 
Twelfth Symposium on Naval Hydrodynamics, (This 
Volume, Section III). 

Preston, J. H. (1945). The effect of the boundary 
layer and wake on the flow past a symmetrical 
aerofoil at zero incidence; Part I, the veloc- 
ity distribution at the edge of, and outside the 
boundary layer and wake. ARC R&M 2107. 


APPENDIX 


The raw data and derived results of the present 
experiments are tabulated in the following so that 
they can be used independently by other investi- 
gators. Table 3 shows the measured pressure and 
shear stress coefficients on Afterbodies 1 and 2. 
Tables 4 and 5 provide the measured static pres- 
sure coefficients across the stern boundary layers 
and near wakes of Afterbodies 1 and 2, respectively. 
Tables 6 and 7 contain the values of measured mean 
axial and radial velocities, three components of 
turbulence fluctuations, and Reynolds stresses 
across the boundary layer and near wake of After- 
bodies 1 and 2, respectively. The experimentally 
derived data on eddy viscosity, mixing length, 
planar and axisymmetric displacement thickness, and 
boundary layer thickness are also given. 


148 


TABLE 3 - Measured Pressure and Shear Stress Coefficients on 
Afterbodies 1 and 2 


6 


AFTERBODY 1, RX = 6.6 x 10 AFTERBODY 2, Ry = 6.8 x 
oe oe 

a c c a P c 

L max p T L max p 
0.7060 0.9690 -0.062 = 0.6000 1.0000 -0.013 
0.7455 0.9267 -0.064 0.00281 0.7000 1.0000 -0.024 
0.7952 0.8423 -0.050 0.00265 0.7455 1.0000 -0.035 
0.8449 0.7192 -0.024 0.00248 0.7952 1.0000 -0.106 
0.8946 0.5480 +0.018 0.00213 0.8449 0.9476 -0.160 
0.9145 0.4633 +0.050 0.00185 0.8946 0.7349 -0.010 
0.9344 0.3636 +0.074 0.00163 0.9145 0.6137 +0.053 
0.9543 0.2396 +0.112 0.00130 0.9344 0.4834 +0.090 
0.9642 0.1900 +0.133 0.00115 0.9543 0.3317 +0.170 
0.9741 0.1394 +0.135 0.00104 0.9642 0.2547 +0.183 
1.0000 0.0000 +0.116 - 0.9741 0.1740 +0.198 
1.0000 0.0000 +0.185 


TABLE 4 - Measured Static Pressure Coefficients Across Stern Boundary 
Layer and Near Wake of Afterbody 1 


x/L = 0.7553 x/L = 0.9144 x/L = 0.9344 

é r-r) = rr) a Tr 
"max Tmax C Tmax vnax Cp Tax "ax Cp 
0.9127 0 -0.0560 0.4633 0) 0.050 0.3636 0 0.0740 
0.9345 0.0218 -0.0530 0.4997 0.0364 0.0604 0.4214 0.0578 0.0821 
1.0283 0.1156 -0.0510 0.5181 0.0548 0.0604 0.4981 0.1345 0.0791 
1.1298 0.2171 -0.0500 0.5508 0.0875 0.0587 0.6102 0.2466 0.0674 
1.2392 0.3265 -0.0480 0.5892 0.1259 0.0570 0.7253 0.3617 0.0682 
1.4736 0.5609 -0.0434 0.6687 0.2054 0.0546 0.8375 0.4739 0.0624 
1.6767 0.7640 -0.0380 0.7397 0.2764 0.0514 0.9497 0.5861 0.0593 
1.8720 0.9593 -0.0350 0.8519 0.3886 0.0492 1.0989 0.7353 0.0518 
2.0908 1.1781 -0.0302 0.9670 0.5037 0.0464 1.2509 0.8873 0.0471 
2.2861 1.371 -0.0287 1.0835 0.6202 0.0434 1.4000 1.0364 0.0421 
2.4736 1.5609 -0.0270 1.1957 0.7324 0.0400 1.5563 1.1927 0.0368 
2.8798 1.9671 -0.0226 1.3093 0.8460 0.0370 1.7054 1.3418 0.0333 
3.2783 2.3438 -0.0214 1.4386 0.9753 0.0338 1.8546 1.4910 0.0305 
3.7079 2.7952 -0.0197 1.5906 1.1273 0.0301 2.0066 1.6430 0.0264 
4.3251 3.4124 -0.0158 1.7426 1.2792 0.0269 2.0904 1.7268 0.0256 

WASOSI W4298F 1OROZ4 1S 5022 i74m2; O09 010235, 

2.0451 1.5418 0.0221 3.3580 2.9943 0.0114 

2.1971 1.7338 0.0206 3.6960 3.3324 0.0099 

2.3491 1.8858 9.0183 4.0355 3.6719 0.0079 


PBHRRWNHRPRPRPRrFRrRRrFPODOCOCCCOCCO0O 


10 


SiOLOlIOLO LOLOrOrO1Oore 
[o) 
i=) 
iy) 
(>) 
S 


BPRWNWRPRPRrRrRrrROOCOOOCOCCoO 


SOLO OLOLOROROL OOF OLOROLOEOVOL OTS: 


TABLE 4 (Continued) 


Sel, S On CyAl x/L = 0.9830 x/L = 0.8462 
ag ey r 16 * me 
= Tax C Tax rnax CS = Yr = 
max max max 
0.1364 0 0.135 0.1164 0 Ost 7/155 0 -0 
0.1600 0.0236 0.1441 0.1278 0.0114 0.1250 0.7450 0.0295 -0 
0.2324 0.0960 0.1400 0.1647 0.0483 0.1249 0.7620 0.0466 -0 
0.3447 0.2083 0.1293 0.2031 0.0867 0.1241 0.7791 0.0636 -0 
0.4569 0.3205 0.1172 0.2471 0.1307 0.1224 0.8018 0.0864 -0 
0.5748 0.4384 0.1058 0.2826 0.1662 0.1211 0.8217 0.1063 -9 
0.6827 0.5464 0.0950 0.3153 0.1989 0.1189 0.8572 0.1418 -0 
0.7949 0.6586 0.0856 0.3892 0.2728 0.1142 0.8572 0.1801 -0 
0.9057 0.7693 0.0781 0.4659 0.3495 0.1090 0.9297 0.2142 -0 
1.0251 0.8887 0.0705 0.5412 0.4248 0.1036 0.9694 0.2540 -0 
1.1373 1.0009 0.0628 0.6136 0.4972 0.0974 1.0092 0.2938 -0 
1.2466 1.1102 0.0574 0.7684 0.6520 0.0858 1.0873 0.3719 -0 
1.3645 1.2281 0.0536 0.8437 0.7273 0.0817 1.1598 0.4443 -0 
1.4796 1.3432 0.0490 0.9162 0.7998 0.0774 1.2322 0.5168 -0 
1.5918 1.4555 0.0450 0.9914 0.8750 0.0735 1.3061 0.5906 -0 
1.7395 1.6031 0.0406 1.1463 1.0299 0.0650 1.3785 0.6631 -0 
1.8901 1.7537 0.0374 1.2968 1.1804 0.0555 1.7000 0.9945 -0 
2.0137 1.8773 0.0350 1.4431 1.3267 0.0505 
1.5966 1.4802 0.0441 
1.7343 1.6179 0.0409 
x/L = 1.00 x/L = 1.0076 sh = i057 Sih & Waiey 
16 r a6 ag 
ie C r C r C Cc 
max p max 2) max p max p 
0.061 0.1295 0 0.0995 0 0.0418 0 0.0181 
0.098 0.1209 0.099 0.0987 0.057 0.0413 0.036 0.0174 
Qobye  OWtO O,292- ©0868 O,152° O,040 —~O,07%41  @.0nz2 
0.249 0.1022 0.327 0.0897 0.247 0.0426 0.149 0.0178 
0.361 0.0940 0.453 0.0815 0.344 0.0424 0.224 0.0189 
0.477 0.0865 0.585 0.0740 0.433 0.0415 0.303 0.0194 
5537 W075 Os720, OL0677%  O.527 > O.0400) 0.637 O.Olee 
0.702 0.0716 0.855 0.0604 0.616 0.0396 0.561 0.0180 
0.814 0.0656 0.980 0.0553 0.831 0.0364 0.716 0.0176 
0.928 0.0604 1.116 0.0503 1.000 0.0353 0.864 0.0170 
TOSS OO 527. 2417 WN OMO46S el l6 ON = ON0333s Og. OL Ol70 
W273) 100464515382) 1004267 91,342) 040316" ol sll6s) © O-ol70 
Loos Oo0205 2.5 O.0K7 MoS ©0298) Tale —@,oeH 
1.724 0.0344 1.724 0.0344 1.724 0.0249 1.464 0.0160 
1.900 0.0230 1.621 0.0150 
2.240 0.0195 1.756 0.0140 
2.580 0.0168 
2.750 0.0150 


149 


150 


TABLE 5 


x/L 


HIH 


NNN FPRP RRP RRP PRR O 


max 


-9855 
.0536 
.1417 
.2141 
.2937 
- 3406 
-4173 
.4968 
.6374 
.9357 
.2085 
-4926 
. 8860 


NNFRrRrROOOCOOCOCOC oO 


RPreroooocna0qcco0co 


Measured Static Pressure Coefficients Across Stern Boundary 
Layer and Near Wake of Afterbody 2 


0.8400 


1e=ae 


NrRPrRPOODOOCOCOCOCOCOCCO 


-0 
-0 
-0 
-0 


=0). 
-0. 
=10)6 
-0. 
-0. 
-0. 
=< 
=. 
-0. 


CHS) Cy(ey(o) Koyo (Se) ey (yoyo yie) (SiS) 


SY LS iSyeC ft SS) (S) (Sy-o(e) Sy )(S) 


x/L 


NrrFrFODOCOCCOCCCOCCOCCOO 


Sila) (ele) (sriere) (2) ee) S) (Ss) 


SOLO sO LOFOLOLOROLOL ORO LOLS 


0.9336 


SLOZOLOLOLOLOLO LOFOlO.OLOLOr©) 


SOLO ,OLOLOLOLO VOLO LO ONOFOVOrOoro 


x/L = 0.9702 

r ETO 

Ts (¢, 

max max p 
0.2419 0.0364 0.1884 
0.2930 0.0875 0.1854 
0.3754 0.1699 0.1740 
0.4379 0.2324 0.1654 
0.5061 0.3006 0.1563 
0.5686 0.3631 0.1482 
0.6623 0.4568 0.1342 
0.7262 0.5207 0.1252 
0.8541 0.6486 0.1110 
0.9961 0.7906 0.0957 
1.1467 0.9412 0.0804 
1.4393 1.2338 0.0604 
eV: Wo Sle Me@Ak3e} 
2.0941 1.8886 0.0340 
2.4038 2.19835 0.0247 
IP SO} 7/ x/il) =) Lls2 

r 
G 10 C 
p max Pp 

0.0471 0 0.0004 
0.0467 0.0426 0.0067 
0.0492 0.2102 0.0095 
0.0503 0.3722 0.0099 
0.0447 0.5313 0.0090 
0.0484 0.6889 0.0120 
0.0450 0.8523 0.0179 
0.0493 1.1250 0.0191 
0.0462 1.7159 0.0172 
0.0439 

0.0368 

0.0340 

0.0260 

0.0234 

0.0232 

0.0198 

0.0181 


TABLE 6 - Measured Mean and Turbulence Velocity Characteristics for Afterbody 1 


x/L = 0.755, r_/r = 0.9127 tan a = -0.0671 
O° max 


es) uy vy Jar2 Jae wi2 pee uty! TR € 
max UG Us U5 U Y we, oF O. Uso 
0.013 O67 O05 O07) O.0883 O47 On08 Ooi O.084 
0.034. 0.719 -0.018 0.069 0.039 0.049 0.146 0.168 0.121 0.0080 
0.08: On Or O.0683 Os083 OLOlG ONS OnIG2 Ole  WrOlAo 
0.074 0.820 -0.019 0.064 0.036 0.043 0.123 0.170 0.264 0.0152 
0.02) OG O01 O08 OsOGS ~O.042 “OLUIO O.NG2 ORIG Oily 
0.110 0.881 -0.019 0.057 0.032 0.039 0.0969 0.167 0.393 0.0149 
0.129 0.908 -0.019 0.052 0.030 0.036 0.0809 0.165 0.460 0.0145 
OMmi1Go! 051) 08020) \0s038) 0N024) 0,052) 00589) sOnuzey)  On6045) o,00911 
0.203 0.983 -0.020 0.027 0.017 0.018 0.0159 0.118 0.725 0.0072 
0.241 1.003 -0.019 0.012 0.010 0.011 0.0022 0.061 0.861 0.00125 
0.280 1.015 -0.017 0.007 0.006 0.006 0.0004 0.024 1.00 0.00017 
0.361. 1.020 -0.016 0.003 0.003 0.003 0.00005 0.019 1.29 
O.“5 ICO O02 0,002 0,002 002 = 
0.6% 1.000 0.008 ©0022 0,008 O00 < 
1,768 1000  -W.008 O.00F  O.002 002 = 
&* $* 8 
2 = 0.0426, = = 00444 Mi ="0R280 
max max max 
x/L = 0.846, r /r_4 = 0.7155, tan a -0.1343 
aT uy vy fie Sie (a2 Aopen -uly! TEXO € 
r U_ Um 5 Y % wa aS Sy Uso" 
max eo} (e) 
O05 O05 Os073  O.067 Oc0k3 ~ WL0AS Oss Oe) OIA 
0.0280 0.660 -0.082 0.066 0.035 0.042 0.110 0.150 0.0800 0.00536 
0.0843 O78 0.080) OGG O.058 0.050 O.108 O16? O.i153°  O,00oo 
0,070 O77 0,08 0,089 @.05 O.088 OOP Osis  O.al7 O,0128 
QnOgS4y) ONS06) 08093), 005400) 0505595 0F036) 0809555. Ohl 79m MOnzelN ss ononad 
01700 O40  O,098 O02  O.052 O05 O00 O.i67 0.283 O06 
O15, 0.664 0.08 @,052 GO.08l § O.055 O.082 OniG7 O6e05  —W.0140 
0.1600 0.890 -0.093 0.050 0.031 0.035 0.0815 0.174 0.457 0.0142 
0,105 O98 0,093 O.042 ©0285 O08 O.0888 0.145 —G.557 —W.0nas 
02168 O98. 40,098 6.055 0.023 0.025 0.0204 O.129  O.G18  O.00me 
0.2250 0.665 =.0S 0.029 0,02, @.025 O03 O07 O.6720 Ooms? 
0.2560 0.073 0.08 0,026 0.018 ©0283 0118 O.094 O.75l O,005i8 
0.2770 OOS, 0,087 0.099 0,014 O07 ©0025 0.07 O.791  O,0May 
0.3341 1.003 -0.084 0.005 0.005 0.006 0.00045 0.046 0.955 0.00053 
0.6040 1.002 -0.063 0.003 0.002 0.002 0.00002 
1,203 1,000 O05 0,002 0.002 @.002 © 
DMOONSHNI O00) 20024) ONOO2MNONOO2NNOROO2M =n0 
6* 6x 6 
- = 0.0489, = 5 0,082, 22 2 O88 


SOLO FOLOL.O OOO lo: 


SISGLEKOIOVOLOLONORoFOto lS: 


oO 

“I 

co 
CriOlrOlOoLOoxOvORONCIe 


SS OLOLOLOLOLOhOL OLOrOlO) 


151 


TABLE 6 - Continued 


x/L = 0.934, Bfte oe 


r-Y, be Wes ut 

U_ U_ Un U 
max fo) fo) fo) 
0.0127 0.425 -0.096 0.055 0. 
0.0511 0.541 -0.104 0.049 QO. 
0.0909 0.613 -0.105 0.047 0. 
0.1704 0.727 -0.103 0.047 0. 
0.2360 0.805 -0.097 0.040 0. 
0.3309 0.884 -0.091 0.038 0. 
0.4105 0.931 -0.087 0.026 0. 
0.5284 0.964 -0.076 0.007 QO. 
0.6477 0.974 -0.066 0.003 0. 
0.8366 0.982 -0.055 0.003 0. 
1.1093 0.989 -0.044 0.002 0. 
1.4470 0.993 -0.034 0.002 0. 
1.7926 0.997 -0.028 0.002 0. 
2.1803 1.000 -0.024 0.002 0. 


x/L = 0.964, A ee = 0.190, tan a 
ae wi ‘e a2 

max US Ue U5 

0.0145 0.294 -0.085 0.045 0 
0.0700 0.428 -0.093 0.045 0 
0.1168 0.509 -0.086 0.046 0 
0.1751 0.586 -0.083 0.045 0 
0.2589 0.682 -0.075 0.044 0 
0.3469 0.768 -0.069 0.042 (0) 
0.4748 0.868 -0.061 0.035 0) 
0.6069 0.936 -0.055 0.021 QO. 
0.7361 0.964 -0.047 0.004 0 
0.8654 0.978 -0.041 0.003 0 
1.0671 0.988 -0.035 0.002 0 
1.2674 0.994 -0.029 0.002 0 
1.4719 0.997 -0.026 0.002 0) 
1.6793 0.991 -0.022 0.002 0 
1.881 0.996 -0.020 0.002 0 
2.1893 0.998 -0.018 0.002 0 

6 * 


= 0.364, tan a -0.2440 


ee wi Loge saa es € 
fo) Yo U5 a* Yr Usep 
034 0.039 0.104 0.182 0.023 
033 0.038 0.0759 0.154 0.0929 Q.00302 0 
032 0.037 0.0740 0.161 0.165 0.00424 0 
032 0.056 0.0700 0.154 0.310 0.00517 0 
028 0.034 0.0556 0.157 0.429 0.00513 0 
025 0.028 0.0300 0.105 0.602 0.00413 0. 
018 0.021 0.0139 0.096 0.746 0.00287 0 
006 0.007 0.00088 0.066 0.961 0.000465 0O 
003 0.003 0.00002 0.003 1.178 0.0000277 0 
002 0.002 0.00003 
002 0.002 
002 0.002 
002 0.002 
002 0.002 
on Gs 6 
aE = O.L064, |= 0.0251 = 0S 
r r r 
max max max 
= =0)..2770 
ait) wy! Tyat Saniitand r-r 
See SS 
- 0 fe) (e) a0 6 P 
028 0.034 0.0767 0.193 0.0186 
.028 0.033 0.0615 0.158 0.0897 0.00176 
032 0.035 0.0654 ORSa 0.150 0.00232 
032 0.036 0.0651 0.150 0.224 0.00286 
.030 0.036 0.0605 0.146 0.332 0.00336 
.028 0.036 0.0528 ON LS7 0.445 0.00323 
024 0.032 0.0354 0.125 0.609 0.00299 
014 0.017 0.00968 0.106 0.778 0.00145 
004 0.004 0.00090 0.005 0.944 0.000304 
003 0.003 0.00004 0.014 Valog 
002 0.002 0.00002: 0.016 1.368 
002 0.002 0.00003 
002 0.002 0.00002 
002 0.002 
002 0.002 
002 0.002 
a oF 
— = 0.2343, — = 0.78 
max max 


[o-) 
R[x 


.0160 
0212 
.0262 
.0308 
.0328 
0371 
0349 
0237 


oooooocooco 


BDO DOO oe Oe) 


153 


TABLE 6 - (Continued) 


x/L, = 1.0076, soa ie = 0. 
r oe Vy a? ne) wie 100 uy 7% € Q 
Tax UG os Uy 5 UN US ae Sy U5°F 6. 
0 0.368 -0.059 0.044 0.029 0.034 0.0251 0.162 0 
0.061 0.385 -0.053 0.039 0.027 0.031 0.0385 0.154 0.067 0.00219 0.0296 
0.101 0.426 -0.047 0.037 0.027 0.030 0.0495 0.165 Onna 0.00209 0.0248 
0.139 0.462 -0.043 0.038 0.027 0.032 0.0524 0.164 ORS S: 0.00227 0.0262 
0.213 OF 535 -0.037 0.039 0.027 0.033 0.0580 0.174 0.234 0.00248 0.0272 
0.288 0.607 -0.020 0.041 0.028 0.034 0.0625 0.173 0.316 0.00292 0.0309 
0.365 0.670 -0.010 0.042 0.029 0.035 0.0610 0.159 0.401 0.00285 0.0305 
ORS 77, 0.781 -0.005 0.042 0.029 0.034 0.0530 0.141 0.524 0.00246 0.0282 
0.589 0.871 -0.005 0.036 0.028 0.032 0.0380 0.130 0.647 0.00239 0.0325 
0.702 0.929 -0.004 0.033 0.023 0.025 0.0106 0.057 OL 772 0.00105 0.0275 
0.811 0.961 -0.004 0.016 0.012 0.013 0.0020 0.035 0.891 0.00038 0.0208 
0.923 0.977 -0.004 0.004 0.003 0.003 0.0009 0.027 1.014 0.000207 0.0182 
1.040 0.993 -0.004 0.002 0.002 0.002 0.0002 
Om oF o 
= = 0.243, = = 10.350, = 0.91 
max max max 
XY / ily = SOS, ie_//ze =Hi0) 
o’ ma 
oe x ie uate) vie wi2 penne -u'y! ae e 2 
Tax aD M5 Uy os i we in oF Us°p Ps 
0 0.519 -0.049 0.036 0.026 0.034 0.0117 0.038 0 = - 
0.034 0.538 -0.043 0.035 0.026 0.032 0.0235 0.080 0.037 0.00236 0.0293 
ORAS 2. 0.609 -0.024 0.037 0.025 0.030 0.0416 0.144 0.165 0.00351 0.0328 
0.232 0.667 -0.015 0.038 0.024 0.029 0.0493 0.137 0.252 0.00396 0.0340 
5 alal OLw22 -0.010 0.040 0.028 0.031 0.0527 0.158 0.338 0.00438 0.0364 
0.388 0.774 -0.006 0.040 0.026 0.031 0.0505 0.156 0.422 0.00420 0.0357 
0.460 0.824 -0.004 0.037 0.026 0.031 0.0460 0.152 0.500 0.00394 0.0350 
0.573 0.894 -0.004 0.037 0.028 0.030 0.0400 0.132 0.622 0.00357 0.0364 
0.614 0.916 -0.004 0.034 0.021 0.027 0.0353 0.235 0.667 0.00336 0.0370 
0.693 0.952 -0.004 0.025 0.017 0.023 0.0186 0.129 0.753 0.00255 0.0357 
0.770 0.981 -0.003 0.018 0.011 0.015 0.00569 0.098 0.837 0.00104 0.0341 
0.847 0.989 -0.003 0.004 0.003 0.003 0.00033 0.051 0.921 0.00029 0.0305 
0.923 0.991 -0.003 0.003 0.002 0.002 0.00010 0.024 1.003 0.000323 = 
1077, 0.991 -0.002 0.002 0.002 0.002 0.00007 = Naka - 
1276 0.993 -0.002 0.002 0.002 0.002 0.00005 = - 
1.459 0.994 -0.002 0.002 0.002 0.002 0 
1.697 0.998 -0.002 0.002 0.002 0.002 0 
Oe é* § 
aE Ost =) = oLeszi == = pep 
r r r 


max max max 


154 


TABLE 6 - (Continued) 


X/il= ale BD lee 


SS) [It IOs Cy ie) (s} SS) (S)'(S) 


x/L = 


> 
~— 
wo 
SLOT OVOLOLOtO Oro eOLO1Oro© 


= 0 
ax 
ux ‘y \ ul2 yi2 Ww 2 1002u -u'v' 
im Tah Vi) Ti ae Taare 2 
U, U, U5 US U5 UZ q 
632) | -OMO1ON) 10-0400 w OK 034) On034 0 0 
644 08008)" 0040) 08054) ONossi  On0055. Os 01s8 
G0 <O.005 O00 OL  O.025 O.0I40 0.055 
63. 0,005  OxMO- Wes OOH D063 W207) 
Te 0,004 O05 C.050 0,052 O,0403  O.ilA 
ks O08 O06 O025  @.050 O00  G.lBR 
BS 7 eAOO03) 10032 = 0025 te OMOZSHEONOS40N)  NOnI46 
RO) =O, Ose “G.0Rl ~~ Os0e5 Os0R90 ~~ Opis 
7 O02 WelRs Onc”  O,02, 0.0200 050 
oo . EOS 'OL0O  “C.0d) O00 0083 O,025 
995  -0.002 0.003 0.002 0.002 0.00006 0.035 
5 <0 O007 , O00  O.00R O.onm = 
GOSH a = OF0 01 ONOOZMENONCOZMMONOO? 0 u 
6* 6* 
— = 0.1543, — = 0.2832, — = 0.95 
r ae) ay 
max max max 
0.914, r/r.. = 0.9145 x/L = 0.977; r,/r,,, = 0.1364 
aes Be 3 ee ace uy : ts 
Tax (e} OY, Tax UR Us 
0.0042 0.448 0.085 0.0073 0.318 0.033 
0.0136 0.482 0.091 0.0128 0.334 0.039 
ONS W520 © W 20M 0.0183 0.345 0.041 
0.0280 0.552 0.096 0.0238 0.360 0.046 
0.0392 0.574 0.104 0.0349 0.381 0.042 
0.0500 0.598 0.107 0.0459 0.400 0.047 
0.0700 0.639 0.112 0.0624 0.428 0.055 
0.0907 0.675 0.108 0.0845 0.459 0.045 
0.1100 0.706 0.111 0.1011 0.483 0.049 
Ost s07/0 nO 7542 OOS 0.1294 0.524 0.049 
DIS - Oona  Wanws Oot 0.855 ~O.085 
0.1772 0.796 0.099 0.1845 0.595 0.045 
0.2058 0.833 0.096 0.2176 0.634 0.044 
0.2344 0.864 0.093 0.2563 0.676 0.040 
0.2687 0.894 0.091 0.2901 0.713 0.036 
0.2980 0.919 0.089 0.3328 0.748 0.033 
0.3324 0.941 0.088 0.3839 0.796 0.035 
0.3782 0.964 0.085 0.4287 0.837 0.033 
0.4067 0.979 0.082 0.4949 0.884 0.030 
0.4475 0.984 0.078 0.5563 0.915 0.028 
0.4876 0.987 0.074 016225) 08947. 0.029 
0.5276 0.990 0.069 0.6894 0.960 0.025 
0.5677 0.992 0.067 0.7501 0.962 0.020 
0.6142 0.992 0.063 0.8170 0.976 0.019 
0.6599 0.992 0.059 0.9494 0.982 0.016 
0.7229 0.993 0.054 1.0218 0.984 0.013 
0.7915 0.994 0.052 Ino) Woes — OsonA 
0.8609 0.995 0.047 1.2487 0.989 0.011 
0.9417 0.995 0.043 1.3253 0.991 0.009 
1.0275 0.998 0.041 1.3984 0.991 0.009 
1.1369 1.000 0.032 1.4756 0.993 0.007 
iezo7iee wIeOOIy 104052 1.5474 0.994 0.006 
1.4522 1.003) 0.027: 1.6253 0.994 0.006 
1.6474 1.004 0.024 1.7026 0.995 0.005 
1.8770 1.004 0.005 1.7743 0.995 0.005 
6+ 6* 6* 
= 0.0772, = = 0.0875 — = 0.1866, = = 0.2469 
max max max 
§ 
= 0.54 tana = -0.2094 = = 0.80 tana = -0.1036 
max 


LS) (SS) (SSS) O) S) Sys) ©) 
Ww 
wo 
N 


x/L = 1.000, De / oe 


SEOLOKOKOLOLORO YO) 
(=) 
fo) 
os 
| 
Oo 


u 
x 


re eee eee oo ooo ooo ooo oo ooo loko nholonolic) 
Oe Go oO 
uw 
ise) 


NO 
Ww 
ito} 
ss 
ale 


-90 


fon 
fos) 
oo 
he ee eee ee ee eo oN ooo ooo oo Molo nooo Noho—moio}) 


ax 


TABLE 7 - Measured Mean 


x/L = 0.840, r/r = 


Cpa a 
¥ U U 
max ie) 10) 
0.018 0.703 -0.091 
0.058 0.8444 -0.109 
0.0935 0.9096 -0.106 
0.1304 0.956 -0.112 
0.1645 0.9911 -0.115 
0.2380 1.037 -0.110 
0.3094 1.042 -0.103 
0.4500 1.038 -0.089 
0.6659 1.028 -0.073 
1.1602 1.013 -0.051 
1.7668 1.004 -0.038 
6* 
up 
r 
max 


x/L = 0.9336, Tatas = 


r-r u Vv 
° RE aac 

Tr U U 

max ° fo) 

0.0102 0.3206 -0.088 
0.02 0.402 -0.091 
0.055 0.546 -0.115 
0.079 0.598 -0.117 
0.112 0.652 -0.119 
0.140 0.6955 -0.122 
0.201 0.776 -0.121 
0.228 0.8063 -0.119 
0.300 0.878 -0.115 
0.330 0.903 -0.113 
0.375 0.930 -0.110 
0.427 0.9524 -0.104 
0.483 0.970 -0.094 
0.555 0.984 -0.090 
0.660 0.995 -0.085 
0.9273 1.000 -0.070 
1.340 1.000 -0.052 


and Turbulence Velocity Characteristics for Afterbody 2 


0.9618 tana = -0.1047 


ane yi2 wi2 10024 -u'y! io 
US Yo Uo Uo a Sy 
0.0723 0.0364 0.0387 0.1366 0.169 0.064 
0.0676 0.0328 0.0368 0.1234 0.176 0.204 
0.0600 0.0320 0.0340 0.0950 0.164 0.328 
Os, | OL0233- Wn0rOS | O07 Oda” O.718e 
0.0402 | 0.0213 0.0240 0.0411 0.155 0.577 
OLOLOL)  (OX0091N) HOn00929) OF 00S35 eu On las moreso 
O00 W005 O.00s j= 1.086 
0.0025 0.002 0.002 2 
0.002 0.002 0.002 = 
0.002 0.002 0.002 = 
0.002 0.002 0.002 u 
6x 6 
OOE%, =— > O08, => 3 OAs 
max max 
0.4839 tan a = -0.3216 
ate yt fyr2 -u'v'-u'yv! BE 
Uae fh URm Uae ome aq tenes 
[e) (0) e} fe) Le 
0/0569)" 0.030" 104035810093. On173. OxOg2 
O06) O04 O04 @L095 Ol) OsOss7 
0.0541 0.0343 0.038 0.0990 0.178 9.098 
0.052 0.032 0.038 0.0801 0.154 0.141 
0.053 0.031 0.037 0.0824 0.160 0.20 
0.049 0.030 0.036 0.0785 0.170 0.250 
0.049 0.030 0.033 0.0759 0.173 0.359 
QOASH wn ON O27HMOROSONNOFOSI751 NO; LO SHONd Or 
0.037 0.026 0.030 0.0495 0.165 0.539 
O02 0.023 0,025. 0.0632 Msl5SS OLS 
0.029 0.019 0.022 0.0167 0.099 0.670 
0.018 0.014 0.015 0.0078 0.105 0.763 
0.011 0.010 0.000 0.0027 0.084 0.863 
0.004 0.002 0.003 0.0007 0.014 0.991 
0.002  O.004 Oc) = 1.179 
0.002 0.002 0.002 0 1.656 
0.002 0.002 0.002 2.393 
é* 6* 8 
— = 0.1126, = = 0.1296, = = 0.560 
max max max 


oo0o0o00 00000000 


OOOO OOOO) OLS) 


0.05 0.0180 
0.0697 0.0250 
0.0780 0.0280 
0.0809 0.0291 
0.0623 0.0224 


La L 
6 / ae, 
r (reo iste 
0205 0.0124 
0260 0.0157 
0322 0.0195 
0336 0.0203 
0420 0.0254 
0401 0.0243 
0451 0.0273 
0454 0.0275 
0442 0.0268 
0421 0.0255 
0355 0.0215 
0299 0.0181 


155 


156 


TABLE 7 - (Continued) 


x/L = 0 


Sf S) Syeoy SS) SS) (Sy (S) ) (>) 


DS) [3 (SESS) SOOO SOQ) S| © 


SEORLOFOROLOLOLOLO LOLOL OlOr© 


SIOLOLOLOROZOLOLOLOnOFOre: 


sGl7Ad)5) Se //ae 
o’ ma 


x 


0. 


OLOLO LOLOL OVOLOTOLOLOVORSre 


SLOLOLOLSLOVOLOLOLOLOR OLS: 


tan a -0.4077 


yi2 w! 
wv. U 
oO fo) 
0.0244 0.241 
0.0304 0.030 
0.0333 0.0352 
0.0324 0.0359 
0.0314 0.0355 
0.0290 0.0354 
0.0271 0.0315 
0.0232 0.0290 
0.0168 0.0219 
0.007 0.007 
0.002 0.002 
0.002 0.002 
0.002 0.002 
0.002 0.002 

6* 

a 0.239 

max 
tan a = -0.3901 
ui2 , Tanta 
uv. Ua 
ie} ie} 

0.0175 0.0230 
0.0289 . 0.0319 
0.0301 0.0344 
0.0314 0.0354 
0.0317 0.0364 
0.0287 0.0339 
0.0219 0.0250 
0.012 0.014 
0.003 0.003 
0.002 0.002 
0.002 0.002 
0.002 0.002 
0.002 0.002 

6* 

oo = 0.233, 

max 


-u'v' -ulv! 
100 
uz q? 
fe} 
0.0531 0.185 
0.0756 0.177 
0.0850 0.164 
0.0872 0.171 
0.0800 0.165 
0.0666 0.152 
0.0545 0.156 
0.0420 0.149 
0.0199 0.126 
0.00173 0.107 
6* 
5 eS 0229); 
max 
-u'v' -u'v! 
100— 
ug ae 
OROZ19 me Omazil 
0.0544 0.152 
0.0692 0.161 
0.0788 0.161 
0.0770 0.156 
0.0625 0.159 
0.0355 0.157 
0.00868 0.124 
0.000417 0,033 
6* 
= 0294), 
Ty 


SO*OVOROROLOLOLOROI©: 


Lis) eyo) S) SY OC) 
Ww 
> 
i) 


SOLO .OLOLOLOROIO 


SEOEOLOLOVORO!S 


.00166 
.00209 
00244 
-00278 
-00287 
-00289 
-00243 
.00154 
.000339 


.78 


Q L 
o. 22 
r J (x, +6,) 2x2 


.0182 
-0215 
-0248 
.9295 
.0334 
.0372 
.0357 
-0326 
.0250 


-0206 
.0236 
-0260 
-0306 
.0356 
.0367 
.0300 
0201 


OLOsOLOLOVOLOLO OS) 


STOLOTOIOs OGIO: OS: 


.0147 
.0174 
.0201 
.0238 
.0270 
.0301 
.0229 
.0264 
.0202 


157 


TABLE 7 - (Concluded) 


x/L = 1.000 x/L = 1.057 x/L = 1.182 
Yr ox xr uy r “x 
r Um 2 U rr. U 
max fe} max fe) max (e) 
0.089 0.3460 0.000 0.4336 0.000 0.5760 
0.102 0.3605 0.026 0.4440 0.040 0.5850 
O25 Oo5705.  Onl0R  WpcW2 0.085 0.5942 
0.135 0.3755 0.216 0.5623 0.124 0.6006 
ONG) ‘OokES ~ Ooe08 O.6729 0.203 0.6216 
0.204 OcI@ O462 O.7729 0.281 0.6764 
0.238 0.4520 0.560 0.8488 0.398 0.7715 
0.279 0.4975 0.675 0.9088 0.519 0.8538 
0.307 0.5300 0.786 0.9595 0.637 0.9211 
OokSh) 0.5083  lalAl W,OR 0.752 0.9581 
0.393 0.6475 1.659 0.9856 0.904 0.9670 
0.434 0.7020 2.151 0.9893 1.172 0.9780 
0.474 0.7485 1.644 0.9887 
0,36  Oocass 50 2.174 0.9988 
0.646 0.8850 

0.722 0.9325 ra Dacehs 8s 

0.877. 0.9700 — = 0.1880 
22 SZO 6* max 

1.692 0.9970 Toe : 

A007 10S Bae bra eee 
2.265 1.0025 re 

8 

$* a= 100 Srila a5 
ep = "0/2336 Tmax Pen eh 

26 max 

max 

6* 

BS 0,870 

a6) 

max 

8 

TF = 0.96 


Theoretical Computation and Model and 
Full-Scale Correlation of the Flow at the 
Stern of a Submerged Body 


A. W. Moore 


Admiralty Marine Technology Establishment 
Teddington, England 


Go 136° Watilalss 


Admiralty Marine Technology Establishment 


Haslar, England 


©British Crown Copyright 1979. 


ABSTRACT 


This paper describes an empirical method devised 
for modifying measurements made at a propeller 
position at the rear of unpowered bodies such that 
the flow at the same position on a full-scale self- 
propelled body may be predicted. ms 

A boundary layer calculation procedure for esti- 
mating boundary-layer velocity profiles at the 
tail region of a body of revolution is discussed, 
and the inclusion of a simple representation of a 
propeller is described. Comparisons between 
velocities measured at Reynolds numbers of order 
10® and calculated velocities show reasonable 
correlation both for unpowered and for powered 
bodies of revolution. It is shown how the results 
of boundary-layer velocity calculations are used 
to derive a method for modifying flow measurements 
at model scale to represent full-scale flow over 
the propeller disc area. Comparisons are made 
between predictions based on this method and 
measurements on powered and unpowered bodies at 
high and low Reynolds numbers. 


1. INTRODUCTION 


For many applications a self-propelled marine 
vehicle has a propeller fitted at the rear of the 
body where it gains in propulsive performance and 
in cavitation performance by operating in the 
relatively slow moving fluid in the hull boundary 
layer. It follows that a fundamental requirement 
for propeller design is a knowledge of the boundary 
layer flow at the propeller position. This infor- 
mation is not usually known since there are no 
theoretical methods presently available for calcu- 
lating the boundary flow at the rear of a powered 
asymmetric body with appendages. An estimate of 
the required flow field can be obtained from 
measurements at model scale but as the Reynolds 
number based on model length is considerably lower 


158 


than the full-scale value, it is necessary to make 
some modification to the measurements to simulate 
the effect of a thinner boundary layer at full 
scale. If the flow field is measured on an un- 
powered model, as is often the case, further 
modification is required to allow for flow acceler- 
ation due to the propeller. 

This paper describes an approximate method 
which has been developed for estimating corrections 
required to flow measurements on unpowered bodies. 
A boundary layer calculation procedure is briefly 
outlined and then compared with data from tests on 
axisymmetric bodies at low Reynolds numbers and 
non axisymmetric bodies at both low and high 
Reynolds numbers. 


2. BOUNDARY LAYER CALCULATION 

The method is based on the work of Myring (1973) 
and only a brief outline is presented herein. An 
iterative scheme is adopted in which a boundary 
layer calculation is done for a given pressure 
distribution over the body and a potential flow 
calculation is done to calculate the pressure 
distribution over the body with boundary layer dis- 
placement thickness added. In the boundary layer 
calculation procedure, an integral method is used 
in which the laminar flow region is calculated 
using the method of Luxton and Young (1962) and the 
turbulent flow is calculated using a method similar 
to that due to Head (1960). The transition point 
must be specified and it is assumed that momentum 
area and a shape parameter are continuous at 
transition. 

An important feature in Myring's method is his 
treatment of the turbulent boundary layer in the 
region of the tail. The usual boundary layer 
assumptions become invalid in this region where the 
ratio of boundary layer thickness to body radius 
tends to infinity so Myring defines a momentum 
area and a displacement area which overcomes the 


problem and which reduce respectively to body 
radius times momentum thickness and body radius 
times displacement thickness far from the tail 
where boundary layer thickness is small. A con- 
ventional momentum integral equation is derived in 
terms of the defined parameters and this is solved 
using an empirical relationship for skin friction 
coefficient which assumes that wall shear-stress 
does not change sign. Therefore the method is only 
applicable to bodies on which the boundary layer 
remains attached. It is also assumed that the 
variation of static pressure across the boundary 
layer is negligible. This latter assumption has 
been found to be incorrect for bodies with blunt 
tails (i.e., cone angles greater than 30°) and an 
empirical modification has been made based on the 
work of Patel (1974) who developed independently 

a method which is similar to Myring's but which 
recognises the importance of static pressure varia- 
tion. The modification introduced in the present 
method is that the predicted velocity distribution 
along the body is changed empirically in the tail 
region, the change being related to differences 
between measured and predicted velocity distribu- 
tions at the rear of a given body with a blunt 
stern. It has been found that this modification 
results in improved correlation between measured 
and predicted boundary layer velocity profiles. 

A simple actuator disc representation of a 
propeller has now been included in the potential 
flow part of the calculation in order to give a 
first approximation to the acceleration effects 
on the flow caused by the action of the propeller. 


PREDICTED 
x MEASURED 
0-7 


LOCAL VELOCITY 
MODEL VELOCITY 
se) 
Ca) 
aT 


° f 2 3 4 
DISTANCE FROM HULL (%) 
BODY LENGTH 


FIGURE 1. Measured and predicted velocities 0.96L from 
the bow of a body of revolution with tail cone angle of 
AS 


159 


— — — PRESENT THEORY 
NO MODIFICATION. 


PRESENT THEORY WITH 
EMPIRICAL MODIFICATION TO 
ALLOW FOR STATIC PRESSURE 
GRADIENT ACROSS THE THICK 
BOUNDARY LAYER. 


x MEASUREMENTS (PATEL, 1973) 


LOCAL VELOCITY 
MODEL VELOCITY 
° 
a) 


nl 1 n 
ie) ! 


en 
4 


2 3 
DISTANCE FROM HULL (70) 
BODY LENGTH 


FIGURE 2. Measured and predicted velocities 0.96L from 
the bow of a body of revolution with blunt stern [Patel 
(1973) ]. 


3. RESULTS AT LOW REYNOLDS NUMBER ON AXISYMMETRIC 
BODIES 


Comparison between Predicted and Measured Results 


The main interest in the present work is in the 
prediction of boundary layer velocity profiles in 
the tail region of a body and the results presented 
in this section relate to model measurements under 
conditions giving a Reynolds number based on model 
length from 1 x 10© to 6 x 10°. 

The velocity measurements shown in Figure 1 were 
made at a station 0.96 L from the bow of a body of 
revolution of length L and having a relatively fine 
stern (cone angle 26°). The measurements are of 
total velocity whereas the calculation method gives 
values of velocity component parallel to the hull. 
The theoretical curve in Figure 1 is obtained by 
applying a small correction to the calculated 
velocities to allow for the difference between 
local flow angle and hull angle. It can be seen 
that the resulting predicted curve gives values to 
within 4% of the measured velocities. Detailed 
measurements at the rear of a body of revolution 
having a blunt stern have been reported by Patel 
et al. (1973) and results for a station 0.96 L 
from the bow are shown in Figure 2. The broken 
line is the theoretical boundary layer profile 
predicted from Myring's method with no allowance 
for static pressure variation across the boundary 
layer. This curve is significantly different from 
the measured velocities which are more than 10% 
less than predicted values in the inner part of 
the boundary layer. Correlation between measured 
and predicted results is improved when the empirical 
modification allowing for static pressure variation 


160 


has been made and the resulting curve is seen in Velocity profiles close to the propeller plane 
Figure 2 to be in better agreement with the measure- with and without propeller operating have been 
ments. reported by Huang (1976). A typical example is 

Theoretical results obtained with the simple shown in Figure 4 where it can be seen that the 
representation of a propeller included in the Myring theoretical prediction for the unpowered 
method indicate that the propeller can produce body gives velocities which tend to be too low. 
large local changes in the boundary layer flow. An Nevertheless the discrepancy is less than 4% of the 
example for which measurements are also available measured values. 


is shown in Figure 3 where results are presented 
for various stations along a body of revolution 


having a fine tail and contra-rotating propellers. Reliability of Harmonic Analyses of Measured Flow 
The measurements are of total velocity and were Fields 
made with rakes of probes fixed to the body, the 
rake at the forward propeller plane being removed The comparisons between theoretical prediction and 
when the propeller was fitted. The velocity pro- measurement indicate that the calculation method 
files on the unpowered body are well predicted gives a good approximation to velocity profiles 
except close to the tail where boundary layer measured on powered and unpowered bodies of 
separation appears to be present: it is noted revolution. The method is not expected to pre- 
earlier that the calculation procedure will not dict velocities to better than 4% in absolute 
predict separation. The changes produced by the terms but this is satisfactory for the purpose of 
propellers are in surprisingly good agreement with deriving a simple means for modifying model measure- 
predicted changes considering that an actuator ments to represent full-scale values. It is re- 
disc representation of the propellers has been quired to obtain a representative flow field over 
adopted. The velocities in a region close to the the propeller disc area and an essential starting 
hull are under-predicted at the two rearmost point is to have reliable model data not only in 
stations and at the station very close to the tail the sense that velocities can be measured accurately 
the velocities near to the edge of the boundary at a given point, but also that, if a Fourier anal- 
layer are also underpredicted. Apart from these ysis is made of the velocities measured during one 
discrepancies the effect of the propeller is well revolution at a given radius, then a good approxi- 
represented. mation to the magnitudes of wake harmonics is 

1-0 ve) + OF O 


PREDICTIONS 

WITHOUT PROPELLERS 
— — WITH PROPELLERS 
MEASUREMENTS 
+ WITHOUT PROPELLERS 
O WITH PROPELLERS 


/ = 0-763 


/, = 0.621 


LOCAL VELOCITY 
MODEL VELOCITY 


X/1 = 0.898 
PLANE OF FORWARD PROPELLER 


ie) | 2 3 4 O ! 2 3 4 


DISTANCE FROM HUB (*/o) 
BODY LENGTH 


FIGURE 3. Velocity predictions and measurements on a torpedo-like body. 


>|> 
EVE 
V/V 
roy Ke) 
4/4 
Ww] Ww 
>\|> 
aia 
<|w PRESENT 
o}a 
o|O O-4 THEORY EXPT. 
4/|= 
+ NO PROPELLER 

0-3 —_—-— ® WITH PROPELLER 

0-2 

O:-1 

fe} 1 (— n J 
° I 2 3 4 
DISTANCE FROM HUB (°%o) 
BODY LENGTH 
FIGURE 4. Velocity profiles immediately ahead of the 


propeller DTNSRDC body 5225-1 [Huang (1976) ]. 


obtained. This information is relevant to the 
estimation of unsteady forces generated by a pro- 
peller. Some tests have been made in a wind tunnel 
to assess the reliability of model measurements at 
a typical propeller position on a three dimensional 
body. Inflow non-uniformity was introduced by 
fitting four struts to the body and the velocity 
field was measured by a single traversable pitot- 
static prove with head 1.5 mm in diameter. Measure- 
ments at a given radius were made on different runs 
with incremental steps of 1°, 2°, and 3° in the 
circumferential position of the probe and 10 repeat 
runs were made with 3° incremental steps. A Fourier 
analysis of each set of results was made and the 
harmonic spectra are summarised in Figure 5. It 
can be seen that the standard deviations in the 
magnitudes of wake harmonics are quite small show- 
ing that misleading information concerning the 
relative magnitudes of different wake harmonics 
would not be obtained on any one run. The differ- 
ences in magnitudes from the runs with 1°, 2°, and 3° 
steps in probe position are also quite small in 
general although a few wake harmonics, such as ll, 
do show significant changes. No consistent trend 
is observed in comparing amplitudes at low harmonic 
number but at harmonic numbers greater than 25 the 
amplitudes obtained from the run with 1° steps tend 
to be higher than those from other runs, the impli- 
cation being that choosing a coarser step size has 
resulted in a small loss in accuracy. 

The amplitudes of wake harmonics at harmonic 
numbers greater than 20 are small (less than 0.005 
times tunnel speed) except for harmonic numbers 
which are multiples of 4. These higher values are 


161 


associated with the wakes from the four struts 
which each produce a 'trough' in the measured flow 
field. The high harmonic amplitudes at high har- 
monic numbers implies a possible inaccuracy in 
results from a Fourier analysis based on the finite 
number of measured points. This was investigated 
theoretically by assuming an idealised wake defect 
giving a triangular waveform as indicated in 

Figure 6. The number of wake defects and wake 
width could be varied and for each assumed flow 
field an exact Fourier analysis was obtained ana- 
lytically and the results were compared with similar 
analyses determined numerically with the waveform 
described at discrete points as specified in the 
measurements. Figure 7 shows results obtained with 
4 narrow wake defects and 120 points specifying the 
velocity profile. Two wake widths are considered; 
when maximum wake width is 9° harmonics above 20 
are in reasonable agreement with the exact solution 
although harmonics below 20 are too low; when wake 
width is reduced to 44° the amplitude of harmonics 
from the exact solution falls slowly with increas- 
ing harmonic number whereas the amplitudes deter- 
mined numerically show no reduction in amplitude. 
In this case, where points are specified every 3° 
and the width of each wake defect is only 4%°, 
‘aliasing' in the numerical results is not un- 
expected. Such pitfalls in numerical analysis are 
well known and Manley (1945) shows that erroneous 
values in analyses of the type described above 
might be expected at harmonic numbers given by 
(N-jK) where N is the number of specified points, 
K the number of wake defects and j is an integer. 

A parametric study for triangular waveforms in 


0:04 _ 
§ 1° steps (1 RUN) 

2° STEPS (I RUN) 

3° STEPS (MEAN OF I! RUNS) 
STANDARD DEVIATION 


— ce 


0 SSS 22 


N 

\ 
ry \ 
w \) 
w \ 
a \ 
uw \ 
4 0-02 N 
Ww 
z \ 
z 
> | 
- \ 
= 
Ww \ 
> o0-Ol N 
e \ 
a \ 
a \ 
Bs \ 
S \ 
© N 
$ © 
> 3 
« 
< 
=x 


FIGURE 5. Harmonic analysis of different measurements 
of a non-uniform flow field. 


162 


AMPLITUDE 


e 277 277 


kK 


FIGURE 6. Theoretical representation of wake defects 
in the flow field. 


which N, K, and wake width were varied showed that, 
in general, errors did not become significant until 
wake width was less than twice the angular spacing 
of the specified points, i.e., 720°/N. 


4. USE OF THE PREDICTION METHOD IN PROPELLER DESIGN 


A knowledge of the flow in the region of a propeller 
is required first, in order to design it, and 
second, to estimate its performance characteristics. 
The former requires an estimate of the unpowered 
mean velocity through the propeller position to- 
gether with the radial variation of mean circum- 
ferential velocity. Of the latter, the prediction 
of unsteady propeller forces in particular also 
requires the detail wake structure at the propeller 
position in the powered condition. 

The theoretical boundary layer prediction method 
outlined in Section 2 cannot be used directly to 
predict the above wake information for practical 
vehicle configurations because of limitations such 
as its restriction to unappended bodies of revolu- 
tion. However, it can be employed indirectly by 
using the method to predict the changes from model 
testing conditions to full-scale vehicle conditions 
and then applying these scale effects to available 
model data. 

The procedure adopted for the predictions dis- 
cussed in the following section was to replace the 
non symmetric, appended vehicle by an equivalent 
body of revolution. Powered and unpowered boundary 
layer predictions were then carried out and, by 
assuming a simple power law for the boundary layer 
velocity profile, the mean circumferential veloci- 
ties were determined for the equivalent model and 
full-scale bodies. In this way it was possible to 
estimate at any position the scale effect upon the 
unpowered wakes, the propeller induction effects, 
and any combination of the two. These effects were 
then applied to all the measured unpowered model 
data to give predictions of both model and full- 
scale powered wakes for comparison with measured 
data. 


5. COMPARISON BETWEEN PREDICTED AND MEASURED RESULTS 
ON NON-SYMMETRIC BODIES 


As part of a programme to investigate the effects 
of scaling and propejler induction on wakes, experi- 


ments have been carried out on two practical vehicle 
forms covering a range of Reynolds numbers, based 
on body length, from approximately 1 x 10” to 

6 x 108. The two vehicles concerned were propelled 
by a single centre line propeller and were fitted 
with a set of cruciform after-control surfaces just 
ahead of the propeller. The afterbody form was 
axisymmetric in both cases, one vehicle having a 
fine stern (vehicle A) and the other a blunt stern 
(vehicle B). 

The low Reynolds number data were obtained in the 
ship tanks at AMTE (Haslar) using small conventional 
pitot static tubes. For body A, measurements were 
made at a position 26 percent of the local control 
surface chord aft of the control surface trailing 
edge. This corresponded to 28 percent of the 
propeller diameter forward of the propeller. The 
measurements were made at 2° intervals over an angle 
of approximately 90° centred on one control surface, 
and at radial distances from the body surface of 
12.5 percent and 25 percent of the propeller radius. 
For body B, data were obtained 24 percent of the 
local control surface chord aft of the control 
surface trailing edge, corresponding to 22 percent 
of the propeller diameter forward of the propeller. 
In this case 6 pitot static tubes were used cover- 
ing a range of radial distances from the hull of 
12.5 percent to 65 percent of the propeller radius. 

The high Reynolds number data were obtained from 
trials carried out at sea on vehicle A using 5 con- 


° 
(@) N=120, K=4, WAKE WIDTH 9. 
“10 


X NUMERICAL 


EXACT 


AMPLITUDE 
° 
ur 


4 6 12 16 20 24 28 32 36 40 44 48 52 56 60 64 
HARMONIC NUMBER 


1 °o 
(®) N=120, K=4, WAKE WIDTH 4/2. 
10 


fo) 
a 


KIX OMX YK OX EN: 


AMPLITUDE 


4 8 12 16 20 24 28 32 36 40 44 48 52 56 60 64 
HARMONIC NUMBER 


FIGURE 7. Comparison between an exact Fourier analysis 
of the theoretical velocity profile and a numerical 
analysis of the same profile specified at a discrete 
number of points. 


ventional pitot static tubes at each of the above 
radial positions. 

The high Reynolds number measurements could only 
be carried out at self-propulsion conditions. 
However, the model experiments in the ship tank 
were run over a range of propulsion conditions, 
the model speed, propeller rpm, and resistance being 
recorded. 


Analysis of Experimental Data 


The low Reynolds number results for vehicle A are 
presented in Figures 8 and 9 while the equivalent 
high Reynolds number trail data is given in 
Figure 10. For body B the available data is 
restricted to that obtained in the low Reynolds 
number ship tank tests and the results are pre- 
sented in Figures 11 to 13. For the sake of 
brevity the velocity profiles given in Figures 11 
to 13 have been limited to those for alternate 
measurement radii. 

It can be-shown that the propeller diffusion 
ratio, defined as the ratio of the mean velocity 
through the propeller to the unpowered mean wake 
velocity through the propeller position, can be 
obtained from the propeller thrust or hull resis- 
tance together with the mean volumetric wake and 
thrust deduction. Thus, using the model powered 
and unpowered resistance measurements and values 


DIFFUSION RATIO 


1-308 ————— SELF PROPULSION. 
1:226 —— — — 
1-130 -—-—- 

0-8 |:000— --—— UNPOWERED 


0-7 


0-6 


os 


LOCAL VELOCITY 
MODEL VELOCITY 


o4 
CONTROL 
SURFACE 


50 40 30 20 10 ° 10 20 30 
ANGULAR POSITION (°) 


DIFFUSION RATIO 

1-308 ——— SELF PROPULSION 
1-226—-——— 

1-130— -— 

1!- OOO— --—UNPOWERED 


0-6 


LOCAL VELOCITY 
MODEL VELOCITY 


Soa CONTROL 
SURFACE 


50 40 30 20 10 fo) 10 20 30 
ANGULAR POSITION (°) 


163 


of wake and thrust deduction obtained from previous 
model tests, the propeller diffusion ratio has been 
calculated for the model propulsion conditions 
pertaining during the experiments. Similar calcu- 
lations have been carried out for the sea trial 
conditions using data obtained from previous pro- 
pulsion trials. The results of these analyses are 
given on Figures 8 to 13, and also in Table 1 which 
summarises the experimental and trial conditions. 

The velocities just ahead of the propeller have 
been averaged at each radius to give the variation 
of the mean circumferential velocities with diffu- 
sion ratio presented in Figures 14 and 15. [In an 
attempt to quantify the secondary flow component in 
the above velocity profiles the ratio of the mean 
peak velocity to the mean minimum velocity has been 
evaluated and plotted in Figures 16 and 17. The 
normal parameter used to specify the velocity defect, 
namely the ratio of the minimum velocity in the 
"trough' to the mean velocity at the edge of the 
"trough' is given in Figures 16 and 18. No values 
are given for the inner radius on body B because, 
as can be seen from Figure 11, the wake defect is 
not clearly defined at this position. The latter 
parameter is also compared in Figure 19 with an 
empirical relationship based on two-dimensional 
data [e.g., Raj (1973)]. 

The results of using the Myring based boundary 
layer prediction method as described in Section 4 
for the powered model and trial conditions are 
also plotted in Figures 14 to 18. 


FIGURE 8. Vehicle A model velocity profiles 
at position 12.5 percent of propeller radius 
from the hull. 


EYe) 


FIGURE 9. Vehicle A model velocity profiles 
at position 25 percent of propeller radius 
from the hull. 


164 


DIFFUSION RATIO 1-160 SELF PROPULSION. 


o-8 
a °fo PROPELLER RADIUS 
z FROM HULL 
ares °. 
>|Vo ." 1 
clo O-7 25 fo 
oa 
O}> 
Ww] yy O-6 
WE 
ajo 
ol 
ro) o5 
as 4 CONTROL 
=) SURFACE 
re 
FIGURE 10. Vehicle A full-scale velocity ° 1 1 | —1__ 1 ———1_____1_____ 
profiles 12.5 percent and 25 percent of 50 40 30 20 10 ° 10 20 30 40 50 
propeller radius from the hull. ANGULAR POSITION (°) 
DIFFUSION RATIO. 
1-233 SELF PROPULSION 
IRe=——— 
Lett 
O7 | -OO0O0— --—_UNPOWERED 
ai 
ole 
glo 
= a) 
ry od 
>|> 
a 
2|8 
ro} fe) 
a)\= 
é CONTROL 
SURFACE 
FIGURE 11. Vehicle B model velocity pro- 
files at position 12.5 percent of propeller 50 40 30 20 ite} ° 1o 20 30 40 50 
radius from the hull. ANGULAR POSITION (°) 
DIFFUSION RATIO 
1-233————_SELF PROPULSION 
OS 6 Ss 
Dols <1 
0-8 
> 
Ele 
oe 
Slo 
gla 
SI> 
4 
dir 
ule 
o|O 
par) > 3 
CONTROL 
SURFACE 
FIGURE 12. Vehicle B model velocity pro- 
files at position 33.5 percent of propeller 50 40 30 20 !o ° !0 20 30 che) we) 
radius from the hull. ANGULAR POSITION (°) 


Discussion of Measured Data 


It is clear from Figure 16 that the relative magni- 
tude of the velocity defect at the two radii con- 
sidered on vehicle A is virtually unaffected by the 
propeller, and is subject to only a very small 
scale effect. The latter gives rise to a reduction 
in the depth of the velocity defect between the 


model and full-scale equivalent to an increase of 
between 1 percent and 3 percent in the ratio of 


minimum velocity in the 'trough' to the mean veloc-— 
ity at the edge of the 'trough'. It can be seen 
from Figure 19 that for vehicle A the actual values 
of the velocity defect are considerably lower than 
predicted by the empirical relationship derived 
from two-dimensional test results. This is not 


DIFFUSION RATIO 


1-233 ————SELF_ PROPULSION 
1-175 -—— — 
1th) ——-— 

1-0 |-OOO— --—UNPOWERED 


0-9 
>|> 
Ele 
1S) © 
alg o-8 
ry 
wis 
LA i 
2 Fr 0-7 
19, 
Sls 
4 CONTROL 
SURFACE 


50 40 30 20 Te} ° 10 20 30 
ANGULAR POSITION (°) 


surprising since the two radii concerned are close 
to the hull and the velocity defect is developing 
in a complex three-dimensional flow field influenced 
by the secondary flow and this is possibly leading 
to a more rapid mixing of the flow. The results 
obtained over a much larger distance from the hull 
on vehicle B support the above hypothesis since it 
can be seen from Figures 18 and 19 that as the 
distance from the hull increases the magnitude of 
the wake defect increases and approaches the two- 
dimensional value. The model results for vehicle 
B shown in Figure 18 tend, in general, to indicate 
a small increase in the depth of the wake defect 
as the propeller diffusion ratio increases. The 
maximum value of this increase in the wake defect, 
between the model self-propelled and unpowered 
condition, is only of the order of 3 percent. This 
change is somewhat surprising since the propeller 
produces a favourable pressure gradient aft of the 
control surfaces, and on the evidence of two- 
dimensional data this would be expected to reduce 
the wake defect. 


® FULL SCALE MEASUREMENTS. 
+ MODEL EXPERIMENTS. 


FULL SCALE PREDICTIONS 


0-8 
‘alte Jo PROPELLER 2 e 
olo RADIUS FROM // - ea 
° S HULL ae a 
J|4 07 a 
wis 
> 
al4 MOpeL. Ma 
< 3 o6 REDICTIONS 
Fe 
z|\z 
5|4 
2\5 
o|% 0-4 
= = FULL SCALE MODEL SELF 
w = encore PROPULSION PROPULSION 
2 

1-O | 1-2 193 1-4 rs 


PROPELLER DIFFUSION RATIO. 


FIGURE 14. Mean circumferential velocity in the measur-— 
~ ing plane for vehicle A. 


165 


FIGURE 13. Vehicle B model velocity pro- 
50 files at position 54.5 percent of propeller 
radius from the hull. 


In contrast to the velocity defect the secondary 
flow can be seen from Figures 16 and 17 to be 
significantly reduced by the presence of the 
propeller, this reduction becoming larger as the 
diffusion ratio increases. Additionally, at equal 
propeller diffusion ratio, the full-scale secondary 
flow is significantly less than measured on the 
model. From the data obtained on vehicle A 
(Figures 14 and 16) it can be seen that, comparing 
the results at model and full-scale self-propulsion 
conditions, the magnitude of the secondary flow and 
the mean circumferential velocity at the two radii 
considered agree to within 2 percent and 3 percent 
respectively. Although comparison between the 
velocity profiles is difficult because of the non- 
symmetry of the trial data, Figures 1 to 3 indicate 
that at these conditions there is also reasonable 
agreement between the velocity profiles. These 
results indicate a possible condition for similar 


1-0 
°lo PROPELLER 

> 0-9 | RADIUS Bra 

S FROM HULL Ie aug 
> 9 = = 

fe) = = 
Ela 0-8 | 65°0+- 4 + 
OU} w A 7 
o> ee ap 7 

- + 

Ny 54°54 +7 
S|5 t 

Oo O-'7 ee 
|= 2 = 
<jw 44-04 on esa 
z if + se 

. = 

ial f° 6 y ES 
| 2 2 a 
| O a ny Oe 
s|= 33-547 we 
=) ° as 
rs} 3 0-5 ee a 
4 w a + a 
(5) 

ar 23-O+ a oe 
z $ 0-4 Z 
| _ ©  +MODEL EXPERIMENTS 

= 7 ———PREDICTIONS 

Osa iene 

; MODEL SELF 
UNPOWERED PROPULSION 


1-0 t+ 1-2 1-3 1-4 
PROPELLER DIFFUSION RATIO 


FIGURE 15. Mean circumferential velocity in the mea- 
suring plane for vehicle B. 


166 


@®FULL SCALE MEASUREMENTS 
+MODEL EXPERIMENTS 


°/o PROPELLER RADIUS 
FROM HULL. 


MODEL PREDICTIONS 
1-3 


MAXIMUM SECONDARY FLOW VELOCITY 
MINIMUM SECONDARY FLOW VELOCITY 
. 


° FULL SCALE 
25 /o __ PREDICTIONS. 


— — —,MODEL PREDICTIONS 
5 


FULL SCALE 
UNPOWERED PROPULSION 


lo Il 1-2 1-3 
PROPELLER DIFFUSION RATIO 


MODEL SELF 
PROPULSION 


1-4 


MINIMUM DEFECT VELOCITY 
VELOCITY AT DEFECT EDGE 


FIGURE 16. Relative magnitude of the secondary flow 
and velocity defect for vehicle A. 


inflow to the propeller at model and full-scale; 
however, it should not be regarded as a general 
conclusion on the basis of this one experiment. 
Additionally although the propeller inflow may be 
similar at self-propulsion the propeller thrust 
loading, as indicated by the diffusion ratio, will 
be different. 

Comparison between the wake defect and secondary 
flow model measurements for the two vehicles 
(Figures 16 to 18) show generally similar magnitudes 
for the former, but a much larger secondary flow in 
the case of the body with the fuller afterbody. 


The latter effect can also be seen in the velocity 
profiles given in Figures 8 and 1l. 


Comparison between Predicted and Measured Results 


It can be seen from Figure 14 that the mean circum- 
ferential velocity predictions for the powered model 
of vehicle A are always higher than measured. The 
maximum differences occur at model self—propulsion 
conditions and are 7 percent and 4 percent for the 
positions 12.5 percent and 25 percent of the pro- 
peller radius from the hull respectively. Both 
the measured data and the predicted velocities can 
be seen to vary linearly with propeller diffusion 
ratio. For the blunter stern, Figure 15 indicates 
that for radial positions between 23 percent and 44 
percent of the propeller radius from the hull the 
predictions of mean circumferential velocity are 
generally in good agreement with the measured data. 
For the two outer radii the predictions tend to 
be high as in the case for body A, the maximum 
errors at model self-propulsion being of the order 
of 4 percent. However, for the innermost radial 
position, the powered predictions are up to 14 per- 
cent below the measured values. t is apparent 
for Figure 15 that, in contrast to the other radii, 
the model results for this position are not linear 
with propeller diffusion ratio because of the low 
velocity obtained in the unpowered condition. Since 
the measured data was linear at a similar radial 
position for body A this suggests that the poor 
powered prediction of velocity for body B is due to 
the low unpowered velocity measurement which is 
used as the datum for the prediction. This low 
measured velocity may be the result of flow separa- 
tion on the vehicle with the blunt afterbody which 
is suppressed by the favourable pressure gradient 
produced when the propeller is operating. 
Comparison between the full-scale and predicted 
mean circumferential velocities in Figure 14 show 
the latter to be less accurate than for the model 
case, the predicted values being 15 percent and 9 
percent high for the inner and outer positions 
respectively. However, correlation of propulsion 
data from sea trials and model experiments on 
vehicle A suggest an equivalent full-scale hull 
Reynolds number of one-tenth of the true value and 


TABLE 1 Experimental and Trial Conditions 


Hull 
Reynolds Diffusion 
Vehicle Conditions Number Ratio Remarks 
A Model 13 xe 1Od mesos Self propulsion 
1.226 
1.130 
A Trial 5.5 x 108 1.160 Self propulsion 
B Model ToD 3 AO? “a Dsis) Self propulsion 
1.175 


abSatalat 


MAXIMUM SECONDARY FLOW VELOCITY 
MINIMUM SECONDARY FLOW VELOCITY 


FIGURE 17. 


+ MODEL EXPERIMENTS 


— — —PREDICTIONS 
2-0 
°lo PROPELLER RADIUS 
FROM HULL. 4.9.4 2-309 
i) 


4-192 PREDICTION FOR 
12°5°lo RADIUS. 


23-0 12-5 


ee) Ua 1-2 1-3 1-4 
PROPELLER DIFFUSION RATIO. 


Relative magnitude of the secondary flow 


for vehicle B. 


MINIMUM DEFECT VELOCITY 


FIGURE 18. 


+ MODEL EXPERIMENTS 
——-—PREDICTIONS 


fo PROPELLER RADIUS 
FROM HULL. 


23 -O4+——__+—__ + | 
33-5 4+——=_ 5§ —+— 
44-O0+—= 
54°5 See ln 
O-8 — — 


a arr 
6510 cay 


VELOCITY AT DEFECT EDGE 
9 
) 


MODEL SELF 


UNPOWERED PROPULSION 


10 Vl 1-2 1-3 
PROPELLER DIFFUSION RATIO 


Relative magnitude of the velocity defect 


for vehicle B. 


167 


MODEL DATA I VEHICLE A 
MODEL DATA 2 VEHICLE B 
1-0 
fo PROPELLER RADIUS 
FROM HULL 
0-9| 33-5 g1 125 & 25-0 


wi 
Zlo 
sige eee EMPIRICAL VALUES 
8 of ‘OR APPENDAGE 
an 
rare GEOMETRY. 

Ww O-7 
air 
o]o 
4 

- 
w)< 0-6} 

> 
= 
3|5 
P 3 ro) O-5 
z aa) 
z|5 

0-4{ 
al sae ie 1 = SS 
° O02 O04 06 O8 10 
DISTANCE FROM CONTROL TRAILING EDGE 
LOCAL CONTROL SURFACE CHORD 


FIGURE 19. Variation of model velocity defect with 
distance from the control surface. 


if this is used for the predictions the above 
differences become + 1 percent and - 2 percent 
respectively. Thus the speed trial and full-scale 
wake data become compatible and both suggest a 
scale effect on the flow velocity for the vehicle 
A with the finer stern much smaller than predicted. 
This may be due to the fact that the full-scale 
vehicle is hydraulically rough at all but the very 
lowest speeds while the prediction method assumes 
hydraulically smooth conditions. 

The process of adding the predicted mean circum- 
ferential velocity changes to all measured veloci- 
ties are described in Section 4 naturally leads to 
a change in the ratios used herein to describe the 
relative magnitudes of the velocity defect and 
secondary flow. For the velocity defect Figures 16 
and 18 show that the predicted magnitude decreases 
slightly with increasing diffusion ratio such that 
at model self-propulsion the relative magnitudes 
are 3 percent higher than measured for body A and 
up to 6 percent for body B. The predicted relative 
magnitude of the wake defect at the full scale 
condition is within 2 percent of that measured, 
although as already noted the absolute velocities 
are 15 percent and 9 percent higher than measured. 
The use of a smaller scaling effect based on the 
equivalent Reynolds number discussed above would 
slightly reduce the above error in predicted 
velocity defect. 

The predicted relative magnitude of the secondary 
flow can be seen from Figures 16 and 17 to decrease 
with increasing diffusion ratio but at a slower 
rate than actually measured on the models. Thus, 
the propeller is having an influence on the develop- 
ment of the secondary flow in addition to the simple 
change in relative magnitude arising from the 
propeller induced velocity. It is clear that at 
model conditions, the difference between the 
measured and predicted secondary flow is much 
greater for the blunter afterbody form of vehicle 


168 


B. At model self-propulsion conditions these 
differences are up to 5 percent for vehicle A but 
60 percent for vehicle B. The secondary flow pre- 
diction for the full-scale conditions on body A 
given in Figure 16 can again be seen to be higher 
than the measured values but only by up to 4 percent 
at the two radii considered. In this case the use 
of a smaller scaling effect would lead to higher 
predicted values such that the differences between 
these and the measured values would increase to the 
order of 6 percent. 

The above results show that the agreement between 
the measured and predicted data has been limited and 
further work is required before the proposed scaling 
method can be regarded as satisfactory. The princi- 
pal requirement is for further high Reynolds number 
data and it is proposed to obtain this by additional 
full-scale trials, together with experiments on 
models in a compressed air wind tunnel. 


6. CONCLUSIONS 


An integral boundary-layer calculation method for 
bodies of revolution is shown to give a good pre- 
diction of boundary layer velocity profile for 
attached flows in the tail region of a body. 

Inclusion of a simple actuator disc representa- 
tion of a propeller in the calculation method gives 
a reasonable first approximation to the effect of a 
propeller on the flow. 

Comparison between results from Fourier analyses 
of measurements from runs repeated a number of 
times and of measurements made with different 
incremental steps in probe position indicates that 
wake harmonics can be determined reliably. from 
measurements at model scale. 

Fourier analyses of idealised velocity profiles 
representing wake defects in an otherwise uniform 
flow field have been obtained analytically. Com- 
parison between these results and numerical harmonic 
analyses of the same profile specified at a dis- 
crete number of points shows no significant differ- 
ences in the amplitudes of wake harmonics at high 
harmonic number provided that the width of the wake 
is not too small. 

The measurements presented herein indicate that 
the velocity defect produced behind a control sur- 
face is only slightly affected by either the 
presence of a propeller aft of the control surface, 
or by the change in Reynolds number from model to 
full-scale. 

Near the hull, where the flow is influenced by 
secondary flow effects, the velocity defect behind 
a control surface is much smaller than predicted 
from two-dimensional data. For positions outside 
the influence of the secondary flow the velocity 
defect approaches the two-dimensional value. 

The velocity defect is of a similar order of 
magnitude for the two bodies examined. However, the 
secondary flow effects are significantly larger 
for the vehicle with the blunter stern. 

The secondary flow produced by the interaction 
of a control surface with the hull boundary layer 
is reduced significantly by the presence of a 
propeller aft of the control surface, and from 
model to full-scale conditions. This reduction 


increases with increasing propeller diffusion ratio. 

By using the unpowered model measurements as 
datum it has been possible to predict the model 
powered mean circumferential velocities to within 
4 percent for radial positions from the hull greater 
than 12.5 percent of the propeller radius. At this 
radius itself, the predictions are within 7 percent 
for the finer stern model and 14 percent for the 
fuller stern; however, the latter may be due to 
separation effects which are not taken into account 
in the prediction method. 

Predictions of the mean circumferential velocity 
at the full-scale conditions for the vehicle with 
the finer stern are high by up to 15 percent. If 
the ship prediction is made at a reduced Reynolds 
number suggested by speed trial results the pre- 
dictions come within 2 percent. Predictions of the 
powered velocity defect are wit in © percent for 
model conditions and 2 percent for ship conditions, 
the latter figure applying to either the true or 
reduced full-scale Reynolds number. Predictions of 
the model powered secondary flow are within 5 per-— 
cent for the body with the finer stern, but up to 
60 percent for the fuller form. However, for the 
full-scale conditions obtained on the finer stern 
the predictions are within 4 percent at the true 
Reynolds number, and 6 percent at the reduced value. 

A practical method of estimating propeller in- 
duction and wake scaling effects has been proposed 
and demonstrated to give limited agreement with 
model and full-scale data. Further experimental 
data are required to refine the method and to this 
end high Reynolds number model experiments are 
planned to be carried out in a compressed air wind 
tunnel, and further full-scale trials scheduled. 


REFERENCES 


Head, M. R. (1960). Entrainment in the turbulent 
boundary layer. British ARC, R & M 3152. 

Huang, DT. T., S. Santelia, HH. LT. Wang, and IN-vICE 
Groves. (1976). Propeller/stern/boundary-layer 
interaction on axisymmetric bodies: theory and 
experiment. DINSRDC Rep 76-0113. 

Luxton, R. E., and A. D. Young. (1962). Generalised 
methods for the calculation of the laminar com- 
pressible boundary layer characteristics with 
heat transfer and non-uniform pressure distri- 
bution. British ARC, R & M 3233. 

Manley, R. G. (1945). Waveform analysis. Chapman 
and Hall, London. 

Myring, D. F. (1973). The profile drag of bodies 
of revolution in subsonic axisymmetric flow. 
RAE TR 72234. (Unpublished) . 

Patel, V. C. (1974). A simple integral method for 
the calculation of thick axisymmetric turbulent 
boundary layers. Aeronautical Quarterly, 

NG Iie ks 

Patel, V. C., A. Nakayama, and R. Damian. (1973). 
An experimental study of the thick turbulent 
boundary layer near the tail of a body of 
revolution. TIowa Institute of Hydraulic 
Research Report, 142. 

Raj, R.-, and B. Lakshminarayana. (1973). Charac— 
terisics of the wake behind a cascade of airfoils. 
J. Fluid Mechs, 61, Pt. 4. 


Experimental and Theoretical Investigation 
of Ship Boundary Layer and Wake 


Shuji Hatano, 


Kazuhiro Mori and Takio Hotta 


Hiroshima University, Hiroshima, Japan 


ABSTRACT 


Characteristics of the boundary layer and wake flow 
of ships are investigated experimentally and at- 
tempts are made to estimate their velocity distri- 
butions. 

Boundary layer characteristics, before the onset 
of separation, are studied; a three-dimensional 
boundary layer calculation is carried out by the 
integral method, while examining the boundary layer 
assumptions and the validity of auxiliary equations 
by direct measurements of velocity and static pres- 
sure profiles in boundary layer as well as skin 
friction distribution on hull surface. 

Assuming that the wake is the domain of influ- 
ence of the boundary layer and consists of three 
sub-regions, i.e., vorticity diffusion region, 
separated retarding region, and viscous sublayer, 
different governing equations for each sub-region 
are derived by local asymptotic expansions. 

Velocity distribution in the vorticity diffusion 
region is estimated in two steps: first, vorticity 
distribution is found by solving the vorticity 
diffusion equation, then velocity distribution is 
calculated from the obtained vorticity distribution 
by invoking Biot-Savart's law. 

Satisfactory agreements are attained between 
calculations and measurements both for boundary 
layer and wake. 


1. INTRODUCTION 
Introductory Remarks 


The prediction of the viscous flow field around 
ship hulls, boundary layer on the hull surface, and 
the wake, is one of the most important problems in 
ship hydrodynamics. Important design-conditions, 
such as estimations of viscous resistance or wake 
distribution on a propeller disk, are all closely 
connected with this problem. Instabilities of ship 


169 


Maneuvering and propeller-excited-vibrations are 
also presently urgent problems in practice; they 
are also fundamentally connected with the viscous 

Calculations of a ship boundary layer have been 
carried out by many investigators during the last de- 
cade; e.g., Uberoi (1969), Gadd (1970), Webster and 
Huang (1970), Hatano et al. (1971), Himeno and Tanaka 
(1973), and Larsson (1975). They have solved bound- 
ary layer equations in integral forms. Cebeci et al. 
(1975), as well as Soejima and Yamazaki (1978), has 
tried to solve them by the finite-difference method. 

Such remarkable progress in ship boundary layer 
calculations are mainly due to studies of two- 
dimensional boundary layers and to the use of high 
speed computers. Though some of them yield good 
results, an absence of experimental examination of 
boundary layer assumptions or auxiliary equations 
can be found when applying them to shiplike bodies. 
Experimental examinations are very important because 
most of auxiliary equations are derived from two- 
dimensional experiments. 

On the other hand, as to the ship wake, many 
experimental studies have been carried out not only 
for ship models but also for full scale ships, e.g., 
Yokoo et al., (1971) and Hoekstra, (1975) mainly 
discussed the prediction of full scale wake charac-— 
teristics based on model wake survey. 

Rational theoretical studies are still more im- 
portant. As to theoretical studies of wake, we 
must retreat to problems of flow behind rather 
simple obstacles like flat-plates, circular cylin- 
ders, or bodies of revolution. Even in such cases, 
most treatments are based on potential theory such 
as free-streamline theory or cavity-flow theory, 
reviewed by Wu, (1972). However, because vortici- 
ties existing within wakes are mainly generated in 
boundary layers of hull surfaces and shed into wakes 
viscously and convectively through separations, the 
prediction of wake flow should be treated in close 
relation to boundary layer flow. 

The previous works by Hatano et al., 


(1975, UIT) 5 


170 


were carried out from this standpoint. But they 
are only the beginning of research on ship wakes 
and many future problems were pointed out, especially 
requirements for further experimental studies. 

The present authors are firmly convinced that, 
for such viscous flow problems, marriages of experi- 
mental and theoretical studies are primarily impor- 
tant in order to make further progress. Because 
of this, the present paper is divided into two parts; 
experimental studies on ship boundary layers and 
wakes (Section 2 and 4), and theoretical studies 
and numerical calculations (Section 3 and 5). 


Coordinate Systems and Models Used 


Two coordinate systems are employed throughout the 
present paper. One is the right-hand linear coordi- 
nate system, O-xyz, whose origin is at midship and 
on the waterplane and the oncoming flow, Ug, is in 
the x-direction. The other is the streamline coor- 
dinate, x]x2x3; the curves of constant x2 coincide 
with potential flow streamlines on hull surface and 
x3 is normal direction to hull surface (Figure 1). 

All quantities are dimensionless by half ship 
length 2 (=L/2), ship speed Ug, and fluid density p, 
unless specified in another form. 

For the present research three ship models, 
GBT-125, GBT-30, and MS-02 were used whose body 
plans with potential streamlines and principal di- 
mensions are shown in Figure 2 and Table l. 

GBT-125 and GBT-30 are practical tanker ship 
models, similar in geometry to each other. GBT-125 
is a double model and was used under submerged con- 
ditions for studies of boundary layer flow. MS-0O2, 
which was used for the studies of wake flow, has a 
rather simple stern form; the framelines are ellip- 
tic and given by the equation, 


CHlOlae 


0.7 
xa 
Yo = bo | 1- (=) 1 (2) 


and bg is the half breadth of the waterplane at 
x = 0.4 (S.S.3) and d is the draft. The remainder, 
(x < 0.4), has a practical hull form. This is be- 
cause the practical stern form produces a very com- 
plicated stern flow, e.g., an intensive longitudinal 
vortex, not suitable for the present investigations. 
Experiments were carried out in the circulating 
water channel and the towing tank of Hiroshima 
University. 


x >a (a=0.4) (1) 


Coordinate systems. 


TABLE 1 PRINCIPAL DIMENSIONS OF MODELS 
GBT-125 GBT-30 MS-02 
1.250/™) 3.000'") 3.000!™) 
.193 -462 -485 
065 o Si 165 
Ch .836 .836 .768 
NOTATION 
L, ship model length and half length 


Q 
b ship model breadth 
Cy, block coefficient of the ship model 
d ship model draft 
p density 
v kinematic coefficient of viscosity 
Ve eddy viscosity coefficient 
g gravity acceleration 
Up velocity of oncoming flow, ship speed 
F, Froude number =U0/VgL 
Re Reynolds number =UoL/v 
€ small parameter for asymptotic ex- 
pansions = Rg 8 
X,Y,Z orthogonal linear coordinates 
X],*%2,%3 orthogonal curvilinear coordinates 


E,n,G distances along X],X2,x3 coordinates 


hj,ho,h3 corresponding metric coefficients 
K,,K2 convergences defined by Kj = 
Sel ohoney, eeu il ohy 
hyh2dx] 2 hyho 9x9 


normalized distances for vorticity 
*) diffusion region, separated re- 
tarding region, and viscous sub- 
layer respectively 
q velocity vector 
Gy, viscous part of velocity vector 
u,v,w velocity components in x,y,z direc-— 
tions excluding uniform flow 
41/92/43 mean velocity components in x) ,x2,xX3 
: directions 
fluctuating velocity components in 
X1],X2,*3 directions 
U, resultant velocity at boundary layer 
edge 
velocity components at boundary layer 
edge in x)],X9,x3 directions 


UTITY LUTY 2 
~ 

SIwsaidr 
x 

WX OUT? 


' ' ' 
G1192793 


U,,V1 Wy 


RorirYirWi| asymptotic terms of normalized mean 


CLOSE velocity for vorticity diffusion 
Cae ) region, separated retarding region, 
IL Aoo oo 


and viscous sublayer region 

af, 07 Wi } asymptotic terms of normalized fluc- 

(Gil pAeb oo) tuating velocity for separated re- 
tarding region 


wW vorticity vector 


Wye 1 Wy 1 We vorticity components in x,y,z di- 
rections 

W1,W2,W3 vorticity components in X],X2,X3 
directions 


asymptotic terms of normalized vor- 
ticity for vorticity 

diffusion region 

p pressure 


STREAMLINE NO. 1 


GBT-125 
& GBI-30 3 


STREAMLINE NO.1 


MS -02 | 
ee ae 


Wes 
@ 
99 (@=11 92, 000) 
6 

* * 
617697911 
991,912,922 


H 


pressure far upstream 
pressure coefficient =(p-py)/z pu2 


al, 


0 
asymptotic terms of normalized 


pressure 

boundary layer thickness 
three-dimensional boundary layer 
thickness parameters defined by 

Eqs. (5)),) (19) 

shape factor of streamwise velocity 
profile =6}/6) 

angle between surface streamline 

and external streamline direction, 
positive in x2 direction 

index for power-law velocity profile 
parameter for wake part of wall-wake 
law in q}, q2 components 
coefficients of wall-wake law 

wall and wake functions of wall- 
wake law defined by Eq. (10) 
resultant skin friction 

components of skin friction in x} 
and x2 directions 

friction velocity 

entrainment function 

parameter for separation 

positions of onset of separation 

and reattachment 

integral region for induced velocity 
gradient vector £ 
symbol of orders f=O(€); lim ¢ =M 
(M:constant) eo) 


2. EXPERIMENTAL STUDIES ON BOUNDARY LAYER 


Kinds of Experiments and Measuring Techniques 


In order to examine the boundary layer assumptions 
and the validity of semi-empirical equations in 
case of ship-like bodies, the following kinds of 


experiments were carried out [Hatano et al., 


(1978) ]. 


7 


FIGURE 2. Body plans and potential 
flow streamlines of models. 


Static pressure measurements on hull surface 


Static pressure holes of 0.6mm were arranged on the 
hull surface along streamlines and the static pres- 
sure was measured by towing ahead and astern. 


Static pressure measurements in boundary layer 


Static pressure in the boundary layer was measured 
by using a static pressure tube. It is 1.2mm in 
diameter with two 0.4mm $¢ holes on diametrically 
opposite sides. A traverser with a micrometer was 
used to move the probe normal to the hull surface. 
The preliminary experiments showed that the static 
pressure was free from incident flows whose attack 
angles were less than 20°. 


Velocity measurements in boundary layer 


A total head probe, made from hypodermic tubing of 
outside diameter 0.28mm and 2.7mm respectively, 

was mounted on the traverser. Total pressure was 
measured after locating flow directions by yawing 
the directionally-sensitive hot film probe. Using 
the measured static pressure, velocity was estimated 
and decomposed into streamwise and crossflow 
components. 


Local skin friction measurements 


Local skin friction on the hull surface was mea- 
sured directly by a floating-element type friction 
meter [Hotta, (1975)]. The floating element is 14mm 
in diameter with gaps of 0.05mm to the mounting case 
and balanced by electromagnetic force. 

All experiments described above were carried out 
using the GBT-125 under submerged conditions at a 
depth of about 6 times the draft of the model. The 
Reynolds number was kept constant at 10°. 


172 


0.4 


0.0 


0.2 


0.0 


FIGURE 3. Static pressure distribution on 
the hull surface (GBT-125). 


Experimental Examinations of Boundary Layer Assump- 
tions and Semi-Empirical Equations 


Boundary Layer Assumptions 


The usual first approximate calculations of the 
boundary layer were carried out under the assump- 
tion that the static pressure is constant across 
boundary layers and is equal to the inviscid flow 
pressure. These assumptions are open to experimental 
examination when the boundary layer thickness is 

not thin, especially in the case of ship-like bodies. 

Static pressure distributions on the hull sur- 
face along streamline Nos. 5, 7, and 11 are shown 
in Figure 3 with calculated potential flow pressures. 
Potential flow calculations were carried out by the 
well-known surface-source method [Hess and Smith, 
(1962) ] representing the hull by 254 x 2 small rec- 
tilinear panels. Static pressures while being 
towed onward are in good agreement with those calcu- 
lated, except near the stern, where pressure has 
not recovered and is slightly low. However, towing 
astern shows good agreements even near the stern. 
This means that displacement effects of the boundary 
layer are appreciable near the stern. 

Figure 4 shows static pressure profiles in the 
boundary layer . It was observed that pressure pro- 
files are almost constant across the boundary layer 
except for some positions where the pressure is mono- 
tonically increasing or decreasing in that normal 
direction. The tendencies of increments are signif- 
icant at S.S.1% or S.S.1% of streamline No. 11. This 
can be referred to the centrifugal force due to the 
small radii of curvature of the bilge keel. On the 
other hand, a decrease can be found:for all the 
streamlines at S.S.% or S.S.4, which may be the 
effect of separation. (As described later, flow 


*Static pressure on the hull surface does not agree with that 
of Figure 3. While the measurements whose results are shown 
in Figure 3 were carried out in the towing tank, those shown 
in Figure 4 were in the circulating water channel. The 
discrepancies are all due to this difference in experimental 
conditions; the cross section of the circulating water chan- 
nel is restricted to 1200mm 
estimated. 


‘ 820mm and pressure is under- 


) 
i 


CALCULATED BY POTENTIAL THEORY 
° MEASURED BY TOWING AHEAD 
+ DO. BY TOWING ASTERN 


6 


STREAMLINE NO. 5 


STREAMLINE NO. 7 


STREAMLINE WO. 11 


can be assumed to have separated near S.S.4.) There 
the concept of boundary layer itself should be dis- 
carded. 


It can be safely pointed out that the pressure- 


Simm) STREAMLINE NO. 5 


aw 


7) 
au 
ae 
e 
Hep 
ecoo 
o 
Pecce 
Pee eegaoee 0% 


— 
t 

L 
(ae 


-0.2 -0.2 -0.1 -0.1 -0.1 0.0 0.0 0.1 
e e 
(mm) STREAMLINE NO. 9 ° le eo) 
60 3 : 2 “4 of 
3 le o. e © 
q 8 : 3 © ) 
1 e 
SeSe) e ° 5 e 
. re es : ais ° 
e ° e e e 3 
e e 
e H ° 5 3 ei 
20 ii if ae 
x es dae SE ( ne) \G 
-0.2 -0.2 -0.1 -0.1 0.0 0.0 0 
(mm) e 
1 STREAMLINE NO. 11 : 416 
e Te ° e 
le e 1 e 1 
1 18 e ° WO eT 
wf = S517 : 2a as 
e e e e e 5 
e s e e 0 e 
tae et Omnia 
20 ! ij i ° e 
/ \ 
fee ee ( Ne 
0 St 
-0 -0.2 -0.2 -0.1 -0.1 0.0 0.0 0.1 
FIGURE 4. Static pressure profiles in the boundary 


layer (GBT-125). 


173 


constant assumption can be employed unless the radii If velocity profiles are represented by Eqs. (3) 
of curvature are not significantly small; the dis- and (4), the boundary layer thickness-parameters 
placement effects are important near the stern and 64, 6,1 and shape factor H are 


should be taken into account in higher order calcu- 


lations [Hatano and Hotta (1977)]. cot nls bag ae 
On = a ff (Uy q,)dt = rik Oley 
e 
Velocity Profiles 1 5 A 
—— - lie = 5 
Mi ip t CU = Cele Sempre 2 
In order to calculate the boundary layer equations e 
by an integral method it is convenient to represent es 
velocity profiles by analytical functions which in- m $1 — nt2 
clude several parameters. 11 n 
The most commonly used formulae are based on a 
1/n-power law and on a wall-wake law. The former and Eqs. (3) and (4) can be written in other forms, 
has a definite merit of simplicity. The latter, H-1 
developed by Coles (1956), has more freedom than = iS H-1 2 
70, = I ] (6) 
the 1/n-power law and can be expected to represent e 0, ,H(A+1) 
velocity profiles more exactly. H-1 
Mager's expression is well known as the three- Dr 
dimensional velocity profile model based on a 1/n- = Cc H-1 ] ~ & H-1 2 
? Cy AU, = reete\3 > Saree [el — ——] . 
power law, Mager (1951). He gave the streamwise ee a H (H+1) ony H (H+1) (7) 
and crossflow velocity profiles as 
> 4 1/n, 3 Tf 944 and 53 are integrated and 8 is determined 
en, a 6 (3) from measured velocity profiles, then velocity pro- 


files represented by Mager's model can be calculated 
17a 2 d with th 
ie Gah wt BG from Eqs. (6) and (7) and can be compared wi e 
92/0, ras ( 9) ) € ) i) measured profiles. 
Figure 5 shows the comparisons of them. It can 


where n is a variable parameter. be safely pointed out that Mager's model is employ- 
(mm) | | 
40 e 
1 | 
STREAMLINE NO. 5 [ 
30 
552 
20 j } 
10) 
q | H Ge/ 
| if Ue pS en — Ue 


40 


STREAMLINE NO. 9 


30 
2 
20 ; 
S.S.8 
10 | 
° 
0 i 
0.0 0.2 0.0 
=o ESTIMATED BY MAGER'S MODEL 
@ MEASURED 
30 
STREAMLINE NO.11 
S50” 


q2 
a= Ne FIGURE 5. Velocity profiles represented by 
0.6 0.6 0.6 0.4 1.0 -0.2 0.0 0.00.2 Mager's model. 


174 


[ 5.5.34 

i 

40 i STREAML INE 
: es 


FIGURE 6. Crossflow profiles in 
the boundary layer on the AFT hull wu, 
surface (GBT-125). 


able for the velocity profiles of ship-like bodies 
as far as streamwise components. 

Figure 6 shows crossflow profiles measured on 
aft parts of a model. As easily observed, there 
are some profiles which have reverse type (S-shaped) 
profiles. For most of remaining parts, the cross- 
flow angles are very small and do not show reverse 
type profiles. Because Eq. (4) has only one inflec- 
tion point, such S-shaped profiles can not be repre- 
sented by it. 

To represent even reverse crossflows, more gen- 
eral polynomial expressions are proposed [e.g., 
Eichelbrenner (1973), Okuno (1977)]. However, they 
require additional equations or boundary conditions 
and it is reported they do not always yield improve- 
ments [Okuno (1977)]. This is because the cross- 
flow does not always have such universal profiles 
near the stern. 

On the other hand, the three-dimensional veloc- 
ity profiles based on Coles' wall-wake law can be 
represented by 


q,/ ec Ce (8) 
a/v = Gr fyi hy sins + 958,05) (9) 
where 
say <u ) = * log, 5 ( =e ) +B, (10) 
eZ) = 5 [1-cos( me )] ; 
and 


Bs 2 
u_ = ( Te / p) . (11) 


fp eer STREAMLINE 
Sine. 


J+ J, are variable parameters, for wake parts, u 

is the friction velocity, and k, B are constants. 
£. given by Eq. (10) is called the wake function. 

Figure 7 shows the existence of such parts in case 


Ue 


of ship-like bodies also. Velocity profiles 
deviate from linearity when approaching the outer 
edge of the boundary layer. Velocity profiles, 
represented by Eqs. (8) and (9), are compared with 
measured profiles. Here parameters g, and Jo are 
determined by the condition that q equals U, and 
dy equals zero at the boundary and u, is determined 
by a least-squares fit to the measured profiles. 
The values of Clauser, 5.6 and 4.9, were used for 
1/K and B respectively. Good reproductions are 
examined except crossflow representations. 

As to crossflow profiles, the situation is not 
much improved from Mager's model; reverse crossflow 
observed in experiments can also not be represented 
by the wall-wake law. The finite-difference method 
may be a possible step toward representation of any 
type of velocity profiles. 


Local Skin Friction 


In the case of turbulent flow, most of the friction 
is due to the turbulence (Reynolds' stress). For 
this reason it is necessary to introduce additional 
equations to determine it in closed form. 

Ludwieg and Tillmann's semi-empirical equation 
for the skin friction [Ludwieg and Tillmann (1949) ] 
is most commonly used; it is 


u_9 0.268 
11 . 
t /pu2 = 0.123x1079-678H ( © —— (12) 
WwW) e Vv 


Because Eq. (12) is obtained from two-dimensional 
experiments, the validity should be examined when 
applied to three-dimensional flow. 

When Coles' wall-wake law is employed for the 
velocity profile, the skin friction can be deter- 


—— ESTIMATED BY 10 
COLES' WALL-WAKE LAW 


@ MEASURED 


STREAMLINE NO.5 


oe? 
10 Sol 1°? 00 voce ah 


STREAMLINE 
NO.9 


FIGURE 7. Velocity profiles represented by Cole's 
wall-wake law. 


mined from the friction velocity. But it should 
also be examined experimentally. 

In Figure 8, three kinds of experimental values 
of skin friction are compared along streamline 


Nos. 9, 11, and 18; directly measured values, those 


Tous 


+ 


STREAMLINE 
NO.9 


eS) 


DIRECTLY MEASURED 
ESTIMATED FROM COLES’ WALL-WAKE LAW 


175 


obtained from Ludwieg-Tillmann's formula, Eq. (12), 
and those from friction velocity, Eq. (11). For 
estimations of the latter two values, measured 


velocity profiles are invoked. Calculated results 
are also shown here for later discussions. 

The values of Ludwieg-Tillmann's formula produce 
fairly good agreements with those directly measured, 
which implies that Ludwieg-Tillmann's expression 
is also good for three-dimensional flow. 


Entrainment Equation 


In streamline coordinates, 
is given by 


the continuity equation 


3 3 aitabee 
dx, (the) + 3x, (12h1) 1 nyh7a 423 =| Oe (13) 


Integrating with respect to X51 from zero to 6, 
gives 


ack 
28 (G6) = 262 
hj, 9x} 1 hgdx2 
5 ee oky 1 9Ue 
=F - (6-67) {—— =2 + —— <=} (14) 
I 
hjh, 3x, Uh, 3x, 
where F is the entrainment function given by 
a a6 06 
b= Wee. v Vee 7 Wal s (15) 


Equation (14) is often used as the third (auxiliary) 
equation when the boundary layer calculation is 
carried out by the integral method. Here, F should 
also be given in someway in closed form. 

In two-dimensional flow, Eq. (14) is reduced to 


F= [U,, (6-63) ]- (16) 


ele adh 
We chai © 
Head (1960) gave a relationship between F and 
(6-67) /01 (=Hg_6%) which was examined by two- 
dimensional experiments. 


© 00. BY LUDWIEG-TILLMANN'S FORMULA 


2.0 88 


Comparisons of local 


3.0 
2.0 
=106 OR 
3.0 Mo lw —caLcuaTen 9S 
STREAMLINE NO.18 v 
° Vv ° 
v 
2.0 ° ° <I ° o % 3 3 8 v ° 
FP 9 8 Te 6 5 4 3 2 1 NLP 


riction (GBT-125). 


176 


Introducing an assumption that the entrainment 
equation of three-dimensional flow is related ex- 
clusively to the streamwise quantities, Cumpsty and 
Head (1967) employed the Head's entrainment function 
for three-dimensional boundary layer calculations. 
This is of course open to criticism. 

In Figure 9, Head's entrainment function and 
experimental values, obtained from Eq. (14), invok- 
ing measured velocity profiles, are compared. It 
can be mentioned that Head's function gives rather 
good mean lines both in relation to Hg-3, - H and 
F-H5-6] - The values of Hs-§1 are not fairly related 
to H in the fore part, where laminar flow may still 
exist, and neither F to Hg-6] in the aft part. The 
former does not seriously effect F. We should bear 
in mind here that the determination of boundary 
layer thicknesses is not clear in the three- 
dimensional case and accurate estimations of their 
derivatives are very difficult. 

Himeno and Tanaka (1973) used the moment of 
momentum equation as the third equation instead. 

In this case, assumptions for the Reynolds" stress 
are also required and significant improvements are 
not always found. 

Summarizing the above discussions it can be 
safely concluded that the integral method, where 
either Mager's model or the wall-wake law is used 
for velocity profile, Ludwieg-Tillmann's equation 
for skin friction, and entrainment equation for 
auxiliary equation, is expected to yield meaningful 
results. Moreover, it can be also pointed out that 
improvements can be attained when the second order 
approximation for the static pressure is taken into 
account near the stern. However, in the region 
where reverse crossflows or large crossflow angles 
exist, although the boundary layer assumption is 
not violated, the integral method is no more 
available. 


BOUNDARY LAYER CALCULATIONS 

According to the preceding conclusions, boundary 
layer calculations were carried out by the integral 
method and compared with experimental results. 
Basic Equations and Auxiliary Equations 


The integrated boundary layer equations are given 
in streamline coordinates by 


FIGURE 9. 
function with experiments. 


Comparison of Head's entrainment 


, 


0611 0842 On Oe 2 
DE + on + cali ea Oem K; (6)) 890) 
=e. f wu 4 (17) 


2021 Ue | 011 We 
0g on U, on Ue an 


(H+1+ 212) 
911 


- 2K)}89] =T Ue pu2 , (18) 


where 9)1, 8,2, 92), and 899 are momentum thickness 
parameters defined by 


2 yee - 
te Pils deh Wisenels. o 


6 
us 912 = f q2(U;-q))daz , 


2 0 = 
We Pan = 4 cin Whissiayiehs p 
2 6 
WE Opp = f q2(Vi-q2)dz . (19) 


The entrainment equation is employed as the third 
equation; 

aU 
al e 


Hol i 
Us 9& 


3E 0 (20) 


3 
(6-67) - noe = F - (6-6}) (-Kj+ 
For the function F, the relation of Head is used, 
which has already been examined. 
If Mager's velocity profiles are employed here, 


boundary layer thickness parameters are given using 
811, H and B, 


851 = 6),E(H)tan® , 695 = 06 1)C(H)tan28 , 
612 = 9110(H)tanB , 65 = 6),D(H)tanB , 
6-6] = 6,,N(H) , (21) 
where 
CH) = - Gaya 
D(H) 16H 


~ 3G) Ges) (Gas) 7 


] 
go eh 


STREAML INE 


° NO.5 
* N0.9 
vy NO.11 


estes by2it> afi 
av) = "Gen Ge) 2 
J(H) = E(H) - D(H) , 
Seoul 
MED = sey = Big ge (22) 


Then Eqs. (17), (18), and (20) are reduced to 
simultaneous differential equations in 36) ]/3&, 
dH/9§, and dR/9E; 


(i=1,2,3) (23) 


of OWNM og Chel 2,98 
dj = JtanB a J 811 tanBs— J8) sec Bon 
611 9Ue 2 
+ — ——" =_ 
(H uN dE + Kj) 6) (1-Ctan“s) 


2 
ap U 
wl Te @ 7% 


a2 = Etanf, b2 = E'8))tan8, co = E6,)sec“B, 


81, 9U 


ta ee ial Anjo ay 2,0H 
do gE ant DE Ctan Bo C'6)),tan Bon 
- 2C611 tangsec?B5= ~ (1+i+Ctan?B)K 0) 1 


2 
+ 
2K)E0),tanB + cee Hf puUctanB 1’ 


(24) 


a3 =N, b3 = N'O)1, c3 = 0, 


d3 = ptangee2 + DOqneenee «. Denmsec2 eo 
5 an on 
n 
1 dUe 
+ F - N6O))(-Ki+ — —, 
Eee NEHGS] a, OG 


(The ' means differentiation with respect to H.) 

If Ludwieg-Tillmann's skin friction formula, 
Eq. (12), is used, all the coefficients of Eq. (24) 
are known at earlier & coordinate. 

This formulation is the same as that of Cumpsty 
and Head (1967). 


Numerical Calculations and Discussions 


Numerical calculations were carried out for GBT-125 
at Re=10°. First, 18 streamlines were traced inter- 
polating the 254 x 2 descrete values of velocity, 
obtained by the surface source method, and x, coor- 
dinates were determined. 

The differentials with respect to nm were numeri- 
cally determined along the n axis which was defined 
by bending short segments orthogonally to the xj 
axis. This is the main difference from Cumpsty- 
Head's original calculations. For such calculations 
as 0Ue/dn, 39911/dn, and so on, the differentials 
with respect to n should be carried out as care- 
fully as possible. Most numerical errors stem from 
‘these terms. 


iby 7/ 


@u/e 102 50 
1.2 
CALCULATED ' 
1.0 e MEASURED oe ? 
3 
0.8 1.0 e e 


STREAMLINE NO.5S 


FIGURE 10. Comparisons of momentum thickness (GBT-125). 


0.5 x 107", 1.4, and 0.0 were used for the 
initial values of 8,,, H and 8 at S.S. 94(x=-0.85). 
These values were obtained from Buri's two- 
dimensional formula assuming the flow is turbulent 
just from F.P. (see Figure 1). Fortunately they 
do not seriously affect the calculations. 

About 200 steps were taken and Eq. (23) is 
integrated with respect to € by Lunge-Kutta-Gill's 
method (five points for each step). 

In Figures 10, 11, and 12, calculated results of 
611, H, and § along typical streamline Nos. 5, 9, 
and 11 are shown along with experimental results. 
The skin friction is shown in Figure 8. Streamline 
No. 5 generates a simple, quasi-two-dimensional 
curve on the hull surface and it may be expected 
the flow can be truly represented by the present 
framework. On the other hand streamline No. 11 
passes through a region where the boundary layer is 
rather thin and also through a bilge corner where 
pressure increments were observed. 

The experimental values of the streamwise momen- 
tum thickness, 6),,, of streamline No. 11 were much 
greater than those calculated around S.S.1. This 
discrepancy can be related to the fact that S.S.1 
of streamline No. 11 corresponds to the position 


2.0 
ef @ STREAMLINE NO. 5 
e 
1.5 SS = 
© oT Temas eaer eae e 


2.0 CALCULATED 
@ MEASURED 
Aun) NO.9 
1.5 OFT OSs es _— 
O e ° e ee 5708 
ee 
2.0 =106 
Re=10 
NO.11 
O a 
@ 
1.5 ee e 5 ° oie ‘ 
ee ° 


FIGURE 11. Comparisons of shape factor (GBT-125). 


178 


ee 
CALCULATED 
P (degree) @ MEASURED ° 
10 5 
O STREAMLINE NO.5 e 
e. e t 


3 STREAMLINE NO.9 
————— 


! ! 


a) Re=10° 


o 


STREAMLINE NO. 11 


FIGURE 12. Comparisons of crossflow angle (GBT-125). 


just behind the bilge keel and the occurrence of 
bilge separation can be suspected. 

Shape factor H, in every case, does not vary 
significantly and agreements between calculations 
and measurements are good except near the stern. 
There, as shown in comparisons of 8, large cross- 
flow angles existed and the present scheme can not 
be employed here. 

It is interesting that large crossflow angles 
can also be observed in experiments near the bow. 
They create a suspicion of the occurrence of bow- 
bilge separation. 

Skin friction Tw, Shows also good agreement. 

It is observed that both experimental and calculated 
values do not decrease. This suggests three- 
dimensional separation differs a little from that 

of two-dimensional where skin frictions vanish. 

As a whole, it can be safely concluded that, 
except near the stern, calculated results show good 
agreements with measured as far as integral quanti- 
ties like 6); or H. It can be also concluded that 
the present scheme, using integrated mementum bound- 
ary layer equations as governing equations, can be 
appreciated in spite of its brevity. 


EXPERIMENTAL STUDIES ON BOUNDARY LAYER SEPARATION 
AND WAKE 


Kinds of Experiments and Measuring Techniques 


The characteristics of separation and separated 
flow of ship-like bodies are dim. Experiments may 
throw light upon them. In order to discuss the 
characteristics of separation and separated flow, 
the following experiments were carried out in addi- 
tion to the previous experiments. All experiments 
were carried out with MS-02 and experiments (c) and 
(d) used GBT-30 also. Experiments were executed 

at the speeds of Fy=0.1525(Re=2.17x10°) and Fn=0.16 


(Rg=2.38*10°) for MS-20 and GBT-30 respectively. 


Flow Observations 


Planting twin tufts on the hull surface, flow di- 
rections near the stern were observed by a submerged 


camera; one tuft was just on hull surface and the 
other was 22mm off, normal from surface. 

Free-surface flow around the ship stern was also 
observed in relation to the separated flow by the 
aluminium powder method. 


Velocity Measurements in Separated Flow Region 


Velocity in the separated region was measured using 
a hot film anemometer. The probe is a conical type, 
2mm in diameter. One horizontal plane of z=-0.02 
was covered where framelines are almost vertical. 
Because the probe was set parallel to the uniform 
flow, the velocity is not quantitatively accurate. 


Velocity Measurements in Wake 


Two five-hole pitot tubes were used for velocity 
measurements in the wake; 8mm-diameter tube for MS-02 
and 10mm-diameter tube for GBT-30. For estimations 
of vorticity, measurements were carried out on three- 
dimensional lattice-points spaced 0.025, 0.015, and 
0.015 in x, y, and z directions respectively. 


Vorticity Estimations in Wake 


The vorticity can be estimated by differentiating 
the measured velocity distributions; 


ow av du ow av du 
‘Om = p = =—= HS a o 


x ay dz “y az teen eB 3x dy - 


(25) 


The differentials were obtained numerically by 
three-point approximation. 


Discussions on Boundary Layer Separation and Wake 
Flow 


Boundary Layer Flow near Separation 


Figure 13 shows flow directions near the stern of 
MS-02 obtained by the twin tufts method. 

It was observed that, very near A.P., both tufts 
are drooping. This means that the velocity is al- 
most dead; in other words, separation has occurred. 

On the remaining parts, the outer tufts show 
almost the same direction as the calculated poten- 
tial flow direction; on the other hand the inner 
tufts differ greatly from them and produce large 
crossflow angles. A reference to the surface pres- 
sure distribution gives a clear explanation that 
flow near the hull surface, whose velocity is very 
low, cannot make further steps against the pressure 
increments and change direction suddenly from the 
external streamwise direction toward the low- 
pressure regions. Significant occurrences of shear 
flow and generation of vortices are assumed which 
correspond to beginnings of three-dimensional sepa- 
ration. 

The above situation can be understood more 
clearly from velocity profiles in the boundary layer 
near separation. Figure 14 shows the velocity pro- 
files of GBT-125 along streamline Nos. 5, 9, and 
11. A sudden large crossflow occurs near S.S.% 
for all the streamlines and, correspondingly, the 
streamwise velocity profile also changes. The 


STREAMLINE NO.72 = —— 


Ne ate 


> 


NO.9 noaooe 
" OUTER TUFT 


Ae sen TUFT 


Slmm) trem ine 5 1] 
of STREAMLINE NO.5 


0 Ape ede 


0.4 0.4 0.4 0.4 


maximum crossflow velocity amounts to about half 
of the streamwise velocity. 

Such behaviors of flow near the stern are not in 
the category of boundary layer flow, therefore, 
boundary layer calculations should be stopped and 
another treatment employed. 


Criterion for Boundary Layer Separation 


It is necessary to introduce some criterion for 
boundary layer separation in order to change the 
governing equations from boundary layer to some 
others. 

There are many criteria mainly for two-dimensional 
separation [e.g., Chang (1970) ]. 

A parameter, I',, defined by 


Speed (26) 


is proposed. 


179 


FIGURE 13. Flow directions near 
stern and isobar lines (MS-O2). 
FIGURE 14. Velocity pro- 


files in boundary layer 
near the stern (GBT-125). 


The proposal is based on the experimental facts 
that the beginning of three-dimensional boundary 
layer separation is closely related to the pressure 
gradients, as discussed in the previous section, 
and that boundary layer flow, such as with large 
momentum thickness and small skin friction, can no 
longer exist. Therefore, flows with large values 
of IF cannot exist in real flow in the sense of 
boundary layer flow. On the other hand, if the 
boundary layer assumptions are kept, the calculated 
values of Ty can increase without any upper bound. 

Figure 15 shows I’, obtained by the boundary 
layer calculations and from experiments. The calcu- 
lated values get increasingly large approaching 
the stern, but experimental values do not and they 
seem to have some upper bound. 

The value of [, = 20 is reasonable as a criterion 
for separation, because, as shown in Figure 14, 
large crossflow angles were observed near x=0.9 
(S.S.4) and the onset of separation is suspected. 

Of course, more experimental data are necessary 
for the present discussion and further experimental 


180 


CALCULATED 


Oo — ESTIMATED 
FROM MEASURED DATA 


SEPARATION 


STREAMLINE NO.5 


SEPARATION 


STREAMLINE NO.9 


pe ae 


° 
SEPARATION 


10 
STREAMLINE NO.11 
0 
5.5.2 1s 1 a AP 
FIGURE 15. Criterion for separation. 


and theoretical studies may give a firmer founda- 
tion for the present criterion. 


Flow Field after Occurrence of Separation 


Once separation has occurred, the flow field differs 
greatly from the unseparated boundary layer flow. 
The existence of the dead region, pointed out in 

the previous section, is one phenomena. 

Figure 16 shows velocity profiles, after the 
occurrence of separation, measured by the hot film 
anemometer. The bars in the figure represent fluc- 
tuations in velocity. The region where the velocity 
fluctuates so intensively and is very low consists 
of a characteristic thin layer, a separated retard- 
ing region. It can be definitely distinguished 


3 (mm) 


100 —t 
INTENSIVE 


FLUCTUATION 
® 


1 
Soe > 


FIGURE 16. 


near the separation position 


Velocity profiles 


(MS-02) . 


from the outer part where the flow does not differ 
greatly from the unseparated flow, The newly— 
generated vortex is confined to this region. 

Figure 17 is free-surface flow of MS-O2. It 
shows more clearly the existence of the above 
mentioned, separated retarding region. The divid- 
ing streamline can be observed which coincides with 
the border line of the separated retarding region. 

In the case of practical ship forms, we have not 
enough information as to whether or not such regions 
exist. But from the velocity profiles of GBT-125 
(Figure 14), their existence can be supposed in 
those .cases also. 

According to the present experimental studies, 
it is implied that any single approximate equation 
of the Navier-Stokes equation completely governs 
the flow field near the stern. 


Eddy Viscosity Coefficient in Wake 


In order to predict turbulent terms in the Navier- 
Stokes equation, there is a concept of eddy viscos- 
ity. It is based on an idea that momentum loss due 
to turbulence can be represented by momentum loss 
due to friction and the coefficient is constant as 
to positions and directions. According to this 
assumption, the Navier-Stokes equation is written, 


ae 2 
q- Vu, - w.Vq = VAY Wr (27) 


where v, is the eddy viscosity coefficient. 

Equation (27) is a kind of diffusion equation 
with vg the diffusivity coefficient. It can be 
determined experimentally; substituting the measured 
values of velocity and vorticity into Eq. (27) 
leaves only Ve as an unknown. 

Using experimental data of the GBT-30, covering 
1.08<x<1-16, ve is determined by the least-square 
method. The estimated values of vg are not unique; 
they differ slightly for each direction, 2.7 x 101, 
2.4\x 107%, and|1.6 x 10>* fox w,, wy, ands.) the 
mean value is 2.2 x iO", and consequently the 
equivalent Reynolds number, based on the eddy vis- 
cosity, is about 1/300 of the real Reynolds number. 


e e @ 
e 

® @ ) 

@ 
5 e © 

e 
3 : ¢ 

5:8, 2 1 

8 S.S. SS mee 14 


FIGURE 17. Free-surface flow near the stern (MS-0O2). 


Subdivision of the Flow Field 


It has been made clear by experimental studies 

that the separated flow has at least two, quite dif- 
ferent viscous regions where no single approximate 
equation of Navier-Stokes equation seems to be 

valid for both. It can be proposed to subdivide 

the flow field near the stern into five regions as 
shown in Figure 18; potential flow region, boundary 
layer region, vorticity diffusion region, separated 
retarding region, and viscous sublayer region. 

Their characteristics are as follows. 

Potential flow region: 

The region where the viscous term can be wholly 
neglected and only displacement effects should be 
taken into account. 

Boundary layer region: 

The region where the boundary layer assumption 
is valid and the backward influence of separation 
can be neglected. 

Vorticity diffusion region: 

The region where the vorticity, which has been 
generated in the boundary layer, is diffused con- 
vectively and viscously. No vorticity is newly 
generated in this region. Because the dividing 
streamline is a kind of free-streamline, the pres- 
sure on it might be constant. 

Separated retarding region: 

The region where the velocity is very small 
and the turbulence is intensive. Because even a 
recirculating flow can be observed, the governing 
equation for this region should be an elliptic type. 

Viscous sublayer region: 

This is the very thin layer region which just 
adheres to the hull surface. The molecular viscos- 
ity is predominant and the velocity profile should 
satisfy the no-slip condition on the hull surface. 


CALCULATION OF VELOCITY DISTRIBUTIONS IN THE SHIP'S 
WAKE 


Approximation of Navier-Stokes Equation by Local 
Asymptotic Expansion 


In order to get appropriate approximations of the 
Navier-Stokes equation for each region, local asymp- 
totic expansions of relevant quantities are made, 
using small parameter e defined by 


3 Ht/8 
SS Bes (28) 


181 


The quantity e€, was first introduced by Stewartson 

(1969) and © << 1 in case of a large Reynolds number. 
If the x3 coordinate can be assumed to be 

linear, i.e., 


h3 = 1, (29) 


the continuity equation and Reynolds equations are 
written in streamline coordinates as follows. 


aq Clee) i 943 


I SKS = 0 30 
hj,dxj h2dx9 0x3 191 22 te) 
Sh Cenk 4. Gea, Sell. _ Oeil fees 
hy 0X] ate ho 9x9 139%, 20192 192 
eye C) Pp 12 a Ot O25 

hy) 0x, ( p q1 ) haox9 q1d2 


) 1 ' ' ' 1 1 
- ~— q3q] + 2Kyqjq5 + K) (q]2-q9") 


8x3 
WY) C) dW3 
+—J— “(ph = 
ho le 202) (31) 
Gal Bah 5, Gh en Ee Oe ae ieee 
iy On he Sea eee, Pe 
aa eae EO atid Wastes on 
re oe tee ES tp Cue 


2 2 
- Ky (q] - 42) + 2K) qa}q49 


re | Gra 7 Bee |e (32) 


eb eh So, BES os LE heh 
hy dx] ho ax9 8X3 3X3 


' J ' 1 ' ' 
= + K +K 
hy ox 4193 hy dX> 9293 19193 29293 
v a 
(hyw = = — (ams) 
hyhp | ax en gee Atos 
(33) 
A. POTENTIAL FLOW REGION 
B : BOUNDARY LAYER 
C ; VORTICITY DIFFUSION REGION 
D ; SEPARATED RETARDING REGION 
SEPARATION f : VISCOUS SUBLAYER 


POSITION 


SHIP HULL 


FIGURE 18. Subdivision of separated flow field near 
the stern. 


182 


where W], W2, W3 are the components of vorticity 
given by 


aq a 
yy Se 


hodx9 ax3 
ed pe ed 
oe 3x3 h) 0x, ‘ (34) 
ee eee q 
3 hy dx, hy dxXo 291 112- 


and they satisfy 
Vew = 0. (35) 


In the reduction of Eqs. (31, 32, 33) from the 
Navier-Stokes equation, conventional predictions 
for turbulent components are used; the velocity is 
assumed to consist of time-averaged terms and 
fluctuating terms. 

On the other hand, if the constant eddy viscosity 
can be assumed, the following equations are derived 
directly from Navier-Stokes equation; 


—— (y 2-wW2q)) + = (w1q3-0341) 
hgdx9 1 dX3 
it) Be a2 3 ey 303 
=v) Gr 2 yO. S ==) 
2 oxo ax3 hj) 0x) hodxo aX3 


9 w2 dW 
= ——— - - + ¥ 
Rep ote) orm ) Seaaeprey |e Se! 


eee = ))) ar eo -W392) 
hj) 0x] W29Q1-W192 aX 3 293-302 
i OS 34 wy 93 
= y Say ( —) 
e né os oats hodx2 hj) dx) dx3 


9 dW) dW. 
7 Tmo, SIC) SS ners ic OO 
(Ky - 2 (wyq3-w3qy) + (Ko- 24 ( ) 
Le tata le oe 2% HaOxoY was 


i pe 1 94 a ew dw 
yy eS re a i ee ara np Ls 
e E x2 = h2 aR) os che OnE x ma 


1 1 2 2 


a dw 
p= (K 4K = (ears 3 
axa MSAD) (Ky 1) Os 


= teonealSaS| 


hodx9 (38) 
where 
1 dh 1 9h, 
K Sao ak Sf en eee, 
1l h2 ox, / Ko9 aD Fis, (39) 
1 2 


Vorticity Diffusion Region (C-region) 


For the vorticity diffusion region, the constant 
eddy viscosity is assumed. 


Introducing non-dimensional curvilinear small 
line segments 3, dn, and dz, we represent the dif- 
ferentiations 


OAD (40) 


Here we assume the derivatives by new variables 
Ebay Gull O(a), alaGion 


nor 


= 5 41 
5 0(1) (41) 


= 0(1) , a 0(1) ' 
on 


Raat 


The origin of the new variables coincides with that 
of O-x)x 9x3 but ==0 corresponds to the position 
of separation. 

We tentatively assume that the asymptotic series 
for velocity and vorticity of C-region have the 
following forms; 


qi /Ug = tig (E02) + eu, (&,n,6) ar eau (Gann) ar G00 G 


epi) = mn Spns) 2 Erm (Ecee) 2 aac 5 


q3/Ug ee e7wy (E,n,c) ap eS (E,n,c) chiateheta , 

Ml = 2 a, (Esme) +, (E,neE) + 

Up/L Ee Ey aa Eo ms cee , 
Sa Aye cca Oe a Le a eect 

Taye = cE herteb) +S OL (Ene) ee 


a 
U/L & (43) 
All the quantities appearing in Eqs. (42) and (43) 
are assumed to be O(1). 

Moreover, we introduce non-dimensional variables 
k) ,k2,k11,k22 by 


IS, = WEP) op ES SR SS OK o 
koo = L*eKo9 p (44) 
whose orders are O(1) for all the regions. 


Substituting Eqs. (40), (42), and (43) into 
Eq. (30) and Eq. (35), we get as leading terms, 


55 
Oy (45) 
3€ 

du 3a 
=e (46) 
an dt 

and into Eqs. (36, 37, 38) we get, 

Ola eats LO (47) 
aA ni to? + aF ci to) O 4 
a M5 Ot Bye kg (48) 
€3 Uh dz2 Epanlel , 
2~. 
BOS ENS ROGUE A (49) 
e3) Uh Wae2m | OE, td 0” : 


In order for the viscous diffusion term to exist, 
Ve/UpL should be at least O(e3). 


We have obtained four equations, Eqs. (46-49), 
for three unknowns, U9 1 Wyys and Orly but it can be 
easily shown that one of them is not independent. 

Changing variables back into the original ones, 
we get, as the governing equations for C-region, 


r) C) 
hp dx W241) ote dxq (3a) = 0) 9 (50) 


a 
= es = ©. 5 (51) 


3 
a oxgccsa) > Ont (52) 


The terms of order 0(1/e2) are neglected in the 
above equations. 


Separated Retarding Region (D-region) 
Introducing normalized variables, E, nN, i for the 


separated retarding region in the same manner, the 
orders of differentiation are assumed, 


a _ a ) a Lee 
hy dx] L Ede hodxo L forehal 
a i @ 
ox3 Le 30g ’ =) 
3 a 3 
a O(a) 5 oa O1) ae = OG) (54) 


Velocity and pressure are assumed to be expanded 
asymptotically, 


q1/Ug = € (0, +0} ) + ©? (ap +09) woo 0 ’ 
qo/Up = E(vitvi) + e2 (oto) +... , (55) 
= een ee We nO 
q3/Up = €> (wi tw)) + e*(Wotwo) +... , 
2 
(p-p,,)/PU, = €P] + Epa + --- , (56) 
where Uy, Usr-++ are all time-averaged variables 
Ag Lad 


and u,, u,,..-. are fluctuating. Here the fluctu- 
ating terms of pressure are omitted because they 
do not appear in the basic equations. 

The vorticity can be also expanded asymptotically, 


ol iL OM ah ON 
SS ae er a OCI) ig 
Ug/L Se he c o 
w 3u au 
ZUG ay OL 2s ore 


U/L €2 ar € 3¢ 


Ug 9 9n (57) 


Under these assumptions, the leading terms of the 
continuity equation are written, 


Ce foN Oo (58) 
9e an ac 


“and the governing equations are 


183 


{a8 a 8 » SB Bae Ac Move tee ae 
ar & pases a AWitre oF Vilage i + Wars ar Wome) Uy 


ao m2 
-koujvj+ mi] 


) at 3 U2 Uist 
_ test 20,05) + a (alve + 09V}) 
O patar aint ann a2 12 
? eben oP u2W}) -2k9ujV} = ky (ay Savalt ) 
u 
Gages 7 (59) 

, ov a Be ng OY 
Wilma o0 Wlme oF WA 

3 ae ap 


8) ae Oe ie SP). 808 
+e aa) + (iipe Wiis! Var Weis © eae 


+kouy2- eat 


1 a 1 @ pSoan fe) 0%) Oo ,.148 
SS Sse | een) oP Sela”) oP Satan) 
é On E an ae 


E 
Oane nen AGP ao ssi 
+ eae + vow]) + k, (a) Na ) - 2k,0,¥v | 
pee a2¥1 + O(e%) (60) 
e6 UpgL a2 ‘ 
3 3 
ete 5 OC) =O, (61) 
lrg dt 


The leading terms of Eqs. (59, 60, 61) yield 


p1(&,n,S) = const. (62) 


Equation (62) means that the pressure is constant 
throughout the present region as far as O(c) is 
concerned. 

Now the second terms of Eqs. (58, 59, 60) yield 


Oe oO BS Big, a Bom 
19E 195 38 aE P2 1 
a ,n0n8 oO ,nias 
= SH) = salen) op (63) 
3 87 
SN ee 
W152 + wigs + Bie = ae t1V¥1) 
a on 2 ) SOO 
SSO a Wi) = ae) (64) 
ee Va ea) 
C) 
a EHC (65) 
Cc 


In the above equations, the molecular viscosity 


184 


disappears but its effects are still existing in- 
directly through turbulence. 

Equation (65) gives the so-called boundary layer 
approximation. But because cross terms of fluctu- 
ating components exist in Eqs. (63) and (64), they 
do not always yield the same type as boundary layer 
equations which can not predict the recirculating 
flow observed in experiments. 


Viscous Sublayer (E-region) 


In the viscous sublayer, E-region, the no-slip con- 
dition must be satisfied on the hull surface. Here 
the intensity of turbulence may be very small and 
all the turbulent terms in the Reynolds equations 
vanish infinitesimally. 

The following asymptotic expansions are assumed 
from the Blasius solution. 


2 


* * 
CHYAUO = SWYy “EWA sP cco py 


2 


q2/Up = evi a vi ut ROSE (66) 


5 


q3/Ug = etwt a? & wh Ti ooo p 


where the orders of each term are all O(1). 
The derivatives are represented by 


Cheb te riliert a. 8 
web Ps SE inte Fy Goa 2 
Gs Bale a8) 
axg | betoe 7 sou 
and their orders are 

3 a a 
—— = (0) p == 0 1 = 5 
dE (1) a5 (1) ac 0(1) (68) 


Substituting the above assumptions into Eqs. 
(31, 32, 33), the leading terms are obtained as 
follows in original variables; 


+ + g3—— 
Why 8x] qh Fax 139 x 
1 98P aq) aS 
p hy, dx, ax2 uf ( ) 
aq 945 Tey) 
Wnjax; hoax, * dx, 
so ase (70) 
P hodxo ax2 i 
DITO. 2 (71) 
dXx3 : 
The continuity equation is 
oq) 8a oq3 (5) 


+ + = 
h) 0x] hodx9 ax3 


Here the quantities of O(e%) are omitted. 


These are the boundary layer equations themselves. 


They must be matched with the solution for the D- 
region in quite the same manner as the conventional 
method of boundary layer calculation. 

The following matching conditions should be 


satisfied when governing equations are solved. 
(i) for upstream; 


Wii Se Sieh) 2 es Se 
Ess0 VOCE Ne) > Un, pee wien e) = ve, 
Teme) == 

Bee wiGomee) = (73) 


whe re Up, Vp, and wp are the velocity components 
in the boundary layer in the x),x2,x3 directions 
respectively. 

(ii) far from the hull surface; the solutions 
should be matched to the solution of the A-region, 
potential flow. 

(iii) between the C- and D-region; 


ug(E,n,0) =0, uw (En,0) 


lim a Gee 8 Oi ne Se 
Pro O1(Esn/o) - S5e Gig (Enc) |=, | . (74) 
miGeonpO) = Se Soak 5 
vi (E,n,0) = oe vil€n,c) , (75) 


wi(&sn,0) = 0, wo(E,n,0) 


lim} * ape 2 2 2 
Poco | W1 (Erno) tpewi (Erm 2) |z_o (76) 
(iv) between the D- and E-region; 
MEO ean) o vi 
a 8 A ee 
En) = pees neo) (78) 
wi(En,0) = 0 , wolE,n,0) = 
lim She ore ea Ae = By 
a Ego) 1S rs wi (Erb) | a6 (79) 


The governing equations for the D-region do not 
close. Some auxiliary equations are required, but 
this problem is left for future work. 


Numerical Calculations for the C-Region 


To solve the derived equations analytically is al- 
most impossible; this is because not only are the 
equations non-linear but also the hull surface, 
where the boundary conditions are prescribed, is 
very complicated in geometry. Instead, they must 
be solved numerically. But it may be still more 
difficult because the calculation should be carried 
out for all the regions at the same time in order 
to satisfy the matching conditions. However, this 
difficulty can be removed by an iteration method; 
the surface consisting of dividing streamlines (DSL) 
is given a priori in the beginning as the inter- 
mediate region between C- and D-regions where the 
matching is carried out. Of course the surface of 
DSL can be obtained finally as a solution of the 
flow field, but the assumption of DSL makes it pos- 
Sible to solve the governing equations in every 
region almost independently and it is expected that 


repeated iterations may bring forth a reasonable 
solution. 

The flow in the C-region can be determined by 
taking a new streamline coordinate system O-x)X5%34 
where the x,-axis coincides with DSL and the x,- 
axis is normal to the DSL surface. 

By the finite-difference scheme, Eqs. (48) 
(49) are transformed into tridiagonal linear equa- 
tions for k > 2; 


and 


wo(i,j,k-1) - 2€(i,j3,k)wo(i,k,k) + wo(i,j,k+t1) 
=Ag(i,3,kK) , (80) 
W3(2,3,k-1) = 2C(i,j3,k)w3(i,5,k) + w3(i,3,k+1) 
= A3(i,3,k) , (81) 


where wo(i,j,k) etc. denote those values at x)=xjj, 


SSH! and X31 
ae Ac? ny, 
C(i,j,k) =1 + vbe Gin (Loa pbs) p 
Agi(i;a 7k) = = Wo(i=1,5,k=1) + 
3 ai Ac? ena Sots 
Aids (GALS) phe) || 2b a Con (st ala pis) || Wy (simak 5) petal) 
VeAE 
Aaa) = = wis\(G—1 a), kK) 
: : Ac2 Z : : : 
2w3 (i-1,5,k) |1 - —— qj (i-1,3,k) | - w3(i-1,5,k+1), 
VeAE 


(82) 


and AE, At are short segments in the x), x3 directions. 
Equations (80) and (81) can be solved by the 
forward marching procedure if the velocity profile 
of q; is given at the separation position. Here the 
value of vorticity at k=l, on DSL, is made equal to 
that at k=2. 
Once the vorticity distributions are obtained 
throughout, the boundary layer and wake, say V, 


0.27 AssuMED 0.2 0.2 
DSL 
UV.W ' 
‘ a an 
0 See 
Ol ae ar) 
(Geena 
-0.2 - 
-0.4 2=-0,015 -0.4 -0.4 
-0.6 
0.2 
0 
-0.2 
o Z=-0.075 
-0.4 slave -0.4 LAS, HA 
-0.6 


2 -0.8 -0.8 


185 


velocity distributions can be calculated as induced 
velocity of vorticity by invoking Biot-Savart's law; 


(x )=V 1 ppp W_ Bh ‘ay'az! 
eh SOGOU * an Vv ie Y] ested 


' 
where w, is the mirror image of w' whose components 


are Wx, Wy, Wz and 
ee (Feo a Ge) ae (aa!) 
ee = (EEE )e => Wa) & (a9) = (84) 


Because Eq. (83) gives the viscous component of 
velocity, the potential component should be added 
to qy;. 

In the present calculation, DSL is determined 
from experiments for the first iteration; it con- 
sists of line segments, departing at x5=0.9 and 
reattaching at x;=1.1 (see Figure 18). The stream- 
wise velocity q,; in Eq. (82) is given by a quadratic 
function of ~ which is equal to U; at the outer edge 
and to 2/3 U; on DSL. The integral intervals for 
x and ¢ are 0.005 and 0.0025 respectively. 

In order to obtain the velocity distributions at 
x=1.025, the region covering from x=0.8 to x=1.4 
is integrated in Eq. (83). Here, 300-times molecular 
kinematic viscosity is used as Ve. 

The boundary layer and the potential flow calcu- 
lations are carried out in the same manner as in 
Section 3. 

In Figure 19, typical calculated results of the 
first iteration for MS-02 are shown compared with 
experiments. The ship speed is Fpy=0.1525 and the 
corresponding equivalent Reynolds number is about 
8700. Here the calculations for the D- and E-regions 
have not been carried out; therefore both regions 
are excluded from the vorticity-integrating region V. 

Satisfactory results are obtained, as far as 
C-region is concerned, especially in u and w. The 
velocity v is always underestimated, in other words, 
overestimated in the negative direction; this may 


uo 
V4 em = 
we —-—-- 
° 
Fp=0.1525 FIGURE 19. Velocity distribu- 
( R,#2.17x106 ) tions in wake at (1/8)L AFT from 
een A.P. (MS-O2) . 


Fy20.1525 


-0.05 ( R,#2.17x108 ) 


---- CALCULATED =} 
—— MEASURED 


MIDSHIP SECTION Pes 


FIGURE 20. Wake distribution at (1/8)L AFT from A.P. 
(MS-02) - 


be because in the present calculations the potential 
components are determined with no attention to dis- 
placement effects. 

Figure 20 shows the calculated wake distribution 
compared with measured. They do not always produce 
quantitative agreement with each other, but compli- 
mental uses of the present calculations with model- 
wake survey may offer a useful method for the 
prediction of full scale wake characteristics. 

It is expected that much further improvement can 
be attained by taking into account the D- and E- 
regions. 


CONCLUSION 


The flow characteristics of boundary layers and 
wakes of ship-like bodies are discussed. The fol- 
lowing remarks can be mentioned as conclusions; 

(i) The pressure-constant assumption of boundary 
layer is a good approximation except near the 
ship stern or bilge keel where there is a 
small radius of curvature. The pressure does 
not recover near the stern because of the dis- 
placement effects of the boundary layer. 

(11) Most commonly used semi-empirical equations 
for velocity profiles, skin friction, and 
entrainment can be safely employed in case of 
ship-like bodies, but the functional expres- 
sion for crossflow in boundary layer has a 
certain limit for large or reverse crossflows. 

(111) The integral method of boundary layer calcu- 
lation may be carried out more effectively by 
a hybrid use of integral and finite-difference 
methods. 

(iv) The three-dimensional boundary layer separa- 
tion is closely related to pressure distribu- 
tion on the hull surface. Its initiation is 
referred to the occurrence of large crossflow. 

(v) The eddy viscosity coefficient is about 300- 
times the molecular one, in the ship's wake. 

(vi) The separated flow region has sub-regions 
which have different characteristics and no 
single approximate equation of Navier-Stokes 
equation is valid uniformly for all regions. 

(vii) The local asymptotic expansion method is 
promising for the separated flow. Further 
experimental investigations as to turbulence 
are necessary. 


ACKNOWLEDGMENT 


The assistance of graduate students of the Faculty 
of Engineering of Hiroshima University, who partici- 
pated in carrying out experiments and numerical 
calculations, is cordially appreciated. 


REFERENCE 


Cebeci, T., K. Kaups, and J. Ramsey (1975). Calcu- 
lations of Three-Dimensional Boundary Layer on 
Ship Hulls. Proc. of First Intern. Conf. on 
Numerical Ship Hydro., 409. 

Chang, P. K. (1970). Separation of Flow, Pergamon 
Press) Lede Oxford, pp.) 139) 

Coles, D. (1956). The law of the wake in the tur- 
bulent boundary layer, J. of Fluid Mech. 1, 191. 

Cumpsty, N. A., and M. R. Head (1967). The Calcula- 
tion of Three-Dimensional Turbulent Boundary 
Layers, Part 1: Flow over the Rear of an In- 
finite Swept Wing. Aero nautical Quart. 18, 55. 

Eichelbrenner, E. A. (1973). Three-Dimensional 
Boundary Layers. Annual Review of Fluid Mech. 5, 
M. Van Dyke, and W. G. Vincenti, ed., pp. 339-360. 

Gadd, G. E. (1970). The Approximate Calculation of 
Turbulent Boundary Layer Development on Ship 
Hulls. Trans. RINA 113, 59. 

Hatano, S., M. Nakato, T. Hotta, and S. Matsui 
(1971). Calculation of Ship Frictional Resis- 
tance by Three-Dimensional Boundary Layer Theory. 
J. of Soc. of Nav. Arch. of Japan 130, 1. 
(Selected Papers from J. of Soc. of Nav. Arch. 
dik n- WA) 

Hatano, S., K. Mori, M. Fukushima, and R. Yamazaki 
(1975). Calculation of Velocity Distributions 
in Ship Wake. J. of Soc. of Nav. Arch. of Japan 
138), 54; Hatano, S=, K. Mori, and T > Suzuka! 
(LOM V2ndsRepoxt, Do, 415) LOE 

Hatano, S., and T. Hotta (1977). A Second Order 
Calculation of Three-Dimensional Turbulent Bound- 
ary Layer. Naval Architecture and Ocean Engi- 
neering 15, 1. 

Hatano, S., K. Mori, and T. Hotta (1978). Experi- 
ments of Ship Boundary Layer Flows and Considera- 
tions on Assumptions in Boundary Layer Calculation. 
Trans. of The West-Japan Soc. of Nav. Arch. 56 
(to be published) . 

Head, M. R. (1958). 
Boundary Layer. 
and Mem. 3152. 

Hess, J. L., and A. M. O. Smith (1962). Calcula- 
tion of Non-Lifting Potential Flow about Arbi- 
trary Three-Dimensional Bodies. Rep. No. E. S. 
40622 Douglas Aircraft Co., Inc.. 

Himeno, Y., and I. Tanaka (1973). An Integral 
Method for Predicting Behaviors of Three- 
Dimensional Turbulent Boundary Layers on Ship 
Surfaces. J. of The Kansai Soc. of Nav. Arch., 
Japan 147, 61. 

Hoekstra, M. (1975). Prediction of Full Scale Wake 
Characteristics Based on Model Wake Survey. 
Intern. Shipbuilding Progress 22, 204. 

Hotta, T. (1975). A New Skin Friction Meter of 
Floating-Element Type and the Measurements of 
Local Shear Stress. J. of the Soc. of Nav. Arch. 
of Japan 138, 74. 

Larsson, L. (1975). A Calculation Method for Three- 
Dimensional Turbulent Boundary Layers on Shiplike 
Bodies. Proc. of First Intern. Conf. on Numerical 
Ship Hydro., 385. 


Entrainment in the Turbulent 
Aeronautical Res. Council Rep. 


Ludwieg, H., and W. Tillmann (1949). Untersuchungen 
uber die Wandschubspannung in Turbulenten 
Reibungsschichten. Ing. Arch. 17, 288. 

Mager, A. (1952). Generalization of boundary layer 
momentum integral equations to three-dimensional 
flows including those of rotating systems. NACA 
Rep. 1067. 

Okuno, T. (1977). Distribution of Wall Shear Stress 
and Cross Flow in Three-Dimensional Turbulent 
Boundary Layer on Ship Hull. Nav. Arch. and 
Ocean Eng. 15, 10. 

Stewartson, K. (1969). On the flow near the trail- 


187 


ing edge of a flat plate II. Mathematika 16, 
106. 

Uberoi, S. B. S. (1969). Viscous Resistance of 
Ships and Ship Models. Hydro-og Aerodynamisk 
Laboratorium Rep. Hy-13. 

Webster, W. C., and T. T. Huang (1970). Study of 
the Boundary Layer on Ship Forms. Jour. of 
Ship Res. 14. 153. 

Yokoo, K., H. Takahashi, M. Nakato, Y. Yamazaki, 
H. Tanaka, and T. Ueda (1971). Comparison of 
Wake Distributions Between Ship and Models. J. 
of Soc. of Nav. Arch. of Japan 130, 41. (Selected 
Papers from J. of Soc. of Nav. Arch. 11, 25.) 


A General Method for Calculating 
Three-Dimensional Laminar and 
Turbulent Boundary Layers on Ship Hulls 


Tuncer Cebeci, 


Ke Gree Chiang, 


and Kalle Kaups 


Douglas Aircraft Company, 
Long Beach, California 


ABSTRACT 


A general method for representing the flow proper- 
ties in the three-dimensional boundary layers around 
ship hulls of arbitrary shape is described. 
use of an efficient two-point finite-difference 
scheme to solve the boundary-layer equations and in- 
cludes an algebraic eddy-viscosity représentation 

of the Reynolds-stress tensor. The numerical method 
contains novel and desirable features and allows the 
calculation of flows in which the circumferential 
velocity component contains regions of flow reversal 
across the boundary layer. The inviscid pressure 
distribution is determined with the Douglas-Neumann 
method which, if necessary, can conveniently allow 
for the boundary-layer displacement surface. To 
allow its application to ships, and particularly to 
those with double-elliptic and flat-bottomed hulls, 
a nonorthogonal coordinate system has been developed 
and is shown to be economical, precise, and compara-— 
tively easy to use. Present calculations relate to 
zero Froude number but they can readily be extended 
to include the effects of a water wave and the local 
regions of flow separation which may stem from bul- 
bous-bow geometries. 


1. INTRODUCTION 


A general method for determining the local flow 
properties and the overall drag on ship hulls is 
very desirable and particularly so with the present 
need to conserve energy resources. Et is difficult 
to achieve for a number of reasons including the 
turbulent nature of the three-dimensional boundary 
layer, the complexity and wide range of geometrical 
configurations employed, the possibility of local 
regions of separated flow, and the existence of the 
free surface. In addition, and although these dif- 
ficulties may be overcome in total or in part, the 
resulting calculation method must have the essential 
features of generality, efficiency and accuracy. 


It makes 


188 


The purpose of this paper is to describe a general 
method which is capable of representing the flow 
properties in the boundary layer around ship hulls 
of arbitrary shape. It is based on the general 
method of Cebeci, Kaups, and Ramsey (1977), developed 
for calculating three-dimensional, compressible lami- 
nar and turbulent boundary layers on arbitrary wings 
and previously proved to satisfy the requirements 
of numerical economy and precision. To allow its 
application to ships in general, and to double- 
elliptic and flat-bottomed hulls in particular, 
an appropriate coordinate system has been developed. 
Previously described coordinate systems, for example 
a streamline system such as that of Lin and Hall 
(1966) or the orthogonal arrangement of Miloh and 
Patel (1972) are limited in their applicability and 
the present nonorthogonal arrangement is similar 
to that of Cebeci, Kaups, and Ramsey (1977). 

The numerical procedure for solving the three- 
dimensional boundary-layer equations makes use of 
Keller's two-point finite-difference method (1970) 
and Cebeci and Stewartson's procedure (1977) in 
computing flows in which the transverse velocity 
component contains regions of reverse flow. This 
is in contrast to previous investigations, for 
example those of Lin and Hall (1966) and Gadd (1970), 
which are limited either to zero crossflow or to a 
unidirectional and small crossflow. It is also in 
contrast to the previous methods of Chang and Patel 
(1975) and Cebeci and Chang (1977) which did not 
have a good and reliable procedure for computing 
the flow in which the transverse velocity component 
contained flow reversal. 

In representing turbulent flow by time-averaged 
equations, a turbulence model is required and an 
algebraic eddy-viscosity formulation, similar to 
that of Cebeci, Kaups, and Ramsey (1977), is used. 
This is in contrast to the two-equation approach 
which Rastagi and Rodi (1978) have applied to three- 
dimensional boundary layers and which, in principle, 
should be better able to represent flows which are 
far from equilibrium. The previous comparisons pre- 


189 


sented in Cebeci (1974, 1975) demonstrated that the geodesic curvatures of the curves z = const and x 
present eddy-viscosity model allows excellent agree- = const, respectively. They are given by 
ment between measurements and calculations but did 
not include comparison with the three-dimensional il t) dhy 
NG) rey lee Unvacrersi8) 5 


boundary-layer measurements of Vermeulen (1971). hyhgsinO | dx Oz 


Since this data includes a strongly adverse-pressure 


gradient case which allows a stringent test of the a 1 ioe Rey Ie dh 3) 
present model, corresponding calculations and com- D hyhgsind | dz 1 ox 
parisons are reported. 

The calculation method is described in detail The parameters Kj)» and K» are defined by 
in the following section which states the three- = 
dimensional, boundary-layer equations in curvi- asy ell i BG) 1 306 
linear, Se ea Meeaheaeae and describes B12 sind | (x: i 1 2) pe cose (x, ei 22) ie) 


and discusses the required initial conditions, a 
turbulence model, and transformations in separate al ( dy ee) i 8 
: ; : S : K = {| Kp =) 6 (Kk), += — 
subsections. Section 3 is devoted to the coordinate ay sind 2 hg dz Soe 1 hy ox Ko9) 
system which is an essential feature of the present 


method. The numerical method is discussed briefly For an orthogonal system 6 = 7/2 and the parameters 
in Section 4 and calculated results are presented Ki, Ko, K,2, and Kj), reduce to 

in Section 5 which includes comparisons with the ss 

measurements of Vermeulen (1971) and demonstrations ear Ste 1 Os eee 1 ohe 7 
of the ability of the method to represent the geom- 1 hyhy 92 20 hyh, 0x (7) 
etry of different hull configurations and to result 

in realistic velocity and drag characteristics. eK on ee (8) 


Summary conclusions are presented in Section 6. 


At the edge of the boundary layer, (2) and (3) reduce 


to 
2. BASIC EQUATIONS 
uy au, W du, 5 
Cee: BAS a 2 
Boundary-Layer Equations h), 0x ho dz Ne COS! i RCS Ce ‘i uae 
The governing boundary-layer equations for three- as esc26 3 Pp cot@cscé 3 P 9 
dimensional incompressible laminar and turbulent er hy exp x ho dz\ po @) 
flows in a curvilinear nonorthogonal coordinate 
system are given by: u_ ow w_ oW 
e e e e 2 2 
a ee a ee I SIONIELS) Se NCTBLCISIONS) ae TL Ay 
Continuity Equation hy 0x hg oz S 5 cas 
5 ; 5 5 _ cotescé 2-(2) _ csce76 (2) (10) 
x (uh 2sin®) + jg (Wh isin®) + peamnns ing) =O (1) hy ax \p ho az \pe 


x-Momentum Equation The boundary conditions for Eqs. (1) to (3) are: 


9 3 3 y = 0: u,v,w = 0 (11a) 
i 4y% Y+yt- K,u~cot6 ap Kow“cscd + K)2uw 
hy ox.) hip) dz oy 
7 = OB u = Ug (x,2), W = We(x,z) (11b) 


ae esc28 Of i cot8csc8 9 /p 
hj dx \p ho dz \p 


Initial Conditions 


) du ; 
Ww By (% 2 Wy ) (2) The solution of the system given by (1) to (3), 


subject to (11), requires initial conditions on 
two planes intersecting the body along coordinate 
lines. In general, the construction of these 
initial conditions for three-dimensional flows on 
cscO + Kp)uUw arbitrary bodies such as ship hulls is difficult 
due to the variety of bow shapes, which may be ex- 
tensive and complicated. For this reason, assump- 
cot8csco al) a esc*8 mG) tions are necessary in order to start the calculations. 
hi ax \P he 3z \p In our study we choose the inviscid dividing 
streamline on which dp/dz = 0 to be one of the 
a ( ow vu") (3) initial data line (see Figure 1). In the case of 


z-Momentum Equation 


a ow ) 
e wl & WV 2s = Kone 


u Aue 2 
hy ox ho dz dy 


cot@ + Kyu 


+ — (v— 
OY oy rectilinear motion of a ship, this streamline runs 
along the plane of symmetry. Because of symmetry 
Here, h, and h2 are the metric coefficients and conditions, w and dp/dz are zero on this line causing 
they are, in general, functions of x and z; that is, (3) to become singular. However, differentiation 
with respect to z yields a nonsingular equation. 
hy = hy (x,z); hg = ho (x,2) (4) After performing the necessary differentiation for 
the z-momentum equation and taking advantage of 
Also, 9 represents the angle between the coordinates appropriate symmetry conditions, we can write the 


“x and z. The parameters K; and Kp are known as the so-called longitudinal attachment-line equations as: 


190 


Zz 


~<| 


INITIAL CONDITIONS ON 
A CROSS SECTION (x = xq) 


DP 


Lug 


FIGURE 1. The nonorthogonal coordinate system and the 
initial data lines for the ship hull. 


Continuity Equation 
3 : ; a F 
9q (un2sin®) + h)sinéw, + oy (vhjhgsin@) =O (12) 


x-Momentum Equation 


u_ du ou 2 
hy ox SV; By cotéK)u 


Us du + 
= a - K,uzcot6 + 2 (8 - wv") (13) 


z-Momentum Equation 


dWy Ww ow 


u 
— ~—— + — + vy —— + 
h, ox ho M oy LOE 


Ue BW i Wze ane peels ( owe = 

ia OE ia 21UgWoe By Coa - (w'v Ne (14) 
Where wz = dw/dz and (w'v')z = d(w'v')/dz. These 
equations are subject to the following boundary con- 
ditions: 


y = 0: u=ve=w, =0 (15a) 


y= 6: WS Tap We SMa (15b) 

The other initial data should be selected near 
the bow of the ship along the line perpendicular to 
the z = const coordinate (see Figure 1). However, 
because of the variety of possible bow shapes, 
approximations are necessary. For a simple, smooth 
bow section, where curvatures are small and no 
separation is expected, the flow along the initial 
line can be successfully assumed to be two- 
dimensional without pressure gradient, and the 
governing two-dimensional equations for a flat 
plate are solved. However, for most general mer- 
chant ships, the bow section is complicated and 
flow separation and reattachment are expected be- 
cause of large curvature variations and adverse 
pressure gradients; as a consequence, the boundary- 
layer calculations can only be performed downstream 
of the attachment line (or point) where turbulent 


5 INITIAL CONDITIONS (z = 0) ON 
ve THE PLANE OF SYMMETRY 


flow is presumed (since it is unlikely that the flow 
remains laminar after separation and reattachment 
with high Reynolds number). Generation of the 
initial data for turbulent flows is much more in- 
volved if there are no experimental data available. 
It requires sound mathematical and physical judgment 
and tedious trial-and-error efforts. We shall 
discuss this aspect of the problem later in the 


paper. 


Turbulence Model 


For turbulent flows, it is necessary to make closure 
assumptions for the Reynolds stresses, -pu'v' and 
-pv'w'. In our study, we satisfy the requirement 

by using the eddy-viscosity concept and relate the 
Reynolds stresses to the mean velocity profiles by 


' ’ du ' ' ow 
UNA NG Ri ty 


Hh et (16) 


We use the eddy-viscosity formulation of Cebeci 
(1974), and define €, by two separate formulas. In 
the inner region, €, is defined as 


au A 
(emi = | (3) 


where 
L = 0.4 y [1 - exp(-y/A) ] (18a) 
v Cae We 
NS 26 wh. S|) (18b) 
T p 
2 2 L 
Trac ) + (2) + 2cos@ (2) (=) (18c) 
Y/ w YY. <n NIMS 


In the outer region €_ is defined by the following 


m 
formula 
Eq = 0.0168 at (una u,) dy (19) 
0 
where 
= (as bP aweh 2 Q)2 (20a) 
Daa S (US We UgW.Cos 
UE = (u2 + we + es) e (20b) 


The inner and outer regions are established by the 
continuity of the eddy-viscosity formula. 


Transformation of the Basic Equations 


The boundary-layer equations can be solved either 
in physical coordinates or in transformed coordinates. 
Each coordinate system has its own advantage. In 
three-dimensional flows, the computer time and 
storage required is an important factor. The trans- 
formed coordinates are then favored because the 
coordinates allow larger steps to be taken in the 
longitudinal and transverse directions. 

We define the transformed coordinates by 


x 

u 

xX =X, Z= Z, dn= (=) dy, Ss) = Jonjax (21) 
1 0) 


and introduce a two-component vector potential 
such that 


uhysing + oun wh ; sing =e) (22a) 
ay oy 
vhyh sing = - (Be ae) (22b) 
ox ox 
where y and » are defined as 
y= (vs jug) thf (x,z,n) sing (23a) 
$ = (v8 )Ug)% (Uo ¢/U,)h) g(x,2,n) sing (23b) 


and Uyef is some reference velocity. 

Using these transformations and the relations 
given by (9), (10), and (11), we can write the x- 
momentum and z-momentum equations for the general 
case as: 

x-Momentum 


(b£")' + m)f£" - mp (£')2 - msf'g' 


+ mgf"g - mg(g')* + mj] 


z-Momentum 
(bg")' + mjfg" - m,f'g' - m3(g')* + mggg” 
ay2 
= fake) (GEN) sp ial 
ag! af ag' ag 
4 @ OB gf OE ; ee os 
= m0 (« es =) wad (s BD We eBeay) ate 


and their boundary conditions as 


i) = Wkap 287 = dbo Gl = WEAtas (26) 


Here primes denote differentiation with respect to 
nN, and 


€ 


The coefficients m, to mj, are given by 


s 3u, s 
m, = L @ Ee — =) fe pee NE See (hy sing) 


2 hyu, 8x hjhosin@ 3x 
S] dU Vref 
a ae = -s,K t 
M5 hug Ox s,K,cote, M3 S)K5 Fi cot¢ 
¥ _ 8a Gress dUe mR Uref 
M, = S)Ko1, ne as ae aon SiRio ou 
e 
Sal Crete 
Me = pects ee a Daa Oe (4 s; hy sing (28) 
h hosing aera dz e Ue 
eS 1 
i 2 
Si ref 
m7 Se, Mg = 51K) ( ) escé 
2 Ys Ue 
u S$} 
Mg = S )K) csc8, mig = rae 
ref 1 


191 


We We 
nn, Stiles oe Ms + m ( 
ef Uref 
w We 2 
m)2 = my m3 ( ) + M9 
ref Uref 
M19 We M7We IWe 
Uref ox u2 if az 


To transform the longitudinal attachment-line 
flow equations and the boundary conditions, we use 
the transformed coordinates given by (21) and de- 
fine the two-component vector potential by 


uhgsin@ = oe, 


: oY) 
hyh =-(— + 
vh;h9sin6o (2 ®) (29) 
with ¢ and still given by (23). With these vari- 
ables, the longitudinal attachment-line equations 
in the transformed coordinates can be written as 


(bE")' + m, ff" = Ty (£1)2 ar mefi"g + Mm) } 


Ss ' 
1 @ O20 a at ) Go 


(bg")' fy mg"£ pa myf'g! - m3 (g!)2 + megg" tas Mg (£')2 
= ag! of 
+mjp = — (= a a) (31) 


The boundary conditions and the coefficients m, to 
m)2 are the same as in (26) and in (28) except now 


Mien 7.8 
N = Neo? gr 
Yet 
m3 = af Suet me =m 
3 ho us , 6 3 
mg = 0, M)}] = m2 
2 
Wze Wze 3 1 dWre 
m)2 = m3 + my oF A 5 (32) 
Uref Uref Ip Uret ox 


In terms of the transformed variables, the alge- 
braic eddy-viscosity formulas as given by (17) to 
(20) become 


2 


i 2 y\ 2 
(a= S01 |, = Seo . 2) (Ey) 
vRy B 
2 L 
u a u 3 
+ aa (g") 4 + 2 nee ogcost (33) 
Ge Ue 
he We 2 We 4s 
(Sale = 0.0168 VR, if 1+ e + 2\ —]cosé 
é Ue Ue 
2 
u 
- ee (Ens 
Ue 
u A 
+ 2 ee eeticose dn (34) 
Ue 
Here R, = u,S}/v and 


Une ‘3 
+2 mamgawewccce (35) 


3. COORDINATE SYSTEM 


Since, in general, a ship hull is a complicated non- 
developable surface, a Cartesian coordinate system 
is not suitable for boundary-layer calculations. 
Most existing merchant and naval vessels possess 

the following features: a flat bottom [y = £(x,zZ) 
is not a single-valued function]; a bottom which is 
not parallel to the water surface; and a bow which 
has a submerged bulb extending toward the origin. 

In addition, the problem is further complicated by 
the existence of a free surface, corresponding to 
the water level of a partly-submerged hull. The 
chosen coordinate system must be sufficiently general 
to allow these various features to be represented in 
the boundary-layer calculations. 

The streamline coordinate system is superficially 
attractive but the determination of the streamlines, 
the orthogonal lines, and the associated geometrical 
parameters requires considerable effort. They are 
dependent on the Froude number, and also on the 
Reynolds number if the displacement effect is taken 
into account. Consequently, and in addition to 
being hard to compute, this coordinate system be- 
comes uneconomical to use when the effect of the 
Froude number and the Reynolds number are to be 
systematically examined. - 

A desirable requirement of a coordinate system 
for the boundary-layer calculations is that it be 
calculated only once. Miloh and Patel (1972) pro- 
posed an orthogonal coordinate system which depends 
only on the body geometry and is calculated once 
and for all. This coordinate system has been applied 
by Chang and Patel (1975) to boundary-layer calcula- 
tions on two simple ship hulls: ellipsoid and double 
elliptic ship. One of the coordinates is taken as 
lines of x = X = constant and the other as z(X,Z) = 
constant, which is orthogonal to x = constant lines 
everywhere on the ship hull, and is obtained from 
the solution of the differential equation 


dz fzts 
ax | 1 + £2 Ee 
Zz 

Here y = £(X,Z) defines the ship hull, and (x,y¥,Z) 
denote the Cartesian coordinates. The major ad- 
vantage of this coordinate system is its simplicity. 
Because one of the coordinates is subject to the 
condition (36), there is no guarantee that the 
boundaries of the ship hull are coincident with the 
coordinate lines. Furthermore, for a ship with flat 
bottom for which y = £(xX,Z) is not a single-valued 
function, one of the coordinates cannot be calcu- 
lated from (36). The coordinate system is limited, 
therefore, to some special geometries only. 

In this study we adopt a nonorthogonal coordinate 
system similar to that developed by Cebeci, Kaups, 
and Ramsey (1977) for arbitrary wings. It is based 
on body geometries only and, hence, it is calculated 
once and for all. In addition, the system can deal 
with the peculiar features of most merchant and 
naval vessels discussed previously. The details 


of this coordinate system are described briefly 
in the following paragraph. 

Now consider the ship hull as given in the usual 
Cartesian coordinate system; that is, x along the 
ship axis, y and z in the cross-plane (see Figure 
1). We select x = x = constant as one of the co- 
ordinates and the other coordinate, z, lies in the 
yz-plane. Because the coordinate system is non- 
orthogonal, we are free to select the values of z 
in the plane to satisfy the condition that the 
boundary lines of the ship hull are coincident with 
Z = constant coordinate lines. There are several 
ways of finding the z-values. Here z is determined 
by mapping each yz crossplane into a half or hull 
unit circle depending on whether the crossplane in- 
tercepts the free surface or is completely submerged. 
The polar angle, normalized by tm or 27 on the unit 
circle, is taken as z-values. The z-values then 
range from 0 to 1 on each crossplane. The advantage 
of the mapping method is that equi-interval, z = 
constant coordinate lines are automatically concen- 
trated in the region of large curvature where the 
boundary-layer characteristics are expected to vary 
greatly. Hence the number of z = constant coordinate 
lines can be reduced without loss of accuracy. 

There are several methods available for the map- 
ping of an arbitrary body onto a unit circle. Here 
we use the numerical mapping method developed by 
Halsey (1977). It makes full use of Fast Fourier 
Transform techniques and has no restrictions on the 
shape of the body to be mapped. To map a smooth 
crossplane onto a unit circle, the procedure is 
fairly easy. If there are inner corner points, or 
trailing-edge and leading-edge corner points (see 
Figure 2) caused by the reflection of the cross- 
plane, they must be removed before mapping is per- 
formed to improve numerical accuracy and to provide 
rapid convergence. The inner corner points are 
rounded off by using Fourier series expansion tech- 
nique and the leading-edge and/or trailing-edge 
corner points are removed by using the Karman- 
Trefftz mapping. For details see Halsey (1977). 

To use the mapping method to find the coordinate 
system, it is only necessary to define the ship hull 
as a family of points in the x = constant planes, 
to locate the intersection of the ship hull and the 
free surface, and to indicate whether corner points 
exist. The data in each plane is then mapped into 
a unit circle as ¥ vs z and z vs z and interpolated 
for constant values of z. Another set of spline 
fits, in the planes z = constant for y vs x and Zz 
vs x, completes the definition of the coordinate 


[aie le aN 


LEADING-EDGE 
CORNER POINT 


W.L. 


TRAILING EDGE f y 
CORNER POINT: 


SHIP HULL 


«(NNER CORNER POINT 


FIGURE 2. Notation of corner points used in the 
Mapping procedure. 


system. The lines formed by the intersection of 
the planes x = constant and z = constant with the 
hull constitute the nonorthogonal coordinate net 
on the surface, and the third boundary-layer coor- 


dinate is taken as the distance normal to the surface 
in accordance with first-order boundary-layer approx- 


imation. 


Since the spline-fitting also yields derivatives, 


the metric coefficient and the geodesic curvatures 
of the coordinate lines can be calculated from the 
formulas given below. 

The metric coefficients: 


D) 2. 
4 i 
mene (24) « (2 (37a) 
ax / 2 SY 
2 (SEN> (ES? 
hp = (=) oF (2) (37b) 
x x 


The angle between the coordinate lines: 
i =) (=) Gy &) 
cose = (2 as ap — = (38) 
hyho oz 5% ox i ox 2 Cr4 . 
The geodesic curvature of the z = constant line: 
fe, 1 oy 22 | _ (ax oz 
1“ huhosind oz ox ox dz 
1 x Zz Zz x 


(2 a4y oy ay 
ox ox2 ox ox2 


oF 29 
‘ @) e “A 
x Zz 


(39) 


az (eS 
az ax2 
x 2 


The geodesic curvature of the x = constant line: 


(ase jaceaee oy) OB. on (oe 82 
a hy hi sind Zz)  \ox/ , ax/ , \dz 
x 


(40) 


The other parameters K}2 and K2] are calculated from 


(6). It may be noted that K; and K» can also be 
obtained from (5). This provides a check on the ex- 
pressions given by (39) and (40). 

In the boundary-layer calculations, we need the 
invisid velocity components along the surface 
coordinates. Let V be the total velocity vector 
on the hull, (Uu,v,w) the corresponding velocity com- 
ponents in the Cartesian coordinates, and (Ue,We) 
in the adopted surface coordinates. As can be seen 
from Figure 3, 


V- ie, = ¥ o ee cos® 

pean sin26 a 
v- ty = Vv 2 t) cos¢é 

eS kee aa: oe 


> => 
Here t; and t2 are the unit tangent vectors along 
x and z coordinates and are given by 


ne kles @)s a) | 
t) = Taq 1+ Cz 5] (2) k (43) 
Zz Z 
= = Lee) se oz\ 2 
tre (=) 4) (2) k (44) 
x x 


193 


FIGURE 3. Resolution of the velocity components. 


With the definition of V and with the use of (43) 
and (44), Eqs. (41) and (42) can be written as 


= {92 
MN Ba (45) 
x 
1 atl |le x) B (2 
= + —< 
“OT Sines ho @ © \G2 ee 


4. NUMERICAL METHOD 


We use the Box method to solve the boundary-layer 
equations given in Section 2. This is a two-point 
finite-difference method developed by Keller and 
Cebeci. This method has been applied to two- 
dimensional flows as well as three-dimensional 
flows and has been found to be efficient and accu- 
rate. Descriptions of this method have been pre- 
sented in a series of papers and reports and a 
detailed presentation is contained in a recent 
book by Cebeci and Bradshaw (1977). 

In using this numerical method, or any other 
method, care must be taken in obtaining solutions 
of the equations when the transverse velocity com- 
ponent, w, contains regions of flow reversal. Such 
changes in w-profiles will lead to numerical in- 
stabilities resulting from integration opposed to 
the flow direction unless appropriate changes are 
made in the integration procedure. Here we use the 
procedure developed by Cebeci and Stewartson (1977). 
In this new and very powerful procedure, which fol- 
lows the characteristics of the locally plane flow, 
the direction of w at each grid point across the 
boundary layer is checked and difference equations 
are written accordingly. At each point to be calcu- 
lated, the backward characteristics which determine 
the domain of dependence, are computed from the 
local values of the velocity. Since the character- 
istics must be determined as part of the solution 
a Newton iteration process is used in the calcula- 
tion procedure to correctly determine the exact 
shape of the domain of dependence. 


To illustrate the basic numerical method, we shall, 


194 


at first, consider the solution of the longitudinal 
attachment-line Eqs. (30) and (31) and then the 
solution of the full three-dimensional flow equations 
as given by (24) and (25). We shall not discuss the 
Cebeci-Stewartson procedure for computing three- 
dimensional flows with the transverse velocity, w, 
containing flow reversal since that procedure will 

be fully described in a forthcoming paper. 


Difference Equations for the Longitudinal 
Attachment-Line Equations 


According to the Box method, we first reduce the 
Eqs. (30), (31), (32), and (26) into a system of 
five first-crder equations by introducing new depen- 
dent variables u(x,zZ,n), v(x,Z,n), w(x,Z,n), 
t(x,zZ,n), and 6(x,zZ,n). Equations (30) and (31) 
then can be written as 


ey yy (47a) 
w'=t (47b) 
(bv)' + 6v - mou* + mj] = mg u =e (47c) 

(bE) 4 + 8 = miuw = m3w- = mou + mj)2 

ow 
= mM) 0 u Bee (47d) 
6' = mju + mew +m au (47e) 
1 6 10 ox 

The boundary conditions (26) and (32) become 
n= 0: u=w= 0 = 0 (48a) 
N = No: u=l1, we Woe/Uref (48b) 


We next consider the net rectangle shown in Figure 
4 and denote the net points by 


x0 


i] 
(o) 
6 

=} 

| 


SS Peesy) an ky 


ll 
(o} 
=) 


no = - 


j-1 + hy a ek Bp acon a 


We approximate the quantities (u,v,g,t,9) at 
points (x05) of the net by functions denoted by 
(uh, v9 wh, th, 65) . We also employ the notation for 
points and quantities midway between net points and 


for any net function SH 


(x) n-1/2 
(xq) n-1 (x,) 


n 


FIGURE 4. Net rectangle for the longitudinal 
attachment-line equations. 


, 


na ph 
aia = 2% * na} yaya = 2 ("3 i "j-) 
nS /2 5 a(n y-1 ) eS pean n 
ei, =" (33 + s3 yEE i fo =D S5 + 5-1) (49) 


The difference equations which are to approximate 
(47) are formulated by considering one mesh rect- 
angle as in Figure 4. We approximate (47a,b) using 
centered difference quotients and average them 
about the midpoint (%y N5-1/2) of the segment P )Po2. 


Olid S sit enya! 
hy (uy j-1) 5-1/2 (50a) 
il soy Gal — Aig 


Similarly, (47c,d,e) are approximated by centering 
them about the midpoint x,-1/2+15_1/2 of the rect-— 
angle P)P )P3P,. This gives 


-1 n n n 
he by) =" (bv ar (Chi) 
: ( XA) ( Deon ( Deep 
a n 2\n _ aril meer 
(m5 a a) (u 5 -a/2 Reey/2 M)] (50c) 
= ak n n n 
ite || Coren = “Cee)) + (Git) 
J 5 5-1 ‘5-1/2 
= n n sd n 2 n = n 2 n 
(m, + O,) (UW) a7 m3 (Ww eee Mg (u Vee 
a, sat we uM yntl wi 
Sn) Y4-1/205-1/2 5-1/2" 5-1/2 
=cut n (50a) 
jai72 ~ ie 
1 n n n n TT 
hj G = 1) -( nj + 2a) 8-1/2 ea 
n-1 
aT (50e) 
Here 
n-1 pa 2,n-1 
jo a ey 
Sil n= n-1 ig —ill 
= hie bv). 
{5 [( v) (bv) 1] = COR. 
=a! ey a. n-l 
2: 3-172 aH a (51a) 
n-1 a ( n-1 
y= SO Bp 
=e n-1 n-1 n-1 
=< he bie)is 3 (VoKe)) a + (6 
o [ ( ); ( 5-1] ( ) 5 1y2 
TAN 4 Dol n-l, 9\n-1 
—m (w ts 2 
3 ewe 9 (u Boy (51b) 
n-1 n-1 
ey. gooey 
Ly ek 1 n-1 n-1 n-1 
£5 (@ i ai) Ty Mile 
TS =i jaa 
6 j-1/2 (Slc) 


(j,n,i-i) (j,n,i) 


(j-1,n,() 


(jn b— 
a 
Ha 
“aA 
(j-1, n—1, i) 
(j-1, n=1, i-1) 


f— Cierra 


FIGURE 5. Net cube for the difference equations 


z(i) 


for three-dimensional flows, wj > 0. 
n-1/2 
w@ 
= (51d) 
n xk 
n n= 


Difference Equations for the Full Three-Dimensional 
Equations 


The difference equations for the full three- 
dimensional equations, as given by (24) and (25), 
are again expressed in terms of a first-order sys- 
tem. With the definitions given by (47a) and (47b), 
they are written as 


(bv)' + @v - mju? - mcouw - mgw? + m)) 


=Mj9 u oy m7w x (52a) 
(bt)' + 6t - myuw - m3w* - mgu2 + m)9 
=™)9 u oy m7w a (52b) 
6' = mju + mew + mo a + m7 se (52c) 
Their boundary conditions, (26) become: 
n= 0: u=w=60=0 (53a) 
in) = wWs3 u=1, W = We/Uref (53b) 


The difference equations for (47a) and (47b) are 
the same as those given by (50a) and (50b): they 
are written for the midpoint [(x) ye (2) 55-770] of 
the net cube shown in Figure 5; that is, 


al Ca - ut ) = yori 


j jon) > “3 ai7a° 
SW ciogah L Gaelpsl\\ = Afalpal 
my G “ie osy2 ee) 


The difference equations which are to approximate 
(52a,b,c) are rather lengthy. To illustrate the 
difference equations for these three equations, we 
consider the following model equation 


(bv)' + 0v + m,)] = mo u oe 4 m7w ~— (55) 


The difference equations for (55) are: 


195 


-l) — — —— n-1/2 
A a ; 8 
te |=, Wa + ( 5 fp + Cn) 5 yp 
u-wu 
a yn 25 n n-] 
10° 4-1/2 3-1/2 k 
Bly 1 ln 
INI / 2 al i-l 
+ See 
O21 7295-12) xy oe 
Here, for example, 
rae al n,i i ynriml n-1,i-1 va) 
J @\9 j j j 
a ak Flat n,i-1l alae ,i-l 
=a + 
a a G j el oP 4; 1 
Sn dh pal. n=), n,i n=1,i 
na 7 a (o! io ene eo ) ee) 
and 
n-1/2 
(m1) 5-1/2 
1 n n n-1 n-1 
S 7 inde w Cina ony Cade |e in) aaa 
Zz) = 0 Fag Sag on eB a a Bo acbop 28 (58) 


Solution of the Difference Equations 


The difference equations (50) for the longitudinal 
attachment-line flow and the difference equations 
for (52) are nonlinear algebraic equations. We use 
Newton's method to linearize them and then solve the 
resulting linear system by the block-elimination 
method discussed by Keller (1974). A brief 
description of it will be given for the streamwise 
attachment-line equations. 

Using Newton's method, the linearized difference 
equations for the system given by (50) are: 


h, 
5 al = 
ORS Oy, Sa Way > Og) = Ga) (59a) 
Ay 
- ‘ -=— me taOiter. = F 59b 
by > Ota oe Oe, Oh, 4) = Ge (59b) 


aOWin oF év., + 200.5 4 NOOR 
(S1) 5 Nea (G2) Ve (63), 5 (Su) 5 5 


1 


+ “OU ou, = ; 5S) 
(os), Bas (Se), Bag 5S) 2 (59c) 


R a (Sh A a sor : : 
(By), 8, + (B2),6t,_, + (Ba), 60, + (8,), 68 


> i 


_ W. _W. gu, + (Bg). cu, 
+ (Bs), 8", + (Be), + (B75 S48, + (Bq); ou 


1 J aaa 


; (59d) 
(ry), 


+ fi A + z s 
(01) 68, (oa) a OO5_ 5 (93). du 


+ (6). 6, 
j Si 5 Oe 


j-1 


+ Wop) Om + Woie)) 4 Ors 5 = (x5), (59e) 


196 


Here we have dropped the superscripts n, i and have 
defined (1K) 51 (%) 5. (8) 54 and (oy) 5 by 


r= Ble - ws to BeVe a75 (60a) 
ro = a - we + Meee iV/e (60b) 
= ann ay 
ea) s = See M)] 
= bv)! + : 
[ ean” SONS, 
(60c) 


- (my + 2) saa 


p= 


wa ase = 
(Ty) 5 Faia 7 2 


- lone) Y ne) = 
[ Veena eo ey Ot On) (UW) 5 a7 


ag 2 = 2 
Hee FOC n7e 
ms! n= 
z ("5-1/2 wj-1/2 “4-1/2 “a7 oe 
_ ari 
ity) 5 T 5-1/2 
. ev SONS Ce ORE pn Petscaja | (60 
fo i 
a eo 
(5 AT ies 8, (61a) 
j 
P51 
(aS Se te Ones (61b) 
j 
(3). = ev 61 
(E5) 7 2 73 ( c) 
(ty). = 2G 
Bia ay Was (61d) 
(e5), =-—- (mp + a )u (6le) 
(56). = =) (GY) oC ae (61£) 
Sie = 
( 1), (21), (62a) 
Bo). = : 
( 2), (2), (62b) 
Fa deep ae 
(83), = ks (62c) 
(Baya it 
Tepe (62a) 
A at il n-1 
(Bs). = 5 (m, + 4 _)u. - m3w. - hon Y5_1/2 (62e) 


(Be) =- = (my, + a_)u i maa 
tg = 
Deana eZ ozs) 
= + a ple n-l 
(87), = (m, + @ Ns mgu. + 5 a W512 (62g) 
(Go, = oS Gu eg. om 
eee) aan ipa ean 
ee ee 62h 
DS S/2 (ny) 
1 
(91), = (63a) 
J 
all 
Kop) Baie (63b) 
J 
(93). = - Ga + 2a_) (63c) 
3) > (mM a c 
(Oy). = - Tr ae 2x01) 63d 
4 5 = 2 1 n ( ) 
(C5) ai 2 63 
Sa > M6 (63e) 
(Ge). =o z m (63£) 
The boundary conditions become 
Sug = Swo = 58 = 0, bus = ow, = 0 (64) 


The solution of the linear system given by (59) 
and (64) is obtained by using the block elimination 
method. According to this method, the system is 
written as 


fA gé=r 65 
aes (65) 
Here 

Ao nS 

Sr eae 

NG 
w= : Cc 
J 3) J 


8 Ey (xy), Oe 
§, Ko (r)5 év 
goa] B [a to | Os || 
: : (5 OE. 
a5 Xs (ae 80, 


The A., By, C. in /A denote 5x5 matrices. The 
solution of (25) is obtained by the procedure 


described in Cebeci and Bradshaw (1977). 


5. RESULTS 


Turbulent Flow Calculations for a Curved Duct and 
Comparison with Experiment 


The turbulence model described in Section 2 has been 
used with considerable success to compute a wide 
range of two-dimensional turbulent boundary layers 
[see for example Cebeci and Smith (1974)]. The 
model has also been used to compute three-dimensional 
flows and again is found to yield accurate results 
[see for example Cebeci (1974, 1975) and Cebeci, 
Kaups, and Moser (1976)]. To further test the model 
for three-dimensional flows, we have considered the 
experimental data taken in a 60° curved duct of rect- 
angular cross section. Figure 6 shows a sketch of 
the flow geometry. The experimental data are due 
to Vermeulen (1971). Here z denotes the distance 
from the outer wall, measured along normals to the 
wall; x denotes the arc length along the outer wall; 
and y denotes distance normal to the plane x,z. 

To test the computed results with the data, it 
is necessary to specify the initial profiles given 
by experiment. This can be done in a number of ways. 
In the study reported by Cebeci, Kaups, and Moser 
(1976) the profiles were generated by using Coles' 
velocity profile formula. That formula, which repre- 
sents the experimental data rather well for two- 
dimensional flows, was not very satisfactory for 
three-dimensional flows. Here we abandon the use 
of Coles' formula in favor of Thompson's two- 
parameter velocity profiles as described and im- 
proved by Galbraith and Head (1975). According 
to this formula, the dimensionless u/ug velocity 
profile is given by 


INITIAL 
CONDITIONS 


MEASURING LOCATIONS 


eee 


FIGURE 6. Coordinate system and notation for the 
. curved duct. 


197 


u 
ei Ye (=) 15 (hy S 57a) (66) 
= €/ inner 


Here yg is an intermittency factor defined by the 
following empirical formulas: 


‘LoS = 
Oia Oe 
y y Z 
405 < °*S4 Oo, = 1 - 2.64214(— - 0.05 
Sn” hes (5 nh ) 
y y d 
poi Ss = PE 
O25 <a 8 Oo ie 4.4053(2 0.5) 
- 1.8502( 4 = 0.5) + 0.5 
89 
y 2 
0.7<*%< 0.95 yg, = 2.64214 (= 0.05) 
50 0 
y A 
nee 0.95 Ys = 0.0 


The dimensionless velocity profile for the inner 
layer, that is, (u/ug)inner, is given by 


y <4 ut = y* 


4< yt < 30 ut = c) + coln yt + c3(1n yt)? 
+ cy (in y*)3 
y > 30 u’ = 5.50 In y’ + 5.45 


Hene en = 4 Si co =" Sey 45y Cape— OP elOF ici 
=0. 767) yu = WY, B= (t,,/2) 2, ut = u/u,, and 6, 
is a parameter which is a function of 6, Ce, and H. 
To find the functional relationship between On 
CE 68, and H, we use the definitions of displacement 
thickness, 6*, and momentum thickness, 9. Substitut- 
ing (66) into the definition of 6*, after some alge- 
bra, we get 


A 
SE i ae red a 
55 Rae 
c Cc 
£ 5* f ) 
= 0.5 + — — - = — 
0.5 5 Ay, In oe A3 Ao In ce 5 (67) 

where 


A, = 50.679, Ap = 1.1942, A3 = 0.7943, Ay = 1.195. 


An expression similar to that given by (67) can also 
be obtained if we substitute (66) into the defini- 
tion of 8. However, the resulting expression is 
quite complicated. For this reason, the expression 
for 8 is obtained numerically, and for a given value 
of © and H, the corresponding values of c¢ and 6, 
are computed from that equation and from (67). 

Equation (66) is recommended for two-dimensional 
flows. Here we assume that it also applies to the 
streamwise velocity profile by replacing u/u, by 
u,/ug, with c, now representing the streamwise skin- 
friction coefficient. 

In order to generate the crossflow velocity com- 
ponent (a)/ug,) + we use Mager's expression and 
define Up/Us, by 


198 


08 


0.2 


FIGURE 7. Comparison of gener- 
ated initial total velocity ° sesh 


profiles with Vermeulen's data. y 0 
Yn Us ; 
aan ee ( - x) tang, (68) 
Se "se 


with the limiting crossflow angle fy, obtained from 
the experimental data. 

Once the streamwise and crossflow velocity pro- 
files are calculated by the above procedure, we 
compute the velocity profiles u/u, and w/wWe in the 
orthogonal directions x and z by the following rela- 
tionships 


u Uu. WwW - 
u s n e 
Tee Gow eMOS feo) 
e Se Se e : 
u u 
Ww Ss Un e 
a (69b) 
We Us, Us, We 


Figure 7 shows a comparison of generated and 
experimental total velocity profiles along the line 
A. As can be seen, the procedure discussed above 
for generating the initial velocity profiles from 
the experimental data is quite good. This is impor- 
tant for an accurate evaluation of a turbulent 
model, especially for three-dimensional flows. 

Here 


Oo DATA 


— PRESENT METHOD 


844 (mm) 
© 
iT 
o 
} 
\e) 
° 
° 
° 
=O 
SS SS SS 
> oa ao 
° 


x (m) 


FIGURE 8. Comparison of computed momentum thickness 
with Vermeulen's data. 


AN 
ANZ 
A10 
oo AN AB 
GENERATED PROFILES 
GENERATED PROFILES 
DATA 
° DATA 
— ) 0 —— 4 ————— 
30 40 0 10 20 40 50 
y (mm) y (mm) 
(0) u2 + w 
Oa wee ee (70) 
— e 


The solution of the boundary-layer equations also 
requires the specification of the metric coefficients 
and the geodesic curvatures. They are calculated 
from the following expression: 


(a: straight section 
hy = 
1 - 2/R curved section 
le fe) 
hy = 1.0, Ko = 0 (71) 
(a : E 
(0) straight section 
K] = 
1/ (Ro-2) curved section 
S 


A comparison of calculated and experimental values 
of streamwise momentum thickness, 8 ),, shape factor, 
H}1, skin-friction coefficient, cf, and limiting 
crossflow angle, 8y, is shown in Figures 8, 9, 10, 
and 11, respectively, along the lines B, C, D, E. 
Here the limiting crossflow angle is computed from 


° DATA 


— PRESENT METHOD 


x (m) 


FIGURE 9. Comparison of computed shape factor with 
Vermeulen's data. 


© DATA 
——PRESENT METHOD 


x (m) 


FIGURE 10. Comparison of computed skin friction 
coefficient with Vermeulen's data. 


We Ate urea) oe 5 f 


EE (72) 
2 " " 
a) (eed a) Se bg 


tanBy = 


Figures 12 and 13 show a comparison of calculated 
and experimental total velocity profiles and cross- 
flow angle profiles along the lines C and E. Here 
the crossflow angle is computed from 


w/oa L ee) 5" ie £"] 


(73) 
9/2, (2/4)? 


sinpy, = 


As in Figures 8 through 11, again the agreement 
between calculated results and experiment is very 
good. The computed results follow the trend in 
the experimental data well and indicate that the 
present turbulence model, as in two-dimensional 
flows, is quite satisfactory for three-dimensional 
flows. 


Results for a Double Elliptic Ship Model 


To test our method for ship hulls, we have con- 

sidered two separate hulls. The first one, which 
is discussed in this section, is a double elliptic 
ship whose hull is given analytically. The second 


DATA 


PRESENT METHOD) 


1.0 15 2.0 25 3.0 
x (m) 


FIGURE ll. Comparison of computed limiting crossflow 
angle with Vermeulen's data. 


nie}) 


one, which is discussed in the next section, is 
ship model 5350 which has a rather complex shape. 
Its hull is represented section-by-section in tabu- 
lar form and contains all the features of most 
merchant and naval vessels. It proves an excellent 
test case to study the computational difficulties 
associated with real ship hulls. 

The double elliptic ship model can be analytically 
represented by 


pi ty UAL Ne le BY? (le 
Y = 2 (s4) = TON ae 1 -{ = (74) 


It has round edges except for the sharp corners at 

x = +L and z = +H. The body of L:H:B = 1.0:0.125:0.1 
together with the nonorthogonal coordinate nets on 
the hull is shown in Figure 14. 

The potential-flow solutions were obtained from 
the Douglas-Neumann computer program for three- 
dimensional flows. To get the solutions, 120 control 
elements on the surface were used, 12 along the x- 
direction and 10 along the z-direction. 

Before we describe our boundary-layer calculations, 
it is useful to discuss the pressure distribution for 
this body shown in Figure 15. As can be seen from 
the figure, the streamwise pressure gradient is 
initially favorable in the bow region and then ad- 
verse up to the midpoint of the body. This is fol- 
lowed by a region of favorable pressure gradient and 
then by a shape adverse pressure gradient very close 
to the stern. The crosswise pressure gradient varies 
in a more complex manner. Near the bow the pressure 
decreases down from the water surface to a minimum 
and then increases as the keel is reached. As the 
flow moves downstream, the location of the minimum 
pressure moves up and reaches the water surface at 
about x/L = -0.80. The minimum pressure remains at 
the water surface to about x/L = 0.80 and then moves 
toward the keel. As a result, near the bow and the 
stern, one may expect flow reversal of the crossflow 
across the boundary layer does not reverse direction 
from the keel to the water surface. This conclusion 
is drawn from considering the pressure gradients only. 
The real situation may be somewhat modified because, 
in addition, there are the upstream effects and the 
curvature effects on the flow characteristics. 

The boundary-layer computation starts with turbu- 


lent flow from X/L = -0.90. We have tried to start 
the computation from X/L = -0.97 and X/L = -0.95. 
However, flow separation was observed at X/L = -0.90 


near the keel due to the sharp curvature and adverse 
pressure gradient in the bow region and can be seen 
from Figure 15. In the previous calculations of 
Chang and Patel (1975) and Cebeci and Chang (1977), 
the flow separation near the bow was not found due 
to the orthogonal coordinate system they adopted in 
which the second net point from the keel is so far 
from the keel that the region of adverse pressure 
gradient is omitted. 

In our boundary-layer calculations, we have used 
40 points along the x-direction and 16 points along 
the z-direction. In the normal direction, we have 
taken approximately 40 points. The nonuniform grid 
structure described in Cebeci and Bradshaw (1977) 
is employed in the normal direction so that the grid 
points are concentrated near the wall where the 
velocity gradients are large. 

Some of the computed results for R; = 10’ are 
shown in Figures 16 and 18. Figure 16 shows the 
spanwise distributions of the pressure coefficients, 
Cp, local skin-friction coefficient, cre, the shape 


200 


0.2 


0.2 


FIGURE 12. Comparison of com- 
puted total velocity profiles ° 1 


C6 0.8 


Cc! 02 


PRESENT METHOD 
DATA 


PRESENT METHOD 


DATA 


3% 20 
y (mm) y (mm) 


E4 


£3 


PRESENT METHOD 
PRESENT METHOD DATA 


DATA 0 


with Vermeulen's data. 


factor, H,;,], the Reynolds number based on the momen- 
tum thickness, Rg, and the limiting crossflow angle 
for x/L = -0.85, 0.0, and 0.75. As can be seen from 
these figures, the boundary-layer parameters vary 
greatly near the keel where the curvatures and the 
pressure gradients are large and remain almost un- 
changed near the surface where the curvatures and 
the pressure gradients are small. Except at x/L = 
-0.85, the limiting crossflow angle is positive. 
This implies that the crossflow near the wall moves 
from the keel to the free surface as predicted from 
the pressure distribution. Figure 17 shows typical 
longitudinal and transverse velocity profiles at 

z= 0.6 for several values of (x/l), and Figure 18 
shows typical transverse velocity profiles at (x/L) 
= -0.2 for several values of z. As can be seen from 
Figures 17(b) and 18, the transverse velocity compo- 
nent undergoes drastic changes in the longitudinal 
and transverse directions under the influence of 
pressure gradient and body geometry. As was dis~ 
cussed before, when the transverse velocity changes 


30 20 
y (mm) y (mm) 


sign across the boundary layer and contains regions 
of reverse flow, numerical instabilities results from 
integration opposed to flow direction unless appro- 
priate changes are made in the integration procedure. 
The new numerical procedure of Cebeci and Stewartson 
(1977) handles this situation very well and does not 
show any signs of breakdown resulting from flow re- 
versal of transverse velocity component. 


Results for Ship Model 5350 


The ship model 5350, unlike the one discussed in the 
previous section, is a realistic tanker model. The 
geometry of the hull is so complicated that it is 
represented in tabular form section by section. 

The model possesses all the special features of 
existing merchant and naval vessels, that is, a 
bottom which is flat and not parallel to the still- 
water surface and an extended bow completely sub- 
merged under the water surface, and consequently 


201 


20) Cé = PRESENT METHOD 
——— PRESENT METHOD 
° ‘DATA 
° DATA 
30 
y (mm) ‘S 
nN 
c 
20 + 
i \ 
\ 
’ \ 
NS 
10 
" Ny \ 
A Sie. . 
9 As as we rt AS a 
0 10 15 20 10 20 30 40 
J (degrees) J (degrees) 
60 
£10 
£16 
E14 
50 5 50 | 
£12 i 
PRESENT METHOD 
fo} fe) 
40 + ° DATA 40 + ?\ \ © 
c6 Es OF o\ O\ 
y (mm) o \ o\ a ~ PRESENT METHOD 
&4 a\ ° aX } 
30 | 0 \ ° } ON ° DATA 
BS om ° ) ° 
y (mm) g o\ ° ° ° 
€1 } g °\ Ne ° 
20 20 + 2 \ 9 ° 
. Boe eens as 
\ ve ‘e aX : s ; 
Na} \e ) ° ° 0 
10 7 QR 10+ \° ° ° ° 
\e So \o f) ° 
hoe o 6 Xe > GJ ° 
<2 SO'g 0 
Bs On } 5 See oe plo oe as . 
0 Ji fan os fu = 5) 0 o— ~Baed.c- 0 6, pa SS 
0 0 0 0 5 10 15 20 0 0 ty) () 0 10 20 30 40 50 
J (degrees) J (degrees) 
FIGURE 13. Comparison of computed crossflow angle with Vermeulen's data. 


serves as an excellent case on which to apply our 
method. 

Figure 19 shows a three-dimensional picture of 
this ship model together with our nonorthogonal co- 
ordinate system. We see from this figure that, as 
a by-product of the mapping method discussed earlier, 
the z = const. coordinate lines are concentrated 
in the bow and corner regions where the curvature 
is large. Figure 20 shows different cross-sections 
(indicated by solid lines) and interpolated values 
obtained by a cubic-spline method (indicated by 
circles) from which the geometric parameters are 
obtained. 

The inviscid velocity distribution for the model 
is obtained by using the Douglas-Neumann method 
treating the model as a double ship model. Figure 
21 shows the pressure distribution for the entire 
ship and Figure 22 shows a detailed pressure distri- 
bution for the bow region. We see from these figures 
that the longitudinal pressure gradient near the keel 


FIGURE 14. Three-dimensional picture of double ellip-— 
-tic ship model with the nonorthogonal coordinate system. 


FIGURE 15. 


GIRTH, % 


Pressure distribution for the double- 


elliptic ship. 


is) 


in) 


Cc, x 103 


Hay 
nN 


(b) 


40 


2.0 


Rg x 10-4 


(c) = 425 


(a) 1.0 


FIGURE 16. 


20 40 60 80 100 
GIRTH, % 


model for Ry; = 10° at (a) x/L = -0.85, (b) 


| 
0 20 40 60 80 100 | 
GIRTH % 4 
KEEL 
zZ 
———) 
100 | 
KEEL 
Computed Cp, Ce, Hii, Ro, and By for 
x/L = 


x/L = 0.0 
x/L = 0.75 
_>y 
x/L = —0.85 
—_-—~Sy 


the double-elliptic ship 
0.0, (c) x/L = 0.75. 


< 


1.0 7 


0.8 


0.6 


0.4 


x/L = —0.85 


4 1 
0) 10 20 30 40 50 60 70 80 


n(=(u,/¥s,) 2y) 


FIGURE 17. 
10’ at z = 0.6. 


is favorable and then later becomes adverse. The 
pressure gradient in the transverse direction de- 
creases rapidly from the keel to a minimum value 

and then increases continuously up to the free sur- 
face. Due to this rapid pressure variation in the 

bow region, preliminary boundary-layer calculations 
showed flow separation and required an approximate 
procedure to generate the solutions for x < 22.5 m. 
After that (x > 22.5), the three-dimensional boundary- 


FIGURE 18. Computed transverse velocity profiles. 


203 


*/L\= 0.25 


x/L|= 0.50 


x/L = 0.75 


—0.05 0 0.05 0.10 0.15 


—0.10 


Computed longitudinal and transverse velocity profiles for the double-elliptic ship model for Ry = 


layer calculations were performed for a given invis- 
cid pressure distribution. The initial conditions 

at x = 22.5 m were generated by solving the boundary- 
layer equations in which the z-wise derivatives for 

a constant z were neglected. 

Figures 23 to 25 show some of the computed re- 
sults for R; = 3 x 108. Figure 23 shows the varia- 
TOM Cr Co, Cp Rg, Hy], and 8, at the cross-planes 
of x = 30 m, 105 m, and 210 m. Typical streamwise 
velocity profiles at x = 105 m and z = 0.2 are shown 
in Figure 24 and typical crossflow velocity profiles 
at x = 60 m are shown in Figure 25. As can be seen 
from these figures, the crossflow velocity profiles 
show great variations and indicate clearly the flow 
reversal that takes place in the crossflow plane. 
This implies that differential methods based on two- 
dimensional and/or small crossflow approximations as 
well as methods based on integral methods are not 
adequate to boundary-layer calculations on ship 
hulls. Other interesting results that emerge from 
these calculations are the sudden jumps of the limit- 
ing crossflow angle from positive to negative, and 
the thickening of the boundary layer in the corner 
region of the crossplanes. The jumps of the cross- 
flow angle indicates the convergence of the flow from 


FIGURE 19. Three-dimensional view of ship model 5350 
with the nonorthogonal coordinate system. 


204 


INPUT SHIP FORM 


° INTERPOLATION BY CUBIC SPLINE FUNCTION 


BOW SECTIONS 


FIGURE 20. 


Body plan for ship model 5350. 


both sides of the corner region and, hence, enhances 
the thickening of the boundary layer. This thicken- 
ing of the boundary layer in the corner region of 
ship hulls has been verified experimentally by Hoff- 
mann (1976). 


6.. CONCLUDING REMARKS AND FUTURE WORK 


According to the studies presented in this paper, 
the three-dimensional boundary layers on ship hulls 
can be computed very efficiently and effectively. 
The turbulence model, as in two-dimensional flows, 
again yields satisfactory results for three- 
dimensional flows. This has been demonstrated 

by Soejima and Yamazaki (1978) who also have applied 
the present turbulence model to compute three- 
dimensional boundary layers on ship hulls. However, 
there are additional studies and problem areas that 
need to be considered and investigated before the 
present method can become a more effective tool to 
design ships. They are briefly discussed below. 


WATER SURFACE 
STERN SECTIONS 


Generation of Initial Conditions on Arbitrary 
Bow Configurations 


In Section 5, we presented calculations for the ship 
model 5350 and mentioned that due to flow separation 
in the bow region, we had to start the boundary-layer 
calculations at some distance away from the bow. 
Additional studies are required to generate the ini- 
tial conditions on the bow. These studies can lead 
to a better design of bow configurations and to 
better handling of bilge vortices, which contribute 
to the total drag of the ship. However, this is by 
no means an easy task. Consider, for example, the 
ship model 5350 discussed earlier. A sketch of the 
bulbous nose with a plausible inviscid streamline 
distribution is shown in Figure 26. We assume 

that the ship is symmetrical about the keel plane 

and there is a nodal attachment point on the bulbous 
nose at B. If the ship is floating, then the water 
line is determined by conditions of constant pressure 
and zero normal velocity. Hence the intersection A 
of the plane of symmetry with the water line and the 


1.0 (WATER LEVEL) 


ia 0.85 
\ 
<[ 
0.50 ‘ 
—0.6 y 
7 
0.30 \ 
Px ry ‘ 
l] 7 0.15 \ 
0.2 \ 
7 1200 (KEEL) \ ‘ \ 
) \ \ \ 


FIGURE 21. Pressure 


entire 


distribution 


for the ship’ model 5350. 
’ 


Yi 


205 


evidence for this is based on a successful scheme 
that we have already worked out for the prolate 
spheroid, Cebeci, Khattab, and Stewartson (1978). 
Other aspects that need further study include the 
condition at the water-line section. It has been 
usual to assume that the normal velocity is zero at 
the undisturbed free surface. This is not quite 
correct and the error may have implications for the 
nature of the solution near A and especially the 
question of separation along BA. Even if separation 
does occur, it may be possible to handle the post- 
separation solution, since it probably extends only 
GIRTH, % over a limited region of the ship, by means of an 
interaction theory, i.e., modifying the inviscid 


0.2 flow by means of a displacement surface. 


Viscous-Inviscid Flow Interaction 
0.4 
The present boundary-layer calculations are done 

for a given pressure distribution obtained from an 
inviscid flow theroy. In regions where the boundary- 
layer thickness is small, the inviscid pressure dis- 
tribution does not differ much from the actual one; 
as a result, the boundary-layer calculations are 
satisfactory and agree well with experiment, see, 

for example, the papers by Cebeci, Kaups, and Moser 
(1976) and by Soejima and Yamazaki (1978). When 

the boundary-layer thickness is large, which is the 
case near the stern region, the effect of viscous 
flows on the inviscid pressure distribution must 

be taken into account. One possible way this can 

be done is to compute the displacement surface for 

a given inviscid pressure distribution and iterate. 
Such a procedure is absolutely necessary to account 
bow is a saddle point with the streamlines of the for the thickening of the boundary layer as was 
inviscid flow converging on A along the line BA and observed by Soejima and Yamazaki (1978). 

diverging along an orthogonal direction. It is 

known that the boundary-layer equations can always 


0.6 


0.8 


1.0 


FIGURE 22. Pressure distribution for the bow region 
of ship model 5350. 


be solved at B but that at A the situation is more Prediction of Wake Behind Ship Hulls 
complicated and furthermore it is still not entirely 
clear what their role is in relation to the general The present boundary-layer calculations can be done 
solution. It is likely, however, that provided no up to some distance close to the stern; after that, 
reversed flow occurs at A in the component of the flow separation occurs. Since one, and probably the 
solution along the direction BA, then separation can biggest, reason why one is interested in boundary- 
be avoided along this line by appropriate choice of layer calculations on ship hulls, is the calculation 
design. Furthermore, if separation does occur, its of drag of the hull, additional studies should be 
effect may be limited. The recently developed Cebeci- directed to perform the calculations in the separated 
Stewartson procedure (1977), however, can be applied region and in the wake behind the ship. Recent calcu- 
to the present problem but there are some hurdles to lation methods developed and reported by Cebeci, 
be overcome. Keller, and Williams (1978) for separated flows by 

Of particular difficulty is the choice of coordi- using inverse boundary-layer theory and recent calcu- 
nate system on which to compute the solution and to lation methods developed and reported by Cebeci, 
join it with the already well-established method Thiele and Stewartson (1978) for two-dimensional 
downstream of CD. We have seen that in the case of wake flows are appropriate for these purposes. 


the prolate spheroid (see Cebeci, Khattab, and 
Stewartson (1978)) it is helpful to have a mesh 


which is effectively Cartesian near the nose and the PRINCIPAL NOTATION 

methods which were used to produce it in the earlier 

study are applicable to any body which can be repre- A Van Driest damping parameter, see 

sented by a paraboloid of revolution in the neighbor- (18b) 

hood of the nose. Now here we have a paraboloid near A, ,A2,A3,Ay constants 

B but not one of revolution, but we believe that the CE local skin-friction coefficient in 
necessary generalization is possible. The mesh now streamwise direction 

has to match with that which has proved convenient C1, ,C2,C3,Cy constants 

downstream of CD. Again we believe that a smooth £ transformed vector potential for wp 
transition can be achieved by building into the g transformed vector potential for 
mesh sides, right from CBA, an appropriate spacing hy, ,ho metric coefficients 

such that the points of a uniform mesh on CD are hy net spacing in n-direction 

also points of this mesh although not, of course, at H,Hj] boundary-layer shape factor along 


~ a constant value of one of the coordinates. Our streamwise direction, 6*/8)} 


FIGURE 
Rg, and 8, for ship model 
for Ry, 


(b) 


x 


2) 


3 
3 


105m, 


23. Computed c 


pr 


10° at (a) 


and 


(c)) x 


Cc 


1.0 a a 


_50 GIRTH, % 


x = 105m 


Cy x 103 
$ 


Hiy 


(b) 


a 


eee 9 


KEEL 


—5° GIRTH, % 


—0.05 4 —l. —| z 


X = 210m 


o> 


SS SS SSS a 


me eee ee EE EEO 


3 

° 
Fad 
m 
m 


-50 GIRTH, % 


10) 50 100 150 200 


nl=(u,/¥s,)"/y) 


100 150 200 


nl=u,/vs,)"/2y) 


FIGURE 24. Computed streamwise velocity profiles for ship model 5350 for Ry, = 3 x 10” along (a) z = 0.2 and 


(b) x = 105m coordinate lines. 


150.0 


nl=(u,/vs,) '/2y) 


QOOOQOGO 


—0.04 —0.02 0 0.02 0.04 0.06 0.08 0.10 


FIGURE 25. Computed crosswise velocity profiles 
for ship model 5350. 


A.D FREE SURFACE D 


_——Se ae Cc 


207 


- 250 


FIGURE 26. Pattern of streamlines near the bow of ship 


model 5350. 


ky net spacing in x-direction 

geodesic curvatures, see (5) 

geometric parameters, see (6) 

L mixing length, see (18a), or refer- 
ence length 


mM] ,Mo,---M)2 coefficients, see (28) or (32) 

p static pressure 

Q total velocity in the boundary layer 

Ry Ry, Reynolds numbers, ugs)/v and u,L/v 

Rgx Reynolds number, ug _6*/\v 

Ro Reynolds number, Us,811/Y 

s arc length along coordinate line 

1 ,t2 unit tangent vectors along x and z 
directions 

u,V,W, velocity components in the x,y,zZ 
directions 

u,V,W velocity components in the Cartesian 
coordinate 

Ug Uy velocity components in boundary layer 
parallel and normal, respectively, 
to external streamline 

u. friction velocity, see (18c) 


Ugo freestream velocity 


ref reference velocity 
XT VIn nonorthogonal boundary-layer coor- 
dinates 
Se Cartesian coordinates 
-pu'v',-pv'w' Reynolds stresses 
B crossflow angle 
By limiting crossflow angle 
) boundary-layer thickness 
6* displacement thickness, 
oc 
Sa- us/Us_,) dy 
Exp eddy viscosity - 
eu dimensionless eddy viscosity, E/Y 
n similarity variable for y, see (21) 
811 momentum thickness, 
(oe) 
al Us/us, (1 = us/Ug,,) dy 
u dynamic viscosity 
v kinematic viscosity 
p density 
rT shear stress 
o,w two-component vector potentials, see 
(23) 
Subscripts 
e boundary-layer edge 
s streamwise direction 
t total value 
Ww wall 


primes denote differentiation with respect to n 


ACKNOWLEDGMENT 


This work was supported by the David. Taylor Naval 
Ship Research and Development Center under contract 
NOO014-76-C-0950. 


REFERENCES 


Chang, K. C., and V. C. Patel (1975). Calculation 
of three-dimensional boundary layers on ship 
forms. Towa Institute of Hydraulic Research, 
Rept. sNo 7.6). 


Cebeci, T. (1974). Calculation of three-dimensional 
boundary layers, pt. 1, swept infinite cylinders 
and small crossflow. ATAA J., 12, 779. 

Cebeci, T. (1975). Calculation of three-dimensional 
boundary layers, pt. 2, three-dimensional flows 
in Cartesian coordinates. ATAA J-., 13, 1056. 

Cebeci, T., and P. Bradshaw (1977). Momentum 
Transfer in Boundary Layers, McGraw-Hill/ 
Hemisphere Co., Washington, D.C. 

Cebeci, T., and K. C. Chang (1977). A general 
method for calculating three-dimensional laminar 
and turbulent boundary layers on ship hulls. 1. 
Coordinate system, numerical method and pre- 
liminary results. Rept., Dept. of Mech. Engg., 
California State Univ. at Long Beach. 

Cebeci, T., and A. M. O. Smith (1974). Analysis of 
Turbulent Boundary Layers, Academic Press, New 
York. 

Cebeci. T., and K. Stewartson (1977). A new 
numerical procedure for solving three-dimensional 
boundary layers with negative crossflow, to be 
published. 

Cebeci, T., K. Kaups, and A. Moser (1976). Calcula- 
tion of three-dimensional boundary layers, pt. 

3, three-dimensional incompressible flows in 
curvilinear orthogonal coordinates. ATAA J-, 
147, hOIO 

Cebeci, T., K. Kaups, and J. A. Ramsey (1977). A 
general method for calculating three-dimensional 
compressible laminar and turbulent boundary 
layers on arbitrary wings. NASA CR-2777. 

Cebeci, T., A. A. Khattab, and K. Stewartson (1978). 
On nose separation, paper in preparation. 

Cebeci, T., H. B. Keller, and P. G. Williams (1978). 
Separating boundary-layer flow calculations, paper 
submitted for publication. 

Cebeci, T., F. Thiele, and K. Stewartson (1978). 

On near wake laminar and turbulent shear layers, 
to be published. 

Gadd, G. E. (1970). The approximate calculation of 
turbulent boundary-layer development on ship 
hulls. RINA paper W5. 

Galbraith, R. A. McD., and M. R. Head (1975). Eddy 
viscosity and mixing length from measured bound- 
ary-layer developments. Aeronautical Quarterly, 
2OMAlbn. AB Shs 

Halsey, N. D. (1977). Potential flow analysis of 
multiple bodies using conformal mapping. M.S. 
thesis, Dept. of Mech. Engg., California State 
Univ. at Long Beach. 

Hoffmann, H. P. (1976). Untersuchung der 3- 
dimensionalen, turbulenten grenzschicht an einem 
schiffsdoppelmodell in windkanal. Institut ftir 
Schiffbau der Universitat, Hamburg, Bericht Nr. 343. 

Lin, J. D., and R. S. Hall (1966). A study of the 
flow past a ship-like body. Univ. of Conn., 
Civil Engineering Dept., Report No. CE66-7. 

Miloh, T., V. C. Patel (1972). Orthogonal coor- 
dinate systems for three-dimensional boundary 
layers with particular reference to ship forms. 
Iowa Institute of Hydraulic Research, Rept. No. 138. 

Rastogi, A. K., and W. Rodi (1978). Calculav.on 
of general three-dimensional turbulent boundary 
layers. AIAA J., 16, 151. 

Soejima, S., and R. Yamazaki (1978). Calculation 
of three-dimensional boundary layers on ship 
hull forms. Trans. West-Japan Soc. Naval 
Architechs, 55, 43. 

Vermeulen, A. J. (1971). Measurements of three- 
dimensional turbulent boundary layers. Ph.D. 
thesis, Univ. of Cambridge. 


Study on the Structure of Ship 
Vortices Generated by Full Sterns 


Hiraku Tanaka and Takayasu Ueda 


Ship Research Institute 


Tokyo, Japan 


ABSTRACT 


Many attempts have been made to measure the vortic-—- 
ity distribution of vessels tested at the Ship 
Research Institute. This led to the successful 
development of the rotor-type vortexmeter and a 
method for its calibration. In order to investi- 
gate the structure of the full ship stern vortices 
and gain an understanding of interaction of the 
vortices and propeller, the wake flow behind two 
geosim models was studied experimentally. 

Using this vortexmeter, detailed diagrams of the 
vorticity distribution are presented for the dis- 
cussion of the structure and scale effects on the 
stern vortices. The authors found the existence 
of a separating vortex sheet in the vorticity dis- 
tribution and indicated that, by using the vorticity 
concentrated on the vortex sheet (Max. line), it 
was possible to simulate the original vorticity 
distribution. With these experimental results the 
relation between the vorticity distribution and 
the propeller performance on the geosim models was 
also analyzed. 


1. INTRODUCTION 


In recent years, the knowledge of the wake structure 
including stern vortices has made it essential for 
the ship builder to obtain a better understanding 
of the stern vibration with full stern forms. 
Nevertheless, the stern vortex characteristics such 
as its geometry and structure as well as the scale 
effect remained obscure. This situation may be 
partially due to the fact that the stern vortices 
do not cause serious problems in the resistance 
augmentation or in the self-propulsion factors. 

To overcome this lack of detailed knowledge, 
systematic investigations have been made concerning 
the problems of full ship models with unstable 
propulsive performance. This research was begun in 
.1975 under the Research Panel SR 159 of the Ship- 


209 


building Research Association of Japan (Chairman, 
Prof. H. Sasajima) which was mainly concerned with 
the following areas: sources of the unstable 
phenomenon, the unsymmetrical flows accompanying 
this phenomenon, and the procedure for testing model 
ships exhibiting this kind of phenomenon. 

Throughout the Panel discussion there was great 
interest in the behavior of the stern vortices as 
the basic approach to understanding this phenomenon 
and this led to the request for quantitative data 
regarding the stern vortices. The major part of 
this paper was completed during the course of this 
Panel's activities in which one of the authors was 
placed in charge of developing a technique for 
measuring the fluctuating stern vortices. Asa 
result of the discussions, a rotor-type vortex- 
meter for obtaining a detailed description of the 
structure of the stern vortices was adopted. 

Needless to say, by obtaining an illustrative 
model of the stern vortices it will be possible to 
develop a mathematical model which will be extremely 
useful for understanding the flow around the full 
ship stern. Various vortex models have been sug- 
gested by Tagori (1966), Sasajima (1973), and 
Hoekstra (1977). The structure of the stern 
vortices can be roughly described by a stream line 
which, flowing upward around the bottom of the 
hull, separates at a separation line formed at the 
bilge. This flow rolls up at the boundary layer 
around the bilge forming a separated sheet with 
vorticity. 

Sasajima has suggested a simplified model of 
conical separating sheets as shown in Figure 1-1. 
He assumed that the separating sheet could be 
described by a triangular plane with which he 
attempted to explain the basic character of the 
stern vortices. This vortex model shown in Figure 
1-1 has a core enclosed with a separating line 
(S-S'), an attachment line (A-S') and the surface 
of the separating sheet. In this model it was 
assumed that the direction, velocity, and vorticity 
of the flow along this developed vortex sheet would 


DEVELOPED 
SEPARATING 
| SHEET 


-ATTACHMENT LINE 
SEPARATION LINE 
7S) ie 


FIGURE 1-1. [Sasajima (19 


FIGURE 1-2. 


SEPARATION LINE 


[Hoekstra (1977)]. 


Illustrative models of stern vortices 


have the same values as they had at the point the 
flow passed on the separation line. 

Hoekstra's vortex model also had a conical 
separating sheet with a cusp as shown in Figure 1-2. 
Although the stream along the separating sheet 
flows upward and rolls inside, it does not touch 
the hull surface to form an attachment line. 

In addition to the study on the scale effect of 
the stern vortices by Huse (1977), the studies 
based on the theory of the three-dimensional 
boundary layer by Okuno and Himeno (1977) has made 
it possible to discuss the detailed structure of 
the stern vortices. However, these studies did not 
pay much attention to the vorticity distribution. 
The authors concluded through their study that the 
prominent features of the stern vortices could be 
revealed by studying diagrams of the vorticity 
distribution. 


2. ROTOR-TYPE VORTEXMETER 


Although numerous efforts have been made to investi- 
gate the stern vortices, the state of the art for 
measuring the vorticity distribution in aft section 
of a model ship remains less developed than the 
techniques for measuring the wake distribution. 

This is evident by the few papers in which the 
complete data of the vorticity distribution has been 
published. This is largely attributable to problems 
in developing vortexmeters for towing tank measure- 
ments. 

In the authors' experience the problems in using 
five-hole Pitot tubes for measuring the vorticity 
have been in maintaining sufficient accuracy through- 
out the measurements. The analysis of vorticity 
distribution which includes finite difference 
methods results in insufficient precision. Besides, 
for one mesh point of a vorticity measurement, it is 
necessary to use the flow velocity data from four 
adjacent mesh points which makes it difficult to 
perform measurements close to the hull surface as 
well as to measure fluctuating vortex flows. 

The study of stern vortices has been greatly 
stimulated by flow visualization developments and 
especially noteworthy contributions have been made 
by researchers using tuft grid observations. 
However, flow visualization for observing the vortex 
flow has a weak point illustrated in the following 
discussion. 


Superimposing an arbitrary irrotational flow on 
a vortex flow, the resulting total flow should have 
the same vorticity as the original vortex. An 
example is shown in Figure 2 which is a velocity 
vector diagram of a circular vortex core super- 
imposed on a parallel flow. Examining this figure, 
it can safely be said that few people would be able 
to estimate an exact geometry or locate the center 
of the vortex from only this vector diagram of the 
total flow (or from a photo or sketch of the tuft 
grid observation). 

One of the authors [Tanaka (1971)] suggested 
adopting a rotor-type vortexmeter for towing tank 
measurements. He applied this technique to analyze 
the stern vortices generated by a submerged body 
running near the free surface. The application of 
the vortexmeter is reported in many aerodynamic 
investigations dating back to the 50's, and it was 
proposed for ship research by Gadd and Hogben [1962]. 

The vital problem in adopting the rotor-type 
vortexmeter for towing tank research lies in the 
accurate calibration of the rotor. This is mainly 
due to the fact that no one has succeeded in gener- 
ating a stable vortex useful for the calibration in 
a steady flow field. 

The rotor-type vortexmeter utilizes the principle 
that four-unpitched vanes mounted on a rotating 
shaft, shown in Figure 3, are not affected by any 
parallel and shear flow and only respond to a 


(A) (B) (A) +(B) 

Parallel circular vortex flow pattern on tuft grid 
flow flow 

FIGURE 2. Tuft grid pattern due to a circular vortex 


and a parallel flow. 


rotational Sheor 
Flow Flow 
(rotor) (Velocity vectors on rotor ) 


FIGURE 3. Principle of rotor-type vortexmeter. 


rotational flow. When the rotating shaft of a 
vortexmeter is parallel with vorticity axis, the 
rotor turns with angular velocity of W-S, where S 
is the slip due to rotational friction of the rotor 
shaft and W is the vorticity in the fluid. At 
present, the slip S can be estimated using the 
following technique. 

Using the simple consideration of the elementary 
wing, the torque Q due to a rotor element having 
small length dr in a radial direction can be deter- 
mined from Eq. (1), where C; , %, and U are lifting 
derivative, cord length of vane, and advance speed 
respectively. 


Q(x) = Ap LURC, gy rar (1) 


where 


ali Ssh ass 
LY Soya me oy souRUa) 


R. and L.F. are the rotor radius and lifting force 
respectively. In the flow with uniform vorticity 
distribution, the magnitude of the torque acting 
on the rotor becomes: 


R 


O(w) = 2 a Q(r)dr = (2/3) pLRPUC, ts (2) 
oR 


calibration - motor 


cl" ? 


(at calibration mode ) 


outer- tube 
inner — shaft 


rotor 
miniature ball bearing 


DIRECTIONS 


angular velocity 
modes 


in fluid rotor 
{ 


calibration 


vortex— 
measurement 


FIGURE 4. Principle of rotor-type vortexmeter calibra- 
tion. 


211 


Then, the slip of the rotor in a rotational flow 
can be determined by following equation, where 
q shows rotational friction of the shaft. 


S = d 


(273) p2R30C, (3) 


af 

For the calculation of S, the rotational friction 
of the ball bearings q should be determined experi- 
mentally. This problem will be briefly discussed 
later. 

As previously stated, since the generation of a 
stable vortex for the calibration is presently not 
feasible, a mechanical calibration was attempted in 
which vorticities mechanically act on the rotor 
through the shaft of the rotor. This principle of 
the calibration is shown in Figure 4 where a newly 
designed rotor shaft is composed of duplicate inner- 
shaft and outer-tubes. The outer-tube is mounted 
on the outer rings of the ball bearings and the 
vanes are fitted on the outer-tube. The inner- 
shaft is connected to a calibration-motor. 

To obtain the slip S, the vortexmeter is cali- 
brated in an irrotational flow in which it travels 
along at a constant speed. The inner-shaft is 
driven by the calibration-motor at an angular 
velocity, w, and the rotor turns at an angular 
velocity, S, in response to the condition of the 
ball bearing's frictional torque and the hydro- 
dynamic characteristics of the rotor. 

From the measured vorticity, Wg, we can estimate 
the vorticity in fluid as w = Wo + S. According to 
the authors' experience, if the frictional torque, 
q, is approximately 10-6 kg-m it is possible to 
consider S = O except in the case of fairly slow 
speed (cf. Figure 25). This means that the 
calibration of the vortexmeter seems unnecessary 
for ordinary test conditions. 

Although ball bearings exhibiting frictional 
torque values less than gq = 0.7 107-®kg-m in air were 
chosen in manufacturing the vortexmeter, there was 
no direct measurement of the frictional torque of 
the miniature ball bearings in water. The frictional 
torque, gq, also can be determined by measuring the 
torque on the outer-tube generated by inner-shaft 
turning in water. According to the results of these 
measurements, it can be said that there is hardly 
any difference between the frictional torque value 
of the bearings when they are used in water or 
air. 

An example of a vortexmeter is shown in Figures 
5 and 6. The diameter and length of the rotor are 
30mm and 18mm respectively, section of the vane 
is lenticular shaped with a thickness ratio t/2 
= 1/q. A transducer for rotating the rotor is 
used in connection with a photo-transistor which 
makes 4 pulses-signals in one revolution. Assuming 
Cry = 0.67, g = 107°kg-m Einel Gf = 155 Wei, ae aig 
possible to make a rough estimate of the vortex- 
meter's precision from the value of slip obtained 
by Eq. (3). From these values, the slip value, S, 
equals 1072r.p.s.- which corresponds to 1% error 
relative to a normal vorticity of w = 1 r.p.s. 

As will be mentioned later, the vortex cores of 
the stern vortex near the hull surface have a very 
steep gradient in vorticity distribution. There- 
fore, it is useful to consider the vorticity values 
measured by the rotor with a finite diameter at 
such boundaries. It is clear from the Eq.(1) that 
a mean value of a torque during a turn due to a 
wind element dr (see Figure 3) corresponds to a 


ise) 
H 
N 


200 
7380 


H 
Cs) 
ies} 
ea) 


Rotor-type vortexmeter. 


mean value of the vorticity in a path of a wing 
element. 


Q(x) = %p2UC, x w(x) dr (4) 
27 277 

Oe) = aah (9,r)d8 (x) = i w(6,r) ae 

= aa Or Ohler DN Oe m 
0 0 


On the other hand, concerning the vorticity gradient 
influence on the radial portion of the rotor, from 
the following equation it can be understood that 

the tip of the rotor has a higher sensitivity: 


R R 
2 ff Q(x) ar = pLUC, F f rw (x) dr (5) 
-R -R 


In practice, using only large models, the error due 
to the finite diameter of the rotor can be eliminated. 
Such a problem is also present when determining 

the mesh interval in the vorticity measurement 

by a five-hole Pitot tube. 


3. EXPERIMENTS AND RESULTS 


The ship models used in the experiments exhibited 

an unstable propulsive performance in ballast 
condition. In recent studies, it has been recog- 
nized that the limiting stream line around the stern 
and the pressure distribution change along with the 
thrust fluctuation in the self-propulsion tests 

of the model ship. The influence of these 

phenomena on the ship design has been reported by 
Watanabe and Tanibayashi (1977) and Watanabe et al. 
(1972) . 

A special feature of this phenomenon was that it 
appeared only in the self-propulsion tests and was 
not observed on the towing tests. Thus, while this 
phenomenon easily appeared in the self-propulsion 
tests at Froude number 0.18 and 65% full displace- 


ment, at the same conditions there was no indication 
of this phenomenon during the tests concerned in 
this report. The body plan of the 4 and 7m geosim 
models are shown in Figure 7 and the principal di- 
mensions are summarized in Table 1. 

The intent of the experiments was twofold: first 
to determine the structure of the stern vortices 
using the rotor-type vortexmeter, and secondly, 
to investigate the performance of the propeller 
working in the presence of these stern vortices. 


FIGURE 6. Rotor-type vortexmeter and stern of model. 


213 


rey 


\ \. \ ak 


FIGURE 7. Body plan of model. 


Also as a reference, the vorticity distribution 
was measured by the five-hole Pitot tube for com- 
parison with the rotor-type vortexmeter measure- 
ments 

The positions where the vorticity distribution 
of the stern vortices was measured, are shown in 
Table 2 in which sq.st. 1/8 correspond with the 
section of the propeller disk. The effects of 
velocities of the model ship are studied at several 
mesh points. 

In order to discuss the scale effect of the 
vorticity distribution, the results of measure- 
ments on both models are shown in Figures 8 and 9, 
and the induced velocity vectors on the Y-Z plane 
which are calculated with the vorticity distribu- 
tion, are shown in Figures 10 and 11. 

The interval of mesh drawn on both diagrams of 
vorticity distributions and velocity vectors, 
corresponds to a non-dimensional length of 0.5% 
Lpp. The values of equi-vorticity contours in 
diagrams of vorticity distribution are non- 
dimensional vortices defined as follows: 


Bue) $2 
E = a ee (6) 


where wy shows the vorticity in r.p.s. which 
corresponds to twice the number of rotor revolu- 
tions. Considering a diagram of vorticity dis- 
tribution as a geographical contour map, the vortex 
core can be compared to a typical plateau. The 


TABLE 1 Principal Particulars of Models 


Model Ship No. M-7 M-4 
Length (m): Lpp 7.000 4.000 

4 Breadth (m) 1.167 0.667 

2 Breadth Draft Ratio 2.760 
Block Coefficient 0.802 
Longitudinal Prismatic Coeff. 0.810 

4, Pitch Ratio (const.) 0.7143 

“ Boss Ratio 0.180 

B Expanded Area Ratio 0.665 

Ai Number of Blades 5) 


TABLE 2 Measurement Positions of Vorticity 


Distribution 
Sq. “Sit 
Model 1/2 1/4 1/8 eb Bexs) 
Port £37) 
= Port 
M-7 Port Port St aeboara or 
eed Port x2. 
Starboard 
Notes 


*]1 Corresponds to propeller position 
*2 Measured by vortexmeter and 5-hole Pitot tube 
*3 Corresponds to Sq.St. -1/8 


fact can clearly be seen in the foreward detections, 
especially sq.st. 1/2 in Figure 12. In this 
connection, the vorticity distribution is sq.st. 1/4 
and A are presented in Figures 13 and 14 
respectively. 

As a reference, the induced velocity vector 
diagram on sq.st. 1/4 is shown in Figure 15. 
Furthermore the vorticity distribution (for 
M.No.M-7) obtained by the five-hole Pitot tube 
(diameter 12mm, angle between center and side 


FIGURE 8. Vorticity distribution of M.No.M-7 at 
Scot lS 


\e) 

Y 2 Y 

io 2 8 6 =o Bee ©) 4 8 

8} 8 
e Zz 

> 6 3 
4 4 
Ot Oe: ) 
1! aon J} ~ 
\ RAO A) |) 

-2} i An 1 Lf of f | 2 z 
% Med ¥ f/f 
- oF LSD) Of LY 

-4 - = 4 Oh da f fy ' 4 

\ \C Tape / 
6 ~. = Si. 6 
mh a XN oe 
x SZ » 
1% InSESS af ee Wie 
et ae 8 


Be co BN 4h 3 . \ 
y LE Lee | j re SS 
2 =~ Mth tn — Sea 
‘ ‘ Uae | é \ | | 
f H a a " ‘ ' 
\ | ea, , 
o} ®, 4 44-4 Se Tele 
i VW AS anes Vy iy PG 
\ \ ae Le), , 
y Hs) = a = /-2 
JN i me IT \ Lt Ww 
Nt ol VENV aie 
Rae ae dele Lie , 
Sa ff Se s | 4 a ea 2 
a Ce 
ae y | \ x ee aS 
Qi Se Ay ye a ee.) 


FIGURE 10. Velocity vectors on y-z plane due to stern 
vortices (M-No.M-7, Sq.St. 1/8). 


“ = zi z e 1 
Pool A gS yh Op phe ete 
PSS Sey Ray Yh ll edhe a RS Le 
Oat oa mt CEN i Wa tea ae ae YB a no 
Diet Diem Sy. | | a Lae reel eee 
6 = a a I fe Te 6 
a Se i Les SoS ~ 
TA -— a ae a ae S% 
/ Naat Alearet (i Soar s ‘ 
2t Thole esi TH ise Lace A A152 
I ate y 14 i Vy 
n | "Ay Te ea a y e 
HL TEEEBREE ah 
\ J z 
3) Ve fyeh | A| | | \ ‘ 
2 ~ = Se 
Bas aC AL AR So d 
; oa Kegel Hd bend MONS rcrsiece 
=4s = | ye Wei, <4 
Tee aan ae Te Rall le eee 
ee Sa Hes wwsSN——- 6 
8~ aa THAN = - «8 


Velocity vectors on y-z plane due to stern 
(M.No.M-4, Sq.St. 1/8). 


FIGURE 12. Vorticity distribution of M.No.M-7 at 
SqSt..- 17/2) 


~ 


—— 


lye 


FIGURE 13. Vorticity distribution of M.No.M-7 at 
SquSteml/4p 


CHIE 
Ha 


NY 


SaS> 41 rees 


‘\ m 


FIGURE 14. Vorticity distribution of M.No.M-7 at 
Seis. As 


holes 25°) is shown in Figure 16 and the wake 
distributions and the velocity vectors by the five- 
hole Pitot tube are shown in Figures 17, LB} p | I), 
and 20 respectively. 


FIGURE 15. Velocity vectors on y-Z plane due to stern 
vortices (M.No.M-7, Sq-St. 1/4). 


215 


A 
APR P 
lg 
uy 
Viet 4 ~ 
Olacmmat + i i 
Y 
8 Bes 
Zz 
6 
\ hs 


Os 


-BL 


FIGURE 16. Vorticity distribution measured by 5-hole 
pitot tube (M.No.M-7, Sq.) Sit.) 178). 


4. DISCUSSIONS AND APPLICATIONS OF THE RESULTS 


Remarks on Vorticity Measurements 


The rotor-type vortexmeter performed as expected. 
As seen from a comparison between Figures 8 and 16, 
the rotor-type vortexmeter is more sensitive and 
can be used to obtain a finer vorticity distribu- 
tion contour than the five-hole Pitot tube. While 
both vorticity distribution diagrams appear to have 
a similar shaped vortex core, they have fairly 
different values. The distinguishing difference 

is mainly in the pattern of the distribution. Al- 
though the plateau-type distribution would be the 
expected form of the typical vortex cores in the 
vorticity distribution obtained by the vortexmeter, 
the plateau-type is broken in Figure 16. It can be 
said that the difference between these results 
indicate the usefulness of the vortexmeter's resolv- 
ing ability. 

Contrary to the general opinion that a geometry 
of the stern vortices is fluctuating, in the 
authors' measurements, the vorticity and geometry 
of the stern vortices were generally quite stable. 
However, there is an unstable vorticity-zone at 
the top of the main vortex core indicated in 
Figures 8, 9, and others. Through these experi- 
ences, it can be shown that the dynamic character 
of the vortexmeter is one of its prominent features. 

While the present diameter of the vortexmeter's 
rotor was selected for maintaining its accuracy in 
measurement, it is possible that the rotor diameter 
is too large for the 4m geosim model (M.No.M-4). 
Furthermore, it appears that there were some 


4 6 8 10 


nm 


“10 -8 -6 Y -4 -2 0 


06! 0.5) 0403! 10.30.4105 6 Oz. 


FIGURE 17. Wake distribution (M.No.M-7, Sq.St. 1/8). 


problems due to the presence of an oblique flow in 
those experiments. It is recommended in further 
work that the characteristics of the rotor ina 
strong oblique flow should be studied. 


Structure of the Stern Vortices 


As stated in the previous section, the equi- 
vorticity contours of the stern vortices can be 
compared with plateaus in geographical contours. 
Furthermore, in examining carefully the diagrams 
of the vorticity distribution, there is a line of 
concentrated vorticity on the "table of plateau," 
which is denoted by the "Max. line" in this paper 
and indicated in the contours. The Max. line 
can be clearly shown in a cross section of the 
diagrams of the contour as seen in Figures 21 and 
22. 

The Max. line can be considered as a kind of 


~10  -8 -6 Y -4 -2 0) 2 4 6 8 10 
; '05 04103 02) 020304 05| | 06 | 


FIGURE 18. Wake distribution (M.No.M-4, Sq.St. 1/8). 


-10 -8 -6 -4 = ce} 2 4 6 8 10 
Pee I, ties a ae eed ae 
| Q__ 20%. of U 
8 L { { X = 
240 ae a aibe So TSS TS ~ 
6 : L- Tk er ae oe =o 6 
i IN oe ay 
s+ a en eee 
te + 


ir 
—, 
N 


Sees | ee 
FIGURE 19. Velocity vectors on y-z plane (M.No.M-7, 
Sqisteawl/A8)) ie 


a ridge on this plateau; it is steep in the forward 
section and becomes gently sloping while shifting 
afterward. It is noticeable that the Max. line 
seems to show the existence of a separating vortex 
sheet. As is well known, stream lines flow from 
under the bottom of a hull up the boundary layer 

at the bilge and turn into part of the vortex 
sheet. Although the vortex sheet previously 
mentioned has only been used as a hydrodynamic 
description, the authors are able to show its 
existence in the flow behind the full stern as well 
as provide quantitative measurements. 

The development of the vortex sheet depends 
mainly on the potential flow and the induced flow 
from the vorticity. Its development is strongly 
affected by each ship form, with effects of model 
ship velocity and the Reynolds number effect 
mainly limited to the diffusion of the vorticity. 
In a comparison between Figure 8 and Figure 9, the 
forms of the Max. line which correspond to a form 


FIGURE 20. Velocity vectors on y-z plane (M.No.M-4, 


SS (Sa.5t., 
a ete \s, STARBOARD 
(Sq.St.4, PORT) / ‘ : 


0 = Vp 0) Sy NP NSS 7182) 
qT OCR OF Zot oS oF 
-20 ke v4 Z.Y=!= 35mm MS.NQM-7 
-40 SNe = 20mm M.S.NOM-4 
-60 


FIGURE 21. Cross section of vorticity distribution 
(M.No.M-7 and M-4, Sq.St. 1/8). 


of the vortex sheet are fairly similar for both 
geosim models. On the other hand, the difference 
in breadth of each model's vortex core seems to be 
due to the effect of difference in Reynolds number. 

Furthermore, the suitability of adopting the 
idea of the Max. line is shown by the following 
facts. Assuming that all the vorticity of the 
stern vortices are concentrated on the Max. line 
for computing the induced velocities, the resul- 
tant velocity vector diagrams are similar to the 
complete flow field velocity. For instance, 
Figure 23 is a diagram of velocity vectors, which 
have the same circulation value as Figure 8 but 
with the vorticity concentrated on the Max. line 
divided by ten, of circular vortices with mean 
strength on the original Max. line. It can be 
seen that both diagrams of the velocity vector, 
Figure 10 and Figure 23, are fairly similar. This 
will allow not only simplified treatment of the 
stern vortices but also should simplify future 
numerical analysis of the stern vortices. 

In order to predict the wake of full stern 
ships, it is necessary to estimate the wake com- 
ponent due to the stern vortices in addition to 
the potential and frictional wake components used 
in Sasajima's wake prediction method. The concept 
of the Max. line in the vorticity distribution 
also may lead to the wake component due to the 
stern vortices. 

In order to discuss the relation between the 
stern vortices and the wake distribution, an 
illustrative model of the stern vortices is 
presented in Figure 24. A stream line flowing 
under the bottom of a ship, separates around the 
bilge and forms a part of the separating vortex 
sheet. The vortex sheet crosses to the hull 
surface near the propeller bossing where the 
authors denote the secondary separation line. 

And at the secondary separation line, the vortex 


217 


(Sq.st. +, PORT) 
Y 


-6  -4 


Z-| 


Z2-2 


FIGURE 22. Cross section of vorticity distribution 
(M.No.M-7, at Sq.St. 1/2 and 1/4). 


sheet makes the cross flow with the limiting stream 
line flowing aft passing through the tunnel of 

the vortex sheet. The crossed flow generates a 
reversed vortex at the secondary separation line 

as seen in the diagrams of the vorticity distribu- 
tion. 

The flow passing through the tunnel of the 
vortex sheet can be found at the section of the 
propeller disk (sq.st. 1/8) which appears as an 
eye in the wake distribution pattern in Figures 17 
and 18. This fact may be proved by the Max. line 
which just covers the eye. 


FIGURE 23. Velocity vectors due to concentrated vor- 
ticity on max. line (M.No.M-7, Sq-.St. 1/8). 


\72 


SEPARATING 
FLOW 
+LIMITING 
STREAM LINE 
SECONDARY 
SEPARATION LINE > cCPARATION CINE 
FIGURE 24. Illustrative model of stern vortices. 


This tunnel vortex sheet is quite different from 
the conical vortex sheet used in the model proposed 
by Sasajima or Hoekstra. Considering the flow as 
passing through this tunnel makes it possible to 
discuss the relationships of the wake flow, limit- 
ing stream line, attachment line, and the stern 
vortices. 

Regarding the wake patterns of vessel with a 
full stern, the authors suppose that if the Max. 
line can be considered independent of the Reynolds 
number, then the "eye" in the ship's wake pattern 
should be in approximately same location as shown 
in Figures 17 and 18. The above mentioned facts 
will lead to further studies for prediction of 
ship's wake, using the potential and frictional 
wake patterns estimated by Sasajima's method. 

Actually, the authors cannot verify the 
relationship between the stern vortices and 
Reynolds number because the range of the scale 
ratio used in geosim models tested is too small 
for a discussion of the similarity of the stern 
vortices. However, it can be said that the 
alternation of the Max. line between both models 
seems relatively smaller than that of the wake 
pattern. Furthermore, the vortex center, which is 
defined as the vanishing point of the induced 
velocity vector due to the stern vortices, has 
shifted a distance corresponding to only 4% of the 
propeller diameter as seen in comparing Figures 10 
and 11. While the model size has comparatively 


for r/R= 0.50 


Vx/U 


Vx/U for r/R= 0.90 


FIGURE 26. Circumferential distribution of 
wake flow on propeller disk, V,/U. 


06 


0.4 


0.2 


0.6 


0.4 


0.2 


Relation between vorticity and velocity 
of model ship. 


FIGURE 25. 


small effect on the shape of the Max. line, the 
model size causes differences in the diffusion of 
the vorticity. Thus, from the calculation of the 
circulation of the vortex cores presented in 
Figures 8 and 9, it was found that the circulation 
of M.No.M-4 was smaller than M.No.N-7. The magni- 
tudes of these differences were 6% smaller on the 
portside and 8% smaller on the starboard side of 
M.No.M-7. However, even for the same model ship, 
the difference in the port and starboard side 
stern vortex circulation was on the order of 8%, 
so it is not possible to reach a definite conclu- 
sion about the significance of the differences in 
the geosim tests. 

Since the authors limited study to vessel speeds 
corresponding to Froude number 0.18, the effects 
on the velocity due to the stern vortices still 
remains obscure. However, the authors can in- 
dicate some examples in which the vorticity has 
been measured at the several mesh points as seen 
in Figure 25. If the free surface effect could be 
neglected, the non-dimensional vorticity €y should 
be constant. Although the cause of the different 
results explicitly shown in Figure 25 remains un- 
known, it may not be said that the rotor-shaft 
friction of the vortexmeter can be safely considered 
as negligible in a range of very slow speeds such 
as ES 0.1. 


r/R =0.50 


Vx/U for (/R*0.70 


Effect on Propeller Operation Due to Tangential 
Stern Vortex Flow 


In the previous section, the authors have mainly 
discussed the structure of the stern vortices 
obtained from the towing experiments. As was 
reported by Hoekstra (1977) it can be considered 
that the structure and geometry of the stern 
vortices is strongly affected by the flow induced 
propeller thrust. However, the authors have studied 
the forces and moments on the working propeller as 
a preliminary problem, assuming the structure of 
the stern vortices is not changed by the influences 
of the propeller suction. 

The forces and moments on the propeller are 
remarkably related to the pattern of the flow 
distribution at the propeller disk location. The 
flow distribution relevant to the present problem, 
is composed of the wake component, V_/U, and the 
tangential components, V_/U, which were obatained 
by the five-hole Pitot tube. The authors assumed 
that the tangential components could be further 
decomposed into the component obtained by the 
vortexmeter, V__/U, and other components. Although 
each component has already been shown in previous 
figures, for convenience the circumferential dis- 
tributions of V_/U, V_/U, and V__/U at 90%, 70%, 
and 50% of the Sisk radius are ENewn respectively 


Vr/U :tangential velocity component 
obtained by five-hole Pitot tube /R =0.50 


219 


in Figures 26, 27, and 28. Furthermore, the authors 
have included the tangential velocity vector com- 
ponent, Viny/ Ue in Figure 29. 

In order to determine the propeller forces and 
moments induced by the stern vortices, the authors 
have performed the following calculations using 
the unsteady lifting surface theory developed by 
Koyama (1975). The authors thus calculated the 
thrust and torque of the propeller, along with the 
vertical and horizontal forces and moments imparted 
by the propeller shaft of the working propeller 
with and without stern vortex flow. The definitions 
concerning the forces and moments are shown in 
Figure 30. 

The authors have assumed for the calculation that 
the tangential flow obtained from the subtractive 
procedure (V_/U - V__/U) simulates one eliminating 
the effect oie the of Som vortices, and a common 
wake flow can be used for both calculations with 
and without the stern vortices. 

Since the results of the calculation for M.No.M-4 
are quite similar to the results of M.No.M-7, only 
the results of M.No.M-7 are shown in Figures 31 and 
32. Figure 31 indicates a comparison of the torque 
and thrust on a blade of the propeller with and 
without the stern vortex flow. Total torque, 
thrust, and other forces and moments on the pro- 
peller (indicating propeller turning angle 0° to 


Top Starboard side 
n 


Top FIGURE 27. Circumferential dis- 


{o) 90 180 


03 


Vrv/U tangential velocity component 
obtained by vortexmeter 


clockwise 


anti - 


tribution of tangential flow on 
propeller disk, Vp/U. 


clockwise 


FIGURE 28. Circumferential dis- 
tribution of tangential flow on 


Top Starboard side Bottom Port side Top 
L 1 i 
{e) 90 180 

@ (deg ) 


360 propeller disk (induced flow), 
Vpy/U- 


220 


ne tl ba 


ih ss eS Vv/U 


YP Lits 
180 5 as 


FIGURE 29. Tangential velocity vector due to induced 
flow on propeller disk (M-7). 


72°) are shown in Figure 32. According to the 
results, main effects of the stern vortices flow 
appeared on the vertical force (F_) and the 
horizontal bending moment (M_) of the propeller, 
but the other components are almost negligible. 


It can be concluded that the effect of the stern 
vortices is fairly limited to a few components of 


forces and moments generated by propeller. The 
results may be attributed to the tangential flow 


around the propeller caused by the stern vortices. 
It is mainly concentrated at the underside near the 


bossing, and does not severely appear on the 
propeller tip as shown in Figure 29. 


5. CONCLUSION 


The authors developed the rotor-type vortexmeter, 


giving careful attention to the calibration method 
of the vortexmeter, and, by using it in these tests, 


showed its high utility. 


PIGURE 30. Definitions of forces and moments due to 
propeller. 


0.070 


0.060 


0.950 


(T7e.n?- D4) 201 


0.040 


0.0070 


ze 


0.0060 


0.0050 


(a /p-r?. 08) 


0.0040) , F 
1 
Top Starboard side Bottom Port side op 


(e) 90 180 270 360 
@ (deg.) 


with stern vortices 


areas without stern 
vortices 


FIGURE 31. Variations of thrust and torque 
coefficients on one blade (M-7). 


—— with stern vortices 
---- without stern vortices 


0.27- 0.029 - 


Thrust 


way 
Pe 
I= 
\a 
EN 
. Horizontal bending Mt. 
Vertical force i 9 
iN 
4 \ 
¢ XN , 


Vertical bending Mt. 
Horizontal force 


0005; , ,-~0:005 


Top @(deg.) 
(et ti) 
0 30 60 72 
clockwise clockwise 


FIGURE 32. Variations of force and moment coeffi- 
cients on propeller (M-7). 


Measuring the vorticity distribution around the 
full stern of the geosim models, the authors 
determined the structure of the stern vortices 
and found the presence of a concentrated vorticity 
line in the vortex core which corresponds to the 
separating vortex sheet of the stern vortex. 

As an application of the results, the effect on 
the propeller operation due to the induced flow of 
the stern vortices has been studied. The effect is 
fairly limited to a few components of forces and 
moments generated by the propeller. Consequently, 
it can be said that the effect of the stern 
vortices on the performance of the propeller and 
propeller excited vibratory shaft forces and 
moments is relatively small. However, in the case 
of this ship model, this effect appears to change 
the direction of the vertical force and the 
horizontal bending moment acting through the pro- 
peller shaft. 


ACKNOWLEDGMENT 


This research program has been carried out mainly 
under the foundation of the Research Panel SR 159 
of Shipbuilding Research Association of Japan. 
The authors would like to express their apprecia- 
tion to Profs. H. Sasajima and I. Tanaka, who led 
the Panel, for their valuable suggestions. 

The materials described in the present paper have 
been kindly prepared by the members of the Ship 
Research Institute among whom the authors are 
especially grateful to Mr. K. Takahashi, Mr. T. 
Haraguchi, Mr. Z. Ishizaka, Mr. N. Sugai, and 
Miss H. Handa, for their helpful support. 


REFERENCES 


Gadd, G. E., and N. Hogben (1962). An appraisal of 
the ship resistance problem in the light of 
measurements of the wake pattern. WN.P.L. Ship 
Rep. 36. 

Hoekstra, M. (1977). An investigation into the 
effect of propeller hull interaction on the 
structures of the wake field. Proceedings of 
Symposium on Hydrodynamics of Ship and Offshore 
Propulsion Systems. 

Huse, E. (1977). Bilge Vortex Scale Effect. Pro- 
ceedings of Symposium on Hydrodynamics of Ship 
and Offshore Propulsion Systems. 

Koyama, K. (1975). A Numerical Method for Propeller 
Lifting Surface in Non-Uniform Flow and Its 
Application. Journal of the Society of Naval 
Architects of Japan, 137, 78 (Japanese). 

Okuno, T., and Yoji Himeno (1977). Distribution of 
Wall Shear Stress and Gross Flow in Three- 
Dimensional Turbulent Boundary Layer on Ship 
Hull. Journal of the Kansai Society of Naval 
Architects of Japan, 165, 83 (Japanese). 

Sasajima, H. (1973). On scale effect of ship 
resistance, Text of Symposium on Viscous 


Resistance. The Society of Naval Architects of 
Japan, 213 (Japanese). 
Tagori, T. (1966). Investigations on vortices 


generated at the bilge. Proceedings of 1lth 
Fen WAC BAG 

Tanaka, H. (1975). A study of resistance of 
shallow-running flat submerged bodies. Selected 
Papers from Journal of the Society of Naval 
Architects of Japan, 13, 15. 

Watanabe, K., K. Yokoo, T. Fujita, and H. Kitagawa 
(1972). Study on flow pattern around the stern 
of full ship form by use of the geosims. Journal 
of the Society of Naval Architects of Japan, 131, 
Se 

Watanabe, K., and H. Tanibayashi (1977). Unusual 
phenomenon of the stern of full ship models. 
Proceedings of Symposium on Hydrodynamics of 
Ship and Offshore Propulsion Systems. 


Session [IV 


SHIP BOUNDARY LAYERS 
AND 
PROPELLER HULL INTERACTION 


TAKAO INUI 

Session Chairman 

The University of Tokyo 
Tokyo, Japan 


eit 
By 


Wake Scale Effects on a 
Twin-Screw Displacement Ship 


Arthur M. Reed and William G. Day, Jr. 
David W. Taylor Naval Ship Research and Development Center, 


Bethesda, Maryland 


ABSTRACT 


The results of a wake survey and boundary layer 
profile measurements on a full-scale twin-screw 
displacement ship are presented. The corresponding 
model-scale measurements are also presented. The 
full-scale wake measurements consist of the three 
velocity components which contribute to the nominal 
wake in the propeller plane, at four radii. The 
full-scale boundary layer profile was obtained at 
three longitudinal locations with and without the 
propeller operating. The model-scale nominal wake 
was determined in a towing tank using five-hole 
pitot tubes while the model-scale boundary layer 
measurements were made on a double model in a wind 
tunnel using hot wire anemometers. 

In order to identify the scale effects between 
the model and ship, the deviation of the velocity 
in the propeller disk from a uniform axial flow has 
been separated into the velocity field due to shaft 
inclination in a uniform stream, the perturbation 
due to the hull and its boundary layer, and the 
viscous wake due to the appendages. The principal 
contribution to this perturbation from the axial 
flow is the effect of inclining the shaft in the 
uniform stream. The perturbation of the flow due 
to the potential flow about the hull is small, as 
are the effects of the displacement thickness of 
the boundary layer of the hull. The proposed 
scheme for predicting the viscous wakes of the 
shaft and struts meets with little success. Never- 
theless, some conclusions are drawn as to how these 
wakes will vary between the ship and model. 


1. INTRODUCTION 

If unsteady propeller force and hull loading pre- 
dictions are to be precise, the inflow to the pro- 
peller must be known accurately. At the present 
time the nominal wake of a model is measured and 
extrapolated to full scale assuming geometric 


225 


Similarity. The extrapolation fails to take into 
account any of the scale effects which may possibly 
exist between model and full scale. This paper 
presents preliminary results from a series of full- 
scale nominal wake and boundary layer velocity pro- 
file measurements on a high-speed transom-stern 
ship. In addition, the corresponding model-scale 
measurements are reported, along with a series of 
analytical predictions, which are intended to 
identify the principal contributions to the wake. 

This is not the first investigation of this 
nature. However, it is the first project to suc- 
cessfully measure the three velocity components in 
the propeller disk of a high-speed twin-screw 
transom-stern hull form. The British have per- 
formed an extensive series of experiments on a 
frigate, [Canham (1975)], and the Japanese and 
Germans have performed flow measurements on 
several full-form ships. The Japanese and German 
experiments were conducted on single screw tanker 
forms and are reported in an extensive series of 
reports [see for instance: Namimatsu et al.(1973), 
Namimatsu and Muraoka (1973), Schuster et al.(1968), 
Takahashi et al.(1970), Taniguchi and Fujita (1970), 
and Yokoo (1974)]. 

While the British measurements were obtained on 
the ship type of interest, a high-speed transom- 
stern ship, only the longitudinal velocity compo- 
nent in the propeller plane was obtained. This 
resulted in the loss of the important tangential 
and radial velocity components. In the case of 
twin-screw transom-stern ships, these velocity 
components are generally very significant due to 
the inclination of the shaft to the direction of 
the free-stream. 

The Japanese, on the other hand, were able to 
measure all three velocity components in the wake, 
but they had to make their measurements in a plane 
ahead of the propeller disk. Due to the full 
sterns of the tankers, the flow into the propeller 
is highly influenced by viscous effects, and as a 
consequence is highly affected by changes in 


226 


Reynolds number. Therefore, while treating a much 
more difficult problem, the results of the tanker 
experiments are not applicable to the scaling of 
the wakes of high-speed hull forms. 

The full-scale velocity component ratios which 
are presented here were obtained at a speed of 15 
knots; the corresponding Froude and Reynolds numbers 
were 0.36 and 4.10 x 108 respectively. The model 
wake survey was conducted in a towing tank at the 
full-scale Froude number. This resulted ina 
model speed of 5.22 knots, and a Reynolds number of 
1.56 x 107. The full-scale boundary layer measure- 
ments were conducted at four speeds between 6.2 and 
16.5 knots. These speeds correspond to Reynolds 
numbers between 1.7 x 108 and 4.5 x 108 respec— 
tively. The model-scale boundary layer measure- 
ments were obtained on a double model in a wind 
tunnel at a Reynolds number of 1.68 x 107. 

Significant differences are observed between the 
model and full-scale velocity components, particu- 
larly in the magnitudes of the radial and tangen- 
tial velocity components. These differences are 
in the regions away from the ship's hull and 
appendages; therefore, these differences do not 
seem to be due to Reynolds number effects. A more 
likely explanation is a lack of ship-model simi- 
larity, possibly due to unexplained differences in 
hull form or initial trim. 

In order to obtain an understanding of the com- 
ponents which contribute most significantly to the 
deviation of the wake from uniform axial flow, an 
attempt has been made to predict the velocity com- 
ponents as seen by the propeller. To make this 


prediction, the velocity field (in shaft coordinates) 


was decomposed into its major components as follows: 


Velocity = Uniform Stream 
+ Perturbation due to Hull 
+ Perturbation due to Hull Boundary 
Layer 
Viscous Wake of Struts 
Viscous Wake of Shafting 


+ + 


The results of this decomposition show that the 
inclination of the propeller shaft to the free 
stream is the most significant factor contributing 
to the deviation of the velocity from a purely 
axial uniform flow. In particular, approximately 70 
percent of the measured radial and tangential flow 
is contributed by the inclination of the shaft to 
the uniform stream. The boundary layer of the hull 
is found to contribute insignificantly to the per- 
turbation of the free stream. Although the viscous 
wake of the shafts and struts makes a significant 
contribution to the nonuniformity of the flow, the 
empirical technique proposed herein overpredicts 
the wake of the struts and underpredicts the wake 
of the shafting. 


2. BACKGROUND 


During the last ten to fifteen years there has been 
a marked increase in the installed horsepower per 
shaft on high-speed commercial and naval vessels. 
This increase in power has led to increased steady 
and unsteady forces on propellers, and increased 
loads on the hull surface. If adequate structural 
designs are to be developed for the propeller, its 
shafting, and the shaft supports; then the un- 
steady forces and moments on the propeller must be 


known accurately. Similarly, if the hull is to be 
habitable and to have minimal vibration, the 
structural design must adequately account for the 
propeller-induced surface forces. The propeller 
forces and surface loads can in turn only be ac- 
curate if they are determined using the full-scale 
flow into the propeller. 

Several theories exist for predicting the un- 
steady forces and moments acting on a propeller in 
a nonuniform flow, and the hull-surface forces 
induced by a propeller. Tsakona et al. (1974) and 
Frydenlund and Kerwin (1977) report on two of the 
theories for the unsteady forces on a propeller; 
Vorus (1974) reports on a theory for predicting 
the hull-surface forces. In these theories, the 
flow into the propeller is used in conjunction 
with an unsteady lifting-surface theory to predict 
the unsteady forces on the propeller and hull as 
the propeller rotates through the nonuniform flow. 

Typically, a propeller is wake adapted, that is, 
designed to the radial distribution of the circum- 
ferential mean velocity. The alternating forces 
are determined by considering the propeller in a 
nonuniform flow circumferentially. The variations 
of the forces and moments in the nonuniform stream 
from those in the uniform stream are then con- 
sidered to be the unsteady forces and moments on the 
propeller. 

The longitudinal component of the velocity in 
the propeller disk is the principal component of 
the velocity on a transom stern ship with inclined 
shafts. Typically the radial and tangential com- 
ponents vary sinusoidally around the propeller 
disk, and have peaks which are 20 to 25 percent of 
the longitudinal velocity component. However, in 
the process of determining the circumferential 
average of the radial and tangential velocity com- 
ponents, these components are reduced to 1 or 2 
percent of the longitudinal velocity component. 
Because of this, the tangential velocity component 
contributes very little to the angle of attack on 
a propeller blade as computed for the propeller 
design. However, in unsteady force calculations, 
the longitudinal velocity component varies from 
its mean by 10 to 15 percent while the radial and 
tangential components vary by 1000 percent from 
their means. Thus the variation in the tangential 
velocity component contributes significantly to 
the changes in the angle of attack on a propeller 
blade as it rotates through the wake. These 
changes in angle of attack in turn result in the 
unsteady forces and moments on the propeller. 

Experiments by Boswell [Boswell et al. 1976)], 
show that the maximum unsteady loads on the 
propeller occur in the area where the tangential 
flow velocities in the propeller disk are at their 
maximum. As will be seen later, it is the tangen- 
tial velocity components that are in least agree- 
ment between model and full scale. It is this 
fact that makes the issue of wake scaling important 
to the accurate determination of the unsteady 
forces on a full-scale propeller. 


3. TRIAL VESSEL AND INSTRUMENTATION 


A number of criteria went into the selection of 

the ship on which the full-scale measurements would 
be made. The hull form and appendage arrangement 
of the ship had to correspond to that which is 
typical of high-speed twin-screw commercial and 


naval vessels. The ship had to be available for an 
extended period of time and a means of propelling 
the ship had to be available. 

Of the ships which were in the U.S. Navy fleet, 
four classes seemed to meet the geometric criteria, 
and a means of propelling them could be identified. 
These were the Gearing Class (DD 710), Forrest 
Sherman Class (DD 931), Spruance Class (DD 963), 
and the Asheville Class (PG 84). However, of these 
classes, only the Asheville Class, which was being 
decommissioned, met the criterion of long term 
availability. As it tourned out, the David W. 
Taylor Naval Ship Research and Development Center 
(DTNSRDC) already had one of these ships under its 
control, the Research Vessel (R/V) ATHENA. 

The ATHENA had the added advantage that an ex- 
tensive series of model- and full-scale correlation 
experiments were already planned. Unsteady blade 
loads, stresses, and pressure distributions were 
going to be obtained full scale. The blade loading 
measurements were also going to be repeated at model 
scale. This blade loading data complement the full- 
scale wake data, and would result in some of the 
most complete correlation data of this type for any 
ship and model. 

The R/V ATHENA is a twin-screw aluminum hull 
CODOG (COmbined Diesel Or Gas Turbine) propelled 
high-speed displacement ship. Formerly designated 
PG 94, the 46.9 meter LWL ship was decommissioned 
in 1975 and placed in service as a high-speed 
towing platform for DTNSRDC. The hull form and 
propulsion arrangements are similar to today's 
destroyers and frigates which are propelled by 


Scale Ratio r 
Block Coefficient 
Prismatic Coefficient 
Length/Beam Ratio 
Beam/Draft Ratio 
Displacement/Length Ratio 


Coefficients 


227 


controllable-, reversible- pitch propellers using 
gas turbines as prime movers. The principal di- 
mensions and form coefficients for R/V ATHENA are 
presented in Figure 1. Figure 1 also shows the 
body plan, and bow and stern profiles of the ship. 
Figure 2 shows a drawing of the propeller. 

The ATHENA is equipped with two Cummins 750 
V-12 diesels for low speed propulsion and a single 
General Electric LM 1500 gas turbine for high- 
speed propulsion. In the diesel mode, the ATHENA 
is capable of speeds of around 14 knots. Under gas 
turbine power, she can attain a speed of 40 knots. 
The ATHENA is appended with twin shafts, struts, 
and rudders typical of most high-speed transom 
stern ships. In addition, she also has two anti- 
roll fins located just aft of amidships. 

Once the ATHENA was selected for the study of 
wake scaling, the question of how to propel the 
ship had to be resolved. The ATHENA is small 
enough that she could be towed by either one or 
two ships at speeds high enough to provide useful 
data, or she could be propelled on one shaft and 
measurements could be made on the other shaft. The 
two-ship tow would have been the most ideal means of 
propelling the ship during the experiments, because 
it would have allowed the ATHENA to be towed with no 
yaw angle, and outside the wake of another ship. 
However, the logistics of this option made it much 
less practical than propelling on one shaft. 

A series of model experiments was instituted, 
aimed at determining whether or not single shaft 
propulsion could provide good course keeping 
ability with minimal yaw angles.’ Flow visualiza- 


Sf (825 
on 0.48 
Ce 0.63 
Wo yee 
B/T 3.89 
A, 7.15 


STATIONS 


FIGURE 1. 


228 


rin riicn SECTION 
LINtaR oer ne SHIP 
cat 


nOoeL 


PROP. MO 01 oMIr 
ar 70 pcy.|az 70 PCr. 


INCHES | qq 5 8.727. |72.000 | 9.697 
cul o710 | 8-250 221.67 |1828.80| 246.30 |2 
INCHES | 47 5 727, [72-000 | 9.697 
Bil 471 | 8-256 221.67 |1828-80| 246.30 


a1 


0 DIA. 
nook shir 
AY 70, PCr. 


46.358 


riven | wunaen F tant 
nario. | or tor | eet awe, | Re | Be | gn, | ance | notation 
Ar 70 pcr. BLADES ate 7 Tie 
7 44 
Beit tata | elcri2 | oO) 


4 119.33 
4.698 
119533 


46.358 3 |40.100 9 
29908 .6 0.775 0.443 EE 0.670 | 0.048 | 0-000 LH. 4950 


ponung 4.598 
pomuagyt 0.670 | 0.048 | o-coo R:H. | 4950 : 


+2 


SHEPT OUTLINE 


FIGURE 2. Controllable-pitch propeller geometry. 


tion studies in the circulating water channel at 
DTNSRDC indicated that yaw angles of less than 
four degrees would provide satisfactory inflows 
to the propeller disk, and still exclude the wake 
of the roll fins. Subsequent self-propulsion 
model experiments using only one shaft, indicated 
that with the rudder set at one degree to port, 
the ship would have less than one degree of yaw 
and insignificant sway. Therefore, the decision 
was made to propel the ship on one shaft rather 
than to tow the ship. 

The instrumentation which was installed on the 
ATHENA consisted of three types. Five- and 
thirteen-hole pitot tubes were used to determine 
the velocity field in the propeller plane on the 
starboard side, and ahead of the struts on both 
the port and starboard sides.* A set of eight 
boundary layer probes were used to measure the 
boundary layer profile at four symmetric locations 
on the port and starboard sides of the ship. 
Finally a piezoelectric pitot tube, a five-hole 
pitot tube with piezoelectric pressure transducers 
mounted on its face, was used to measure the time- 
varying flow ahead of the operating propeller. 

The locations of the pitot tube rakes and bound- 
ary layer probes are shown in Figures 3 and 4. The 
location of the struts and the shape of the after 
stations are shown in Figure 5. As can be seen in 
these figures and in Figures 6 and-7, which show 
photographs of the actual pitot tube rakes mounted 
on the ship, two rakes of four pitot tubes each 
were mounted on opposite sides of the propeller 
hub. These rakes were attached to the crank disks 
for two of the propeller blades. The details of 
one rake with pitot tubes mounted are shown in 
Figure 8. 


*For the details of the instrumentation design and operation 
see Troesch et al. (1978). 


Q.25GR._ IN. 
6-gk- An 


2.727 IN. 
69.3 7H 
2 
6 


1.697 _ IN. 1.27 
a3. Ah hag 


FOR HUB ¢ PALM DETAILS 
SEE P-4#7/09// Sh. 2 
ALSO FOR HUB EXTENSION (wor smounm) 


Figure 9 shows a close-up photograph of one of 
the full-scale boundary layers probes. These 
probes, which extended 0.46 meters from the hull, 
contained 13 pitot tubes. Ten of the pitot tubes 
were total head tubes, and three were Prandtl 
tubes. 


4. CORRELATION MODELS AND INSTRUMENTATION 


The model correlation experiments were performed 
using two fiberglass models designated DTNSRDC 
Models 5365 and 5366. These models, which were 
built to the lines of the ATHENA, had a scale 
ratio of 1 to 8.25; the principal dimensions of 
these models are listed with the ship dimensions 
on Figure 1. A full set of appendages including 
shafts, V-struts, rudders, roll stabilizer fins, 
and a centerline skeg were fitted to each model. 
Model 5365 was a ship model which was used for the 
correlation wake surveys performed in the towing 
tank to investigate the scale effects between the 
model and ship wake surveys. Model 5366 was a 
mirror image double model obtained by reflecting 
the lines of the ATHENA about the mean water line 
corresponding to a full-scale speed of fifteen 
knots. This model was used for the boundary layer 
correlation experiments which were made in a wind 
tunnel. 

The model-scale wake survey was made on the ship 
model, Model 5365, using five-hole pitot tubes. 
The pitot tubes were mounted on a rake, the shaft 
of which was placed through the strut bossings and 
stern tube on the model. Figures 10, ll, and 12 
show the model which was used for the wake 
surveys, and the details of the pitot tube rake 
mounted on the stern of the model. Two papers, one 
by Hadler and Cheng (1965) and the other by Hale 
and Norrie (1967), give a thorough description of 


229 


-sayer zeAeT Aaepunog pue soeyeT aYEM JO UCTRIeETTeISUT HuTmoys TINY JO META STTFJOAq “€ FaNSIA 


‘@310N SS3INN S3T0H S S38NL LOL!ld WV 310N 


13) MOUS 06 38 01 JIVD | 


ay 
u | 


3903 4314v 


3903 OvvANOs 


y30gny 
w3oon 


373) 310m Ss 
1O1ld TOM Cl — 


(Lud) 


(ouvos 


wy aE Ns 
FROWd WAV] AMVONNO® 


Wows wv) Awvonnos 
a 


su 


as + +—t 


a 


j++ +++ 


he 


(AINO) LuOd 
TNGV) OWlLvEIeD 


HIT wULNIO~ 


(ATWO) Lii0d HON 


230 


9 8 7 6 5 ‘ 3 2 1 Ww 0 


. centering 


Boundary Layer Positions scale. s16tto1-o" 


FIGURE 4. Plan view of hull showing boundary layer rake locations. 


Ruup @ 9.727 ft. (0.222 mD 


L A 
WNY 
Zs 
Va aK IAN 
fp y 
; He ns 
as eae] r/R = 0.456 
a ——! r/R = 0.633 
r/R = 0.781 
if r/R = 0.963 
ae R>ROPELLER ~ 7°00 ft (0-91 m ) 


FIGURE 5. 


Afterbody sections of ATHENA hull showing radii of wake measurements. 


231 


FIGURE 6. Starboard side view of R/V ATHENA 
in drydock. 


FIGURE 7. Port side wake rakes and propeller 
on R/V ATHENA. 


FIGURE 8. Close-up view of five-hole pitot 
tube rake on starboard shaft on R/V ATHENA. 


nN 
WwW 
to 


ae 2 


~ 


FIGURE 9. Close-up view of boundary layer rake Peres 
on R/V ATHENA. eae 


FIGURE 10. Fitting room photograph of DTNSRDC 
model 5365 representing R/V ATHENA. 


8g 10 124 


ERS 


4 6 
CENTIMET) 


FIGURE 11. After end view of DTNSRDC model 
5365 fitted with a rake of five-hole pitot 
tubes on the starboard shaft. 


233 


FIGURE 12. Afterbody profile view of DTNSRDC 
model 5365 fitted with a rake of five-hole 
pitot tubes on the starboard shaft. 


FIGURE 13. Double model installed in DTNSRDC 
wind tunnel. 


the use and calibration of five-hole pitot tubes. 

The boundary layer velocity profile measure- 
ments on the double model, Model 5366, in the wind 
tunnel were obtained using a hot wire anemometer. 
The model was mounted on its side and the anemom- 
eter was moved in the horizontal direction by a 
rack and pinion drive. The rack and pinion, with 
its stepping motor, allowed the position of the 
anemometer to be set to within a fraction of a 
millimeter. 

Figure 13 shows the double model mounted in the 
wind tunnel. The vertical strut at the stern of 
the model is the support for the anemometry, and 
the bottom horizontal bar is an arm to steady the 
strut. The top horizontal bar is the traversing 
arm on which the hot wire anemometer is mounted. 

A close-up of the hot wire anemometer is shown in 
Figure 14; a centimeter scale is shown in the 
background of the photograph. 


FIGURE 14. Hot-wire anemometer probe used for model 
wind-tunnel boundary layer profile measurements. 


° 
wo 
x 


x 


0.456 RAD. 


FULL SCALE DATA 
MODEL SCALE DATA 


oe 


FIGURE 15. Velocity component 
ratios for R/V ATHENA and DTNSRDC 
model 5365 at 0.456 radius. 


-20 0 20 


5. FULL-SCALE WAKE SURVEY AND BOUNDARY LAYER 
MEASUREMENTS 


The full scale trials were run in the Atlantic 
Ocean off the Florida Coast near the mouth of the 
St. Johns River. The conditions for the trials 
were excellent as is shown in Table 1, which gives 
the trial agenda and sea conditions. The full-scale 
measurements were divided into four trials. Trial 
1 consisted of a wake survey in the propeller disk, 
and ahead of the struts on the port and starboard 
sides.* The objective of the measurements ahead of 
the struts was to determine the differences in the 
wake both with and without the propeller operating. 
Trial 2 consisted of a repeat of the wake survey in 
the propeller disk. However, for this repeat 
trial, the two rakes ahead of the struts on the 
starboard shaft were removed to eliminate any 
possibility of interference in the measurements. 
Trial 3 consisted of boundary layer profile measure- 
ments on the port and starboard sides of the hull. 
Again, the purpose of these measurements on both 
sides of the ship was to determine the effects of 
propeller induction on the development of the 
boundary layer. Trial 4 consisted of measurements 
of the time varying pressures in a plane ahead of 
the operating propeller. The results of Trial 4 
are discussed in Appendix A. 


*Note: The data from the wake surveys ahead of the struts 
and in the propeller disk at a lower speed are not re- 
ported in this paper, but will be reported in the future. 


80 190 120 140 165 130 


200 220 240 


IN DEGREES 


60 280 309 320 340 360 380 
ANGLE 


The pitot tubes on the rake in the propeller 
plane were located at non-dimensional radii (local 
radius divided by propeller radius) of 0.456, 
0.633, 0.781, and 0.964. The angular position of 
the rake was adjusted by turning the entire shaft 
using the jacking gear. The shaft could be rotated 
through approximately 230°, and because of this an 
overlap of 50° could be obtained in the data 
around 180°. 

The data from the wake survey at 15 knots are 
given, along with the corresponding model data, on 
Figures 15 through 18. This ship speed corresponded 
to a Froude number of 0.36 and a Reynolds number of 
4.14 x 108. The data are presented as velocity 
component ratios, where the velocities are given in 
cylindrical coordinates centered about the pro- 
peller shaft. The longitudinal velocity component 
(VX) is positive for flow toward the stern. The 
tangential velocity component (VT) is taken to be 
positive in the counterclockwise direction when 
looking forward on the starboard shaft. The radial 
velocity component (VR) is taken as positive in- 
ward. The angles are defined positive in the 
counterclockwise direction, with zero directly 
upward. The conventions for the angles and the 
directions of the velocity components are shown on 
Figure 5. These conventions are those of Hadler 
and Cheng (1965), except that the data is presented 
on the starboard shaft rather than on the port 
shaft. Therefore, the angles increase in the 
opposite direction from Hadler and Cheng, as do the 
tangential velocity components. 


0.633 Radius 


N 


i=) 


QO 
w 


2 
| 4 


VX/V 
i=) 
@ 


235 


oO 
N 


FULL SCALE DATA 
MODEL SCALE DATA 


Qo 
an 


ow 
de 


x 


t=) 
uo 


i=) 
i 


oO 
w 


=) 
N 


oO 


VT/V 
ro) 


D 
Oo 


1 
i=) 
ft 


yes 


FIGURE 16. Velocity component 


40 150 186 
GLE IN DESREES 


-20 0 


20 


There are not sufficient data at any one radius 
or circumferential position to adequately define 
the limits of accuracy for the full-scale measure- 
ments. A comparison between two different pitot 
tubes at any one radius may be made in the region 
between 150 and 200° where the data overlap. At 
all radii the longitudinal velocity component 
ratios show the greatest scatter in the full-scale 
data. In particular at the innermost radius 
(x/R = 0.456), the scatter in the longitudinal 
velocity component ratios is greatest, approxi- 
mately plus or minus ten percent. The scatter in 
the longitudinal velocity component ratios at other 
radii is significantly less than that, more nearly 
plus or minus five percent. The increased scatter 
in the longitudinal velocity component ratios is 
due to the computation procedure which uses the 
average of the longitudinal velocity components 
from both the radial and tangential velocity 
computations. 

The full-scale wake survey provided a unique 
opportunity to study the development of a turbulent 
boundary layer on a ship, and also the effects of 
propeller action on the boundary layer. The full- 
scale boundary layer was measured at the eight 
locations which are shown on Figures 3 and 4, at 
four speeds. These speeds were 6.2, 9.1, 14.8, 
and 16.5 knots; these speeds correspond to Reynolds 


ed eh 
200 220 240 260 280 360 320 340 360 330 


ratios for R/V ATHENA and DTNSRDC 
model 5365 at 0.633 radius. 


numbers of 1.74 x 108, 2.56 x 108, 4.14 x 108, and 
4.63 x 108 respectively. 

The data obtained at location 1, for all four 
speeds, are plotted on Figure 19. Except for the 
data at 6.2 knots, which show a great deal of 
scatter, the data are quite consistent with the 
fullness of the boundary layer increasing as the 
Reynolds number increases. The data obtained at 
14.8 knots (RL 4.14 x 108) for location 1, 2, and 
3 are plotted in Figures 20, 21, and 22 along with 
the corresponding model data at the same Froude 
number. The data from Locations 1, 2, and 3 are 
plotted again in Figures 23, 24, and 25 along with 
the data for the corresponding locations on the 
port side with the propeller operating. 


6. MODEL-SCALE WAKE SURVEY AND BOUNDARY LAYER 
MEASUREMENTS 


For the model-scale wake survey, Model 5365 was 
ballasted while at rest to the drafts corresponding 
to those of the ship during the full-scale wake 
survey. The model was then towed at 5.22 knots 
(2.685 m/s), the Froude-scaled speed which corre- 
sponds to 15 knots full-scale. The velocity com- 
ponent ratios were measured with a rake of five-hole 
pitot tubes at radii corresponding exactly to the 


N 
w 
or) 


OnGSileRAD: 


VX/V 


VR/V 


FIGURE 17. Velocity component 
ratios for R/V ATHENA and DTNSRDC -20 0 20 640 
model 5365 at 0.781 radius. 


full-scale wake survey radii, slowing a direct 
one-to-one comparison of the data. The data from 
this wake survey are plotted on Figures 15 through 
18. 

It is customary to perform wake survey experi- 
ments in the towing tank by towing the model at a 
speed corresponding to the Froude-scaled speed of 
the ship. In order to investigate the effects of 
Reynolds number on the model-scale wake, a second 
wake survey was run at an increased speed. This 
second speed was the highest speed for which steady 
data would be obtained, 13.5 knots (6.9 m/s). For 
this second wake survey, the sinkage and trim of 
the model were kept the same as at the 5.2 knot 
condition. This was done in an attempt to separate 
the effects of sinkage and trim, which is dependent 
on Froude number, from other speed effects. 

The data from the model-scale wake surveys at 
5.2 knots (Fp = 0.36, R, = 1.56 x 107) and 13.5 
knots (Fh = 0.93, Ry = 4.04 x 107) are presented 
in Figure 26. The longitudinal and radial velocity 
component ratios at these two speeds show no dif- 
ference. However, the tangential velocity compo- 
nent ratios obtained at 13.5 knots have peaks which 
are 4 to 6 percent lower than those obtained at 
5.2 knots. This is contrary to what might be 
exepcted, in that the increased Reynolds number 
should produce a thinner boundary layer and there- 
fore, a flow which more closely approaches the 


108 120 140 160 180 200 220 240 260 280 300 320 340 360 330 
ANGLE IN DEGREES 


potential flow around the hull. This anomalous 
result is probably due to the increased Froude num— 
ber and the corresponding change in the wave pattern 
around the model. 

The model-scale boundary layer profile measure- 
ments were made in a wind tunnel using hot wire 
anemometers. The double model was manufactured so 
as to take into account the dynamic trim of the 
ship. Although this cannot take into account the 
effects of the free surface, it does account for 
the angle of the shafting to the free stream, which 
contributes significantly to the radial and tangen- 
tial velocity components. 

The model scale boundary layer profile was 
obtained at a Reynolds number of 1.68 x 107, which 
was intended to equal the Reynolds number of the 
model in the towing tank at a Froude number of 0.36. 
The Reynolds number in the wind tunnel in fact 
turned out to be about 8 percent higher than the 
Reynolds number in the towing tank. However, this 
was not considered to be critical to the correlation 
of the model and ship data. 

The boundary layer profiles obtained in the wind 
tunnel, without the propeller operating, at Loca- 
tions 1, 2, and 3 are given in Figures 20, 21, and 
22; where they are plotted against the full scale 
data at the corresponding locations. The data 
obtained at the same locations with and without 
the propeller operating are plotted against the 


"9.7 © FULL SCALE DATA } 
ee ae MODEL SCALE DATA | 
0.5 + { {+} 4 1 __} x) eed Ee 


168 120 140 165 180 200 220 
ANGLE IN DEGREES 


40 635 


Velocity Profile Data From R/V ATHENA 
and Wind Tunnel Model 5366 
Location |, x/Ly, = 0.90 


E © RUN 209 m/s, Re= 1.74 x10° 
= ; | G RUN 210 m/s, Re = 2.56 x10° 
w © RUN 2il m/s, Re = 4.14 x 10° 
i a] O RUN2I2 m/s, Re= 4.63 x10° 
2 a 
S 
a a 
Es 3) 

x = 
50 
400 
o 
z 
5 20 
2) 
300 4 
=) 
=) 
25 
= 30 
{2} 
a 
200 
WW 
Z 20 
fas 
a 
a 
elo an 
zx 10 
= 
a 
fo} 
2 
oO Oo 
fo} O02 04 o6 o8 1.0 12 
U/Ue ° Ue = SHIP, MODEL SPEED 
FIGURE 19. Velocity profile data from R/V ATHENA 


measured at four speeds at location 1. 


240 


— 
250 285 300 320 340 350 335 


FIGURE 18. Velocity component 
ratios for R/V ATHENA and DTNSRDC 
model 5365 at 0.963 radius. 


corresponding data from the ship in Figures 23, 
and 25. This is the extent of the model scale 
boundary layer data. 

The accuracy of the model scale measurements with 
five-hole pitot tubes is known reasonably well. 
Model wake survey data have been repeated in past 
experiments, with the circumferential mean longi- 
tudinal velocity components repeating within 0.01 
of the free stream velocity. The velocity component 
ratios for the model data are repeatable to within 
plus or minus one percent, except in areas where 
steep velocity gradients occur. In the areas where 
high velocity gradients exist, such as behind the 
shaft struts, the five-hole pitot tube has much 
lower accuracy. Experiments with hot wire anemom- 
eters have shown that they are at least as accurate 
as five-hole pitot tubes. In fact, in regions where 
there are steep velocity gradients, hot wire anemom- 
eters may be an order of magnitude more accurate 
than pitot tubes. 


24, 


7. COMPARISON OF MODEL- AND FULL-SCALE DATA 

A study of the velocity component ratios presented 
in Figures 15 through 18 shows that the degree of 
scatter of the full-scale data is higher than that 
of the model data. This is due to the higher 
variations in both pressure measurement and ship 
speed. In particular, the full-scale data for the 
longitudinal velocity component ratio at the inner- 
most radius (r/R - 0.456) show the largest scatter, 


Velocity Profile Data From R/V ATHENA 
and Wind Tunnel Model 5366 
Location | x/Ly, * 0.90 


. EXPERIMENTAL MEASUREMENTS 
7 | OO MODEL SCALE U, = 38.1 m/s, Re =1.68x107 
Ww 
| FULL SCALE Us=7.6m/s, Re=4.14x10° 
a ’ 6 
a oO EQUIVALENT BODY OF REVOLUTION CALCULATIONS 
a n ———— MODEL SCALE 
a a -——- FULL SCALE 
a 6 
me 
* = 
50 
400 
& 
2 
=) oy 
a 
300 4 
=| 
=) 
Be 
s 30 
° 
« 
200 & 
8 
z 20 
<q 
EF 
n 
a 
loos 
=z lo 
= 
& 
oO 
a 
° ° 


te) o2 o4 o6 os 1.0 12 


U/Ue * Ue = SHIP, MODEL SPEED 


FIGURE 20. Measured and calculated boundary layer 
velocity profiles for R/V ATHENA and wind tunnel 
model 5366 at location l. 


Velocity Profile Data From R/V ATHENA 
and Wind Tunnel Model 5366 
Location 2 x/Ly, * O.77 


iE EXPERIMENTAL MEASUREMENTS 
E | MODEL SCALE Us = 38.1m/s Re = 168x107 
E Ww 
| = FULL SCALE -U, = 7.6 m/s, Re = 4.14x10° 
oa a EQUIVALENT BODY OF REVOLUTION CALCULATIONS 
a o ————MODEL SCALE 
tr =I - —FULL SCALE 
WwW 
a re) 
x 
7 = 
50 
400 
Ww 
< 
ire 
5 40 
yn 
300 4 
=) 
=) 
Be 
= 30 
3 
© 
200 
Ww 
Z 20 
q 
= 
Cy 
(=) 
100, 
=z 10 
= 
ac 
fo} 
z 
() i) 


ie} 02 04 o6 08 10 le 


U/Ue | Ue = SHIP, MODEL SPEED 


FIGURE 21. Measured and calculated boundary layer 


velocity profiles for R/V ATHENA and wind tunnel model 


5366 at location 2. 


Velocity Profile Data From R/V ATHENA 
and Wind Tunnel Model 5366 
Location 3 x/Ly, = 0.98 


5 EXPERIMENTAL MEASUREMENTS 
e | MODEL SCALE U, = 38.1m/s, Re = |.68 x 107 
ve} 
| 2 FULL SCALE U, = 7.6m/s, Re=4.14x10% 
ce Oo EQUIVALENT BODY OF REVOLUTION CALCULATIONS 
a My ———-—MODEL SCALE 
3 my -—— -FULL SCALE 
Ww 
a ) 
se 
x = 
50 
400 
Ww 
< 
w 
= 40 
a” 
300 4 
— 
5) 
= 
s 30 
fo) 
a 
zoo & 
8 
+2) | (40) 
<q 
(re 
ca) 
a 
fete) © TF 
a io 
= 
x 
3 
z 
to) ( 


fo) o2 04 o6 0.8 1.0 12 
U/Ue * Ue = SHIP, MODEL SPEED 


FIGURE 22. Measured and calculated boundary layer 
velocity profiles for R/V ATHENA and wind tunnel model 
5366 at location 3. 


Velocity Profile Data From R/V ATHENA 
and Wind Tunnel Model 5366 
Locations | and 8 x/Lw, * 0.90 


Ie EXPERIMENTAL MEASUREMENTS 
E | © MODEL SCALE U,= 38.1m/s, Re = 168x107 
Ww OO w/O PROPELLER 
| t| @ > FULL SCALE U,=7.6m/s, Re=4 14 x10° 
Ww S|] @ W/O PROPELLER 
= 7) 
S 
o a 
& ° 
I 
at = 
50 
400 
3 
rn 
5 a0 
Ww 
300 4 
a) 
5 
I 
= 30 
ro) 
[vq 
200) es 
Ww 
S 20 
fas 
7) @® 
fas) @ 
100, 
a 10 
= 
us e 
z e@ 
6 6) See ye (CE ef ae 
Cy 02 04 06 o8 10 12 


U/U. | Uo = SHIP, MODEL SPEED 


FIGURE 23. Measured boundary layer velocity profiles 
for R/V ATHENA and wind tunnel model 5366 with and 
without propeller at locations 1 and 8. 


Velocity Profile Data From R/V ATHENA 
and Wind Tunnel Model 5366 
Locations 2and 6 x/Ly, * 0.77 


E EXPERIMENTAL MEASUREMENTS 
E | | GO MODEL SCALE U, = 38.1m/s, Re = 1.68 x10" 
Ww e) w/O PROPELLER F 
| a) FULL SCALE U,=7.6mM/s, Re® 49.14x10 
w S| e W/O PROPELLER 
=! ao 
S 
a wu 
a 
£ ° 
Fs = 
50 
400 
6 
E 
5 40 
n 
300 4 
=) 
=) 
a5 
s 30 
° 
v4 
200) se 
o 
eco) 
ras 
Ca) 
a 
100) 
a lo 
= 
& 
fo} 
cA 
° ° 
° 02 04 06 08 1.0 12 
U/Ue > Ue = SHIP, MODEL SPEED 
FIGURE 24. Measured boundary layer velocity profiles 


for R/V ATHENA and wind tunnel model 5366 with and 
without propeller at locations 2 and 6. 


and the greatest deviation from the model-scale 
wake. 

In part, this scatter is also due to the fact 
that the longitudinal velocity component ratios 
presented are an average of the longitudinal velocity 
component as measured in the tangential plane and in 
the radial plane. Therefore, any scatter error in 
either the tangential or radial plane measurements 
will influence the calculation of the longitudinal 
component. Another factor which probably contrib- 
uted to increased scatter at the innermost radius 
is the close proximity of the pitot tube to the 
strut bossing. 

The longitudinal velocity component ratio at the 
innermost radius is about 10 percent lower for the 
ship than for the model, while the peaks of the 
tangential and radial velocity component ratios are 
about 10 percent higher for the ship than for the 
model. Although there are undoubtedly scale effects 
on the shafting and strut bossing at this radius, 
another significant factor is that the bossing on 
the ship is proportionately much longer than on 
the model. This is due to the collar to which the 
pitot tube rakes ahead of the struts were attached. 

At the outer radii the longitudinal velocity 
component ratios for the ship are 2-4 percent lower 
than those for the model. The peaks of the radial 
and tangential velocity component ratios at the 
outer radii are 8-10 percent higher for the ship 
than for the model. At the two innermost radii, 
the shift in the radial and tangential velocity 
component ratios indicate that there is a stronger 
upflow on the ship than the model, in the region 
under and outboard of the propeller hub. This 
effect is much weaker, and has shifted to the inside 


239 


on the two outer radii. One possible cause of the 
shift at the outer radii is the fact that the full 
scale trial was performed with a propeller operating 
on the port shaft, while the model data were col- 
lected without the propeller present. However, the 
most likely source of the increased upward flow is 

a difference in attitude between the ship and model. 

The models were run at a number of Reynolds num- 
bers in the towing tank and wind tunnel and the 
longitudinal velocity component was measured at a 
single location near the hull for these various 
Reynolds numbers. The results of these measure- 
ments are plotted in Figure 27. These results in- 
dicate that for a Reynolds number greater than 107 
there is very little effect of either Reynolds num- 
ber or Froude number on the longitudinal velocity 
component. Therefore, in cases where it is desir- 
able to obtain accurate longitudinal velocity 
component measurements, the model should be run at 
the correct Froude trim, at a Reynolds number 
greater than 107. 

A comparison of the boundary layer profiles 
presented in Figures 20, 21, and 22 shows that, as 
might be expected, the model velocity profile is 
not as fully developed as the full-scale velocity 
profile at Locations 1 and 3. This is clearly a 
consequence of the one decade difference in Reynolds 
number between the model and ship. However, at 
Location 2, the model- and full-scale boundary 
layer velocity profiles almost coincide. This is 
clearly an anomalous situation, particularly be- 
cause even at 0.46 meters from the hull full scale, 
the velocity has not reached the free-stream 
velocity, let alone the potential flow velocity 
which is even higher. The most likely explanation 
for the low full-scale velocity profile is a mal- 


Velocity Profile Data From R/V ATHENA 
and Wind Tunnel Model 5366 
Locations 3 and 7 x/Ly, =0.98 


E EXPERIMENTAL MEASUREMENTS 
E Fj MODEL SCALE U, = 38.1m/s, Re = 1.68x107 
Ww W/O PROPELLER - 
| zl FULL SCALE Us: 7.6m/s, Re=4.14 x10 
Ww ss W/O PROPELLER 
s 7) 
S 
a rr 
a 3 
= 
Fs = 
50 
400 
& 
z 
5 40 
12) 
300 4 
a) 
=) 
= 
s 30 
° 
[rg 
200 & 
8 
2 20 
es 
& 
(=) 
TOO Mma, 
=z 10 
= 
x 
° 
ra 
0 0 
) 02 04 06 08 1.0 12 
U/Ue : Ue = SHIP, MODEL SPEED 
FIGURE 25. Measured boundary layer velocity profiles 


for R/V ATHENA and wind tunnel model 5366 with and 
without propeller locations 3 and 7. 


240 


FIGURE 26. Velocity component 
ratios for DTNSRDC model 5365 at 
0.633 radius for model speeds of 
5.22 knots and 13.5 knots. 


function in the instrumentation but a check of the 
data records indicated no obvious errors in the 
data. 

The results of boundary layer profile measure- 
ment with the propeller operating, plotted in 
Figures 23, 24, and 25 indicate that the data at 
positions 1 and 8, just ahead of the propeller, 
show a slight increase in velocity profile due to 
the propeller suction. The increases are about 
the same at both model- and ship-scale. The data 
at positions 3 and 7, behind the propeller, show 
rather significant increases in the velocity pro- 
file for both scales. This is undoubtedly due to 
the wake of the propeller. From the model-scale 
data, at Locations 2 and 6, there is no noticeable 
difference in the data obtained with or without 
the propeller operating. This is consistent with 
the separation between the boundary layer probe and 
the propeller. There is no ship-scale data ahead 
of the operating propeller at location 6 due to the 
failure of that boundary layer probe. 

In order to evaluate our ability to predict the 
boundary layer of the hull, a series of boundary 
layer calculations were instituted. For these 
calculations, the ship was approximated as a body 
of revolution, and the boundary layer was calculated 
using the standard DTNSRDC method for bodies of 
revolution [Wang and Huang (1976)]. Two methods 
for generating the bodies of revolution were tested. 
In one, the body was generated with radii equal to 
the square root of twice the sectional area of the 


$c 103 120 149 160 
ANGLE IN DEGREES 


135 205 220 240 250 280 309 320 340 360 33¢ 


ship; and in the other, the body was generated 
using circumferences equal to twice the girth of 
the ship. The boundary layer calculations using 
the body of revolution based on sectional area 
agreed best with the experimental data. 

The results of the equivalent body of revolution 
calculations are plotted with the experimental data 
on Figures 20, 21, and 22. The calculations for 
the ship at Locations 1 and 3 agree reasonably well 
with the full-scale data. However, at the model- 
scale, the calculations do not agree nearly as 
well. This is probably due to the fact that at 
lower Reynolds numbers, the boundary layer is much 
more sensitive to errors in the flow velocity and 
pressure gradient than at higher speeds. As stated 
previously, the data at Location 2 is anomalous, 
as is shown by a comparison with the calculated 
boundary layer profile. 


8. PREDICTION OF NOMINAL WAKE 


Although the model- and full-scale wake of the R/V 
ATHENA both agree qualitatively, there are some 
substantial quantitative differences between the 
model- and full-scale velocity components. To 
develop an understanding of the origins of these 
differences, it was necessary to predict the wake 
of both the model- and full-scale ship analytically. 
Since the hull of the ATHENA showed no separation, 
it appeared that the presence of the hull could be 


MODEL 5365 REPRESENTING R/V ATHENA 
WAKE SURVEY SCALE EFFECT 


4) AT O° 


r/R = 0.63 


lw =18 67 FT 
Ls =1540 FT 


O SHIP MODEL IN TOWING TANK 
Q DOUBLE MODEL IN WIND TUNNEL 
OD SHIP TRIAL 


0.8 


LONGITUDINAL VELOCITY COMPONENT RATIO (Vx 


5x10° 10" 5x10" 10' 
REYNOLDS NUMBER (Re,) BASED ON SHIP LENGTH 


dealt with primarily by potential flow techniques, 
combined with calculations of the boundary layer 
displacement thickness. It was also assumed that 
the viscous flow about the appendages could be 
dealt with empirically. 

The velocity in the propeller disk, expressed in 
shaft coordinates, was decomposed as follows: 


Velocity = Uniform Stream 

Perturbation due to Hull 
Perturbation due to Boundary Layer 
Viscous Wake of Struts 

Viscous Wake of Shafting. 


oe eee 


The principal factor contributing to the radial and 
tangential components of the velocity in the pro- 
peller plane is the inclination of the shaft to the 
free stream. The shafting of the ATHENA makes an 
angle of 8.9° with the baseline. In addition, at 
15 knots Ga = 0.36), the ATHENA takes a bow-up 
trim of 0.3° as indicated by model experiments. 
Thus, the propeller shaft is inclined a total of 
9.2° to the incident stream. The effect of resolv- 
ing the incident stream into shaft coordinates is 
shown on Figure 28. 

The effects of perturbing the incident stream by 
the presence of the hull were obtained by means of 
potential flow calculations. For the purposes of 
this study, the free surface was represented by the 
zero Froude number condition, and the calculations 
were made for a double model in an infinite fluid. 
The hull was reflected about the mean waterline at 
15 knots, and flow about the resulting body was 
computed using the DTNSRDC potential flow program 
[Dawson and Dean (1972)]. The results of this 
computation are also shown on Figure 28. As can 
be seen, the effects due to the perturbation of the 
incident flow by the hull are small, on the order 
of two percent of the ship speed. 

The effects of the displacement thickness of the 
boundary layer were considered next. The intention 
was to increase the thickness of the hull by the 


displacement thickness of the boundary layer, and to 


repeat the potential flow calculations. However, 
at its thickest point, the model scale boundary 
layer determined from the equivalent body of 
revolution calculations, would only have increased 


241 


FIGURE 27. Longitudinal velocity component 
ratio at O-degree position of 0.633 radius 
10? as a function of Reynolds number based on 
hull length. 


the thickness of the hull by 1 percent of the beam. 
The full-scale boundary layer would have increased 
the thickness even less. Since the complete hull 
potential flow had only a two percent effect, the 
revised potential flow was not computed for such a 
small change in effective hull shape. The error 
due to neglecting the displacement thickness of the 
boundary layer is probably much less than the error 
incurred by making the zero Froude number approxi- 
mation for the potential flow calculations. There- 
fore, the velocity component ratios based on only 
the first potential flow calculations are presented 
in Figures 29 through 32. 

The velocity defect caused by the struts was 
predicted using an empirical scheme based on data 
from aerodynamics. The velocity defect was com- 
puted using the following formula from page 584 of 
Goldstein (1965). 


(7R=0633 


Sao 3-D POTENTIAL FLOW 
UNIFORM FLOW 


fe) MEASURED VALUES 


VELOCITY COMPONENT RATIO 


rt 

° 40 80 120 160 200 240 280 320 360 
ANGLE IN DEGREES 

Velocity Component Ratios Predicted and Measured Full-Scale 

Trial 2, Vs = 7.87m/s 


FIGURE 28. Effect of shaft inclination and hull po- 
tential flow on velocity component ratios for R/V 
ATHENA at 0.633 radius. 


242 


T/R=0456 


—-—-—-PREDICTED VALUES 
O MEASURED VALUES 


VELOCITY COMPONENT RATIO 


° 40 80 120 160 200 240 280 320 360 
ANGLE IN DEGREES 
Velocity Component Ratios Predicted and Measured Full-Scale 
Trial 2, Vs = 7.87 m/s 


FIGURE 29. Predicted and measured values of velocity 
component ratios for R/V ATHENA at 0.456 radius. 


Ie 
= 3/18a,x* 
rend Oe Lo Gee 
and 


No 3/2a,, = 10D/n PU, 


where Umax is the velocity defect, U, the free- 
stream velocity, n_ is the nondimensional wake half 
width, x is the nondimensional distance from the 
strut, D is the strut drag, and p the fluid density. 
These formulas predict the longitudinal velocity 
defect in terms of the strut drag, wake thickness, 
and distance behind the strut. 

The shaft struts on R/V ATHENA are Navy EPH 
sections with a chord-to-thickness ratio of 6. 


r/R =0633 


V/V 
oo 
®o0° 


— — -0.2 
5 oo 


—-—-—-—PREDICTED VALUES 
O MEASURED VALUES 


VELOCITY COMPONENT RATIO 


° 40 80 120 160 200 240 280 320 360 
ANGLE IN DEGREES 
Velocity Component Ratios Predicted and Measured Full-Scale 
Trial 2, Vs = 7.87m/s 


FIGURE 30. Predicted and measured values of velocity 
component ratios for R/V ATHENA at 0.633 radius. 


r/7R=0.78l 


—-—-—-—PREDICTED VALUES 
O MEASURED VALUES 


VELOCITY COMPONENT RATIO 


° 40 80 120 160 200 240 280 320 360 
ANGLE IN DEGREES 
Velocity Component Ratios Predicted and Measured Full-Scale 
Trial 2, Vs = 7.87 m/s 


FIGURE 31. Predicted and measured values of velocity 
component ratios for R/V ATHENA at 0.781 radius. 


Assuming that the drag on the EPH section would not 
be too different from the drag on an elliptic 
section of the same thickness-chord ratio, many data 
for a number of elliptic sections were collected. 
These data are plotted on Figure 33 as a function 
of Reynolds number. 

The nondimensional wake half-width was predicted 
using Equation (4) from Silverstein et al. (1938): 

os 45 5 
n = sets) (G6 ce 415) 
(0) D 

In this equation n_ is again the nondimensional half 
width of the wake, X is the nondimensional distance 
from the strut, and C_ is the drag coefficient per 
unit length of the strut. 

Using the strut Reynolds numbers based on chord 
length, of 1.46 x 10° for the ship, the correspond- 


r/R =0.963 


Vy/V 
ie) ° 
=r ©" ~-o- ~ 688 


’ SHIP 
i 


—-—-—-—PREDICTED VALUES 
© MEASURED VALUES 


VELOCITY COMPONENT RATIO 


° 40 80 120 160 200 240 280 320 360 
ANGLE IN DEGREES 
Velocity Component Ratios Predicted and Measured Full-Scale 
Trial 2, Vs = 7.87m/s 


FIGURE 32. Predicted and measured values of velocity 
component ratios for R/V ATHENA at 0.963 radius. 


=6.0 WILLIAMS & BROWN (\937) 
=40 
£30 Fuinosey 11938) 


2394 
745 


+ WARDEN (1934) 


en ET AL (1929) 


Cp = DRAG PER UNIT LENGTH /P/> cv2 


104 5x10% 10° 5x10° 10° 
REYNOLDS NUMBER (Re,) BASED ON STRUT CHORD LENGTH 


ing drag coefficients are found to be 0.050 and 
0.018 for the model and ship, respectively. Sub- 
sitution of these drag coefficients into the above 
formulas from Silverstein, et al. (1938) and 
Goldstein (1965) yields the velocity defects which 
are shown on Figures 29 through 32. 

These computed velocity defects due to strut 
wake are significantly greater than the velocity 
defects which were observed at either model or full 
scale. The cause of this over-prediction is 
probably the fact that the formulas from Goldstein 
are derived by assuming that the wake is being 
calculated far enough downstream that the cross 
flow terms in the momentum equation can be neglected. 
This is an assumption which is undoubtedly violated 
in the region near the struts, where the wake has 
been predicted. 

Although the empirical method for predicting the 
wake of the shaft struts was not successful, it 
does at least provide some insight into how the 
wake should vary with Reynolds number. Both the 
width of the wake of the struts and the velocity 
defect in the wake of the struts are proportional 
to the square root of the drag coefficient of the 
section. Therefore, the velocity defect and the 
width of the wake should both decrease (like the 
square root of the ratio of the drag coefficients) 
as the Reynolds number increases. However, the 
full-scale wake survey data were not collected at 
angular increments spaced closely enough to confirm 
this scaling law. 

The empirical method for predicting the wake 
behind an inclined shaft is not as well defined as 
the methods for predicting the wake behind the 
struts. Following the methodology of Chiu and 
Lienhard (1967), it was assumed that the separated 
flow behind a yawed cylinder is a function of the 
component of the velocity normal to the cylinder. 
Following the method of Roshko (1955) and (1958), 
an estimate of the velocity defect in the wake of 
the shaft was developed based on the pressure 
coefficient at the point of separation and the 
Strouhal number. i 

Data showing the base pressure behind a circular 


*Note: The base pressure is not necessarily the pressure at 
the separation point because there is usually some pressure 
variation in the separated region. 


5x10° 


243 


FIGURE 33. Drag coefficients of elliptical 
7 section struts as a function of Reynolds 
number based on chord length. 


10 


cylinder have been collected, and are presented as 
a function of Reynolds number in Figure 34. Based 
on this data and the Reynolds number based on cross 
flow velocity, the pressure coefficients for the 
model (R_ = 1.63 x 10*) and ship (R_ = 4.26 x 10°) 
were found to be -1.1 and -0.2 respectively. These 
pressure coefficients resulted in a predicted veloc- 
ity defect, perpendicular to the shaft axis, of 
0.25 for the model and 0.10 for the ship. However, 
when resolved back into the direction of the flow, 
the shaft wake is less than two percent of model 
speed and one percent of ship speed. This is 
significantly less than than the velocity defect 
which is measured for either the model or the ship. 
In fact, if the velocity defect in the direction 
normal to the shaft were 100 percent of the forward 
speed, the velocity defect in the wake would only 
be seven percent, still less than the velocity 
defect measured experimentally. 

These results are not surprising when one con- 
siders the discussion in Chiu and Lienhard (1969). 
In this discussion, data are presented which point 
out that the wake of an inclined shaft is in general 
not parallel to the shaft. This is due to the 
axial component of the flow along the cylinder which 
develops a boundary layer which separates. The 
Reynolds number for separation in the axial direc- 
tion on the shaft is independent of the Reynolds 
number of the flow normal to the shaft. In addition, 
the data from Bursnall and Loftin (1952), show that 
as a circular cylinder is inclined further and 
further to the flow, the transverse Reynolds number 
at which separation takes place becomes lower and 
lower. 


9. CONCLUSIONS 


Significant differences have been found in the 
tangential and radial velocity component ratios 
between the ship and the model wake surveys. In 
particular, the full-scale tangential velocity 
component ratio has a peak amplitude approximately 
eight to ten percentage points higher than that at 
model scale. Similarly, the ship radial velocity 
component peak is higher by six to eight percentage 
points. These differences cannot be attributed 

to scale effects. The most likely cause seems to 


© BURSNALL & LOFTIN (195!) 
Q_FAGE 8 FALKNER (193!) 
© ROSHKO (1953) 


0.8 


2 
° 
b 


Cp, = BASE PRESSURE /P, v2 
° ° 
+ fo} 


° 
o 


10% 5x10* 108 5x10° 
REYNOLDS NUMBER (Rep) BASED ON SHAFT DIAMETER 


be a difference in trim between model- and full- 
scale. Because the model was ballasted to the 
draft of the ship, further work will be required 
to identify the source of these differences. 

The longitudinal velocity component ratios for 
the full-scale trial show a much greater scatter 
than the tangential and radial components. For 
this reason it is unclear that any difference is 
shown by these data, when compared to model-scale 
data. The innermost radius (r/R = 0.456) does show 
that the high longitudinal velocity component 
normally measured at these inner radii is not found 
full scale. This may not be the result of scale 
effects on the shafting and strut bossing, but the 
fact that the full-scale bossing is longer than the 
model-scale bossing. This is a result which will 
have to be investigated by further model experi- 
ments. 

The results from model experiments in both the 
wind tunnel and in the towing tank, and from the 
full-scale trial indicate that for a circumferential 
position near the hull, there was little difference 
in longitudinal velocity component ratio for speeds 
corresponding to Reynolds numbers greater than 10". 
Therefore, when measuring only the longitudinal 
velocity component ratios experimentally, the model 
should be run at the trim corresponding to that of 
the Froude-scaled speed and at a speed high enough 
to yield a Reynolds number of greater than 107. 

The attempt at predicting the wake for this high- 
speed displacement ship showed that the most im- 
portant contribution to the variation in tangential 
and radial velocity component ratios was the shaft 
angle to the flow. The calculation of the potential 
flow around the hull and the resulting velocity 
components showed that the effect of the perturba- 
tion due to the hull was small. The effects of the 
boundary layer of the hull on the wake were also 
shown to be small. 

In summary it may be stated that the full-scale 
and model wakes differ by approximately ten percent 
of the ship speed. These differences cannot be 
adequately explained at this time. Further work 
on wake of appendages is recommended as one step in 
improving the understanding of these differences. 


Chiu, W. S., and J. H. Lienhard (1967). 


FIGURE 34. Base pressure coefficients of cylin- 
10 drical shafts as a function of Reynolds number 
based on shaft diameter. 


ACKNOWLEDGEMENTS 


This work was performed under the controllable- 
pitch propeller research program sponsored by C. L. 
Miller of the Naval Sea Systems Command (NAVSEA 
0331G) administered by the David W. Taylor Naval 
Ship Research and Development Center (DTNSRDC) . 

The authors wish to express their appreciation 
to personnel of the Ship Performance Department of 
DTNSRDC, the University of Michigan, and the crew 
of R/V ATHENA from MAR Inc. for their assistance 
in conducting the full-scale trial and model experi- 
ments which provided the data for this paper. CHI 
Associates, Inc. and Rosenblatt Inc. are also 
acknowledged for their assistance in the prepara-— 
tion of this paper. 


BIBLIOGRAPHY 


Boswell, R. J., J. J. Nelka, and S. B. Denny (1976). 
Experimental Determination of Mean and Unsteady 
Loads on a Model C. P. Propeller Blade for 
Various Simulated Modes of Ship Operation. 
Eleventh ONR Symposium on Naval Hydrodynamics, 
London, England, VIII 75-110. 

Bursnall, W. J., and TR. K. Loftidne (1952))5" qxpemim 
mental Investigation of the Pressure Distribu- 
tion about a Yawed Circular Cylinder in the 
Critical Reynolds Number Range. NACA Tech. Note 
2463 

Canham, J. J. S. (1975). Resistance, Propulsion 
and Wake Tests with HMS Penelope. Trans. Royal 
Inst. of Naval Arch., 117, 61-94. 


Flow Over Yawed Circular Cylinders. Jour. Basic 
Engineering, Trans. ASME, Series D, 89, 851-857. 

___(1969) Discussion to: On Real Fluid Flow Over 
Yawed Circular Cylinders. Jour. Basic Engineer- 
ing, Trans. ASME, 91, 132-134. 

Dawson, C., and J. Dean (1972). The XYZ Potential 
Flow Program. NSRDC Report 3892. 

Denny, S. B., H. A. Themak, and J. J. Nelka (1975). 
Hydrodynamic Design Considerations for the 


On Real Fluid 


Controllable-Pitch Propeller for the Guided 
Missile Frigate. Naval Eng. Jour., 87, 2; 72-81. 

Fage, A., and V. M. Falkner (1931). Further Experi- 
ments on the Flow Around a Circular Cylinder. 
British A.R.C. Reports and Memoranda No. 1369, 
186-208. 

Frydenlund, O., and J. E. Kerwin (1977). The 
Development of Numerical Methods for the Compu- 
tation of Unsteady Propeller Forces. Norwegian 
Maritime Research, 17-28. 

Goldstein, S. (1965). Modern Developments in Fluid 
Dynamics, Vol. II. Dover Publications, Inc., 

New York, pp. 331-702. 

Hale, M. R., and D. H. Norrie (1967). The Analysis 
and Calibration of the Five-Hole Spherical Pitot. 
ASME Paper 67-WA/FE-24. 

Hadler, J. B., and H. M. Cheng (1965). Analysis of 
Experimental Wake Data in Way of Propeller Plane 
of Single- and Twin-Screw Ship Models. Trans. 
Soc. Naval Arch. and Mar. Eng., 73, 287-414. 

Lindsey, W. F. (1938). Drag of Cylinders of Simple 
Shape. NACA Report 619, 169-176. 

Namimatsu, M., and K. Muraoka (1973). The Wake 
Distribution of a Full Form Ship. JIHI Engineer- 
ing Review, 7, 3. 

Namimatsu, M., K. Muraoka, S. Yamashita, and H. 
Kishimoto (1973). Wake Distribution of Ship and 
Model on Full Ship Form. J. Soc. Naval Arch. 
Japan, 134 (Japanese), 65-73. 

INoKEIS7 So Mey Ws Wo @s iMG Wee bisdeyelovexen Jel 126 
Young, and D. A. O'Neil (1977). Increased 
Profits for Gas Turbine Container Ships by 
Unique Applications of Combustion Technology and 
Hydrodynamics. Trans. Soc. Naval Arch. and Mar. 
Eng., 85. 

Rains, D. A. (1975) DD 963 Power Plant. Marine 
Technology, 12, 1; 1-24. 

Roshko, A. (1953). On the Development of Turbulent 
Wakes from Vortex Streets. NACA Report 1191, 
801-824. 

___(1954) On the Drag and Shedding Frequency of 
Two-Deminsional Bluff Bodies. NACA Tech. Note 
3169. 

___(1955) On the Wake and Drag of Bluff Bodies. J. 
MERA, Behe, 2A, NAAM. 

Schuster, S., H. Grothues-Spork, H. Thieme, and 
H. Schwanecke (1968). Meteor-Messfahrten 1967. 
Jahrbuch der STG, 62, 159-204. 

Silverstein, A., S. Katzoff, and W. K. Bullivant 


APPENDIX A - TIME-VARYING PRESSURE MEASUREMENTS 
AHEAD OF AN OPERATING PROPELLER 


During Trial 4, the time-varying pressures at the 
head of a piezoelectric pitot tube were obtained 

as a function of shaft position. The pressures 
were measured for each six degrees of shaft rota- 
tion for the pitot tube at a fixed angle. Measure- 
ments were obtained at four angular positions of 
the pitot tube and at two ship speeds. 

It should be noted that due to the fact that the 
pitot tube is approximately one diameter of the 
propeller forward of the propeller disk, the 
amplitude of the pressure oscillation is only 1 per- 


245 


(1938). Downwash and Wake Behind Plain and 
Flapped Airfoils. NACA Report 651, 179-206. 
Takahashi, H., T. Ueda, M. Nakato, Y. Yamazaki, 
M. Ogura, K. Yokoo, H. Tanaka, and S. Omata 
(1971). Measurement of Velocity Distribution 
Ahead of the Propeller Disk of the Ship. J. Soc. 
of Naval Arch. West. Japan, 129 (Japanese), 
153=J'68r 

Taniguchi, K., and T. Fujita (1970). Comparison of 
Velocity Distribution in the Boundary Layer 
Between Ship and Model. J. Soc. Naval Arch. 
Japan, 127. 

Troesch, A., V. A. Phelps, and J. Hackett (1978). 
Full-Scale Wake and Boundary Layer Survey 
Instrumentation Feasibility Study. Dept. Naval 
Arch. and Mar. Eng. Report, Univ. of Mich. 

Tsakonas, S., W. R. Jacobs, and M. R. Ali (1973). 
An Exact Linear Lifting-Surface Theory for 
Marine Propeller in a Nonuniform Flow Field. 

Jip Shipy Res, Li77, VI6—20)7 i. 

Vorus, W. S. (1974). A Method for Analyzing the 
Propeller-Induced Vibratory Forces Acting on 
the Surface of a Ship Stern. Trans Soc. Naval 
Arch. Mar. Eng. , 82), 186—21'0). 

Wang, H. T., and T. T. Huang (1976). User's 
Manual for FORTRAN IV Computer Program for 
Calculating the Potential Flow/Boundary Layer 
Interaction on Axisymmetric Bodies. DITNSRDC 
Ship Performance Dept. Report 737-Ol. 

Warden, R. (1934). Resistance of Certain Strut 
Forms. British A.R.C. Reports and Memoranda 
LG), BUY SQN 

Wennberg, P. K. (1966). The Design of the Main 
Propulsion Machinery Plant Installed in the 
USCGC HAMILTON (WPG-715). Trans. Soc. Naval 
Arch. and Mar. Eng., 74, 29-69. 

Williams, D. H., and A. F. Brown (1937). Experi- 
ments on an Elliptic Cylinder in the Compressed 
Air Tunnel. British A.R.C. Reports and 
Memoranda 1817, 103-112. 

Yokoo, K. (1974). Measurement of Full-Scale Wake 
Characteristics and Their Prediction From Model 
Results--State of the Art. Symposium on High 
Powered Propulsion of Large Ships, Wageningen, 
pp. XI.1-28. 

Zahm, Ao Be, Re He Smith, and FA. louden) (11929) 
Forces on Elliptic Cylinder in Uniform Air 
Stream. WACA Report 289, 217-232. 


cent of the mean pressure signal. Due to the 
failure of two of the pressure transducers in the 
head of the pitot tube, it was impossible to obtain 
any data on the variation of the flow velocity with 
angular position. 

During each of the runs with the piezoelectric 
pitot tube, data were collected for a period of 
time totalling between 5 and 10 minutes. All of 
the data points for each angular position of the 
shaft were then averaged to obtain a mean level for 
each signal. Figures A-1, A-2, and A-3 show these 
averaged pressure signals as a function of angular 
position. Runs 209 and 205 were both obtained at 
the same ship speed (15 knots) and shaft speed 


246 


i i th 
0 40 80 120 160 20 240 | 280 20 360 
T Wali aT I All T T 
144-00 
ie x os 9 As as Poa @ 
s ® i \ A @ 
13-95 i A o e 
N [es St. ry Wad Odi @ 9%e J @ AR 
13-90 L. r y b ® ° . Ve 
5-30 | . - 
me Pe e ry oes . 
5-25 @ vee Seta ee { 28 al f wee 
12 is \ ® e °, @ «6 ee, 9 ry ed J 
} e 
5-20 8 . 8 x 
14-30 = 
° pee a a ee ro” a* oot 
a), ees p” * As Went 7 Ale SAR a Seay | 
OU ° ves Vv, wee ® 8 
14-20 A + 
+ v of 2, + 
97:0 v . + 
ISS Oo She ° * a 
A i +, + + 
R2 96:5 |_ . 5 O86 4 + x 
° ar * A * 
96-0 |_ + * ‘7 + : ‘ pase 
2-10 HH nee t orton ns TF HH eget Met eeee tht Meee teatangs ete reget ttt tenn et 
FIGURE A-l. Circumferential dis- ii 
tribution of piezoelectric pres- c  &05 i | | =| | | ! 
sure transducer signals for 15- Q o) 0 129 @ ™m a » am ca 
knot run 209. ott) BLADE POSITION @ IN DEGREES 360P= OME REVOLUTION (0.17SEC) 
(345 rpm), but with the pitot tube in different signals. For Run 209, the amplitude of the eighth 
angular positions, 180° for Run 205 and 300° for harmonic was 3 percent and those of the other har- 
Run 209. Run 208 was obtained at a speed of 8.9 monics were generally less than 10 percent of the 
knots and a propeller speed of 245 rpm, with the eighth harmonic. The only exception to this is the 
pitot tube at 300°. sixteenth harmonic which is again of increased mag- 
As can clearly be seen from the data obtained nitude. Although the magnitude of the harmonics 
during Run 209, the pressure signal from the three from Run 205 were lower than those from Run 209 the 
operating pressure transducers is periodic. There saem results apply. There is not obvious periodicity 
is an obvious periodicity at twice blade frequency in the data from Run 208. However, a harmonic 
(eighth harmonic in shaft frequency). A Fourier analysis of this data shows that the twelfth har- 
series analysis of the data from the two 15-knot monic is dominant, although not nearly to the same 
runs showed that the second harmonic in blade extent as in the cases of the high speed runs. 
frequency was the dominant harmonic in all three 
0 ® 80 1 10 m 200 0 30 360 
i T mal al Ty TR Im iperaneee| 7 a1 
im oe °@ / ° 4 
ign oe ba eWieiues eo%e f Se gt 8 a Pe ae ore R. Hee: 
Tl 96 ahs OY © 6 \ =e e s © oe” » eo e Z| 
a0 - 4 
3°40 Le = 
e ° CV, 2 @ cy a ? a -? 
7] 5B - simtiialiselve TS idee 0% Wa ee) FMI SPs =| 
ee » 8 Oy ‘ oe” so, Poa 3 ae Lome id 
5D el 
BE re e 9 e 9 9 ee ae @ @ ° Pier aame 
| . Q ° eo @ 0 ) g R 
a) 13-D ee | a ea) eS) oN Oy Se Q a4 
1 lat Yo ah ie aN os 6 %e o on 
13-25 aU 0 é : 0 6 3 A 
th 4 
e 
e 
R2 = e on e ~J| 
Le e e © e : e ae e? | 
0 O eo 
e e 
26 me 6 eh s ° 
e ° ee 
C 26-55 ite . SMe. Oras 
v en Oi fe: FIGURE A-2. Circumferential dis- 
aD ry \ i ¢° ane Le f iL 1 i tribution of piezoelectric pres- 


0 i) 80 120 1 °° 20 0 320 360 sure transducer signals for 15- 
BLADE POSITION 8 IN DEGREES 360° = ONE REVOLUTION (0.17 SEC ) knot run 205. 


WW 


12 


RI 


R2 


0 40 80 120 160 z 40 280 320 360 
14-49 = ommlps y aneareaia a ae 
Q oo 
og r coc, Gogo" 2 Q 9 \ J ° reg o 
1448 = \y bord * ? doe Fie wig Nees 0" 
° o o © og og 
14-42 ia 
a ) 
Q Po ° o%, 970900%> 
7:28 a o® Poo 920 
6 42 9%go oo @® Q eo a Q 
ey er Bog 3° % 2s bg \ c:) 
ry eo og 
= 
i 9 
pa a ¢ o 
pooo® PES) og Pa 9 2 a ow a 
16:35 e009 24% eS ee, yf? Beoe™\ OH 
“o. @ | eo / a 
#9 o \9 Pa og <j »? 
16°34 Ta 
lass Cc) 
5 ® o °° © ) ® 
25:37 e ee @ ® @ 
5 OMe ® cc) aires of - OF ® oO. 
e 
25-36 ee AOS O16 0 . 
: @ ® ©00°e 2% 
ree | al ieee [eee | | 
0 40 8 120 160 20 240 280 320 360 
BLADE POSITION 8 IN DEGREES 360°= ONE REVOLUTION (0.17 SEC) 


FIGURE A-3. Circumferential 
distribution of piezoelectric 
pressure transducer signals for 
8.9-knot run 208. 


247 


Influence of Propeller Action on 
Flow Field Around a Hull 


Shunichi Ishida 


Ishikawajima-Harima Heavy Industries Co, Ltd. 


Yokohama, Japan 


ABSTRACT 


Flow field in the vicinity of a hull is analyzed 
by using acceleration potential, and an approximate 
calculation method is derived. The present method 
can calculate the change of pressure on the hull 
caused by a propeller action. Numerical results 

by the present method are shown with experimental 
results. 

Wake far from a ship is analyzed by using Oseen's 
approximation, and an optimum condition is given 
for wake energy recovery by a propeller. This 
condition is examined by the results of the self- 
propulsion tests and the wake survey measurements 
at distant positions behind a ship. 


ibn INTRODUCTION 


When a hull is towed in still water, a flow field 
is induced around the hull. This flow field is 
very complicated, and becomes more complicated by 
propeller action. Many researchers have studied 
experimentally and theoretically the phenomena 
caused by the interaction of the hull and propeller, 
[Yamazaki et al. (1972) ]. Unfortunately, however, 
the number of practical uses of the study results 
is less than those derived in other fields of naval 
hydrodynamics. One of the reasons is because the 
various suggested methods are themselves complicated 
owing to the complexity of the phenomena. 

It has been popularly known that both the equa- 
tions and the boundary conditions which describe 
flow field can be simplified, and analyzed easily 
if disturbance by an object in the flow is a small 
quantity of the first order. One of the typical 
examples is the method of acceleration potential 
in inviscid flow fields used for propeller theory 
[Tsakonas et al. (1973) ]. Another example is 
Oseen's method in a viscous flow field used for 
the separation of hull resistance components [Baba 
(1969) ]. 


In this paper, the above-mentioned concept is 
applied to analysis of flow fields induced by the 
interaction of the hull and propeller, and the 
author derives practical methods relating to the 
propeller-induced pressure change on the hull and 
wake energy recovery by the propeller. Section 2 
explains coordinate systems used in this paper. 

In Section 3, the author applies the method of 
acceleration potential for analysis of inviscid 

flow fields in the vicinity of the hull, and derives 
a method which can be used to calculate the change 
of pressure induced by a propeller on a hull surface. 
In Section 4, the author applies Oseen's method for 
analysis of wake far from the hull, and derives a 
method to predict recovery of wake energy by the 
propeller. Then, this method is examined by the 
experimental results obtained from self-propulsion 
tests and the wake survey. Section 5 concludes 

this paper. 


Bo COORDINATE SYSTEMS 


We assume that a ship with a single propeller is 
moving with a constant speed on the free surface of 
still water. At first, we define a coordinate 
system O-XYZ fixed in space and a coordinate system 
o-xyz fixed on the hull as indicated in Figure 1. 
The coordinate system O-XYZ is an orthogonal coor- 
dinate system, in which the XZ-plane coincides with 
the still water surface and the positive direction 
of Y-axis coincides with an upward vertical line. 
The coordinate system o-xyz is a moving coordinate 
system in which the origin o is moving on the X-axis 
in the negative direction with a constant velocity 
U, and this sytem satisfies the following relation- 
ship with 0O-XYZ: 


Mako WE WS we 4S Bp (1) 


where t represents time. 
Next, we define two more coordinate systems 


FIGURE 1. 


Coordinate systems. 


0)-x1y1]Z and 0)-x)r8 related with the propeller 

as indicated in Figure 1. In the coordinate system 
01-X1]¥1Z1, the origin 0) coincides with the propeller 
center and we assume that the x]-axis coincides 

with the propeller axis and is parallel to the x- 
axis. Further, the coordinate system 0]-x ]y]2Z] has 
the following relationship with the coordinate 

system O-xyz: 

SF Op Oe SoS A oP Alp BS Bio (2) 
where (x,, -f, 0) are the coordinates of the pro- 
peller center on o-xyz. Moreover, the following 
relationship is satisfied between 0]-x y ]Z] and 
0)|-x)4r8: 


X] = X1, y) = © cos®, 2] = r sind. (3) 


3. PRESSURE ON A HULL SURFACE AND ACCELERATION 
POTENTIAL 


Pressure generated on the hull surface in the towed 
condition differs from that in the self-propulsion 
condition because of the influence of propeller 
action. The time-independent part of this change 
corresponds to the pressure component of the thrust 
deduction and the time-dependent part corresponds 
to the propeller-induced surface force. Now, with 
conventional methods devised to calculate these 
forces, numerical procedures tend to be extremely 
troublesome. Consequently, a great deal of calcu- 
lation time is required, especially in calculating 
propeller induced velocity, and it is hard to apply 
to a practical hull of a complicated form. Hence, 
an easy method with which the calculations of pro- 
peller influences can be reduced is needed. 

In this chapter, the method which can calculate 
change of pressure induced by a propeller on the 
hull surface is explained. This method can be 
obtained by using acceleration potential. 


Fundamental Equation 


In this section, we assume that the flow field 
around the hull is inviscid. This assumption may 
be considered reasonable in solving the problem of 
pressure on the hull surface when the boundary layer 
on the hull surface is thin. 

At first, let us examine the flow field around 
the hull in the towed condition. Denoting the 
velocity potential of disturbance due to the hull 


249 


by $,(x,y,2), the velocity potential for the over- 
all flow field can be expressed by U*x+},, and the 
following equation must be satisfied for $<: 


2 2 2 
anes r OS a cats 
ax2 ay2 az2 


=o. (4) 


Boundary conditions are given as follows. On the 
hull surface, S,, the following equation must be 
satisfied: 


(o + 2) + sae) £ ais 0 , (5) 
ny om Orn = : 
x dy y 0z Zz on & 


where Ny, Ny, and nz, represent xX-, y-, and z- com- 
ponents of the outward normal unit vector on S.. 

On the free surface, we have two boundary conditions. 
One of them can be obtained from the Bernoulli's 

law and the condition of constant pressure there, 

as follows: 


1 (See p es; fs (2) t Oe ‘ | ee 
2 | \ox dy dz ox SE eget 0 


(6) 


where (,(x,Z) represents the vertical displacement 
of the free surface, namely, wave height. Another 
boundary condition on the free surface is the 
kinematical condition as indicated below: 


a6 af ao ag ot 
(o + 2)". s Ss Ss Ss 


oe ho 8a... BE mies (7) 
y=Ss 


At infinity, the following boundary conditions 
might be given: 


Derivatives of bs > 0 when Vx? + y2 Foo => Cd 


C5 * O when Vx* + 22 > . (8) 


Next, let us examine the flow field around the 
hull in the self-propulsion condition. We assume, 
similarly to the towed condition, that the velocity 
potential of disturbance exists. Then, we can 
express the velocity potential of the overall flow 
field by Uex + >, + o*. Here, the 9*(x,y,z;t) xrep- 
resents the change of the velocity potential due 
to the propeller action when the moving condition 
is changed from the towed condition to the self- 
propulsion condition, and $* must satisfy the 
following equation: 


2 2 2 
ra) o* te A) o* ey A) o* A 
ax? dy? 2° 
We can also obtain the boundary conditions under 
the self-propulsion condition in the same manner 
as under the towed condition. In this case, however, 
time derivatives appear in some conditons by the 
influence of propeller rotation. On the hull sur- 
face, the following boundary condition is given: 


( Oy ag * (Ss ey 
Wise ee ee OS De pe 


OO. a 
lee vie) te 


Dio (9) 


on So. =O. (10) 


250 


On the free surface, the condition of constant 
pressure and the kinematical condition can be given 
as follows: 


/ Id*\ 9 3 ao* p) ap* 
al ted Cte ea a 
2 ax ax dy dy az dz 
36. ee) ao* 
ee — = ; 11 
+ U (= tae) eae oul \ lees (0) (11) 
Gis 2) es 
u ox ax ox dy by 
cle) ap* at ot 
s sp sp 
a Sie = 5 12 
a3 (= + 22 ) Ozie w moe vaca. me oe 


where Ssp represents the wave height in the self- 
propulsion condition. At infinity, the following 
boundary conditions might be given: 


Derivatives of ¢, + >* > O when Vx2 + y? +2270, 


G+ O when Vx? + 22 +m . (13) 
sp 
Finally, using the equations derived under the 
towed condition and the self-propulsion condition 
described above, let us derive the equation and 
boundary conditions for $* which express the change 
of the flow field around the hull due to the pro- 


peller action. At first, $* must satisfy the Laplace 


equation (9). Next, let us obtain the boundary 
conditions for ~*. On the hull surface, the fol- 
lowing relationship is given from (5) _and (10): 


CUES ab* , OWE = 6 
Fay es oy y i dz oz ee Ss, me oe 


On the free surface, the following equation is 
given from (6) and (11) in correspondence with the 
condition of constant pressure: 


oe (ee 
2 ax ax ay dy az az 


(ee es age 
F Ui einer) ites ak OY, leer 


8d_\2 (26.\2 /2¢.\ 2] a9 
Ee ley a a 
5 te ) *\ay Nae ee ON ace, 


(15) 


And, using (7) and (12), the following equation is 
given in correspondence with the kinematical 
condition: 


* 
Fg Alas Mas ag tance 
\ ax ox ax oy dy: 


ot 
Pst eae Weed sD Sp 
az oz / az at Yop 


Uh ilies: BUN eect ea RAD We nce 
5 \ ax ) ax oy Oz 


dd * 
(Zs Ch) Oo. 


aan eee (16) 


At infinity, the following boundary conditions 
might be given: 


Derivatives of »* > O when Vx? 4 22 aa Cie, 
c* + 0 when Vx? + z2 +0 , (17) 


where C* (x,z;t) represents change of wave height 
due to the propeller action and the following 
relationship must be satisfied: 


Ge -SGSpe Gs. (18) 


Acceleration Potential and Approximate Calculation 
Method 


Acceleration Potential 


The purpose of this section is to indicate that the 
equation and the boundary conditions for $* derived 
in the previous section can be expressed in the 
terms of acceleration potential on the assumption 
of thin hull. 

At first, using the assumption of thin hull, we 
express the shape of the hull as follows: 

Z = Of (o0,y,) in See p (19) 
where € represents a small quantity of the first 
order and S_* represents a projected plane of the 
hull surface, Ss, in the xy-plane. And, it seems 
reasonable to develop all our quantities in powers 
of €, as follows: 


o, = 1 + S40n' soe ; (20) 
o* = ed] + [765 +... ; (21) 
SoS en Seige ee ap (22) 
a Boi PE Ba cbo 7 (23) 


Thus, €* can also be developed as follows: 
* 2 * 
Ge 0 Gn ar EG Poco oo (24) 


Next, we proceed to obtain the equation and 
boundary conditions for $,* which correspond to the 
first order of € by substituting the development 
(20) ~(24) into the equation and boundary conditions 
for ~* in the previous section. The following 
equation in Q* can be obtained from (9) and (21): 


24k 24k Jak 
a2gt a2ge 92g 


(25) 
aa ay2 * 922 


Let us consider the boundary conditions for )*. 
First, using Eq. (19), we can estimate the magni- 
tude of nye ae ne in the Eq. (14) as follows: 


mim SO), mM = OS), mm = O@) - (26) 
x y Zz 


where O denotes the order symbol. In addition, we 
obtain from (19) and (21) 


ap* 
wie | Se + O(e*) , 


2=€f£ (x,y) =) 


(27) 


ap* oi 2 
al = 6 7 Ole) 5 (28) 
Y z=ef (x,y) Y z=0 
ag* 
* 
2 | ay = | + o(e2) (29) 
ca z=ef (x,y) z=0 
Hence, by substituting (26)~(29) into (14), we can 
obtain 
* 
do 
== = 0) sin SG (30) 
dz s 
z=0 
Further, for the boundary conditions on the free 
surface, the following equation can be obtained by 
substituting (20)~(24) into (15): 
* * 
d¢1 91 A 
bes so = al 
y=0 
And, in correspondence with the Eq. (16), the follow- 
ing equation is also obtained: 
* 
os 9b. eal | = 0 (32) 
ox Oy Oe rf ; 


* 
Hence, eliminating f, from (31) and (32), we can 
obtain the boundary conditions on the free surface: 


2k 
3 $5 
9x2 


2 a> o* L oO gi 


ag* 
a 1 
u2 9t2 


+ 
U2 ay 


= 0 (33) 


i BeBe eo 


Moreover, at infinity, boundary conditions are given 
as follows by (17), (21) and (24): 


A ee ee 
Derivatives of $, > 0 when Vx2 + y2 + 2250, (34) 


* —_——___—— 
en => © Winein (EOE ee a co 5 (35) 


Now, let us denote the pressure of the flow field 
in the towed condition and that in the self-propulsion 
condition by pg(x,y,z) and Psp (*,¥,2;t) respectively. 
By substituting (20) and (21) into Bernoulli's 
expression, we can obtain 


ag 

Ps 1 

SS = Ce 2 

ae gy € ax OMES) i, (36) 

) * * 

Psp _ - -ey( vy paki) = a + O(e2 
Or SY ox ox at (er) a (S) 
where 9, represents fluid density. Hence, the 


pressure change, i)(x,y,z;t), due to the interaction 
of the hull and propeller is given by the following 
equation: 


1 
a = a 
ig if 


This equation shows that the magnitude of Wy is of 


( aor a) 
sp ir P,) Geis Cae at /- (Ss) 


251 


the order of € Moreover, W/P. can be considered 
as an acceleration potential as is obvious from the 
relationship with $y: 

Finally, we proceed to convert Eqs. (25),(30), 
(33), (34), and (35) for b} to equations for \) by 


using the relationship (38). Using (25) and (38), 
y must satisfy the following equation: 
a2y . a2p a2 
Pal BY Bae 7S (22) 
On the hull surface, So’ we can obtain from (26) 
the following equation: 
3 
in | eae gona cil 
on s y °y < on S 
s s 
25 a | + O(e7) in sé (40) 


z=0 


On the other hand, if (x,y) is a point on See the 
following equation can be obtained from (30) and 


(38): 
a a 362 
1 
Chie 
3 F) o* 
-c +5): pew (cae alam © 


* 
Thus, the hull surface condition for $, can be con- 
verted to that for | as follows: 


(41) 


(42) 


Similarly, the free surface condition (33) for )* 
can be converted to that for W as follows: 


2 2 2 3 
a*w my Bo ha) Peal oa) a & v = 6 (43) 
9x? U dxdt Ul dts Udy y=0 
Moreover, for the boundary condition at infinity, 


the following equation is given from (34) and (35): 


W > O when Yx2 + y2 + 22 30. (44) 


Integral Equation 


We proceed to seek the solution of which is the 
harmonic function in the region bounded above by 
the plane y=0 and elsewhere by the hull surface and 
satisfies boundary conditions (42), (43), and (44). 

At first, we separate the solution into the two 
parts and write it as follows: 

W(x,y,Z;t) = V(x,y,z;t) + W(x,y,z;t), (45) 
where both V and W are the harmonic functions in 
the region as indicated above. Moreover, let W 
represent the pressure induced by a rotating pro- 
peller moving straight ahead with a constant speed 
in still water and a free surface. Now, we have 


252 


many formulas for W*(x,y,z;t) which represents (52), (53), (55), and (56) with the method of Green's 
the pressure induced by an N-bladed propeller moving function: 
in infinite space. One of the formulas for W* is 
given on the assumption of thin blades as follows ay (2) gy (2) 
[Jakobs et al. (1972) ]: (e) ss ( ¥ Y ) 
anv *°" (Q,) = eS) (Sa 2 a) @ & (OpO5) 
. 2 JvNt ns 
Wty mW (a y771Z) ae 
v=0 v y(t) 
+(V 
en ‘ ee (ans ‘ v ) 5am, i, a, (2; Q ) 5 (7) 
= 25 1 fas Da L.(E",6,89)e> ese se at ; 
v=0 47 a q=1 =0 rR where Q, denotes a point outside Ss, Q denotes a 
e p Pp (46) point on S,, and the suffix, (i), means the inside 
where 8g = Zu (GN) (47) of the hull surface. Then, we seek a solution of 
N vy, (2) which satisfies the boundary conditon on S., 
as follows: 
with j = imaginary unit, 9% = angular speed of the gy (2) 
= 3W 
propeller, Sp = lifting surface of propeller, L)' = v | eek | (58) 
pressure jump across Sp, (&',0, 89) = point on Sp, dn Bie on Sats 
is = normal unit vector at Spr and R = distance s s 
between (&',9,89) and (x,y,z). Hence, using W* and ; . ‘ 
the method of the mirror image, we can obtain a W Then, using (53) and (58), we can obtain an internal 
which satisfies the boundary conditions (43) and solution as follows: 
(44). Then, we can write W as follows: 
we? Sth, « (59) 
© }V. 
evince) See (48) 
Therefore, by substituting (51), (58), and (59) 
5 g e : 
Next, let us consider V. Then, we assume that into (57), the external solution WS must satisfy 


Vv and can be developed in correspondence with the 


development (48) as indicated below: ( (e) ) 9 / 
ATV (e) = pees Ts 4 
ay ey) GSA aon ie ang a, (0:06 ) 
ya 8 ch Cegrpaen (49) S (60) 
veo" *r¥s x eS 
co JUL F 4 . 2 
V = LOVy (ery, Ze & (50) Finally, we have the following equation by adding 
v= B 


anw, (Q.) to both sides of the Eq. (60): 
Hence, using (42) and (44), we have 


(e) az (2) (oy eelton 
av, (e) | OW, a Amp (Q.) St MMO) ap ds (2) on5 ) (Q79,) : 
22 on g Coons 8. (61) 
Ss Ss 
7 eS Olwhen Woe v2 pecs (52) In this equation, letting Q, be the limit of Q, on 


oa we can get 


where the suffix, (e), means the outside of the 


hull surface. In the same manner, from Eq. (43), (e) 1 ) 
( ares da (e) Sees : 
we have YS) to8) 20 S YW) 6 G,, (0790) 
ONG, av, ov ee 
+ = + Ka 2 i W, | = 05 (53) 
gx? dy ox =0 
= 2W (Q.) 62 
where yee 2 ce 
: 202 
Cia i 2gwk -_ _ EV 
Kos ee BS) = pO 5 : (54) because the singularity of first order exists in 
U W U the ey Se chis Wle) (9 ) is exactly the change of 
Now, we suppose that we know the functions G (&, pressure on the hull surface caused by the propeller 
TVA ECAVIZ) ie U= Ole 2 een ) such that the Gare which we intend to calculate. If W. and G, can be 
harmonic functions for n<O except at (x,y,z) where given a priori, Eq.(62) can be considered to be 
G have a singularity of first order, and G., satisfy * an integral equation for the unknown )(e) (Q5)- Thus, 
he boundary conditions: the problem of calculating the change of pressure i 
on a hull surface caused by a propeller changes to 
once dG 0G the problem of solving an integral equation. 
Vv 
+ Ko— + Ki— + KG =o, (55) 
ae? an 0g ws 
Vie Time-Independent Change of Pressure On the Hull 
As 2 2 Zip 
S,) Suen Vx Pe ihe ame are Meare KC (56) Now, we proceed to give Wo and Go for a steady case 


(v=0). Go(&,n,0;x,y,2) can be written as follows 
Then, we can obtain the following equation by using based on a wave making theory: 


1 
V(E-x)2 + (n-y)2 + (G-2)? 


Go (EN, Gi X,Y 12) = 


T 
— foe) 
2 
ie k exp (k nty + ikp') (63) 
T ae Aye Bees ss inn ee 
k - Kgsec-6 
= 0 
2 
where p' = (&-x) cos @ + (t-z) sin 0. (64) 


We can get Wo by using Eq. (46) as follows. The 
first step is to rewrite the integrated term in the 
right side of Eq. (46) by the transformation 


pe Ey EOL lay (65) 
n 


Then, using the rewritten expression, we can obtain 
Wo as follows: 
Wo(x,y,Z) = Wa (x,y 2) ae (66) 

It should be understood from the above expla- 
nation that L)' must be given to calculate Wo. In 
order to obtain L)' precisely, we must consider the 
boundary conditions on the propeller surface which 
have been disregarded in the discussion up to this 
step. To do so, however, requires complicated 
calculations as seen in the conventional methods 
for the problems of the hull-propeller interaction. 
The complexity of the calculations have caused the 
conventional methods to be impractical as described 
in the Section 1. Hence, the author introduces the 
following approximation. The steady change of 
pressure on a hull surface which we are now examining 
corresponds to a pressure component of thrust de- 
duction. We can consider that obtaining the thrust 
deduction is the same as obtaining pressure on the 
hull surface as a percentage of the mean propeller 
thrust, To. Hence, the relationship between un- 
known Lg' and known Tg can be given as follows: 


N 
- 2 J esti eGo) I = @o > (67) 
q=l es 


Ss 
Pp 


where [ |]. denotes the component in x direction. 
Now, the L,' cin be considered as the jump of the 
pressure change across the propeller surface due 

to the interaction of the hull and propeller, and 
consequently, Eq. (67) may be considered as the 
approximate boundary condition on the propeller 
surface for vfey(g.). By giving an arbitrary 
function, L,', which satisfies the auxiliary Eq. (67) 
and calculating Wo by (46), (65), and (66), we can 
solve the integral equation, (62). This is the 
approximate calculation method proposed in this 
paper. 


253 


Numerical Procedure 


The purpose of this section is to describe the 
numerical procedure for the method explained in 
the previous section. Here, for convenience' sake, 
let us denote §@) in (62) by wr. 


Numerical Calculation 


The integral equation, (62) is an integral equa- 
tion of Fredholm type of the 2nd kind. Generally, 
it is impossible to obtain analytic solutions of 
the integral equation for S_ in an arbitrary form. 
Thus, various approximation methods have been 
suggested. In this paper, a definite integral is 
approximated by a finite sum, the equation is con- 
verted to a linear equation, and this equation is 
solved numerically. 

At first, the following approximations are used: 
(i) A hull in an arbitrary form is replaced by a 

polyhedron. The form of each surface named 
"element" is a plane quadrilateral. 
(ii) On each element, the unknown function Va (Or) 
is assumed to be constant. 
Using this approximation, the continuous function 
VEMOR) is replaced by the discrete quantities, vF 
(UST Di te rcketers , M), for the total number, M, of the 
elements. A control point, Q , where Wo (Q_) must 


be calculated, is selected fof each element. Thus, 
we have the following transformation: 
‘ pee ¥ 3G, 
ds os Seeds Oar v, CB ang? iQue (68) 
Ss aL element 


where ase no, and Q' denote values on the elements. 
The definite integral in the right side of this 
equation is an influence function from point Q to 
point Q 5 and we denote this function by Ag ,Q- 

On calculating Ag,,Q, the existence of a singular 
point, a so called doublet, becomes a problem. How- 
ever, there are many numerical calculation methods 
for this case. In this paper, the Hess-Smith method 
is used [Hess and Smith (1967)]. Further, selection 
of a control point is also a problem. However, for 
this problem various methods have also been suggested 
in the analysis of potential flow field. In this 
paper, each element is selected to be similar toa 
rectangle, and the point of intersection of its 
diagonal lines is employed as the control point. 
Finally, the hull surface after St.11/2 is taken 
into consideration, and it is divided more nar- 
rowly near stern in the longitudinal direction and 
approximately equally in the depth direction. 

Thus, each element, Aj, i'(i,i'= pera c.ccghy) 4 
which corresponds to Ag,,Q can be calculated and 
Wo (Qo), can be calculated for each control point. 
Then, the integral equation of unknown function, 
W*(Q_), is converted to a linear equation of un- 
known, i*- 

Now, in the calculation of Wo(Q_), the author 
uses the approximation that the number of propeller 
blades is infinite. Then, in correspondence with 
(46), (47), (65), and ((6), we can get the following 
relations: 


254 


ro PAL 
. 1 3 ~ 
Wo(Q) =, race dr de@n?r Ce) anne , 
(69) 
rE ie} 


Go (rv, 93%,¥7Z) = Go(xp, xr cos@ -f, sinO; x,y,z), (70) 


(71) 


where H represents a mean pitch of the propeller 
blade, and r, and rg represent respectively radius 
of the boss and radius of the propeller. Moreover, 
['(r,8) represents the thrust per unit length in the 
radial direction of the propeller blade elements 
and can be developed as follows: 


P(e,8) = 42 Tie . (72) 


We can also get the following equation in correspon- 
dence with the Eq. -(67): 


a2 
° 
To = dr NI'p (x) 6 (73) 
J i 
B 
Further, for the calculation of Wo in Eq. (69), it 


is approximated that I) is an elliptic distribution 
against r, and [,, To ....are disregarded. 


Examples 
The numerical calculations are performed in the 


case of two combinations of the hull and propeller 
shown in Table 1. Figure 2 shows the body plan of 


hulls. In order to examine the correctness of the 
TABLE 1 Particulars and Operating Condition 
U 
SHIP Lop He B/T Cc D Z Ba Ou 
L 6-00) (650) 92/586 SBW2 oS S 2505 oAo7 Oa'55 
a 7.00 6.00 2.63 BYE) 52l@  aboA7 ailSs) Sos 
Lpp = Length between perpendiculars (meter), B = 
Breadth, 
T = Draft at mid-section, Cp = Block coefficient, 
D = Propeller diameter (meter), Z = Number of 
propeller blades, 
U = Ship speed (meter/second), Fn = Froude number, 
nq = Propeller's number of revolution per second. 


Ne 


T ship L ship 


FIGURE 2. Body plan. 


approximation used in the calculation of Wo, the 
procedure as follows is performed. First, perfor- 
mance of the propeller in the nominal wake is calcu- 
lated to obtain IT), and L,~. Next, by using the 

five combinations of distribution forms of I), L)~ 
and the number of the propeller blades as follows: 


(al) SN Eeimsktey asain Geli hy lly iereen beget Fy ltge eatin (SS) 
(03) INP alweiotiee, meshing Waplng goss , IT7 in (69) 
(c) N; finite, using Lo' only in (66) 

(d) N; infinite, using Ig only in (69) 

(e) N; infinite, using Ip: elliptic in (69) , 


The Wp are calculated. Then, by substituting these 
Wo in (62), the pressure changes, W*, are calculated 
and indicated in a non-dimensional form in Figure 3. 
As shown in Figure 3, the * barely differ due to 
the distribution form of I!,L', and the number of 
propeller blades. Hence, the approximation of the 
elliptic distribution is reasonable. 


Experiment 


The experiment was performed at the towing tank of 
IHI by applying a standard hull surface pressure 
measurement [Namimatsu, (1976)]. For the ships 
indicated in Tdble 1, pressures on the hull surface 
are measured under both the towed and the self- 
propulsion condition. Differences of the measured 
pressure between the towed and the self-propulsion 
condition are used for the experimental values of 
the pressure change caused by the propeller. 

Figure 4 shows the comparison of the experimental 
values to the calculated values, which are obtained 
by approximating Ip as the elliptic distribution. 
In addition, Table 2 shows the pressure component, 
tor of the thrust deduction fraction, t, which is 
the sum of the pressure change. The comparison in- 
dicates better agreement for the L ship (a thinner 
ship). 


Discussion 


The calculation method in this paper is derived by 
expressing the equations and boundary conditions 
(which determine the change of the flow field due 
to the interaction of the hull and propeller) in 
the form of an acceleration potential. For this 
reason, this method nominally requires calculations 
of pressures induced by the hull and propeller, 
while the conventional methods, which express flow 
fields in the form of a velocity potential, require 


L ship 


y=|.-61.3 


mn 


150mm 100 50 aCp =276.9 
I 
Mg 

2 
I 


elliptic 
p* 


Pe 


ACp= 


100 


Zz 
150 mm 50 


(y,z)=coordinate of a point on hull surface U(ship speed)=2.05m/sec 


FIGURE 3. 


calculations of pressures and velocities induced by 
the hull and propeller. Generally, the calculations 
of induced pressure require less time in comparison 
with the calculations of induced velocity. Thus, 
when the present method is used, the time required 
for numerical calculations can be reduced to a 
practical value. This method can also be applied 
for the calculation of propeller-induced surface 
forces [Ishida, (1975) ]. 

It is anticipated that the results derived by 
this method may be worse as the calculation point 
moves closer to the stern, because, in this method, 
the assumption of a thin hull is used, propeller 
boundary conditions are simplified, and the rudder 
is disregarded. When the actual experimental values 
are examined, it seems that the anticipation may be 
correct. However, it iS more appropriate to con- 
sider that the majority of the error is due to the 
fact that the flow field around the hull is assumed 
to be inviscid. 


4. WAKE ENERGY RECOVERY BY A PROPELLER 

A towed hull pulls still water forward, but when 

the hull is self-propelled, the propeller acceler- 
ates this forward flow toward the back, and thus, 
the propeller recovers wake energy. Hence, it is 
important for the improvement of propulsion effi- 
ciency of a ship to know how the wake energy can be 
recovered effectively. The present, self-propulsion 
test method can give information for the wake energy 
recovery as a propulsion factor. This method is, 
however, insufficient to tell us how wake energy 


255 


T ship 4c, 


y= [£139.8] mm =326. “4 
Ye erate LEON, 5 
200mm 150 100 50 : 


(E186-5] 


-0.1 
pal 
———_ + + 0 
200mm 150 100 50 
AC 
P 
(233-2) © 200mm 150 
-0.2 39555 -0.3 


2 + = 
200mm 150 100 50 


(y,z)= coordinate of a point on hull surface U(ship speed)=1.27m/sec 


Numerical calculation of pressure change on a hull. 


should be effectively recovered. This is due to 
the fact that the balance of force is a basic prin- 


ciple of analysis in the method, in which the balance 


of energy is not given sufficient consideration, 
and further, because almost no information on the 
flow field can be given. To cover the fault of the 
self-propulsion test method, a knowledge of the 


overall flow field is necessary and the distribution 
In the vicinity 


of energy in the flow must be found. 
of the propeller, however, the flow field is so 
complicated that experimental measurement and 
theoretical analysis are difficult. 
consider, as a practical approximation, an attempt 
to estimate wake energy recovery by a propeller 
through an analysis of the wake at a position far 
from the propeller. 

In the next section, the phenomena of the inter- 
action in a distant wake are analyzed by the use 
of Oseen's approximation to determine under what 


Hence, we might 


TABLE 2 Thrust Deduction Fraction 
SHIP t t ie 
Pp Pp 
L .- 166 .140 -109 
ay oOul .-160 -200 


t,, is obtained from pressure measurement. 
tp* is calculated by present method. 


256 


L ship 


* 150 mm 


[E184.5] 


CALCULATION 
=o EXPERIMENT 


3 150 mm 100 SO 0 


(y,2)= coordinate of a point on hull surface U(ship speed)=2.05m/sec 


FIGURE 4. 


conditions the wake energy is effectively recovered 
by the propeller. 


Fundamental Equation 


In this section, we assume that a ship is stationary 
in a uniform flow of speed U. We proceed to examine 
the balances of force and energy between the ship 
and the flow field. 

Now, for the surfaces where force and energy are 
surveyed, we define six rectangular cross-sections 
in addition to the hull and propeller surfaces. 
These six rectangular cross-sections are indicated 
in Figure 5. Two vertical planes are in right angle 
to the direction of the uniform flow at the front 
and rear of the hull. The free surface and the 
bottom of the water are held between the two vertical 
planes, and two more vertical planes are parallel 
to the uniform flow at infinite distances to the 
right and left of the hull. Further for simplicity, 
we assume that the flow field is independent of 
time even if a propeller exists and a coefficient 
of diffusion, Ue, due to viscosity or turbulent 
flow is constant. Moreover, notations used here 
have the same meaning as those in Section 3. 

At first, let us examine the input and output 
of momentum at the individual surveyed surface in 
the towed condition. Then, as a result, the total 
resistance, R_, can be given by the integration on 
the rectangular cross section, S., in the rear of 
the hull as follows: * 


T ship 


y= [E1328] nm 


*700 mm150 100 


=186.5 


“200 mm 150 


=0.2 37305 023) 


Zz 0 gal 
200 mm 150 
ACh 
-0.4 Ss 
-279.9 ——————— 
200 mm 150 


-0.3 


Zz 
200 mm 150 100 


(y,z)=coordinate of a point on hull surface U(ship speed)=1.27m/sec 


Comparison between calculated and experimental value. 


du 


ds[po-p+2u, ae a 


u_(Ut+u_) | 
s s 


(74) 


b 
aL 2 
+ > Pd AZEGSi, 
-b 


where u, v, and w represent x-, y-, and z-components 
of disturbance velocity and b represents the half 
width of S. at the free surface. Further, po repre- 
sents the pressure at x = -®. Moreover, when the 
energy balance is examined, kinetic energy lost 

when the uniform flow passes along the hull must 

be equal to the sum of the energy dissipated to the 
outside through the surveyed surfaces by heat and 


work. Thus, we can obtain the equation as follows: 
p 2 
£ 2 —=—= 2 2 
= ds[u* - (Utug + v_ + w.)](U + u_) 
on 
= av @(@) = ds {pop or) 
Vv Ss 


it [ du. 
re a(u + 7) Ea 

(Ge du 4 
tig pe 


2) 


where V. denotes the flow field surrounded by the 
Survey Surfaces, and 


; dus) 2 av, \2 aw.) 2 
® (e) = ue (—*) + =) + (5) 
aw av.) 2 ow ane 
(52 +3) (yet ae 


(2s vs | 
Vp EET 


By using (74) and (75), the effective horsepower, 
EHP, can be expressed as follows: 


p 
iis d(e) + cau dS (ug+vgtw2) (U+ug) 
Sp 


WE 


(75) 


N 


(76) 


EHP = 


UP 5 
Ps Gigs dS ug (Pg-Po) 
=) Sa 

jh du. (Se ou. 

a Ws 4 dS |2u, cra a? Wes <a ay ) 
A 

(ts **s) 

qr Ws Deas a 6 (77) 


The first term on the right side expresses heat 
energy, the second term expresses the increase of 
kinetic energy, the third term expresses the in- 
crease of potential energy, and the fourth and fifth 
terms express work toward the outside of Vs. This 
equation, (77) gives the work, EHP, transmitted 
to the fluid through the hull when the hull is 
towed in still water. 

Next, the self-propulsion condition can be 
considered in the same manner as the towed condition. 
The equation for the balance of forces is as follows: 


du 
= Bis 
AR = ds E Days 5 P gu iy (OR ) 

or 

b 
i 2 

7 a PS dz Crs! (78) 

-b 


where the subscript sp denotes the self-propulsion 
condition and AR represents the skin friction 
correction which is used for the ordinary propulsion 
test at the towing tank. When the energy balance 

is considered, we can get the following equation: 


FIGURE 5. Survey surfaces. 
DHP -{ dv oe (e) 
Wee 
Pe 
-=— as |u2 =\ uraget + v2_ + w2 (Utu__) 
2 s sp s sp 
Sa 
ie oe (Utu__) 
Sa 
Ons gives Usp 
ar Be BUG) 5 + ul x oe ) 
3 
" (2 P “sp ) (79) 
sp ax az 
where 
du 2 av 2 ow 2 
= LG) (Ge) Ge) 
O55) = E ( re 12 ae 
ee 262) @e: 53)? 
Ne! OBE UAE ky TE 
XG sp)? 
a3 ( ox Oz (So) 


Then, using 


equation: 
DHP + UAR = 
Vv 
p 
~ ds(u2_ + 
2 Ss 
iS) 
A 
b 
Up .g 
pet fe ce + 
sp 
-b 


dv © (e) 
sp 
ie 
Cb OF VW 2 m_)) 
sp 
ds Bap Bao 2) 
Sa 


PST 


(78) and (79), we can get the following 


ae Os ; 
+a. ( ax -] 22) ji (o7)) 


This equation reveals that the work transmitted to 
the fluid by the ship moving in still water with a 
constant speed, U (sum of the delivered horsepower 


and the work UAR caused by skin friction correction), 


changes in the fluid and is dissipated as heat, 
kinetic energy, potential energy, and work through 
the surface si: 


Oseen's Approximation and Problem of Variations 


We assume that the hull is thin and S_ is placed 
sufficiently far behind the hull. Then, the inte- 
grations on S which appear in the right side of 
the Eqs. (74), (77), (78), and (81) can be approx- 
imated as indicated in the Appendix. Hence, the 
following equations can be obtained: 


WwW 
2 2 2 
Oe eee 
ae om ds ay x ox 
Sa bY 
b 
Og) 5 
a if dztc, (82) 
—b 
AR = Pd J estes - 4H ey 
Gd) 
Pe i (Es) ( ‘sp )? (Es) 
+ ds + = 
2 S dy dz ox 
A 
b 
Pg 2 
rn £ dz csp , (83) 
7 
-b 


where Ho, H_, and He represent the total head as 
follows: P 


Po u2 
Hn = — + ay (84 
0 Pd Y 2g ) 
B 2 
HO ==—+y += (UFO + v2 + w*) A (85) 
s [oye s 
£ 
12) 2 


= SS sy Ee (U+u vanes we) (86) 


Further, W represents the sectional area in which 
Ho-H is not equal to zero at S,- 


And, 
Po 
EHP = av ® (e) + aS(Ho - H_)? 
Ss Ss 
V, mn) 
pies o (eye Hee) (“s\? 
2 dy az ax 
Sa 
b 
P-gU 
4 azine (87) 
2 s 
-b 
Po 2 
DHP + UAR = av a) + aS(Hjp - H_) 
Ww 
Ve 
U a 3 3 
lee e (ea (Gen) ( ‘ep)?| 
2 ay az ax 
ox 
b 
 -JU 5 
a) dz Usp (88) 
-b 


In the Eqs. (82) and (83) for the balance of force, 
the forces Rt and AR, given to the fluid from the 
outside are divided into the force related to the 
viscosity expressed by the first term and the force 
related to the wave making expressed by the second 
and third terms. In Eqs. (87) and (88) for the 
balance of energy, the energies EHP and DHP + UAR 
given to the fluid from the outside are independently 
divided into the first and second terms which repre- 
sent the energy related to viscosity and into the 
third term and the fourth term which represent the 
energy related to wave making. 

Now, using (87) and (88) which show that the 
viscous energy and the potential energy are indepen- 
dent of each other, it is obvious that the condition 
for minimizing the viscous energy in (88) is a 
necessary condition for minimizing the DHP. We 
proceed, therefore, to obtain the minimum condition 
of the viscous energy which corresponds to the 
optimum condition for the energy recovery by the 
propeller. For this discussion, we assume that in 
the right side of Eq. (88), the first and second 
terms change independently or that the increase 
and decrease of the second term have, at least, a 
positive correlation with the increase and decrease 
of the first term. Based on this assumption, let 
us consider the conditions required in minimizing 
the following function: 


a (89) 


Using (83), the following equation is obtained: 


da = = - ; 90 
PI S(Ho Bao AR Rg, (90) 
W 


where R_ denotes a wave making resistance under a 
self-propulsion condition. This R_ might be 
approximately equal to a wave making resistance 
under the towed condition. Furthermore, AR can 
also be given by the total resistance under the 
towed condition. Hence, it can be considered that, 
under the self-propulsion condition, the following 
equation is given: 


g dS(Hy) - H_) = C, (91) 
sp 


where C is constant and can be decided by the towed 
condition. Thus, the problem of minimization of E 
is converted to the problem of variations for 
minimization of E given by (89) under the constraint 
condition (91). It is obvious that the following 
solution exists for the problem of variations: 


Ho - H = constant. (92) 
sp 

Furthermore, although it is omitted here, at least 
the conditions that the ship speed and displacement 
are constant are implicitly required in addition 
to this constraint condition. 

Let us consider the meaning of Eq. (92). Since 
Ho - Hs and Ho - Hsp are proportional to the viscous 
wake in a position far from the hull as indicated in 
the Appendix, (Ho - Hg) 2 and (Hg - Hsp) * are propor- 
tional to the kinetic energy of the viscous wake. 
Hence, the minimization of E corresponds to the 
minimization of the kinetic energy of the viscous 
wake. And, it can be considered that the condition 


259 


(92) is the condition for minimizing the kinetic 
energy left in the wake by recovering the kinetic 
energy of the viscous wake with the propeller. 

The optimum condition for this energy recovery 
is obtained under the assumption that the constant 
C of Eq. (91) is given as the constant decided by 
the towed condition. In other words, it is con- 
sidered that condition (92) gives only the condition 
for the propeller to accelerate flow effectively 
under the assumption. If, however, the wave making 
resistance is zero under a purely self-propulsion 
condition, then (AR=0) C can be expressed as C=0 
regardless of the towed condition. Therefore, it 
can be considered that this fact indicates condition 
(92) applies not only to the optimization of the 
flow acceleration by the propeller but also to the 
optimization of the hull-propeller combination for 
effective recovery of the wake energy. 

The author proceeds to examine the correctness 
of this condition in the following sections by 
using results of the self-propulsion tests and 
wake survey measurements. 


Experiment 


Total head at a wake far from the hull was measured 
at the towing tank of IHI. The measurements were 
performed for the ships and operating conditions 
indicated in the Table 1 under both the towed and 
the self-propulsion conditions. The measurement 
cross-sections which correspond to plane Sag were 
three vertical cross-sections of 0.3Llpp, 0.5Lppr 
and 0.7Lpp behind A.P. Figure 6 shows the total 
head loss distribution of the towed condition in 
the non-dimensional forms and also shows H,* which 
is the change of total head loss by the propeller 
action. Here, Hp* is obtained as follows: 


Bie = (ig Cel.) iy = 1) 6 (93) 
Pp sp s 


We observe that in the towed condition the wake of 
the T ship spreads to the relatively lower region 
of the flow field. Further, we can see that the 
peak of the total head distribution in the towed 
condition agrees well with the peak of the change 
distribution for the T ship, but not for the L 
ship. In addition, Table 3 shows results of the 


TABLE 3 Self-propulsion and Towed Test Data and 


Wake Survey 


Sieg 1 te 
A Ww Ths te Rg, AR EE a 

L 5287 collGG aml DoW $557 .@O8 Agia, 5.42 150 

ae 18S .20 OS 1.58 8.92 ails@ 1660 So os’) 

Ww = Effective wake, Re = Total resistance from towing 
test (kg.), 

Ry = Wave resistance from wave analysis at towed condi- 
tion (kg.), 

AR = Skin friction correction (kg.), 

fs = ptg J dS(Ho-Hg) at 0.7 LIpp behind ship in towed 


condition (kg.), 


f£sp = p£9 J ds(Ho-Hgp) at 0.7 Lpp behind ship in self- 
propulsion condition (kg.). 


260 


L ship 


323.1mm ——-+ 


(A) 5 Zuni 


—— TOWED CONDITION 
SS=oss CHANGE OF TOTAL 


U=2.05m/sec 


ey Oo Sim 


1 
g(Hp-Hs) /50" 


dt 
es bee 
gH, [ru 


U=1.27m/sec 


FIGURE 6. Total head 
distribution far from 
a ship. 


self-propulstion test and the towing test, and 
viscous resistances obtained from the wake survey. 


Discussion 


By analyzing the wake at a distant position behind 
a ship, an estimate of the recovery of the wake 
energy by the propeller is made, and the optimum 
condition (92) is given. Table 3 shows that hull 


MEASUREMENT SECTION 
0.3Lpp behind A.P. 


efficiency is better for the T ship than for the L 
ship. Results of the self-propulsion test, therefore, 
indicate that the energy recovery by the propeller 
is better for the T ship. On the other hand, results 
of wake survey measurement far from a ship indicate 
that for the T ship, the peak of the head change 
distribution agrees well with the peak of the head 
distribution in the towed condition. Hence, it can 

be considered that the propeller of the T ship makes 
the wake flatter in order to adapt the conditon (92). 


261 


TOWED CONDITION 
—--—-—- CHANGE OF TOTAL HEAD 


U=2.05m/sec 


583. 3mm 


U=1.27m/sec 


Thus, condition (92) is not contradictory to the 
results of the self-propulsion test. 


Bo CONCLUSION 


From the theoretical and experimental studies for 
the interaction of the hull and propeller, the 
following conclusions are derived: 

(i) Flow field in the vicinity of a hull is 


MEASUREMENT SECTION 
0.5Lpp behind A.P. 


FIGURE 6. (continued). 


analyzed by using acceleration potential, and the 
approximate calculation method is derived. This 
method can be used to calculate the change of 
pressure on the hull and has a higher practical 
applicability than conventional methods. 

(ii) For the analysis of the wake at a distant 
position behind a ship Oseen's approximation is 
used, and the optimum condition is given for the 
wake energy recovery by the propeller. This 
condition is examined by the results of the self- 
propulsion tests and the wake survey measurements. 


|L ship 


| 
eee 
of 
N 
~ 
| U=2.05m/sec 
ess 
| % Ss SSE ae 
| TOWED CONDITION ~s.>>--=72-~7 
Sa CHANGE OF TOTAL HEAD 
T ship 


— 443. 3mm— 


FIGURE 6. (continued). 


ACKNOWLEDGMENT 


Before closing this paper, the author would like 
to express his deep thanks to Prof. R. Yamazaki of 


Kyushu University who kindly examined the contents. 


The author also thanks Prof. T. Jinnaka and all 
members of IHI Yokohama Ship Model Basin who pro- 
vided him with kind guidance and cooperation. 


MEASUREMENT SECTION 
0.7Lpp behind A.P. 


REFERENCES 


Baba, E. (1969). Study on Separation of Ship 
Resistance Component. J. of the Society of Naval 
Architect of Japan, 125, 9. 

Hess, J. L., and A. M. O. Smith (1967). Calculation 
of Potential Flow about Arbitrary Bodies. 
Progress in Aeronautical Science, 8. 


Ishida, S. (1975). On an Approximate Calculus of 
the Propeller-induced Surface Force. J. of the 
Society of Naval Architect of Japan, 138, lll. 

Jacobs, W. R., J. Mercier, and S. Tsakonas (1972). 
Theory and Measurements of the Propeller-Induced 
Vibratory Pressure Field. J. of Ship Research, 
N55 Bp Mac. 

Namimatsu, M. (1976). 
Pressure Resistance and Its Application. 


A Measuring Method of Hull 
Wigs Che 


APPENDIX 


Let us examine the definite integral in Eqs. (74) 
and (78) for the balance of force and the definite 
integral in Eqs. (77) and (81) for the balance of 
energy. At first, we denote these integrals by F 


E 
and Ee as follows: 


= du 
FE as [bo 2 ? Qe qa p-u(U + u) 
& 
Sa b 
1 2 
AP > Pg dz c 1 (94) 
-b 


Hy 
(0) 
WW 
S| mo} 
Hh 
n 
Q sy 
Q 
n 
(as ™ 
(S 
nN 
+ 
<q 
i) 
+ 
<= 
nN 
Sq 
+ 
ft 
+ 
Ss 
7 
ue 
oO 


(95) 


‘ Up _g B 
du dv du dw du 8 2 
a Ue E + aa + =) + o( 2 + 22) + =f aziGe. 


-b 


If the terms to which uw is related are assumed to 


be small, Ee and Ea can be rewritten as follows: 


p 
= i me 2 2B 
Fe = Pg dw (Ho 130) ar 5 das (v°+ w-u*) 
Wy S, 
0.9 
+ = a Gy (96) 
=b 
p,U 
sae = “2,9 dw- u(Hg - H) + Sam das (v>+ we-u-) 
Ww - oN 
epi tae éa Gg (97) 
2 
-b 


263 


the Society of Naval Architects of Japan, 139, 13. 

Tsakonas, S., W. R. Jacobs, and M. R. Ali (1973). 

An "Exact" Linear Lifting-Surface Theory for a 
Marine Propeller in a Nonuniform Flow Field. 
J. of Ship Research, 17, 4; 196. 

Yamazaki, R., K. Nakatake, and K. Ueda (1972). On 
the Propulsion Theory of Ships on Still Water. 
Memoirs of the Faculty of Engineering, Kyushu 
University, 31, No. 4. 


where H represents total head as follows: 


= 2 
Fo a Gar ke eS . (98) 
Pd 2g 


Now, using Oseen's approximation, the following 
relationship can be written: 


Meee bw we Os Wn Wi SS Se db ity A (99) 
ox az 


where ~ represents velocity potential, and u', v', 
and w' represent velocity components of rotational 
motion which are zero at other than W. Then, pres- 
sure, p, and wave height, C, can be expressed as 
follows: 


cate a6 (100) 
= SPA! = WAU eeein 


SBE 6 oy (101) 
12) 


where Tf, is due to a potential motion and [' is due 
to a rotational motion. 

Substituting (99), (100), and (101) into (96) and 
(97), we can get 


dz Gar 


-b -b! 


+ (Vo)? + ow | (102) 


a 
b b! 
0 _gU 0 -gU 
f 2 iB 2 
+ d ' 
+ 5 dz Ss 5 s Cc 
-b —b\ 
b' 
auiees ime 8 ty (103) 
+ p,gU dz SAC ds 2 
=b' W 
where 
0 _U 
£ 12 12 12 cy 9d 


p ie) 
al 2 OO 4 ies a Foe? Ae ROD A 02 GOD} 


13> 19 19d 
+ (its v2 swt) |, (104) 


and b' represents the half width of W at the free 
surface. If only the largest terms in W are kept 
in the definite integral in Eqs. (102) and (103), 
the following approximate equations can be obtained: 


SN 
b 
Pg 2 
+ — Ch (Gs) = (910 dw u' , (105) 
2 £ 
-b 
Ww 
OU 2 2 2 
£ at.) (22 & 
————— a ae ae = 
Le 2 5 \@ 4 az ox 
Sx 
b 
Crake 2 2 
a azinG + 0U) 6lit) mm! 
2 is) £ (106) 
=b W) 


Since the following relationship is approximately 
satisfied in a wake far from the hull: 


eo Sy = GG 8), (107) 


Eqs. (82), (83), (87), and (88) can be obtained from 
(74), (77), (78), and (81) by substituting this 
relation into (105) and (106). 


Prediction Of the Velocity Field in 
Way of the Ship Propeller 


Igor A. Titov, Alexander F. Poostoshniy, and 


Oleg P< Orilov, 


Krylov Ship Research Institute 


Leningrad, U.S.S.R. 


ABSTRACT 


The paper covers the problems involved in determin- 
ing the velocity field in way of the ship propeller. 
The analysis is given for both the structure of the 
stern viscous flow and its change due to the ship 
propeller operation. 

The method is offered for scaling the nominal 
field of axial velocities based on the use of both 
the semi-empirical theory of the boundary layer and 
theory of free turbulence, and the engineering method 
of estimating the action of the working propeller 
upon the velocity field. 

As an illustration, the data of studying the 
influences of the scale effect and the working ship 
propeller upon the velocity distribution and total 
wake flow are presented in reference to a moderate 
displacement tanker. 


1. INTRODUCTION 


The need for a reliable definition of nonstationary 
loads acting on the propeller blades and shafting, 
and also of the intensity of hull vibration and 
cavitation phenomenon, has placed the wake flow 
problem among the most important problems of ship 
hydromechanics in the last few years. Though this 
problem first originated mainly in connection with 
the building of large full ships, it is of no less 
importance in the design of modern high speed con- 
tainer ships and some other classes of ships. In 
this sphere of hydromechanics shipbuilders are facing 
two main problems: a) prediction of the velocity 
field in way of the propeller for a ship of given 
lines as based on geosim model test results and 
b) finding solutions which provide a more favorable 
distribution of the wake flow. The rationalized 
formation of the afterbody wake is also one of the 
possible reserves of ship propulsion which do not 
yet appear to be fully realized. 

At present, the problem of the afterbody wake 


265 


and particularly its prediction attracts the atten- 
tion of a growing number of specialists in research 
centers of the advanced shipbuilding nations in- 
cluding the USSR. In view of the extreme complexity 
of the afterbody flow pattern in the presence of the 
propeller-induced disturbances, the problem of the 
wake flow is still far from being solved. The laws 
regulating the development of wake flow and also 

the dependence of the velocity distributions at the 
propeller disk upon the shape of the afterbody lines 
are not quite clear. The test methods of defining 
the ship model wakes and model-to-ship correlation 
methods are as yet imperfect. Therefore the ac- 
curacy of the flow nonuniformity data obtained in 
way of the propeller and used as a basis for calcula- 
tion of the abovementioned hydrodynamic character- 
istics does not satisfy the requirements of modern 
practice. Hence, a detailed investigation of this 
phenomenon is needed. 

In our opinion the most important tasks are as 
follows: First, comprehensive physical studies of 
the afterbody velocity field. These would allow for 
better understanding and proper evaluation of the 
effects of different factors on the formation of 
wake flow in that region and help create a flow 
model exhibiting the main features of the phenomenon 
and capable of being investigated by analytical 
methods. At this stage the theoretical studies are 
essential primarily for a better understanding and 
more proper analysis of the test results, as well 
as for improving the general knowledge of both the 
flow laws and the scheme of breaking the wake into 
components. Second, the results of the experiment 
and the qualitative theoretical conclusions should 
be the basis for the development: 

- methods for simulation of the nominal wake or 
methods for theoretical estimation of the scale 
effect at early stages of designing; 

- methods for experimental definition of the 
effective wake and approximate methods for the evalu- 
ation of propeller effect using the nominal velocity 
field data. Since the velocity field in way of the 


266 


propeller is normally defined in the idealized con- 
ditions of the towing tank, it is absolutely neces- 
sary to evaluate and take account of the effect of 
operating conditions, i.e., the effect that increas- 
ing the roughness of the hull surface as well as the 
ship motions and drift have on the extent of flow 
nonuniformity at the afterbody. There are also 

some additional tasks, such as improvement of the 
method used for definition of the ducted propeller 
velocity field, estimation of a possible change in 
the wake flow over the propeller axial length, and 
thinking over the practicability of the methods of 
disturbing action upon the flow pattern with preset 
requirements. The methods of experimental defini- 
tion of the flow velocities in the vicinity of the 
hull model are no less important. It is impossible 
to cover the results of all the above studies ina 
short report like this, so we shall restrict our- 
selves to the following traditional problems: the 
scale effect of the velocity field and the propeller 
effect on the flow formation at the stern. 


2. SCALE EFFECT OF THE NOMINAL VELOCITY FIELD 


The decrease of the mean wake in a model--ship 
correlation with sufficient accuracy can be at- 
tributed to variation in total frictional losses. 
The problem of simulating the local wake is far 
more complicated. The flow in way of the propeller 
is a combination of two three-dimensional flows: 
the boundary layer in the upper part of the after- 
body with intensive secondary flows characteristic 
of this region and the initial part of the wake de- 


model - ship correlation, the approximate methods 
of the semiempirical theory of turbulent boundary 
layer and of the free turbulence theory are of 
great importance; also important are comprehensive 
physical investigations of the afterbody flow which 
are necessary for the refinement of the flow model 
and formulation of the simplifying assumptions. 
Such investigations should cover the whole of the 
viscous wake region (Figure 1 and 2) and not be 
limited to the disk propeller area as is usually 
done in practice. 

The phenomenon being too complicated, a general 
approach to simulating the flow seems to be unat- 
tainable at present. Therefore, it is expedient 
to discuss some particular models of the flow. Some 
of the flows may be considered as the most common 
types which can easily be investigated. These are: 

a) the velocity field of a single-screw ship of 
moderate fullness with V-shaped or U-shaped frames 
where the contribution of bilge vortices is not 
Significant; 

b) the velocity field of high-speed, twin-screw 
container ships; 

a more complex pattern and more complex scaling laws 
are characteristic for 

c) the velocity field of full ships (6 > 0.8) 
with U-shaped frames where the intensive bilge 
vortices are formed; 

d) the velocity field of the very full ships with 
the boundary layer separation at the afterbody. 


Model "a" 


The calculation data obtained for a three-dimensional 
boundary layer lead to the conclusion that with moder- 
ate transverse flows the variation in characteristics 


veloping behind the hull which may contain discrete 
vortices resulting from the boundary layer separa- 
tion in way of the bilge where the flow lines from 


under the bottom are extending to hull side sur- 

face (Figure 1 and 2). As shown by experiments, 

the contribution of each of these factors depends 
on afterbody fullness, stern frame form, buttock 

angles, and some other parameters. 

The distributions of the relative axial veloci- 
ties Uy/Ys(y/é;Rn) are different for the boundary 
layer, the wake, and the vortex effect region, and 
largely depend on the afterbody lines and the 
history of the flow. The solution of the scale 
effect problem by a purely experimental way is not 
practicable, so when the general laws of variation 
in the flow characteristics are established for 


of the main flow accounting to Rn does not differ 
markedly from those obtained for a two-dimensional 
boundary layer. Hence, for practical estimation of 
the axial velocity field in the upper part of the 
afterbody (Figure 1) we can use, without introduc- 
ing large errors, the boundary layer correlation 
schemes developed to fit the two-dimensional flow 
on the basis of the logarithmic law and the velocity 
defect law. For simulating the wake flow use can be 
made, with some assumptions, of the known Prandtl 
asymptotic solution for a two-dimensional flow 
which was obtained on the assumption that the flow 
is barotropic and that the velocity defect, AU, is 


FIGURE 1. Nominal velocity 
field in the propeller plane 
for a model of tanker with 
moderate block coefficient, 
cy = 0.73 (model 1). 


art 
/ 
VS SWS WV LS &/ 
S SSS SNS SOS SS i 


SS BSS SRS SS! 


BSNS SSS SSS SSS 


& lye 
Fale ONS SN WSN SONS SSS 
a ESSA SAS SSNS v4 
7 =S 
NN EN / >See 
St AANA 
APE ww My WAVE 


267 


SS 
INN 
. ae ~~> SS 
ESR ESRISS 
ae = 
we Q S ‘ 
I? ~ ‘ \ 


peor | 
) 
ys 


FIGURE 2. Nominal velocity 
field in the propeller plane 
of a "Krym"-type tanker model, 
cy = 0.83 (model 2). 


insignificant as compared to the velocity at the Such a scheme of simulation makes it possible 
boundary of the wake flow: to take into account the variation in both the 
wake thickness and the form of the nondimensional 


Wr = Wi 4 
— 6 ae L profile U,/Us. 
a Us een (C,/A%) S ASP2) ) Model - ship correlation data for a tanker of 
Bela wm * (in (2) 
them eae D = 1 
Z, = 2,/R = 0.875 Ae 

where AX is the relative distance between the body 0.85 Pee oi 
trailing edge and the wake flow section under study. 


Naturally, these relations do not provide a reliable 
qualitative definition of the flow characteristics x 


at the initial part of the three-dimensional wake IS) Osh 

which develops with the longitudinal pressure gra- 

dient. However, the above relations are considered 

to be quite suitable for simulating the wake field 014) ae 
velocity because the deviations due to the effect ) 1 


of some factors ignored here can be mutually com- 

pensating. The practical method of correlation is 

based on the assumption of a negligible effect of 

the potential component and of a free streamline 0.8 
flow around the hull. The effect that the varia- 

tion of the transverse velocity component has upon 

the axial flow with the increase in Rn is also con- x 
sidered insignificant. The initial experimental 
data for the model are defined in the Cartesian 
system as velocity or wake distributions against 
the transverse coordinate, y = y/L, with the dif- 0.4 i ay 

ferent constant values of %. The coefficients, Kj, 0 1 2 3 

and Kj, in Eqs. (1) and (2) are assumed to be (Y/L) X 10? 


constant in the geosim horizontal sections of the = 
Z. = 0.438 
wake. 
cn Se | 


= if = 
W Wa Cr (BNg) /Cag (RT, A aioe 37/0) const (3) 


© © O— According to Equations (1)—(4) 


@ @ @— Jaking Account of the Boundary 
Layer Scale Effect 


= ID 
b. = bo/l, = Cc Rn C 
s s/ 5 by ao ad Fo {En,,) (4) 
where “i me att | 
1 a2 3 
Co = frictional resistance coefficient (Y/L) X 10 
in two-dimensional flow; 

b = width of the wake; FIGURE 3. Velocity distribution in wake extrapolated 
WwW = frictional wake ship, model. to full scale. 


268 


medium displacement are shown in Figure 3 as an 
illustration. Isotaches (lines U = const) plotted 
in Figure 1 show that the upper part of the propel- 
ler disk is in the hull boundary layer region and 
here the flow contraction will take place almost 
normal to the constant velocity lines rather than 
to the longitudinal center plane. In this connec- 
tion an attempt was made to evaluate the variation 
of the flow velocities in the upper part of the 
propeller disk using the approximate method re- 
ported at the 13th ITTC, which provides quite a 
good agreement with the full-scale test data, and 
those obtained by calculation of the three- 
dimensional boundary layer [Boltenko et al. (1972) ]. 
The results of the refined model--ship correlation 
for this model within the propeller disk practically 
coincide. Velocity deviations of 3-4% V are ob- 
served only in the vicinity of the viscous wake 
boundary in its upper sections (outside the propel- 
ler disk), Figure 3. However, in some cases (e.g., 
with pronounced V-shaped afterbody frames) the hull 
boundary layer can play a more significant role in 
the formation of the wake flow, and in that case 

its effect should additionally be taken into con- 
sideration. Similar practical methods based on more 
general assumptions with respect to regularities 

in the variations of the axial velocities were given 
by the towing tanks of Europe and Japan [Sasajima 
and Tanaka (1966), Hoekstra (1977), Dyne (1974) J. 
For comparison Figure 4 shows the model--ship cor- 
relation results obtained by the Japanese method* 
for some specific profiles of the wake of the model 
under consideration. As is seen, this method leads 
to a greater contraction of the wake in model--ship 


correlation and does not take into account the varia- 
tions of the velocity defect in the centerline plane. 


However, apart from some limited regions in the 
vicinity of 6 = 0° and 180° the circumferential dis- 
tribution of axial velocities U,, (78) calculated by 
both methods differs slightly (Figure 5). For the 
above reasons substantial discrepancies in the 
vicinity of 6 = 0° and 6 = 180° can give rise to an 
appreciable change in the harmonic spectrum of the 
field especially in the amplitudes of the even har- 
monics. 

At present it is difficult to find an acceptable 
practical method of simulating the transverse ve- 
locities, though the semiempirical theory indicates 
the possibility of a noticeable scale effect of the 
secondary flow velocities in the three-dimensional 
boundary layer of the ship. 


Model "b" 


The flow nonuniformity in way of the propeller of 
the twin-screw ship is mainly due to the hull bound- 
ary layer and the additional loss of velocity in the 
wake behind appendages 


WV eaMO PF Se SAU Stls ES Oe Os 


*The method of Japanese researches was used as described by 
Dyen (1974). 


il 


According to 
Equations (1)—(4) 


---- According to 
Sasajima and 
Tanaka, 1966 


(Y/L) X 102 (Y/L) X 102 


FIGURE 4. Full scale wake predicted by different 
methods. 


where 
mana potential component of the wake; 
Fo; viscous wake due to the effect of the hull 


boundary layer; 

Aw. = additional losses of velocity in the wake 
behind the appendages; 

U = horizontal local velocity 


U = horizontal local velocity in the "bare" 
hull boundary layer. 


The investigation of the wake scale effect for a 
twin-screw ship, with a probable interaction between 
the wake components, involves a number of complex 
hydrodynamic problems. They include that of the 
hull three-dimensional boundary layer, also the wake 
behind the propeller shaft fairing placed at an 
angle of attack to the flow inside the boundary 
layer, in which case not only is the mean velocity 
Vy (y) changed but also the extent and the scale of 
the "outside" flow turbulence. Then there is also 
the wake--boundary layer interaction problem and, 
finally, oblique flow around the circular cylinder 
(shaft) placed in the turbulent boundary layer. 
Many of the above problems are concerned with some 
insufficiently known aspects of hydrodynamics of 
viscous fluid and, therefore, cannot be completely 
solved for the present. As with the previous case, 
approximation schemes can be used for practical 
estimations. By way of illustration let us con- 
sider the model--ship correlation data for a twin- 
screw ship equipped with propeller-shaft fairings. 


1D 
1 — Model R,, = 1.3 x 107 
2 — ShipR, =1.5X 10° 
(According to Equations (1)—(4)) 
3 — ShipR, =1.5x 109 
(According to Sasajima and Tanaka, 1966) 
1.0 PP 
See SSX 2 
a y 
0.8 va \ 
\ . 
/ \ 
x 
i) 0.6 


FIGURE 5. Full scale circumferential velocity 
distribution predicted by different methods. 


The experiments show that, in the vicinity of the 
heavily-loaded blade sections which are at a dis- 
tance from the hub, the interaction between the 
boundary layer and the wake behind the fairing can 
be considered insignificant; the effect of support- 
ing vortices at the fairing junction is also negli- 
gible or not found at all because provision is 
usually made for a smooth transition of the fairing 
to the shaft body. This enables simulation of each 
component of the viscous wake Wro(Rn) and AW, (Rn) 
to be investigated separately with the total scale 
effect to be determined by the method of superposi- 
tions. Here it is expedient to make measurements 
in the Cartesian system of coordinates as well. For 
the model--ship correlation of the wake behind the 
hull the method described by Boltenko et al. (1972) 
is used. When simulating a component of the wake 
AWp caused by the flow around appendages, use can 
be made of the relationships of the free turbulence 
theory (1) and (2). According to data of the flow 
visualization, it can be considered with an accuracy 
sufficient for practical purposes that the stream- 
lines on the fairing are arranged equidistant to 
the hull surface, and that in evaluating the scale 
effect the strip theory can be used. Then 


U —_& 
AW. = A v a 
RS Wom _HS Cre) Com icone $a) a const 
Unm 
y= Zi const (6) 
b. = oe Coe (7) 


269 


where 

C.. = coefficient of the fairing resistance at 
section at a given distance ¥% from hull 
surface (Figure 6); 

Uh = velocity in the hull boundary layer at a 
given distance % from its surface; 

b = width of wake behind the fairing at the 
propeller. 


From the model-ship correlation data shown in Fig- 
ure 6 it is seen that the flow nonuniformity varies 
almost equally due to the scale effect of the hull 
boundary layer and the wake behind the shaft fairing. 
The mean circumferential axial wake is reduced ap- 
proximately by one half. 


Model "c" 


The discrete vortices, which develop due to separa- 
tion from the bilge, with their axes oriented in 
the direction of the main flow may have, in some 
cases, especially where the flow is around the U- 
shaped stern frames, a noticeable effect on the 
afterbody flow pattern. Generally there are two 
vortices arranged symmetrically in relation to the 
center plane; however, sometimes more complex vor- 
tical systems can be observed in the flow around 
full ships. The development of the bilge vortices 
leads not only to redistribution of the tangential 
velocities at the propeller, but to the additional 
nonuniformity of the axial wake as well due to 

a) redistribution of the velocities of the main 
flow in the hull boundary layer and in the wake 
behind the hull under the action of the vortex- 
induced transverse velocities and 

b) variation of the axial velocities in the 
vortex turbulent cores, the transverse dimensions 


Diagram 


Z/L Gh) 


4X/0 


~) 

NS 

LLL 

Model (R,, = 1.0 X 10’) Ship (R,, = 6.0 X 10°) 
9° 6° 

260 0 100 


FIGURE 6. Scale effect estimates for nominal velocity 
distribution at propeller of a twin-screw ship with 
shafting fairings. 


270 


of which can be rather large as shown in Figure 2. 

Thus the whole flow field containing bilge vor- 
tices can be divided into three parts: 

1) the region of turbulent core, 

2) the region of vortex effect on the hull bound- 
ary layer and 

3) the region of nondisturbed flow in the bound- 
ary layer or in the wake (Figure 7). 

The laws for changing the relative velocities in 
each of these regions are different in model-ship 
correlation. 

Evaluating the scale effect of disturbances in 
the boundary layer is rather a complicated task 
partly due to the difficulty of distinguishing these 
disturbances in the nonuniform three-dimensional 
boundary layer of the hull. Therefore, at the ini- 
tial stages of investigation the principal attention 
was paid to the specific features of such kind of 
flow in simplified conditions, i.e., under the as- 
sumption that artificial vortex systems were pro- 
duced by means of profiles of small aspect ratio 
at the boundary layer of a flat surface [Poostoshniy 
(1975) ]. For such simpler flows one can use the 
approximate methods of evaluating the scale effect 
of axial velocities in the region where influence 
of the vortex is observed. These methods will be 
based on a combination of experiment and theory or 
approximate semiempirical schemes, which is most 
important for having a general idea of the phe- 
nomenon. : 

Extra losses of axial velocities in the vortex 
cores are rather high for some ship models (reach- 


U = constant 


Region of Vortex Influence 


. . ; SS 
on the Boundary Layer ~— Y Undisturbed Flow Region 


Vortex Core -/ 


Uo 


7 Velocity Distribution 
in the Core 


Circulation Distribution in the Bilge Vortex 


“Core of Tanker Model (4 = 60000 t), T= 1.1 m2/s 


2.0 a 


Distribution of Circulation in a Vortex 
Core of Free Flow, To= 0.07 m*/s 


= 
1.0 2.0 


Ig(r/r,) +1 


FIGURE 7. Velocity field in the boundary layer with 
longitudinal discrete vortices. 


ing 20-30% of the mean wake value); these losses 
are also to be studied in detail. 

As shown by the experiments (Figure 6) the 
circulation distribution law for the cores of bilge 
vortices is similar to that for the vortex cores in 
the free flow. So, in order to evaluate the scale 
effect of a relative defect of the axial velocity 
in the core, i.e., the core allowance, use can be 
made of the theoretical relationships derived for 
linear turbulent vortices. 

Calculated results which are based upon rather 
a small amount of data on the variation in eddy 
viscosity coefficients with Rn obtained during model 
tank tests and fall-scale hydrodynamic experiments 
lead to the conclusion that a model-ship correlation 
involves relative decrease of the core size. How- 
ever, far from decreasing, the wake allowance, unlike 
that for the boundary layer, may even be markedly 
growing. Some additional variation in the distribu- 
tion of axial velocities in the core caused by an 
increase in Rn may also be due to an increase in the 
longitudinal pressure gradient at the stern owing 
to the reverse effect of the hull boundary layer on 
the external potential flow both on model and ship. 

It is impossible at present to develop a flow 
model of this complexity, define the component ve- 
locities changing under different model-ship corre- 
lation laws and, finally, determine these laws; in 
other words it is impossible to develop a well- 
founded method for simulation of a three-dimensional 
wake flow with discrete vortices. The results of 
the above-mentioned preliminary studies are of 
qualitative character and need experimental verifi- 
cation. A series of comparative model and full-scale 
tests carried out mainly by Japanese researchers 
[Namimatsu and Muroaka (1973), Taniguchi and Fujita 
(1969) ] confirm the existence of bilge vortices in 
full-scale conditions as well, though the data re- 
ported in the above papers are inadequate to judge 
the quantitative aspect of the phenomenon. We can 
only observe that the disturbances induced by the 
vortices in the flow around a ship are less notice- 
able, i.e., the flow is cleaned up. Therefore the 
attempt to use a more generalized model (model "a") 
seems to be justified also in this case, i.e., in 
the presence of developed bilge vortices, or at 
least an attempt to establish limits for the appli- 
cation of this’ flow model should be made. Compara- 
tive data obtained from model and full-scale tests 
are a decisive factor here. 

Unfortunately no data of nominal wake distribu- 
tion at the propeller are available. For an indirect 
evaluation of the scale effect of nominal wake we 
shall make use of the test data obtained in Japan 
for a 36000 t (displacement) tanker and its 1/37- 
and 1/20-scale models [Taniguchi and Fujita (1969) ]. 
The measurements were taken in the boundary layer 
near the sternpost at a distance of 1.1D from the 
propeller disk. In laboratory conditions the ve- 
locity field was measured both during the towing 
tests and self-propelled tests. The tests performed 
with the model (\ = 1:20) allow the propeller effect 
at the measurement plane to be considered as negli- 
gible (~0.05 V) and practically constant within the 
region equivalent to the propeller disk area. The 
comparison between the velocity distribution in the 
wake transverse section for % = 8, (where %, = 
propeller axis level) and the circumferential dis- 
tribution of the axial velocities (Figures 8 and 9) 
for this tanker and those for a "Krym"-type tanker 
shows that the simplified method of model-ship cor- 


271 


3. PROPELLER EFFECT UPON THE WAKE DISTRIBUTION 


Consideration of the wake scale effect when using 
the nominal velocity field as initial data will not 
always improve the agreement between the calcula- 
tions and full-scale measurements of nonstationary 
loads acting on the shafting and, particularly, of 
the constant bending moment component defined by 
the analysis of the first harmonic. Systematic 
model basin test results indicate that signigicant 
variations of the velocity distribution at the stern 
may be due to the propeller performance. Several 
factors are to be taken into account when analysing 
(Y/L) X 102 the causes of this phenomenon. The most important 
LWL among these are the propeller-induced acceleration 
= of flow and, hence, the decrease of the layer thick- 
(+) ness upstream, and the effect of propeller-induced 
radial velocity in the immediate vicinity of the 
propeller. 

Thus it becomes necessary to investigate the 
ship-hull boundary layer and the wake taking into 
account the transverse pressure gradient. Semi- 
empirical theories do not permit this problem to 
be solved and are adequate only for the most ap- 
proximate estimations of the flow history. There- 
fore, just as in studying some features of the 
nominal wake flow mentioned above, preliminary 
theoretical investigations of the velocity field 
under simplified conditions are of great importance 
here. Although these results are not directly 
applicable to the ship, they may be useful for a 
better understanding of the main relationships of 
0.8 1.6 2.4. the phenomena under study and for the devleopment 
of practical methods to obtain the effective wake. 
In this connection one cannot but mention the 
important contribution of American scientists to 
the investigation of the axisymmetrical problem, 
particularly, the latest works by Huang and Cox 
(CL) Dc 

To obtain approximate estimates of the effective 


“KRYM"-Type Tanker 
© 0 0 —Model, Experiment 


eer Correlation 


—— 
0.8 1.6 2.4 


Taniguchi and Fujita 
Experiment, 1969 


oO 0 O — Model, Experiment 


SS Ship, Experiment 


(Y/L) X 10? 


FIGURE 8. Comparison of velocity distributions for 
model and ship wake. 


relation reveals the characteristic features of 
variation in the velocity field and its harmonic 


spectrum. However, these conclusions cannot be “KRYM"-Type Tanker Tanker, Taniguchi and Fujita 
considered reliable enough; they need further veri- Experiment, 1969 
fication. 


Model "d" 


Several years ago, simulation of the velocity field 
in the case of afterbody boundary layer separation 
attracted the special attention of researchers in 
connection with the development of very large tankers 
with high block coefficients and a tendency to de- 
crease the length-to-breadth ratio. Although this 
problem has lost its vitality by now, studies in 

this field are being continued. The attempts in 
Japan and in the Soviet Union to theoretically and 
experimentally evaluate the scale effect of separa- 
tion of three-dimensional and even two-dimensional 
boundary layers do not yet allow any definite con- 
clusions to be made, even regarding the qualitative 
aspect of the phenomenon, or the development of the 
most approximate scheme of variation with Rn number, 
not only in the velocity distribution, but also in 
the mean value of the wake. Thus the problem of N N 
simulating the characteristics of flow at the stern 

with the boundary layer separation remains one of FIGURE 9. Circumferential velocity distribution and 
the unsolved problems in ship hydrodynamics. harmonic spectrum for model and ship. 


a_ X 102 


272 


wake, both in our practice and in the practice of 
other model tanks, use has been made in recent 

years of engineering procedures based on the results 
of nominal velocity field measurements and propeller 
theory relationships [Hoekstra (1977), Raestad 
(1972), Nagamatsu and Sasajima (1975) ]. 

If we assume that the propeller effects are 
mainly due to the factors mentioned above, the 
propeller can be thought of as having a large diam- 
eter when evaluating the mean wake field. 

This assumption will result in a decrease of the 
wake coefficient. The decrease of the frictional- 
resisted wake due to the propeller effect can be 
taken as inversely proportional to the square root 
of the diameter. Then, 


- 4 
Ware = Wey/ (lL + wi/2V,) (8) 


where 


Con = (pv2/2)F 


To define the potential component Wpe alte) Sk} 
reasonable to apply the known propeller theory 
relationship 


= am (fea/2)) €s V 
Woe Won ( o/ ) (w/ ) 
to = 
= iio yal as G = Al 
WON 2 [ Th ] (9) 
where j 
W No experimentally defined potential component 
P of the nominal wake field, 
tj = thrust deduction at zero velocity of 


model. 

Allowing for the smallness of the 2nd term in (9), 
the thrust deduction fraction undergoing only minor 
changes can be assumed for single-screw ships to be 
tg = 0.07-0.10 (the last figure relating to fuller 
hull shapes) . 

The final expression for the mean effective wake 
field (taking into account the scale effect) has 
the form, 


We = eon tF EGQ/QGAL Cr - 1)] 


nh 
mh Won Cao (RM) 
fo) NS ES eee 
V1/2(vV7l + C. +1 CEA (Ene) 
VW/2(vE+ Cy + 1) Spon (aif 
where 
Cro = frictional resistance coefficient in two- 


dimensional flow. 


Relationship (10) displays good agreement with 
the model test data (see Table 1) and W, values 
close to those obtained from the full-scale test 
analysis. 

As can be seen from the Table, all known approxi- 
mate methods yield practically the same results. 

By making some additional assumptions, similar 
methods can also be applied for an approximate 
estimation of the circumferential distribution of 


the effective wake, and in the main they correctly 
reflect the variation trends of the flow at the 
stern while the propeller is in operation. However, 
they do not permit: taking into account and evalu- 
ating some qualitative changes in the hull boundary 
layer, which may take place due to propeller opera- 
tion, such as variation in circulation of bilge 
vortices and their positions in relation to the 
ship hull; the possibility of preventing or reduc-— 
ing the separation about the stern zone with the 
propeller in operation; and, on the other hand, 

the possibility of the boundary layer separation in 
the vicinity of the stern above the propeller. 
Therefore, when performing a quantitative analysis 
of the effect the propeller has on the wake and the 
harmonic spectrum of the velocity field, these 
methods, in spite of their relative simplicity and 
convenience, should be applied rather carefully, as 
for most tentative estimates. 

At the present stage of the wake problem in- 
vestigation the development of experimental methods 
is of decisive importance. 

Both for the improvement of the general knowledge 
of propeller effects on the flow pattern at the 
stern and for the solution of problems associated 
with ship form design, the accumulation of data 
on the effective velocity fields for ships of 
various types and the improvement of model test 
methods is of great importance, especially those 
taking account propeller induced velocities or 
eliminating the same from measurement data. 

A practical method for estimating the effective 
velocity field, Uy, by way of flow velocity mea- 
surements at some distance ahead of the propeller 
in "open water" and behind the hull, was given in 
Titov and Otlesnov (1975). For measured data 
analysis the quasi-steady theory was accepted. 

When the hydrodynamic flow angle, 8), of a 
propeller blade section for the propeller operating 
in "open water" is equal to that behind the hull, 


— ' = te " a 
tgBy WE oF Wi) Jot (Ura Wi) / (wr Use) (11) 
where 
W' and W" = axial induced velocities ahead 
= of the propeller in "open water" 
and behind the hull 
U = circumferential component of the effective 


Be velocity field 


The axial component of the effective velocity field 
ahead of the propeller is determined from the 
relation 


Comparison of the Mean Effective Wake 
Calculated by Approximate Methods With 
That Obtained from Self-Propelled Tests 
(Model No. 1) 


TABLE 1. 


Titov - Poostoshniy method 0.345 
Nagamatsu - Sasajima method (1975) 0. 340 
Roestad method (1972) 0.355 
Self-propulsion test data 0.350 
Nominal wake 0.390 


Ne —— 


Wa + wi) = Wl We. ar wi) (12) 


following from the equality of forces on the pro- 
peller blade section. 

However, another approach to the problem of 
experimental determination of the effective wake 
is also possible based on the data analysis of 
measured flow velocities and total head pressure 
immediately ahead of the propeller and behind it. 
In this case, measurements are taken only with the 
propeller in operation behind the hull. 

As is known, the circumferential induced velocity 
at propeller section, Wg, in "open water" is pro- 
portional to the jump in the total head at the pro- 
peller disk 


PUTW, (78) = Ho(t0) —- Hy (Té) (13) 


It can be shown that this relationship is also 
valid for the propeller behind the hull, if the 
variation of the circumferential induced velocity 
of the hull wake, Ug, is negligible within the axial 
length of the propeller or between the sections 
where measurements are taken. In this case total 
head pressures at sections 1 and 2 (see Figure 10), 
ahead of the propeller and behind it are, respec— 
tively, equal to 


p 2 2) 2 
= +> + US, + U 
in (ie) al 2 Oe 61 11! 
p -2 2 
8) = Po +> + (W + U 
Ha Mus) 2% a LOL) (Woy 82) 
+ (W_, + U_,)7] (ale) 
T2 T2 
where 
Wh. = UES + Wo = axial flow velocity at 
i i v section 1 
us = UA + We = axial flow velocity at 
2 2 2 section 2 
We G Wo and Whe = propeller induced velocity 
z components at respective 
sections 


Theoretical investigation results of propeller in- 
duced velocities and test data make it possible to 
linearly approximate component variations of the 
induced velocity, W3(x), within the limits of the 
propeller axial length. It is believed that the 
axial component variation of the wake in this re- 
gion is small and also obeys the linear law. 

With the above assumptions, in order to determine 
the design effective velocity, U at section X9 
where the condition 


xOU 


a 


WaCg) = 


(15) 


is observed, we obtain the following set of equa- 
tions: 


U Xo) + wo 7/2 


x 
ee EE SS 
Eoen WT — wy /2 + Us ee) 
W 
a. W,/2 (17) 


2 tgBy 


273 


0 90 180 


FIGURE 10. Circumferential distribution of velocity 
components in way of propeller (r = 0.590). 


w 
a 
wi (.) Su (% ) +See 
Ok 0? sa ) 2 
U =. Gj 
Teen ete (18) 
= SSS OK 
x} AX 1 
where 
8; = hydrodynamic flow angle of a propeller 
blade section 
AX = distance between sections 1 and 2 
AXg, = Xo =) Xa = distance between section 1 


and the point of calculation 


In propeller theory it is generally taken that the 
above condition is met at the propeller disk plane 
corresponding to the midspan section of the blade, 
and, in the case of blade rake, corresponding to 
the midsection of the blade at a relative radius, 
Te O.76 

However the calculation results of variations in 
the anomalous induced velocity, W,(X), of the pro- 
peller with the finite axial length indicate that 
in fact the point must be found upstream of the pro- 
peller disc plane. 

This conclusion is confirmed by the experimental 
investigation results of the propeller velocity field 
in open water. Taking account of these data it is 
more reasonable to assume the point of calculation, 
corresponding to condition (15), to be on the lead- 
ing edge of the blade. 


274 


Application of this procedure can be illustrated 
on a medium size tanker model (Model 1). 

Experimental studies of the velocity field for 
the operating propeller were performed during free- 
running model tests with the operational relative 
speed, Fn = V/VgL) 1 = 0.22. Wake characteristics 
ahead of and behind the propeller were measured at 
equal distances from the propeller centre with a 
6-point probe [devised at our model tank, Otlesnov 
(1969) ], which enables simultaneous measurements of 
total head pressure, (H), static pressure, (P), and 
flow angles in the horizontal and vertical planes 
in the immediate vicinity of the propeller. When 
processing the measured data and analysing the nom- 
inal wake, use was made of calibration relationships 
which took into account the interference of flow 
angles in the vertical and horizontal planes with 
the readings of the probe. Figures 10 and 11 il- 
lustrate the initial data and the calculated induced 
velocities for the starboard-side of the propeller 
disk (right-hand rotation) in the region where 
sections experience maximum loading. 

Comparison (Figure 12) of the nominal velocity 
field with the effective velocity field calculated 
from Eqs. (11)-(12) and (13)-(18) shows the pro- 
nounced effect the propeller has on the wake at the 
lower part of the propeller disk and the minor ef- 
fect at the upper part of the same. This may be 
accounted for by a better possibility for momentum 
exchange between the external flow and the viscous 
wake under the action of radial induced velocities 
in a relatively thin wake at the lower part of the 
propeller disk, and a worse possibility at the upper 
part where the thickness of the viscous wake is much 
greater (see isotachs in Figure 1). 


0 90 6° 180 


FIGURE 11. Circumferential distribution of velocity 
components in way of propeller (r = 0.756). 


=) 

0 90 6 180 
6 
1 — Nominal Field 
2 — Effective Field (Titov and Otlesnov, 1975) 
3 — Effective Field (Proposed Method) 
4 — Effective Field (Hoekstra, 1977) 
x 
ID 


FIGURE 12. Influence of propeller operation on 
velocity distributions. 


The above two methods for defining effective 
field axial velocities yield results which, as a 
whole, show satisfactory agreement. However there 
are some systematic discrepancies in the regions 
of @ » O° and 6 * 80-160°, and additional analysis 
is required to explain these. 

Besides, the velocity distribution data obtained 
on the basis of measurements ahead of and behind 
the propeller in operation make it possible to find 
the thrust distribution (load coefficient of pro- 
peller, Cm,,) over the propeller disk area 


We ar Wy 2 


Cr G8) GG ) are (19) 
xe 


Figure 13 and the equivalent system of singularities 


Q(68) 5 Wa ZU (20) 


In its turn, the knowledge of this system of singu- 
larities allows one to calculate the induced ve- 
locities over the total wake region ahead of the 
propeller, and perform a more detailed analysis of 
the effect the nonuniformity of load distribution 
over the disk has on thrust deduction. 

The following conclusions can be drawn from the 
comparison of Fourier transform coefficients for 
the circumferential distribution of axial velocities 
of the nominal field obtained for the model and ship 


90° sB 


180° 
Down 


FIGURE 13. Load distribution over the propeller disk 
(based on effective velocity field measurements) . 


(model-ship correlation), as well as of the velocity 
field in model tests taking account of propeller 
effects. The amplitudes of harmonics determining 
the nonstationary hydrodynamic forces and moments 
(Figures 14 and 15) may vary several times under the 
influence of the above factors. 

It should be mentioned that no definite regular- 
ity could be observed here. With some relative 
radii the amplitudes increase, with others they 
decrease. 

As the variation in harmonic spectrum of the 
velocity field is of rather a complicated nature 
let us illustrate the effect the variation of axial 
velocities due to scale effect and propeller opera- 
tion has on the constant component of the hydro- 
dynamic bending moment in the vertical plane which 
is mainly defined by the first decomposition har- 
monic [Voitkunskiy (1973) ]; 


e 
M. = = oS 
Yo 7 “30 ~ Pyo ° 2 (2) 
where 
i = 
ey = IC, /4S _ ft[a,+(1/FT) (J+2K,/C))a,, lat (22) 
TO 
1 
Po = -JCoS— f/t[b + (1/FT) (J+2 iT 
70 2S—F/T[b 1+ (1/FT) ( Ko/C,)b, lat (23) 
oa V,/nD 
tT) = relative radius of propeller hub 
KprKo = thrust and torque coefficients at 
design speed 
a,,b, = Fourier transform coefficients for 
the cosines and sines of the first 
harmonic of axial velocity on a given 
radius 
agg = Fourier transform coefficients for 


the cosine and sine of the first 


275 


harmonic of tangential velocity on a 
given radius 

2 = coefficients 

f = coefficient depending on radius 

e = distance between the design propeller 
shaft section and the propeller disk 


The distributions of transverse relative ve- 
locities Uo = Ug/V were taken as equal. 

Table 2 shows the design estimates of relative 
values of the constant component, Myg/Kg, as based 
on various initial data. 

As can be seen, the calculated results based on 
the nominal velocity field data may differ (even 
qualitatively) from those obtained with considera- 
tion for the scale effect or the effect of operat-— 
ing propeller. Although the local variations of 
the nominal field due to the scale effect or pro- 
peller operation are quantities of the same order 
(see Figures 5 and 12), the constant component 
values of the bending moment in the vertical plane 
determined from the effective field prove to be 
4-5 times as large. Physically this may be due to 
the fact that, in contrast to the scale effect, 
the propeller effect on the viscous flow in the 
upper parts of the propeller disk differs from that 
in the lower part. In the upper part of the disk 
(8 = 0 - 90°) the effective field distribution of 
velocities in way of the heavier loaded blade sec— 
tions differs only slightly from the nominal field 
distribution, while in its lower part (@ = 90-180°) 
the effective field velocities are much in excess 
of the nominal field velocities (by a factor of 
1.5-2). This increases the asymmetry of circum- 
ferential distribution of the effective field axial 


} 
[ Ship, Nominal Field (Correlation Based | 
on Equations (1)—(4)) | 

0.1 [ 


y- Model, Nominal Field 


Model, Effective Field 
(Hoekstra, 1977) 


Model, Effective Field 
(Proposed Method) 


Model, Effective Field (According 
to Titov and Otlesnov, 1975) 


FIGURE 14. Influence of scale effect and propeller 
operation on harmonic spectrum. 


276 


FIGURE 15. Influence of scale effect and propeller 
operation on harmonic spectrum. 


velocities and results in an increase of the con- 
stant component of the moment in the vertical plane. 
The Myo/Ko values calculated from the effective 
velocity field approximate those observed for full- 
scale ships of this type under operational condi- 
tions. This fact confirms the importance of taking 
into account propeller operation when simulating 
the velocity field at the propeller. The propeller 
effect upon the velocity field is dependent on the 
load, ship hull form and afterbody shape, initial 
nominal field, and the relationship between pro- 


10 a, 


10 a, 


10 a, 


Model, Nominal Field 
Ship, Nominal Field (Correlation Based 
on Equations (1)—(4)) 


a — 


Model, Effective Field (Titov and Otlesnov, 1975) 
Model, Effective Field (Proposed Method) 


0.6 F 


peller screw size and wake thickness, i-.e., on the 
propeller immersion into the viscous wake. 

The full-scale conditions of effective field 
formation are likely to differ from the model ones. 
Hence, the next step in studying the prediction of 
the flow velocity field in way of the propeller will 
be the development of procedures which enable simul- 
taneous consideration of both the scale-effect and 
the effect of propeller operation on the wake at 
the stern. 


TABLE 2. Variation in the Constant Component of Bending Moment Depending on the 
Velocity Distribution at the Propeller (Model 1) 
Model. Esti- Model. Model. 
mation of Experiment Experiment 
: Propeller Consideration Consideration 
Model. Model-ship Effect of Propeller of Propeller 
Experiment. Correlation According to Effect by Effect by 
Initial Nominal Using Equa- Hoekstra Using Equa- Using Equa- 
Data Field tions (1)-(4) (1977) tions) ((22)i=(@i2)) stafon's (dis) =1('8)) 
M_ /K 0.04 -0.07 -0.08 -0.35 =(0) 57355) 


REFERENCES 


Boltenko, V. P., O. P. Orlov, and A. F. Poostoshniy 
(1972). Determination of the boundary layer 
characteristics of ships from model test data, 
13th ITTC, Material of Interest, Berlin/Hamburg. 

Dyne, G. (1974). A study of the scale effect on 
wake, propeller cavitation, and vibratory pres-— 
sure at hull of two tanker models, Transactions 
the Society of Naval Architects and Marine En- 
gineers 82, 162-185. 

Hoekstra, M. (1977). An investigation into the 
effect of propeller hull interaction on the 
structures of the wake field, Symposium "Hydro- 
dynamics of Ship and Off-shore Propulsion Systems," 
Oslo, March. 

Huang, T. T., and B. D. Cox (1977). Interaction 
of afterbody boundary layer and propeller. Pro- 
ceedings of Symposium of Hydrodynamics of Ship 
and Offshore Propulsion Systems, Oslo, March 
20 =2)5y. 

Nagamatsu, T., and T. Sasajima (1975). Effect of 
propeller suction on wake. Symposium on Pro- 
pulsive Performance of High Block Ships, Tokyo, 
June. 


Namimatsu, M., and K. Muraoka (1973). Wake distribu- 


277 


tion of full form ship. 
7a Sp Massa 

Otlesnov, Yu. P. (1969). New means for measuring 
unsteady characteristics of ship wake ina 
towing tank. Proc. 12th ITTC, Rome. 

Poostoshniy, A. F. (1975). Features of flow around 
the hull and full hull resistance structure, in 
Problems of Ship Applied Hydromechanics, I. A. 
Titov, ed., Izd. Sudostroeynie, 54-92 (in Rus- 
sian). 

Raestad, A. E. (1972). Estimation of a marine pro- 
peller's induced effects on the hull wake field- 


IHI Engineering Review, 


scale effects on hull wake field. Report No. 
72-3-M, Det Norske Veritas. 
Taniguchi, K., and T. Fujita (1969). Comparison 


of velocity distribution in the boundary layer 
of ship and model. Proc. 12th ITTC, Rome. 

Titov, I. A., and Yu. P. Otlesnov (1975). Some 
aspects of propeller-hull interaction. Swedish- 
Soviet Propeller Symposium, Moscow. 

Sasajima, H., and I. Tanaka (1966). On the estima- 
tion of wake of ships. Proc. llth ITTC, SNAJ, 
Tokyo. 

Voitkunsky, Ya. I., R. Ya. Pershits, and I. A. 
Titov (1973). A manual on ship theory. Sudo- 
stroyenie, Leningrad. (In Russian). 


Recent Theoretical and Experimental 
Developments in the Prediction 

of Propeller Induced Vibration 
Forces on Nearby Boundaries 


Bruce D. Cox and Edwin P. Rood 
David W. Taylor Naval Ship Research and Development Center 


Bethesda, Maryland 
William S. 


Vorus 


University of Michigan 


Ann Arbor, Michigan 


John P. Breslin 


Stevens Institute of Technology 


Hoboken, New Jersey 


ABSTRACT 


This paper concerns recent advances in the theory 
and numerical solution of propeller induced pressure 
forces acting on ship hull surfaces. The analysis 
is formulated in terms of the diffracted potential 
flow about general three-dimensional hull boundaries 
in the presence of a free surface. The influence 

of the propeller is derived from lifting-surface 
theory, explicitly accounting for finite blade 
number, blade thickness and skew, and radial and 
chordwise loading (steady and unsteady, but sub- 
cavitating). Two methods have been developed to 
calculate the periodic forces. In the direct 
approach, time-dependent source singularities are 
distributed over the body surface with the strengths 
determined for a prescribed propeller onset flow. 
The force is then found by applying the extended 
Lagally theorem. In the second approach, based on 

a special application of Green's theorem, the force 
is obtained by finding the velocity potential at 

the propeller generated by the boundary executing 
simple oscillatory motions. 

A towing tank experiment is described in which 
blade frequency forces were measured on a body of 
revolution adjacent to a propeller operating in 
virtually uniform flow. The simplifications of 
body shape and propeller loading provided a physical 
model which could be treated in a reasonably exact 
fashion by the theory. The body consisted of two 


parts. A heavy afterbody, attached to the towing 
strut, acted as a seismic mass at all but very low 
frequencies. The forces were measured on a light, 


rigid forebody supported from the afterbody by a 
specially designed strain-gaged flexure assembly. 
Tests with two propellers differing only in blade 
thickness revealed the separate contributions of 
blade loading and thickness and the results obtained 
agree favorably with the analytical predictions. 


278 


1. INTRODUCTION 


Propeller induced ship hull virbration continues to 
be a major source of uncertainty and, indeed, 
frustration to the naval architect. Today we witness 
a trend toward larger and faster ships with higher 
power being delivered to the propeller. These 
designs are inherently more susceptible to propeller 
related vibration problems, as has been learned 

from bitter and usually costly experience and this 
situation has focused renewed attention on the need 
for improved methods to predict propeller exciting 
forces - methods which are both reliable and practi- 
cal for application during the design process. 

Two distinct, but related types of propeller 
exciting forces (and moments) produce hull vibration. 
Unsteady blade loads developed by the propeller 
operating in the nonuniform ship wake and trans-— 
mitted to the hull directly through the propeller 
shafting are termed bearing forces. Periodic 
pressure forces acting on the surface of the 
hull, arising from the propeller unsteady veloc-— 
ity and pressure fields, are called surface forces. 
Various approaches have been developed to predict 
these forces from model tests. For example, bearing 
forces are measured on a model propeller in a water 
tunnel using wake screens to simulate the flow at 
the ship stern. Surface pressures can be obtained 
from measurements of transducers distributed over 
the surface of the model hull afterbody. Alterna- 
tively, the entire hull afterbody can be cantilevered 
on a flexure assembly instrumented to measure the 
total surface force [separated stern technique, 
Stuntz et al. (1960)]. 

The foregoing experimental techniques, and 
others [most notably Lewis (1969)], have proven to 
be costly and difficult to carry out in practice. 
Moreover, a large number of experiments would be 
required to examine all the pertinent physical 
parameters, including hull form, propeller clearances, 


blade geometry and loading characteristics. Con- 
sequently, researchers are attempting to develop 
theories and numerical procedures for calculating 
propeller exciting forces. An analytical approach 
offers a means to economically evaluate competing 
propeller-hull design concepts as well as to diagnose 
at-sea vibration problems and identify corrective 
measures. 

The present paper concerns recent advances in 
the theory for propeller induced surface forces. 
A general three-dimensional boundary intercepting 
the propeller disturbance field poses a formidable 
diffraction problem. As a first step, it is 
necessary to determine both the time-average and 
unsteady loading on the propeller. All of the 
components of loading, together with blade thickness, 
contribute to the propeller induced flow impinging 
on the hull and the resultant unsteady pressure. 
Fortunately, as a result of much past work in the 
analytical prediction of bearing forces, there now 
exist powerful theoretical methods for calculating 
unsteady propeller loading in a prescribed nonuniform 
flow. The analysis rests on a lifting-surface 
representation of the propeller, explicitly account- 
ing for number of blades, radial and chordwise 
distribution of loading, thickness, and skew. While 
further refinements and improvements, such as the 
prediction of transient blade surface cavitation, 
are needed, the calculation of blade loading can 
now be done with sufficient accuracy to address the 
surface force analysis. Also, as these improvements 
in the propeller calculation become available, they 
can be incorporated into the surface force calcula- 
tion without fundamental changes. 

Previous analyses of the surface forces are 
formulated in terms of the diffracted potential 
flow about the solid boundary in the presence of 
a given propeller onset flow. To facilitate the 
analysis, it was necessary to introduce simplified 
representations of both the propeller and the 
boundary as outlined by Breslin (1962) and more 
recently, Vorus (1974). For example, analytical 
expressions for the vibratory force produced on a 
long flat strip and a circular cylinder adjacent 
a propeller in uniform flow were derived some years 
ago [Tsakonas et al. (1962) and Breslin (1962)]. 
These investigations provided useful insights regard- 
ing the importance of propeller tip clearance and num- 
ber of blades. However, such approximate treatments 
neglect what are now known to be certain essential 
physics of the propeller-hull interaction. The net 
force on a long boundary may be deceptively small 
because of cancellation of large out-of-phase force 
components developed fore and aft of the propeller. 
On a hull which terminated in the immediate vicinity 
of the propeller, such cancellation will not occur. 
Also, the components of unsteady blade loading at 
or near blade frequency can produce much larger 
surface forces than those arising from the steady 
loading and thickness. Components of blade loading 
at higher frequencies, while relatively smaller in 
amplitude, generate field pressures which decay 
much more slowly, encompassing a large portion of 
the hull afterbody and resulting in a significant 
integrated force. For this same reason, an experi- 
mental determination of the total surface force by 
measurement of pressures at selected positions on 
the hull boundary can be disastrously misleading. 
In view of these circumstances, it is now generally 
accepted that a satisfactory theory must represent 
the hull boundary in a reasonably exact fashion, 


279 


accommodate the presence of the free surface, and 
account for all constituents of propeller loading. 

This paper sets forth a comprehensive theory 
for propeller-hull interaction and describes proce- 
dures for calculating the periodic forces acting 
on the hull surface. The paper is divided into 
five sections. In the first section, the problem 
for the diffracted potential flow about the hull 
is formulated, in which the propeller unsteady 
disturbance is assumed to be of small amplitude 
and high frequency. In keeping with the desire for 
first order results, the high frequency linearized 
free surface conditon applies. However, the zero 
normal velocity condition is satisfied exactly at 
the hull boundary. Formulae for the surface pres- 
sures and forces may then be expressed in terms of 
the propeller velocity potential and the unknown 
diffraction potential. The following section deals 
with the representation of the propeller. Dipole 
singularities with strenths related to the blade 
pressure loading and thickness are distributed over 
helicoidal surfaces approximating the geometry of 
the actual blade surfaces. Based on this model, 
expressions for the field point velocity potential 
arising from loading and thickness are developed. 
Examination of these formulae and their asymptotic 
behavior at large distances reveals important prop- 
agation characteristics associated with the unsteady 
blade loading components at and near blade frequency. 

In the subsequent sections, two methods of solu- 
tion are developed for determining the surface 
forces. The direct approach consists of distributing 
time-dependent source singularities over the hull 
surface with the source strenths determined for a 
prescribed propeller onset flow using a modified 
Douglas-Neumann calculation [Hess and Smith (1964)]. 
The force on the body is then found by applying 
the extended Lagally theorem to the hull singulari- 
ties. In an alternative approach, based on a 
special application of Green's theorem, the force 
is obtained by finding the velocity potential at 
the propeller produced by the hull boundary executing 
simple oscillatory motion. 

In the final section, a towing tank experiment 
is described in which blade frequency forces were 
measured on a body of revolution adjacent to a 
propeller operating in uniform flow. The simplifi- 
cations of body shape and propeller loading provided 
a physical model which could be treated in a reason- 
ably exact fashion by the theory. Despite these 
simplifications, certain classical problems were 
encountered in the design of the experiment including 
the measurement of a relatively small force, avoid- 
ance of system resonances in the frequency range of 
interest, and retrieval of the force signal from 
background noise. A two-part body design was 
developed, similar in concept to the separated 
stern technique mentioned earlier. A heavy after- 
body attached to the towing strut, behaved as a 
seismic mass at all but very low frequencies. 

Forces were measured on a light rigid forebody, 
supported from the afterbody by a specially designed 
and dynamically calibrated straingaged flexure 
assembly. 

Tests were performed with two propellers differing 
only in blade thickness in order to reveal the 
separate contributions of loading and thickness. 

The measured forces (amplitude and phase) were 
obtained for a range of speeds and advance coeffi- 
cients and for two positions of the propeller 
relative to the test body. The results agree 


280 


favorably with the theoretical predictions. It is 
recommended that this experimental technique be 
extended to study the effects of nonuniform flow and 
intermittent blade surface cavitation. 


2. FORMULATION OF THE PROBLEM 


Consider a ship moving at constant speed U through 
otherwise undisturbed water. We seek to determine 
the periodic forces and moments exerted on the 

ship hull surface arising from the unsteady propeller 
velocity and pressure fields. The fluid is con- 
sidered to be incompressible and inviscid and within 
the domain bounded by the free surface, the hull 
boundary, and the propeller blades (and trailing 


vortex wakes), the flow is assumed to be irrotational. 


Under these circumstances, a fluid velocity potential 
exists which can be expressed in terms of steady 
and unsteady components as 


o(x,t) = Ux + d5(x) + 6) (Kt) + bp (x,t) 


Here, x = (x,y,z) is a cartesian coordinate system 
fixed to the ship with the x and y axes in the me 
waterline plane, and z-axis directed upward. 9, (x) 
is the steady disturbance flow about the bare hull 
in the presence of the free surface, $,(x,t) is the 
propeller potential, and bp Gx, t) is the potential 
of the flow arising from the propeller-hull inter- 
action, often termed the scattering or diffraction 
potential. It should be noted that the presence 
of the viscous, rotational wake of the ship is 
ignored in the diffraction problem, i.e., it is 
assumed that the unsteady pressure forces on the 
hull can be derived from potential flow considera- 
tions alone. 

The propeller potential is periodic in time and, 
by virtue of the symmetry of identical, equally 
spaced blades, may be expressed as a Fourier series 
with harmonics in blade passage frequency as 


a = 1 a! 
bp (x,t) = a bp, Gee Net (1) 


with 6 being the complex amplitude of nth harmonic. 
(In this and all subsequent expressions involving 
einNwt the real part is understood to be taken.) 
Similarly, the diffraction potential will be of the 


form 
co 


bp Ge, t) =) ip, Ce (2) 


n=0 


We now consider the boundary value problem for 
the potential $ = $, + $p, assuming the fluid 
disturbance velocities to be small compared to the 
ship speed, i.e., |V>| and |V¥$,| <(U. Within the 
fluid domain, the potential must satisfy Laplace's 
equation 


V2o,(x) = 0 (3) 


At large depth and distances upstream of the hull 
and propeller the disturbance must vanish 


x > -© 


Vo, > 0 
>0 Za OF (4) 


and at large downstream distances, x > + 1d. 
satisfies a suitable radiation condition. 

The boundary condition on the hull surface, 
denoted by S, requires that the fluid velocity must 
be tangent to the surface, or 


a a = 


n° Voy (x) = 0 x on S (5) 


a being the outward unit normal vector to the sur- 
face (see Figure 1). Here we have assumed the 
hull to be rigid and stationary with respect to the 
translating coordinate system (i.e. hull motion 
and deformation due to propeller excitation is 
ignored) . 

The linearized free surface boundary condition 
May be written in the form 


2 
, 9b 3 oa 3b 
-(nNw)“o_. + (2inNwU) —— + U 7 ej ——— =_ 
n ax 2 
ax dz 


on, z9= 0 (6) 


In order to establish the relative magnitude of 
terms the equation is recast in nondimensional form 
using the ship speed U and propeller radius Rg for 
reference length and time scales, obtaining 


a 326 gR 3a 

n 2 n oO D igh 
ae + —_ — = = 
ax € uD + G2 € NE O on z (0) 


=) ar 2ie 


where € = J/mnN, J being the propeller advance 
coefficient. It may now be observed that typical 
propeller applications, « <<_1 and the first term 
will dominate. Thus, as a first approximation the 
free surface boundary condition (6) reduces to 


bn (x) = 0 on z =0 (7) 


This completes the statement of the boundary value 
problem for the diffraction potential as summarized 
in Figure 1. It should be noted that by virtue of 
(7), the function bn (x) can be analytically continued 
into the upper half plane, z > 0, in a straight- 
forward manner. As will be shown in subsequent 
sections a solution can be constructed in terms of 


VELOCITY POTENTIAL 

Fz DXL) UX + QK)4+h (Kt) 

= = >, inNwt 
PRE Piwe 


v2 py =0 
RV, | ,70 
Pa| 20) 


B=0 


lybal-= 0, |X|-e 00 
E<O 


FIGURE 1. 
problem in propeller-hull interaction analysis. 


Coordinate system and boundary value 


appropriate “images" of the propeller and hull 
singularity systems. 

Upon solving for the velocity potential, all 
other quantities of interest can be determined. 
The linearized, unsteady component of pressure is 
given by* 


v6) , Wa Gc) = iu + Vo, Gx) 
(8) 


p(x,t) = -p (4449, 


or from (1) and (2), 


inNwt 
p(x,t) eee 


I 
1 
me} 


[inNwd + Ue - Von] 


(9) 


i] 
to 
2} 
zy 


where the Py (x) are amplitudes of harmonics of the 
unsteady pressure. The periodic force, F(t), and 
moment, M(t) acting on the hull surface (see Figure 
1) may be written as 


F(t) = - pnds (10) 
Ss 
and 
M(t) = - pxxnds (11) 


Inserting the expression for p, one obtains the 
amplitudes of the force and moment harmonics, as 


a —S => 
r= 2 (inNwd, + Vs * Von)n ds (a) 
s 
and 
—> A — oy —. 
MO =o0 (inNw$, an Wa O Wun) BS 2 inh ols} (AES) 
s 


Until now, the propeller’potential has been regarded 
as a known function. Before proceeding with the 
surface force analysis, it is appropriate to discuss 
the analytical representation of the propeller and 
the velocities and pressures induced at arbitrary 
field points. 


3. REPRESENTATION OF THE PROPELLER 


The primary source of propeller exciting forces is 
the spatially nonuniform wake of the hull in which 


*To be strictly consistent with the high frequency approxi- 
mation, the convective pressure term should be discarded. 
However, this term adds no serious burden to the ensuing 
analyses and by retaining it, numerical calculations can 

be used to demonstrate that the contribution from this term 
is, in fact, negligibly small. 


281 


the propeller operates. As viewed in a coordinate 
system rotating with the propeller, the flow 
approaching the propeller consists of time-average 
or circumferential mean component and an oscillatory 
component. The oscillatory component gives rise to 
unsteady loading on the blades in a manner analogous 
to a hydrofoil encountering a sinusoidal gust. This 
unsteady loading, summed over all the blades, yields 
periodic shaft forces at blade frequency and integer 
multiples. In contrast, the periodic pressure 
forces acting on the hull surface arise from the 
induced velocity and pressure fields from both the 
mean and unsteady components of loading, as well 

as the blade thickness, because of the varying 
aspect of the rotating blades relative to the fixed 
hull boundary. 

Propeller theory for unsteady flow has developed 
as a logical extension of linearized lifting-surface 
theory for hydrofoils. It is assumed that the 
oscillatory components of the wake velocities are 
small compared to the mean, and can be resolved by 
Fourier analysis into "wake harmonics," the funda- 
mental harmonic being the shaft rotation frequency. 
Each of these harmonics, within the linear approxi- 
mation, will produce a component of unsteady blade 
loading with the same frequency. By virtue of the 
propeller's symmetry, upon summing over all the 
blades, only certain harmonics of the loading will 
contribute to the net force on the shaft. However, 
all the harmonics of loading contribute to the 
forces on an individual blade, and, as will be seen, 
to the radiated pressure field of the propeller. 

The propeller lifting-surface theory developed 
by Tsakonas et al. (1973) is adopted in the present 
work. This analysis and associated computer pro- 
grams have been successfully applied in recent 
propeller designs to minimize bearing forces, e.g., 
Valentine and Dashnaw (1975). In addition, the 
analysis has been extended to compute field point 
velocities and pressures, including the contributions 
from the image of the propeller arising from the 
presence of the free surface. As the details of 
the development of these formulae have been largely 
reported in the literature, we shall not burden 
this paper by recounting them, being content to 
outline the procedure. 


Blade Loading Potential 


The linearized equation of motion for unsteady flow, 
referred to a non-rotating cylindrical coordinate 
system (x,r,f) centered at the propeller axis 
(Figure 2), may be written 


36 36 
Sent p 19) 
2 i Dem Os (re) 


=o (15) 


where p is the pressure induced by the loadings on 
the blades due to camber and incidence and p', for 
later convenience, denotes the fluid density. Here 
the angles of attack are produced by each axial 

and tangential spatial harmonic of the nominal hull 


282 


FIGURE 2. Propeller coordinate system-projected 
view looking upstream. 


wake which is presumed to be known from wake survey 
measurements. 

The pressure induced at a field point by a single 
blade is given by the following distribution of 
pressure dipoles 


M 
LUpaat -ikut 9 1 
epasy pe) = a a | Apr(&,p)e np Ros 
Sp A=0 


Pp (16) 


where Ap, is the complex amplitude of the pressure 
loading on the blade arising from the wake harmonic 
order X and, as illustrated in Figure 3, 


Sp is the surface of the blade, represented ap- 
proximately by the helicoidal surface & = U/w 


Ny is the distance directed normal to the surface 
S 


R = [(x-&)? + r2 + o2 - 2rp cos(§ +a -¥Y)] is 
the distance from a point (&,0,89 + a) in the 
surface Sp to the field point (x,r,/) 


6 = - wt is the angular position of the blade 


We note that the representation of the blade is 
only approximate for a wake adapted propeller, 
being correct for a constant pitch propeller in 
uniform flow. Here we also assume that the pressure 
jumps on the blades, Ap), have been previously 
calculated by the unsteady lifting-surface theory 
such as developed and programmed by Tsakonas et al. 
(1973). ; 

To place the harmonic content of 1/R in evidence, 
the following identity can be used 


Rea Jp pita lee 9 ars co 10 
ana | (17) 


where the amplitude A | is given by 


, 


Tq) 10K) yy (Le lx) p<xr<o 


Tog) (HE[=)K) gy (110) 0<xr<op (18) 


Im and K, being the modified Bessel functions of 
the second kind of order m. 

To secure the pressure field for an N-bladed 
propeller, the blade position angle 9 is replaced 
by 8 + 2m/N and the sum over n from n = 0 to N - 1 
carried out. This sum yields a factor N and the 
constraints on the frequencies A and m, given by 
X =m = RN with & = 0,+1,4+2,+3...i.e., products of 
terms for which A - m # 2N will sum to zero. ‘The 
total induced pressure at any field is secured by 
summing over 2 from -~ to +”. 

Upon use of (15), (16) and (17) and looking after 
the shifted time variable, using 8 = -wt which shifts 
to -wt + w/U (x-x'), one obtains the velocity 
potential in the form 


co 


N y i2Nwt 
bp (x,r,P,t) = ~ ou S 


M Q=-0 
A , 12) rU,TG7Se ds 
) a Py (E,0) Pq (Xr /PH E70) ae 
A= Sp 


in which the propagation function, Pr is given by 


an2 


aro “7 } ak (20) 


where for each 2, m= A - &£N, and M is a practical 
upper bound of the wake harmonic order number beyond 
which the amplitudes of the wake harmonics are so 
small as to render negligible values of Ap, for all 
A > M. (A value of M = 8 is reasonable). Details 
of further reductions of the integrals involved in 
(19) and (20) may be found in Jacobs and Tsakonas 
(1975). 


BLADE 
REFERENCE 
LINE 


p(t) —~+ 


7 HELIX: f= Ge 


FIGURE 3. Propeller coordinate system-expanded view 
of blade section at radius p. 


To account for the presence of the free surface 
which, at the frequencies of interest acts as a 
zero potential surface (see Eq. (7), we merely add 
to (19) the potential dp; =- bp (Yur pyr t) in 


which 
a = ly 24+ (2d-z.)2 =r when z =4d (21) 
1 p p p 
a | 
fy = tan aE = 7” when Zp = d and 
= Wf Sia S UPznP Leo) = © (22) 


where d is the distance or depth of the propeller 


axis below the free surface; y,, Z, are the transverse 


and vertical coordinates of any field point (Figure 
2). Thus, the total potential arising from the 
loadings on an N-bladed propeller in the presence 
of the free surface (neglecting the feed-back on 
Ap, from the free surface) is 


N = 
=-—— 2Nwt 
%D; e'U iy Srey 


dp + 
Q=-00 
M 
a A Ap) (€,0) [Pm (x,r,776,0) - 
A=0 S 
Pm(x,ri,fii&,0)] Aas (23) 


and the spatial derivatives of this function yield 
the velocities induced by the propeller and its 
negative image in the free surface. Clearly bp + 


oH = © seers Enlil Se Ehatel Yp for Zp = Glo 


Blade Thickness Potential 


The potential, $,, induced by blade thickness may 
be constructed from a distribution of dipoles (with 
axes tangent to the helical arc along the blade at 
any radius) whose strenths are given by V_, V being 
the local relative resultant velocity and T the 
local thickness provided by the expanded blade 
section drawing. Using the helical geometry as 
before, one can obtain 


Ro on 
| ‘i U2 + (wo)? t(p,a) 
Ry Oo 

) 


iL 
aaa da dp (24) 


or (x,r,f,t) = Te 


where ae (Pp) and a;(p) are the angular coordinates 
of the blade leading and trailing edges. 

To allow for the free surface, 1/R is replaced 
by 1/R - 1/R; with Ry being the distance from the 
reflection of the dummy point in the free surface 
to the field point on or below the water surface, 
making use of relations (21) and (22). Again, to 
place the harmonic content of 1/R and 1/R; in 
evidence and to facilitate integrations over the 
blade surface, the Fourier expansion (17) can be 
applied. 


283 


Asymptotics of the Loading Potential 


The fact that the disturbances induced by each of 
the pressure jumps Ap, are propagated by widely 
different functions of the space variables x,r, ~ 
must be emphasized as these behaviors have a most 
significant impact on the pressure, velocities, and 
the resultant forces generated on the hull. These 
diverse characteristics can best be illustrated by 
examining the asymptotics of the potential for 
upstream locations which are large only with respect 
to the x-wise extent of the blade surface. The 
x-wise extent of the blades is given by the (chord) 
5 Satin bpp being the local pitch angle which, in 
the radial region of heaviest loading, is normally 
of the order of 25° For merchant ships, the blade 
chord in this region is of the order of one-half 
the radius and, hence, the x-wise extent of the 
significant position of a propeller is only about 
0.2 radius. Thus, for axial distances of the order 
of one diameter, the x-wise extent of the important 
region of the blade can certainly be neglected in 
an asymptotic analysis. 


Using the expansion of Rae given by 


Ree © Qm—a72 ') eae) 
T pr 


m=-0 
where Q is the associated Legendre function, and 


MAGE) eistelaeet tO 
2pxr 


and retaining only the leading term in the expansion 
of Q for large Z, one can arrive at the following 
behaviors for the ae ae of the loading poten- 
tial, i.e., dp = Oe op being the part 
associated ten oe OR and To) being that 
arising from the torque-producting loading in the 
forms: 


Siete -i2Née 
an2 p'w 
Q=-2 
een 
Ro 
r 
m : anya | 
Rh 
(25) 
eas czy a 


San MIME eee 
(x24r2402) |™ Eve [x?+r°+p*+4d (a-z,,) ] ae 


inl -i2ne 
pred | ct ae Ne oe 
woe a 
ae An2p'w* oe 
Q=-0 = 
(m=A-2N) 


continued on page 284 


284 


R 
oO 
aN 2 
HEE alta 
Rh (26) 
em? "3 elm fi a 
L(x? +2407) Ree [x2424p?+4a (a-z,,)} mI#1/2 
where 
e flea Oct Cn ae a AZ) e_2)  yeecp cheer 
<Im| I (|}m| +1) 
function (27) 
and 
ee 
LA) =o Ap) (Pp, 0) asus da, the load density 
(28) 
ae 


Here the effect of the free surface is included by 
the last terms in each integrand. To exclude the 
free surface, take the propeller depth of submergence 
d=, 

Limiting our attention to blade rate (£ = 1, - 1), 
we see that, although the mean pressure jumps Apo 
(A = 0) are much larger than those at all other 
wake harmonics, the propagation functions for m = 
X - &N = +N exhibit extremely rapid decay with 
increasing x. In addition, we observe that the 
radial loading for m = +N obtained from Apy is 
weighted by the oscillatory function eiNa which 
has the effect of producing an Iy(o) which is 
inversely proportional to N. In contrast, the 
contribution fon Ae— Ni, ase, m=O} as ofthe 


form 
(N) 
Lo = ie da ' 


which has a "non-destructive" weighting function of 
unity. Another feature which reduces the mean 
loading contribution to the generation of forces 

on the hull (wherein integration over the athwart- 


ship variable yp is involved) is the presence of 
the space angular function 


: R -1 
oiNy aya iN (tan Yp/Zp) 


yielding pressures and velocities at different Yp 
which are not in phase. In strong contrast in the 
propagation mode for the blade frequency loading 
Apy (for which m= 0), dp has no dependence on p 
or p4, and all yp locations receive velocities and 
pressures which are in phase with each other. On 
the other hand, the coefficient C|,| is large for 

} = 0 (being 6.5 for a 5-bladed propeller), whereas 
Cc. = 1, the multiplier for the contribution from 
the blade frequency Ap's. 

These observations are succinctly summarized in 
Tables 1 and 2 for the case of a 5-bladed propeller, 
displaying the rate of decay with x, the variation 
of the influence coefficients C]m| and mC|m|/1+2 m|, 
and the dependence on the angular space coordinates 
pand ~;, without and with the free surface effect 
for the dominant terms at blade frequency arising 
from the loading at wake harmonics i} = 0, N- 1, 
Nand N+ 1. 

One may observe in Tables 1 and 2 that the effect 
of the free surface does not generally increase the 
rate of attenuation of the potentials with x except 
at or near all points in the vertical plane yp = 0 
with the exception of the bp (N) and > (N) arising 
from blade frequency loading on the blades, i.e., 

} =N and m =0, which show a change from x72 to x74 
and x7! to x-3 everywhere, respectively. 

A dramatic contrast in the force-generating 
capabilities of the pressure field components arising 
from the mean (the largest) and the blade-frequency 
loadings on the blades can be found by integrating 
the pressures 


-p! Og) eng 59 94,7, (N) 
ot the 

over a rectangular region of half-breadth b arranged 
symmetrically z, units above the propeller and 
extending from -f radii forward to s radii downstream 
of the propeller plane. Upon defining the coeffi- 
cient of the vertical force on the rectangle as Ze (A) 
= FZ) /o'n2p4, we can arrive at the following 


TABLE 1. ASYMPTOTIC CHARACTERISTICS OF BLADE FREQUENCY COMPONENTS 
OF THE THRUST-ASSOCIATED POTENTIAL ¢, FOR A 5-BLADED PROPELLER 
FOR LARGE AXIAL DISTANCES 


Wake Propagation Influence Relative 
Order Order Coef. Loading* 
r m C mil A) 
2pQ, 
0 =5 8.48 26.7 
N-1=4 -1 4.71 4.8 
N=5 0 3.14 1 
N+1=6 1 4.71 2.1 


Dependence on x, py and ¥, 


Without With With 
Free Surface Free Surface Free Surface 
(yp =9) 
x et ise x (cae 1) 26d (d-Zp)x 
Ix|3 ix |!3 x5 
x tiv x (el? -e “4 10d(d-z,)x 
Ixi Ix [x7 
x 6xd(d-zp) 6xd(d-zp) 
ixP Ixf Ix 
xe? x(eriP—e 10d (d-z),)x 


7 
ix/5 xf |x| 


285 


TABLE 2. ASYMPTOTIC CHARACTERISTICS OF BLADE FREQUENCY COMPONENTS 
OF THE TORQUE-ASSOCIATED POTENTIAL ¢g FOR A 5-BLADED PROPELLER 
AT LARGE AXIAL DISTANCES 


Wake Propagation Influence Relative 
Order Order Coef. Loading* 
r m mC) | LO) 

1+2|m| 2p), 

0 -5 -3.86 26.7 

N-1=4 -1 -1.57 4.8 
N=5 0 0 1 

N+1=6 1 IES; 2.1 


Dependence on x, y and ¥; 


Without With With 
Free Surface Free Surface Free Surface 


(yp =9) 
eisy eid y_ li 22d(d-zp) 
Ix }x |} |x ls 
eiv civ _ eli 6d (d-zp) 
IxP [xP Ix 
ou 2d(d-zp) 2d(d=zp) 
Ix! Ix? Ix 
eiv paigmemGi 6d(d-zp) 
IxP [xP Ix 


*These are relative values as obtained from calculations of a 5-bladed propeller 


using the wake of the SS Michigan. 
** L O) 


im 


2 p04, 204, 


a 


~ a4, 


expressions for the moduli of the blade-frequency 
forces, viz., 


DO Nate 
C. \(2e4b ) 


[Zp] = sin [ (Nt) 7] 
an2p 'n2R2 
(o) 
1.0 
N+1 : 
fe) Ap, (p) sin Noy, 
0,2 (29) 
—(2N+ ~(2N+ 
{ (Vezeztep® NN) _ (Veep? "ONY yap 
where 
Gy, = dbeSaBoos (NEI) ig 2 tem! ica 
N Zo 
2 (N+1)! 
for the contribution from the mean loading, (i = 
0), and 
10 
N N 
|z ( ai | 4p, | 2n 
T 202 N 
81p'n“R 
) 
0.2 
(30) 
(b+ [s*4+22+p2+b2) 
SSE do 


(b+, [£24+22+p24b7) 


for the contribution from the blade-frequency loading 


on the blades. 


b 5 
Ap, (p,a%) eM da = Apy (p) ar 
b 


sin may, 


for Ap, independent of a 


Evaluations of (29) and (30) were carried on a 
hand calculator for various integration lengths f 
forward of a propeller using assumed radial distri- 
butions of Ap, and Apy and representative values 
from computer calculations for a 5-bladed 22.5 ft 
propeller in a single-screw ship (model) wake. The 
calculations were made for a flat-bottomed hull of 
half-breadth b = 2 Ry at 2, = 1.5 Ry (25 percent 
tip clearance) and a stern overhang s = 1. Results 
shown on Figure 4 show dramatically that the force 
arising from the blade-frequency (b-f£) loading is 
(asymptotically) 65 times larger than that from the 
mean blade loading when the free-surface effects 
are omitted (note that Ap, = 40 Ap;). Furthermore, 
the total force due to b-f blade loading rises very 
slowly to its asymptotic value as the integration 
length is increased and even the force from mean 
blade loading requires integration of the pressure 
to three radii forward of the propeller. 

To allow approximately for the effect of the 
free surface, one can subtract terms of the same 
form as (29) and (30) with z,* replaced by z,* + 
4dh with d being the depth of submergence of the 
propeller axis and h the hull draft in way of the 
propeller. The reduction in force for d = 3.5 and 
h = 2 is significant for Zp (N) but is found negli- 
gible for the smaller force. As expected, the 
asymptotic value Zp (N) (£09) is more quickly achieved 
due to the presence of the free surface, but, never- 
theless, requiring that one integrate to some 8 
diameters to achieve the final value. 

These results tell us that the current practice 
in European model basins (in which b-f pressures 
are measured on models in the vicinity of the 
propeller and these are integrated in an attempt to 
secure the b-f hull force) is highly suspect because 
the slowly decaying pressures from b-f blade loadings 
contribute large sectional force densities far from 
the propeller. This effect is exacerbated by the 
"growing" cross-sectional shape as one integrates 
forward which is not accounted for in the constant 
beam "ship" used in the foregoing analysis. 


286 


As an order of magnitude formula, one might use 
(30) for £ = ~ with the correction for the free 
surface included. This reduces to the complex 
amplitude 

1.0 


Zn = pAp. Qn 


(b+ s*+z *+p°+b?) 


do 


(b+ \s2+z 20 2 +4b2+4dh) (eat) 


(which must not be used for hull drafts in way of 
the propeller, h, which are small, as clearly Zep (N) 
*+0O as h-+0O). In practice, Apy = ay(P) cos NO + 
by(p) sin N8, ay, by being the chordwise average in- 
phase and quadrature blade pressures given by the 
unsteady lifting surface calculation. 

With the foregoing considerations of the propeller 
in mind, we now return to the surface force problem 
for a general three-dimensional hull boundary and 
prescribed propeller onset flow. In the following 
section, a procedure is described for determining 
the diffraction potential and the surface pressures 
and forces in terms of singularities distributed 
over the surface of the hull. 


x10 


FIGURE 4. Approximate moduli of B-F forces on 
barge-like ship from pressures emanating from 
mean and B-F loadings on a 5 bladed propeller 
(in a single screw ship wake) as a function of Oo 
integration length forward of propeller. 


, 


B-F FORCE COEFFICIENTZ 


(EFFECT OF FREE SURFACE NEGLIGIBLE) 


5 10 15 20 25 30 
INTEGRATION LENGTH, f, FORWARD OF PROPELLER IN RADII 


4. A DIRECT APPROACH FOR DETERMINING SURFACE FORCES 


A "frontal attack" on the problem of predicting the 
vibration forces generated on an arbitrary hull by 
the induced flow of the propeller, (and its free sur- 
face image) is to construct the potential of the hull 
in the presence of these onset flows. This procedure 
was first applied by Breslin and Eng (1965) toa 
realistic hull form. At that time, however, only 

the mean loading and the blade thickness were 
accounted for in the flow impinging on the hull and 
the computer time was observed to be excessive. In 
contrast to these earlier efforts, the propeller 

flow is now composed of all constituents of loading 
and the (high frequency) images arising from the 
presence of the free surface. 

A solution for the potential, $,, which satisfies 
equations (3), (4), and (7), is constructed by 
distributing source singularities, on (x)einNut, 
over the surface of the hull, such that 


dn) = - A o,f") (—— - —4— Jas) 
Ss |x-x" |x-x! | 
ab 

aC Dae O (32) 


where the region of integration is over the submerged 
portion of the hull and x; is the distance from an 


ASYMPTOTIC VALUE 


FOR f—=@ a 


6.28 


1 ARISING 
FROM B-F LOADINGS Op, ON PROPELLER BLADES 


WITHOUT FREE SURFACE 


WITH FREE SURFACE CORRECTION 
(d=3.5, h=2) 


a NOTE EXPANDED 
Z, —B-F FORCE ARISING FROM SCALE ———= 


MEAN LOADING (4 P= 40 Ap, !) 


WITHOUT FREE SURFACE 


“image” hull point to the field joteulics abo, alse 

x' = (x',y',2!), then xt =X AV Ze ee Source 
strengths o,(x) can be determined by applying the 
hull boundary condition (5) yielding an integral 


equation 


(x) 

oO (x 

n 1 — al iL 

5 > it oy, (x")n o W ae SES as 


+ n(x) “(Yop + Yop; ) = 0, Sons (33) 
n 


The integral teria gives the contribution from all 
source elements other than at the point of interest 
on the hull. The contribution from the source at 
that point is given by the first term, 0, (x)/2. 

Equation (33) with n- [Vop_ + Vop; _] as a known 
i 7 ; n in 9 
input is solved numerically by the generalized 
Douglas-Neumann program [Hess and Smith (1964)]. 

In practice, the hull surface is divided up into 
quadrilateral elements over which o, is considered 
constant and the integral equation is replaced by 

a set of simultaneous algebraic equations. Care 
must be exercised to insure that the sizes of the 
elements are small compared to the spatial "wave 
length" of the propeller-induced velocity field. 
This is particularly the case for field points just 
downstream of the propeller since the velocity 
components rapidly become proportional to sines 

and cosines of N(w/U x-Y) so that the wave length 
of these signatures is A =2mU/Nw, which, for J~1 
and N = 5, becomes A = 0.4R,. In order to obtain 
representations of an entire cycle, it is necessary 
to take element lengths of one-quarter of this 
length or about 0.10R,. Upstream, the induced flow 
is monotonic in x and the element sizes can be made 
much larger without loss of accuracy. 

It is acknowledged that the above-described pro- 
cess does not, in principle, completely solve the 
problem since the feedback of the hull sources on 
the instantaneous flow experienced by the propeller 
is not included in the propeller loadings Ap). To 
do this would require joining the integral equation 
for the propeller loadings (with input from the 
propeller generated hull sources) to Eq. (33) to 
form a pair of integral equations for Ap) and oy, 
which, when solved interatively to convergence, 
would yield the complete solution. For the present, 
we are content to ignore the hull feedback on the 
propeller. 

Once the source densities on the hull surface 
are found, it is convenient to determine the force 
induced on the hull in terms of simple integral op- 
erations on these sources. Although the Lagally 
theorem and its extension by Cummins (1957) is known 
for submerged bodies, it is necessary to develop a 
form which is suitable for use for floating bodies 
beset by high frequency flows. 

The force as given earlier by Eq._ (12) may be 


considered as the sum of two terms F, ) and F, (2) 
given by 

ae = ipnNw [fe nds (34) 
and Ss 


F,(2) = 9 Jf Vz * Vb, nds (35) 


287 


Since $y = 0 on z = O, the region of integration 

in (34) may be extended to include the hull water- 
line plane S, (see Figure 1), thus forming a closed 
surface about the volume ¥ inside the submerged 
portion of the hull, and 


F (1) = ipnnw bn nds (36) 


S+S5 


where the symbols ( )* and ( )~ are used to denote 

a quantity evalutated on the outside and inside of 
the surface of integration respectively. Noting 

that for $,(x) given by (32), bn? = bn (i.e. the 
potential is continuous across a surface distribution 
of source singularities), and using the vector 
identity n=n- Vx, one obtains 


st 


F,(1) = ipnNw dan * V¥ds(x) (37) 


S+S, 


By means of Green's reciprocal theorem applied to 
the volume ¥, (37) becomes 


Sy — 
x 


xn ° Vo_ dS(x) (38) 


° 


since V + V(x) = 0 and V*$," = 0 in ¥. A fundamental 
property of a surface distribution of source singu- 
larities relates the jump in the normal derivative 
of the potential to the local source strength, viz. 


Bl) 


no) Vout = ni) Vduo cy, (39) 


But since n° Vont = 0 on S by virtue of the boundary 
condition (5), Eq. (38) may be written as 


—— — —_ —_. 
Fy, (1) = -ipnNw X On(x) dS(x) + 
S 
aS OO => 
ipnNw x a ds (x) (40) 
Ss 
oO 


The first term in (40) has the same structure as 

that derived by Cummins (1957) for submerged bodies 
generated by internal singularities. The second 
term arises from the capping of the volume by 
extending the free surface through the ship (proposed 
originally by Breslin in 1971). For the important 
case of the vertical force, Fone we obtain 


FZ (1) = ipnw i Op (x) aS (x) (41) 


Ss 


A similar analysis can be applied to the convec- 
tive term F, (2) (see appendix A) to obtain 


F,(2) = - p ile Wop. + op, ) eS) + 
n 


s 
3 


fe) I Vo5 = ds 


So 


(42) 


in which again the first term exhibits the same 
form as for a submerged body and the second term 
accounts for the intersection with the free surface. 
If it is assumed that 96 /dz = 0 on z = O (rigid 
wall free surface condition for the steady flow 


about the hull, i-e., low Froude number approximation), 
then from (42) 
FE, (2) = -9 o i108 (dp + op, ) ds (43) 
=o S dz n in 
S 


and the total vertical force, F 
becomes 


; an 
=> Res, Oi. oF Oma a + . 
“Bg ff ee ca tr 
Z 


As noted earlier, the first term under the integral 
will dominate because of the large multiplying 

factor nNw. This will be confirmed in the calculated 
example to be presented subsequently. First, how- 
ever, we outline an alternative approach for 
determining the vibratory hull force which avoids 

the need to solve for the diffraction potential. 


Zn + ) 


ds 


(44) 


5. AN ALTERNATIVE METHOD FOR DETERMINING THE 
VIBRATORY HULL FORCES 


Vorus (1971, 1974, 1976) has developed an alternative 
procedure for determining the vibratory hull surface 
forces which eliminates the need to solve for the 
hull diffraction potential in the presence of the 
propeller onset flow. The ith oscillatory force 

or moment, Fin, exerted by the pressure on the 

hull may be written from (12) and (13) as 


: =x a Sas 
Ech = (0) (inNwd, + Vs °* Von) my ctr ds (45) 
Ss 
> a 
where the a, are defined as 
= —> a = > 
a, =i Cy, = yk =z 9 
=> — > aS = 
a2 = j Ch = 64 al be Us 
a3 =k OF Gea) i ee (46) 


Vorus has shown that the solution for Fjy, with no 
additional approximation, is given by the formula 


1 /Nw 


-inNwt 


F. =— dt e ds(&,p,9 + a) 


-7 /Nw 
SPyy 


1 72 2 O A 
p /U + (wp) ™ Wate 


E (47) 


All of the variables in (47) pertain to the propeller 
except Hin- Hy, is the amplitude of the fluid 
velocity potential due to the bare hull travelling 
backwards with speed U across the water surface 
and oscillating with unit amplitude in the ith 
direction and at the frequency nNW. Since the 
details of the derivation of this formula may be 
found in the cited literature we will only outline 
major steps as follows. 
The second term in (45) can be rewritten using 

the following vector identity 
Wig 9 Woe) (ao i) SS Wo Ga) Wg 2 ta) + 

> > > == = => 
Vx[¢n(@; x Vs)] * n - oy Vx (aq x Vs) * n (48) 


Only the last term contributes to (45), because We 


*n = 0 (steady flow hull boundary condition) and, 
by Stokes' theorem 
ae > a <=> > > 
ff tes 1) (eno Ve)il om 6 =o tne x V5)d% = 0 
Ss (49) 


where the line integral is taken along the hull 


waterline on which $, = 0. Consequently, Eq. (45) 
becomes 

FR — — a —> 
Fin = 0 J fr [inNwo, - Vx(aj x Vs)] + n ds (50) 


S 


and, upon introducing the function Tela which satis- 
fies 


Uh = © in fluid domain (51) 
H. =0 z= 0, outside S (52) 
in 
= > 4. es —>) —_ 
OY isla hae [inNw a; - Vx(a4;XVg)] on S(53) 
Vu, +0 as [x| > =,z< 0 (54) 
in 
equation (50) is given by 
— > 7 
Fo? Oat AV TH eas (55) 


s 


This form can be identified as one of the terms in 
Green's theorem applied to the functions $y, and 
Hin in the fluid domain bounded by the hull surface 
S, the free surface z = 0, and the surfaces of the 
propeller blades Spyr and slipstream, Spy, which 
yields 


1 /Nw 


N -inNwt 
Fin a ra dte {f Hin 
SPy 


—7/Nw 


ap + db - 
ae eae) we, Ae 0 ; 
an an lS}. ap (bp bp )n. Vv Hi ds 
Bp Pp 
SPyytSWyy (56) 


where dp - bp is the jump in the propeller potential 
across the blade and slipstream surfaces. The two 
terms in (56) can be identified as the contributions 
from blade loading and thickness, and with further 
manipulation can be brought into the form of (47). 

Equation (47) indicates that the velocity corres- 
ponding to the potential Hj, is evaluated over the 
propeller blades and slipstream. The propeller 
representation by distributions of dipoles directed 
normal and tangential to the blade pitch surface is 
the same as previously discussed. In the formula, 
the velocity induced by the bare hull, VH;,, is 
resolved into components in the directions of the 
dipoles, multiplied by the dipole strengths, and 
the products integrated over the blade and slipstream 
surfaces. The first integral in (47), in time, 
extracts the nth Fourier harmonic. Both the blade 
position and the dipole strengths are functions of 
time. 

In the case of vertical force analyses, an 
approximation to the improper integral in (47) has 


been found to yield acceptable results. Let I be 
defined as 
inNw 
a WEE |) a 
I= e im. © Wile lis 0 (57) 


'o 
p 
3 


E 


If the oscillating exponential varies more rapidly 
than VHjy, then the argument of the exponential can 
be considered as "large" and I can be expanded in 
an asymptotic series. VHj,, should vary relatively 
slowly aft in the propeller slipstream for vertical 
oscillation of the bare hull and an asymptotic 
evaluation should therefore be valid. (Such a 
treatment may not apply to an athwartship analysis, 
for example, where a rudder is involved in the bare 
hull oscillation.) To proceed with the asymptotic 
representation, (57) is integrated by parts yielding 


inNw 
go ae ee) 2 
=- ma OW a. | 
inNw Ny win 'é 
inNw 
(57) 
y e us a im oO W SI dé' 
inNw ost I ~ alfa) 


For the conditions stated, the integral term is 
higher order. Hence, to one term, 


intw “P ~ y He) (5) 


and (47) reduces to 
1 /Nw 


Nw -inNwt 


E. = — dt e dS [o'vtn, 


-7 /Nw SP, 


289 


Ap as 
inNw np! iY Ha (5?) 


in which the induced flow is evaluated exclusively 
on the surface of all N propeller blades Spy. 


6. COMPARISON OF THEORY AND EXPERIMENT FOR A BODY 
OF REVOLUTION 


An experiment was conducted to measure the periodic 
forces on a body of revolution adjacent to a propel- 
ler loading provided a configuration which could 
be treated in a reasonably exact fashion by potential 
flow theory. As such, the experiment was intended 
as a fundamental check on the theory and computer- 
aided numerical procedures. However, it is believed 
that the experimental technique can be extended in 
the future to study more general hull geometries 
and the effects of unsteady propeller loading and 
transient cavitation. 

In the following sections, the experimental 
apparatus and procedures are described and the 
force measurements are compared with the analytical 
predictions. 


Test Body and Propellers 


The experiments were performed in the DTNSRDC Deep- 
Water Basin [(22 feet (6.7 m) deep, 51 feet (15.5 m) 
wide, and 2600 feet (792 m) long)]. Both the body 
and propeller were supported and towed from Carriage 
II which has a drive system capable of maintaining 
speed to within 0.01 knot. 

Forces were measured on the forward half of an 
ellipsoid of revolution with a length/diameter 
ratio of 5.65. This "half body" was mounted by a 
specially designed strain-gaged flexure assembly 
to the forward end of a massive streamlined after- 
body, attached to the towing carriage by a single 
strut. The propeller was driven by the DTNSRDC 
35-horse-power dynamometer, separately supported 
from the towing carriage and positioned so that 
both the propeller shaft and body axes were aligned 
parallel to the direction of flow as illustrated 
in Figure 5. 

ne half body consisted of a 0.25 inch (0.64 cm) 
thick fiberglass shell measuring 36.0 inches (0.91 
m) in length and 12.75 inches (0.324 m) in maximum 
diameter. The shell was filled with polyester foam 
in order to minimize the mass and obtain a high 
natural frequency, sufficiently above the propeller 
blade rate frequency range to reduce nonlinear 
resonance effects. The aluminum, free-flooded 
afterbody, together with its support strut had a 
low natural frequency to prevent mechanical vibra- 
tions from the propeller dynamometer gears and 
shafts passing through to the body force dynamometer. 
The towing strut was attached to a large frame, 
mounted on the propeller dynamometer structure. 
Slotted pads supporting the frame permitted trans- 
verse and longitudinal adjustment of the body 
location and orientation. Vibration isolating 
mounts were placed in the framework to further in- 
hibit “pass through" vibrations. 

Vibratory forces were measured for two propellers. 
DTNSRDC propeller 4118 is a 3-bladed, 12-inch (0.305 
m) diameter aluminum propeller designed for uniform 
flow. Propeller 4119 is identical to 4118, except 


290 


FIGURE 5. Experimental 
arrangement. 


AFT VIEW 


that it has twice the blade thickness (and a slight 
difference in pitch to correct for the added thick- 
ness). The principal design characteristics of 
the propellers are listed in Table 3. 
were designed by lifting-surface methods and both 
open water performance [Denny (1968) ] and field 
point pressure measurements [Denny (1967)] have 

been reported. It should be noted that the theoret— 
ical predictions of field point pressures agree 

very well with the experimental measurements (at 
design advance coefficient) and the same propeller 
theory is applied in the present surface force 
calculations. 


The Force Dynamometer 


A dynamometer was developed to measure the horizon- 
tal component of the unsteady forces produced on 
the half body by the propeller. The half body is 


cantilevered from the afterbody on five (5) flexures. 


Forces are determined by measuring the strain in 
one flexure, while the other four flexures absorb 
the vertical force and moments as illustrated 
schematically in Figure 6. The measurement flexure 
transmits vertical forces and moments with miminal 


The propellers 


PROFILE VIEW 


stress while resisting a large part of the horizontal 
force (calculated to be over 90 percent). 

Two competing requirements governed the flexure 
design - the need to resolve small forces and the 
desire to maintain the natural frequency of the 
flexure-half body system far above the propeller 
excitation frequency. Also the flexure was expected 
to experience large (static) forces arising from 
flow misalignment and hydrostatic loading. 

From the relationships for stress and stiffness 
of a simple cantilevered beam, it is known that 
for a given force, the flexure should have a low 
stiffness in order to produce maximum strain. This 
in turn would require a small body mass to keep the 
natural frequency high. However, if the body is 
too small, the resulting propeller force signal 
becomes difficult to retrieve in the presence of 
background noise. Although sophisticated techniques 
were employed to reduce electrical noise and boost 
signal power, it was not possible to completely 
eliminate mechanical noise generated by the rumbling 
carriage. With these compromises in mind, the 
flexure was designed for a frequency ratio of 0.5, 
producing minimally acceptable stress levels of 
1000 psi (6.9 uPa) for the one pound (0.454 kg) 
force in this experiment. 


TABLE 3. PROPELLER GEOMETRY 
4118 419 


DIAMETER, INCHES 

NO OF BLADES 

PITCH RATIO (0.7Ro) 
EXPANDED AREA RATIO 


BLADE THICKNESS FRACTION 


NACA MEANLINE 


—SUPPORT FLEXURE (4) 


DIRECTION 0.750 IN. X 0.035 IN. 
OF MEASURED (1.90 CMX 0.084 CM) 
FORCE MEASUREMENT FLEXURE 


O.500IN. X 0.005 IN. 
(1.27 CM X 0.0127 CM.) 


FIGURE 6. Schematic diagram of flexure arrangement. 


For simplicity and economy, the flexure consisted 
of conventional steel shim stock clamped between 
the half body and the afterbody by sets of wedges. 
The flexures were pinned and epoxied to the wedges 
prior to insertion into the dynamometer plate. 
Before assembly, eight strain gages were mounted 
and waterproofed, with one gage placed at each 
corner of the two large faces of the flexure. The 
gages were electrically compensated for tension 
(or compression) and torsion. In order to check 
vertical alignment to the flow, two of the support 
flexures were also strain-gaged. 

Calculations indicated that the measured strain 
in the flexure due to dynamic forces would be 135 
percent of the strain due to a static force with 
the same amplitude, assuming small damping. Also, 
the phase angle of the strain relative to the applied 
force would be affected by the large ratio of 
excitation frequency to the natural frequency. 
Consequently, the experiment incorporated an inter- 
nally mounted electromagnetic voice coil to calibrate 
the measurement flexure as a function of force 
amplitude, frequency, and forward speed. Initally, 
with a series of known static forces applied to the 
body, a current was applied to the coil to return 
the body to its unloaded position, as indicated by 
the strain output from the measurement flexure. 
These static calibrations revealed that the coil 
current varied linearly with applied force and that 
the flexure strain was virtually independent (less 
than 2 percent variation) of the axial location of 
the applied force. 

Dynamic calibrations of the dynamometer were 
performed using a frequency generator and amplifier 
with the known sinusoidal current directly input to 
the coil. (It is assumed that in the low frequency 
range of interest, O to 60 Hz, the applied force 
is independent of frequency). The response amplitude 
(relative to the applied current or force) was 
found to vary linearly with the applied force. By 
averaging the data, the transfer function for each 
frequency and forward speed was determined as shown 
in Figure 7. These results revealed anomalous 
behaviour for frequencies of 20 Hz and 50-60 Hz, 
which were later identified as resonant frequencies 
associated with the towing structure. 


Instrumentation and Data Acquisition 


During each data run the following physical quanti- 
ties were measured (see Figure 8): the force on 
the half body, the surface pressure at two locations 
on the body, the distance between the body and the 
propeller (tip clearance), propeller blade angular 
position and rotation speed, the forward speed of 
the towing carriage, and the horizontal accelerations 
of the afterbody. 

Pressures were measured by metal diaphragm solid- 
state gages (KULITE XTMS-1-190) flush mounted to 


Zo 


the half body surface. The propeller tip clearance 
which varied slightly with forward speed, was 
determined by measuring the distance between the 
35-horsepower dynamometer body and the test after- 
body at two axial positions using linear variable 
differential transformers (Schaevitz 1000 HCD). 
These low friction devices recorded relative move- 
ment without transmitting mechanical vibration. 

The propeller blade angular position and rotation 
speed were measured by a Baldwin Shaft Position 
Encoder mounted on the 35-horsepower dynamometer 
tachometer shaft, generating one interrupt per 
degree of revolution and another interrupt once per 
revolution. During the experiments each data channel 
was sampled for each six degree increment of propel- 
ler rotation, thus providing 20 samples per cycle 
for blade frequency quantities. (The time lag 
between successively sampled channels and the delay 
between the encoder interrupt and capture of the 
sample, together amounting to several degrees of 
rotation, were later accounted for in the data 
reduction). Analog data output from the measurement 
tranducer was digitized and stored on magnetic tape. 
Data for each angular position of the propeller 
were summed and averaged over several hundred 
revolutions in an attempt to reinforce the signal 
of interest while self-cancelling random noise. 

In order to determine the blade-frequency com- 
ponents of the unsteady force (and pressure) on 
the half body, a Fourier analysis was applied to 
the averaged data to yield the coefficients of the 
series 


8) 
35 
F(0) = a + a, cos mO + be sin m0,-T7 < 0 <T 
m=1 
S) 
a 
=O! = 
= 5 + Co cos (md vey) (60) 
m=1 


in which 6(t) is the blade position angle (Figure 
8). For the three-bladed propellers, the nondimen- 


O KNOTS (Om/s) 


ro) 
° 26.6 © O © 


4 KNOTS (4.1m/s) 


ry 00 Oo 
fc) 


6 KNOTS (3.1m/s) 


Ooo V2®o 


TRANSFER FUNCTION 
(~STRAIN GAGE VOLTAGE OUTPUT / COIL CURRENT ) 


8 KNOTS (4.1m/s) 


° 
q 9 © 


“10 20 30 40 50 60 70 
FREQUENCY, Hz 


FIGURE 7. Force dynamometer amplitude response as a 
function of frequency for several forward speeds. 


nN 
oO 
nN 


FIGURE 8. Schematic diagram of 
experiment. 


sional amplitude and phase of the blade frequency 
force F3, are given by 


lF3| 5 5 
is ahs vja3t + b4 (61) 


Cc = 

on2D 
0 Se tea Gye) (62) 
a = e ein 3/a3 


where the phase angle, Ope is the position of the 
reference blade when the force is a positive maximum 
or, from Figure 8, 8, is the angle by which the 
force leads the blade position. 


Experimental Results 


Force measurements with propeller 4118 located 16.0 
in. (6.3 cm) aft of the nose of the body and with 
a nominal tip clearance of 3.0 in. (1.18 cm) are 
given in Figure 9. The force generally increases 
in amplitude and lags further with higher propeller 
loading. The data points at design J (0.83) for 
speeds of 4 and 8 knots show good agreement. In 
Figure 10, the blade frequency pressure induced on 
tne body in the plane of the propeller [x = 16.0 in. 
(6.3 cm)] shows a monotonic increase in amplitude 
with increased propeller loading and repeats well 
for different speeds. 

Force measurements with propellers 4118 and 4119 
positioned 10.0 in (3.94 cm) aft of the nose of 
the body [4.5 inc. (1.77 cm) tip clearance] are 
shown in Figure 11. Over the range of propeller 
advance coefficient, the force amplitude tends to 
increase with increased propeller loading and the 
effect of thickness is demonstrated. 

The data exhibit some scatter for reasons not 


yet fully understood and further calibration experi-_ 


ments and data runs are needed. The variation in 
the data for different speeds (and hence different 
propeller excitation frequencies) is particularly 
disturbing. It may be noted that a post-test 
examination of the raw (unaveraged) data for the 
flexure, displacement, and afterbody accelerometers 
revealed three specific sources of difficulty. 
First, low amplitude data, particularly for speeds 
of 6 knots and a blade frequency of 35 Hz, was 


AFT VIEW IN PROPELLER PLANE 


PRESSURE 
SDF TRANSDUCERS Z 12.75" 


VD 


be AFTERBODY 
PLAN VIEW 


difficult to process. An example of this type of 
run and comparison with a good data run is shown 
in Figure 12. Generally, the low amplitude data 
resulted in force coefficients much below the 
values obtained from the higher amplitude data. 
Second, for certain runs the data were overscale on 
the individual records, but not in the averaged 
plot. These overscales, if abundant, produced 
anomolies. Third, structural resonances of 18-20 
Hz and 55-60 Hz grossly distort data for blade 
frequencies with these values. To the extent 
possible, data contaminated by these problems were 
discarded and are not in the results presented. 


DESIGN ZERO THRUST 
(J = 1.16) 


n 
Ww 
WwW 
oa 
oO 
Ww 
a 
be 
fo} 
lu 
nn 
<= 
Fe 
a 
° 
4.0 
oO 
° 
t= Sol) 
>< Oo 
wu 
= 2.0 CALC. METHOD 
Breslin (1964-1971) 
Vorus (1974) 
1.6 


0.0 0.2 0.4 : Ws 


J = U/nDd 


Ay We ee! 


FIGURE 9. Calculated and measured blade frequency 
force for propeller located at 2 = 16.0 in. with 
tip clearance C = 3.0 in. 


SYM | SPEED 
knots 


PHASE ©, , DEGREES 


N 
ala 
— i= 
a 
i} 
a 
oO 
0.0 : 0.4 0.6 . 4 
J = U/n 
FIGURE 10. Blade frequency induced pressure on body 


with propeller 4118 located at % = 16 in. anda tip 
clearance C = 3.0 in. 


Application of the Theory 


Direct Approach - Extended Lagally Theorem 
(Breslin and Eng, 1965) 


The test body surface was divided into 154 elements 
as shown in Figure 13 with finer subdivisions made 
in way of the nearest approach of the propeller 
blades. Panels 93 through 100 were used to close 
the body. The geometry of these elements, together 
with the normal velocity induced by the propeller 
due to loading and blade thickness formed the input 
to the generalized Hess-Smith program which inverts 
Eq. (33) to yield the source densities on each of 
the panels. 

A typical velocity variation, as given in Figure 
14, shows that, downstream of the propeller, the 
loading contribution is oscillatory, requiring 
great care as the body sections are becoming larger. 
This test case presents a somewhat difficult appli- 
cation of this technique for this reason. In the 
ship case, there is only a small portion of the 
hull downstream of the propeller, and the sections 
are generally becoming smaller. As a result of 
this non-ship arrangement, @ifficulty was encountered 
in securing an accurate answer, requiring several 
adjustments of the size and location of the source 
panels. 

A calculation for a single set of conditions, 
specified by the geometry of DTNSRDC Propeller 4118 
set at a tip clearance of 3.0 inches (1.18 cm) at 
an axial distance of 16.0 inches (6.3 cm) downstream 
of the nose of the body gives a blade-frequency 
force coefficient Cp = 3.4 x10-3 anda phase angle 
Op = -2.0°. These results are quite close to the 
measured values shown in Figure 9. It should be 
remarked that the evaluation included the Lagally 
force corresponding to the integral of the convective 
pressures, i.e., the action of the transverse pro- 
peller velocity component on the sources which 
generate the body in the uniform axial flow. This 
contribution, as expected, is indeed small yielding 


293 


only 1.0 percent of the force arising from the 
time rate of change of the potential. This surely 
justifies the order of magnitude argument given 
earlier. 


Alternative Approach - Oscillatory Body Potential 
(Vorus, 1974) 


In order to apply Eq. (47) to the experimental 
configuration, it is convenient to consider the 
velocity potential Hj, of tne body travelling back- 
wards and executing simple vertical oscillations, 
so that a.=03 =k in Eq. (53). The free surface 
condition Hj, = 0 on z = 0, Eq. (54), can be satis- 
fied by reflecting the body surface into the upper 
half space and satisfying the body boundary condition 
additionally on the image surface, S;- In Appendix 
B it is shown that the vertical force induced by 
the propeller on a ship in the free surface is 
equal to the force on the "double hull" deeply 
submerged. If we make the further assumption that 
the force due to the convective pressure can be 
omitted, the problem for H;, now reduces to 


7H sO amy (63) 
in 
= — 
nav, He = inno (nek) oneSetaS: (64) 
in i 
Vu, +0, {x| >@ (65) 


in 


where ¥ is the whole space outside the “double-hull" 
sumtacel,) Sica 

This method is particularly convenient in the 
present application because the velocity potential 
of an oscillating spheroid is well known, e.g. Lamb 


(1932). With slightly modified notation 
SYM 
DESIGN ZERO THRUST 
(J = 0.83) (J = 1.16) 
hoe: CBW RIE 
rm 
[aa 
enes Rais LO aig & 4119 
md a 4118 
oe 
20 U 
2 fe) 
a= 
a 
-40 
®@ 
4.0 a2 
@ 4119 
8 ‘ 
= 3.0 o> ack 
=e EXP. rs a 


ct 2.04118 - Open Symbols ey avis 
4119 - Solid Symbols a Le 


CALC. 


=—-= Vorus(1974) 


QO Ws Oa MO POs EO sass 
J = U/nD 


FIGURE 11. Calculated and measured blade frequency 
force for propellers 4118 and 4119 located at 2% = 10.0 
in. with tip clearance C = 4.5 in. 


294 


Good Data Run 


W 
v 
Cn 
fe) 
> 
~ 
a 
© 
c 
on) 
“4 
Ww 
) 
1 
u 
fe) 
= High Frequency Contamination 
oer aa | 
ie 
pee Vel des 2 a 
too F = —s = 
FIGURE 12. Examples of force measurement me 20 so 2c 12e «saad «21a 
flexure signal output - data averaged over 
several hundred propeller revolutions. Shaft Angle, Degrees 
H = ae 2 al Vardl Y ; F f : 5 
in (Orbe?) = -inNwe)c9S ,/l - u > on aan any cosf and (x,r,/) is a cylindrical coordinate system with 
Y Bed the origin at the center of the spheroid and the 
(66) major axis extending from x = L/2 to + L/2. The 
5 constants, c] and c2 in (66) are readily determined 
Here y = ,/1 + ¢% and (t,u,P?) are the spheroidal in terms of the spheroid's maximum diameter/length 
coordinates defined by ratio, 6, as 
oo 2 33 2 
BS Say ie WS Sy = ee rhe We, ay? 
. oh Ise. 
with = = 5 , Cy = focal length (67) 
OS Gs ep ot sme 1,30 2) <= an 
a) sg Se 
PROPELLER PLANE 


FIGURE 13. Schematic of expanded surface of 
DTNSRDC ellipsoidal test body divided into 154 
source panels (dimensions in multiples of pro- 
PANELS USED FOR CLOSURE OF BODY peller radius) . 


SRNR TE fo sees 


0.004 


THICKNESS CONTRIBUTION 


REAL PART 


-2.0 -1.0 
IMAGINARY PART (SINE COEF) 


CONTRIBUTION FROM 


MEAN BLADE LOADING Ww 
U 


il 2t1+ 1-621 _ 5, EBS") Vi-62 


c2 2 § §2 


(68) 


and by a suitable coordinate transformation from 
(u,t) to (x,xr), the velocity V Hj, can be calculated 
at an arbitrary point on the propeller blades. 

In general the propeller dipole strength repre- 
senting blade loading is a function of blade position 
O@(t), i.e., Ap = Ap(p,atO(t)). However, in the 
present experiments the inflow to the propeller is 
uniform so that the loading is steady and Ap = Ap 
(9,a). The blades of propellers 4118 and 4119 
employ NACA a = 0.8 meanline sections. For this 
section, and assuming a radially elliptical distri- 
bution of bound circulation, the pressure jump 
across the blade is given by 


8T hoo 


Ap(p,a) = * F(a) (69) 


0.9(ay- a)) ™(R,? - RN 


in which 


295 


20 


FIGURE 14. Variation of blade frequency verti- 
cal velocities induced by 3-bladed DTNSRDC 
propeller 4118 at r = TOR and > = 


and T is the steady propeller thrust. 

The calculated values of the forces produced on 
the spheroid for conditions corresponding to those 
in the experiment are summarized in Table 4 showing 
the separate contributions arising from blade loading 
and thickness as well as the total forces. The 
latter are also displayed in Figures 9 and 11 and 
agree quite well with the measurements. 

Additional parametric calculations were performed 
to study the effect of propeller location on the 
force produced on an ellipsoid arising from propeller 
mean loading and thickness. In Figure 15 the 
attenuation in force (amplitude) with increasing 
tip clearance is illustrated. (The phase was found 
to be essentially independent of tip clearance) . 
Calculations are presented in Figure 16 for a series 
of axial positions of the propeller with the tip 
clearance held fixed. As the propeller is moved 
aft from the nose of the body, the force increases 


TABLE 4. FORCE CALCULATIONS USING METHOD OF VORUS (1974) 


4= 10.0 IN. 
C= 4.5 IN. 


4=10.0 IN 
C= 4.5 IN. 


PROPELLER LOCATION CONTRIBUTION | S, x103 


MEAN LOADING 
THICKNESS 
TOTAL 
MEAN LOADING 0.88 
THICKNESS 1.58 
TOTAL 1.52 
MEAN LOADING 
THICKNESS 
TOTAL 


1.50 
Bou 
2.90 


0.88 
3.16 
2.97 


PROPELLER 4118 
= 16.0 IN 
J = 0.83 


C/R) = 0.25 


IN EXPERIMENT 


FIGURE 15. Modulus of blade-frequency force on ellip- 
soid as a function of propeller tip clearance [calculated 
using method of Vorus (1974)]. 


rapidly, largely due to the thickness contribution. 


CONCLUDING REMARKS 


The analytical methods given in this paper can be 
applied to a wide range of problems in which it is 
desired to determine the unsteady pressures and 
forces generated by a propeller on a nearby boundary. 
The formulation is quite general, being applicable 
to arbitrary hull (and appendage) geometries, and 
propeller locations, geometry, and loading charac- 
teristics. The assumption of high frequency 
propeller excitation, which greatly simplifies the 
treatment of the free surface, is not at all 
restrictive in most cases of practical engineering 
interest. A severe limitation, to be sure, is the 
restriction to subcavitating propellers. However, 
researchers are actively pursuing this subject and 


/ 
aN GaN! 
- L f \i I Ip 4 
Hew ase 
SHO 
' Ny iN pot 
WA Se NS 
C 3 Rees 


MEAN penne 3 


{e) 
0.0 0.25 0.50 0.75 1.0 
AXIAL POSITION OF PROPELLER, */(L/a) 


FIGURE 16. Modulus of lateral blade-frequency force 
produced on an ellipsoid of revolution (L/B = 6.0) as 
a function of propeller axial position (constant tip 
clearance, C/R_ = 0.25)-calculated for DTNSRDC pro- 
peller 4118 using method of Vorus (1974). 


as procedures for predicting transient blade cavity 
geometry and the attendant pressure field become 
available, this important feature can be incorporated 
into the analytical representation of the propeller 
and the analysis of induced forces. 

As with any theoretical development of this kind, 
the usefulness and limitations can only be fully 
ascertained by comparison with a sufficient number 
of experimental measurements. The comparisons 
presented in this paper for the simple case of a 
body of revolution adjacent to a propeller in uniform 
flow represent an encouraging first check. This 
experimental technique can be extended to examine, 
in a systematic manner, the effects of nonuniform 
flow (unsteady blade loading and cavitation) and 
more general body shapes. For example, wire screens 
selected to produce certain wake harmonics can be 
towed upstream of the propeller. At the same time, 
the need is evident to undertake calculations for 
comparison with results of the many experiments 
reported during the past several decades. 


ACKNOWLEDGMENTS 


This work was jointly supported by the American 
Bureau of Shipping (ABS) and the Maritime Adminis-— 
tration (MarAd). The continued interest and 
encouragement by ifr. S. Stiansen and Dr. h. H. Chen 
(ABS) and Mr. R. Falls (MarAd) is greatly acknow- 
ledged. Prior support of the Office of Naval 
Research, Fluid Dynamics Division, enabled the 
development of the velocity field program. The 
authors are also indebted to Mr. D. Valentine and 
Dr. S. Tsakonas of the Division Laboratory for their 
painstaking effort in developing the programs and 
to Messrs. H. Saulant and M. Jeffers (DTNSRDC) for 
invaluable assistance in the design and conduct of 
the experiments. 


REFERENCES 


Breslin, J. P. (1962). Review and Extension of 
Theory for Near-Field Propeller-Induced Vibratory 
Effects, Proceedings Fourth Symposium on Naval 
Hydrodynamics, ACR-92, Office of Naval Research, 
Washington, D. C. 

Breslin, J. P., and K. Eng (1965). A Method for 
Computing Propeller-Induced Vibratory Forces on 
Ships, Proceedings First Conference on Ship 
Vibration, Stevens Institute of Technology, 
Hoboken, Wew Jersey; available as DTMB Report 
2002. 

Cummins, W. E. (1957). The Force and Moment on a 
Body in a Time-Varying Potential Flow, J. Ship 
Research, 1, 1: 7. 

Denny, S. B. (1967). Comparisons of Experimentally 
Determined and Theoretically Predicted Pressures 
in the Vicinity of a Marine Propeller, NSRDC 
Report 2349. 

Denny, S. B. (1968). Cavitation and Open-Water 
Performance Tests of a Series of Propellers 
Designed by Lifting-Surface Methods, NSRDC Report 
2878. 

Hess, J. L., and A. M. O. Smith (1964). Calculation 
of Nonlifting Potential Flow about Arbitrary 
Three-Dimensional Bodies, J. Ship Research, 8, 
a9 BA. 


Jacobs, W. R., and S. Tsakonas (1975). Propeller- 
Induced Velocity Field due to Thickness and 
Loading Effects, J. Ship Research, 19, 1; 44. 

Lamb, H. (1932). Hydrodynamics, Dover Publications, 
6th Edition. 

Lewis, F. M. (1969). Propeller Vibration Forces in 
Single Screw-Ships, Transactions Society of Naval 
Architects and Marine Engineers, 77, 318. 

Lin, W. C. (1974). The Force and Moment on a Twin- 
Hull Ship in a Steady Potential Flow, Proceedings 
Tenth Symposium on Naval Hydrodynamics, ACR-204, 
Office of Naval Research, Washington, D. C. 

Stuntz, G. R., P. C. Pien, W. B. Hinterthan, and 
N. L. Ficken (1960). Series 60 - The Effect of 
Variation in Afterbody Shape upon Resistance, 
Power, Wake Distribution, and Propeller Excited 
Vibratory Forces, Transactions Society of Naval 
Architects and Marine Engineers, 68, 292. 


Tsakonas, S., J. P. Breslin, and W. R. Jacobs (1962). 


The Vibratory Force and Moment Produced by a 
Marine Propeller on a Long Rigid Strip. J. Ship 
Research, 5, 4; 21. 

Tsakonas, S., W. R. Jacobs, and M. R. Ali (1973). 


297 


An "Exact" Linear Lifting-Surface Theory for a 
Marine Propeller in a Nonuniform Flow Field, J. 
Ship Research, 17, 4; 196. 

Valentine, D. T., and F. J. Dashnaw (1975). Highly 
Skewed Propeller for SAN CLEMENTE Ore/Bulk/Oil 
Carrier Design Considerations, Model and Full- 
Scale Evaluation, Proceedings First Ship Technol- 
ogy and Research Symposium, Society of Naval 
Architects and Marine Engineers, Washington, D. C. 

Vorus, W. S. (1971). An Integrated Approach to the 
Determination of Propeller-Generated Vibratory 
Forces Acting on a Ship Hull, Department of 
Naval Architecture and Marine Engineering, 
University of Michigan, Report 072. 

Vorus, W. S. (1974). A Method for Analyzing the 
Propeller-Induced Vibratory Forces Acting on 
the Surface of a Ship Stern. Transactions 
Society of Naval Architects and Marine Engineers, 
BAip AUK 

Vorus, W. S. (1976). Calculation of Propeller- 
Induced Forces, Force Distributions and Pressures; 
Free-Surface Effects, J. Ship Research, 28, 2; 
OME 


APPENDIX A 
THE LAGALLY FORCE ON A FLOATING BODY REPRESENTED 
BY A SURFACE DISTRIBUTION OF SOURCE SINGULARITIES 


The force F, (2) arising from the convective term 
of the linearized unsteady pressure, Eq. (35), is 
given by 


= + 
vot + Vo, n ds (A-1) 


Ss 


where, as before, the symbols ( ) einel ( ye denote 
quantities inside and outside the hull surface, S. 
We assume that the solutions for We and $y are 
known in terms of distributions of source singular- 
ities and images over the surface S as 


bs (x) =- ma (ee face 4 G(x,x')|dsS 
ese? |] fee? a | 


vy 

& ‘S) 

Vz = iU + Voy (A-2) 

i 1 = 1 1 
n(x) = - aa Oy(x"') TSG Toate | Cay op, 
x-x' | x-x' | 

S) 

+ op, (A-3) 

1n 


in which Og (x) and Op (x) are the source singularity 
strengths, x! is the image point of x, and G(x,x') 
is the "wave potential" of a source located at x! 
and is regular in the half plane z < 0. The deriva- 


tives of these functions on eacn side of the surface 
S are related to the source strengths in the form 


<> + a +E > 
Wa = We 7 neg (A-4) 
—_ > > , 
Vo = Von + no, x on S (A-5) 
from which it follows that 
— SS fs (A-6) 
Ye OVO Sg = Wn = Os Ga 
= 


since vt Cin = Wear 0 ® = 0. 

We now apply Green's theorem to the functions 
Vs and Vd, in the closed volume ¥ surrounded by 
the surface S and So, where Sp is the hull water- 
line plane, obtaining 


+ 


=> a >_ 
Viet O- Wisay saVelis) = V(Vg - Von )dv¥ 
+ 
S+S5 @ 
>_ 3S = > _ 
= [IVs > V(Vo, ) + Von * Wg ] d¥ (A-7) 
¥ 
since V x We = Vx Vo, = 0 in ¥. Using Gauss' 
theorem and the fact that V* V, =V-° Vo,- = 0 


in ¥, (A-7) may be written as 


Ah 9 - Vo, n dS = 


S+S, 


ane (Vo Vo" +n) + Von (VE + n)] as (A-8) 


St+So 


and hence 


J 
oe 
5, 
5 
a 
n 
i 
‘ab 
a, 
a 
os 
oa 
5 


s S 
+ V- (Vg » n)] ds 
+ Jf [ve (Vo + n) 
So 
— >_ —— 
Ore Wctuen 1) h Sa Victi Geren aS 


(A-9) 


The last two terms in the integral over S, combine 
to yield 
= aS 


- Vo. WS +n) =o, + Wy 


— 
n(Vg ° Voy ) 
~s -_=> 
(Go, 23 WY) x bn Vo 
(A-10) 

The first term on the righthand side of (A-10) 
vanishes since $n = 0 on So. The second term also 
vanishes, since by Stokes' theorem 


[fe 2s Gee = ¢ dx x $y Vg = 0(A-11) 


iS) Co 


where the contour cy is taken as the hull waterline. 
Consequently, using (A-6), (A-9), and (A-10), the 
expression for the force becomes 


Fy” = p [Ve Woes a) + Vor We + n) 
S 


> — S 
- dg on n] dS + p Ve (Voy * n)dS (A-12) 


So 
— —= 


The contribution from the free stream, iU(in Vs), 
vanishes since 


iU(Vy, * n)dS = iU V7, a¥ = 0 


StS, ¥ 


Also noting that Voy ° n=-o 
(A-11) reduces to 


= (2 ns aS 
F(?) = - p [o, Vos + 65 Voy + og on n] AS 
Ss 
+p Vos Vor «nas (A-13) 
So 
or, upon defining 
vot + Voz 
V6 2 Ss S 
a 2 
Vo+ + Vor 
A — n 
Whey = 2 (A-14) 


(A-13) becomes 


Be sree Ga Vis = Te Vp,)4s 
Ss 
au 165 ane ds (A-15) 
So 


The first term has the same structure as the steady 
flow Lagally force derived by Lin (1974) fora 
linearized source sheet representation of a slender 
strut piercing the free surface. The second term 
arises from the intersection of the hull with the 
free surface in unsteady flow. 

In the low Froude number approximation, 0$</dz 
= 0 on z = O (rigid wall representation of the 
free surface), and G(x,x') > 0. In this case (A-2) 
and (A-3) yield 


— 
2 1 = -x! ees 
Woks = oP = Og (x") + ds 
at > =>/3 ats ESS 3 
DR | |x-x", 
i 
Ss 
and 
—s > 
Ty = OS (x') ES Mees | a 
ha 40 Sines = 5,3 s 3 Ss 
|x-x'| |x-x', 
al 
‘S) 
+V +V 2a-16 
oP, op. ( ) 
n 
for x on S, and where the integrals are to be 
interpreted in the principal value sense. Inserting 


these expressions into (A-15) and performing the 
integrations, the equation for the force reduces to 


Pasay =—- 6) Og (Vop + Von, a's: 
n tn 
S 
ab 
+ 0 Veg —,- a8 (A-17) 


So 


which is the result given as Eq. (42) in the text. 
The reduction in the first term reflects the fact 
that there is no net force arising from the mutual 
interaction of the body sources. 


APPENDIX B 
REDUCTION OF THE ANALYSIS OF PROPELLER INDUCED 
VERTICAL SURFACE FORCE TO AN INFINITE FLUID 
PROBLEM 


The linearized unsteady pressure at a point, x, 
on the ship hull surface is given by (8) as 


espe) = => (9 3 ie an Wis (G39 ved) (B-1) 


and the vertical force acting in the hull, from 
(10), is 


F(t) = - p(x,t) n+ k as (B-2) 


In the high frequency approximation, ¢ = O on the 
free surface, z = 0, and this condition can be 
satisfied by constructing an image of the hull 
surface and a negative image of the propeller in 
the upper half space and allowing the fluid domain 
to extend to infinity in all directions. The 
negative image propeller is identical to the propel- 
ler proper, but rotates in the opposite direction 
and the signs of the dipole singularities represen- 
tating the effects of loading and thickness are 
reversed from those of their images in the lower 
half space. 

The image hull surface, S;, is identical geomet— 
rically to S, but the signs of the singularities 
on Sj required to diffract the unsteady flow from 
the "two propellers" will be reversed from these 
on S due to the symmetry. The magnitudes of the 
Singularities at image points will be equal. 

The steady flow about the bare hull, Vc, in the 
low Froude number approximation will satisfy the 
rigid wall free surface condition Vs * k = 0. In 
this case, the steadily moving hull can be reflected 
into the upper half plane with a positive image 
singularity system, i.e., the singularities on the 
image surface, S; will be of the same sign as the 
singularities on S to diffract the velocity iU. 

Because of the assumed linearity, the unsteady 
potential may therefore be considered as the sum of 
contributions from the propeller and hull and their 
respective images. 


O = Oe © Oe Pe Ong 2 One (B-3) 
where 
bt) = — bys Gy, t) 
— oN 
bp (x,t) = - py (xy ,t) 
— 
So SOS pip) (B-4) 


ab 


for all (x,y,z) outside the surface § + Sz: If we 
define bpy = bp + by then it follows that 


(x,t) = opy (xt) - dp, Gg,t) all ¥ (B-5) 


Therefore, the complete unsteady potential in the 
fluid beneath the zero potential free surface can 
be obtained entirely from consideration of the 
propeller and the double-hull in an infinite fluid. 

The unsteady pressure at a point on the hull 
surface S is now given by 


ab ie) Eas —_> — 
at Gant) ee Vg (x) . Vopy (x,t) 


p(x,t) =- p 


299 


f) 
Bc TNE mel ig. | | 
at (x5, ) g (x) gua | 
9 => = 
Now if Vg = (U + Ugr Vor Wo), the symmetry of Vo 


is such that Uc and vg are even in z, while weg is 


odd in z. It follows that 

a — eS => 

Vg (x) Vobpy (X47 t) = Vo (x, ) Vopy (x47) 
and hence 

a => 
p(x,t) = Ppy(x,t) - Ppy(x;,t) (B-6) 
in which 
ap 
PH = 
Bay = 2 Fe ge ° We 


Thus, the unsteady pressure at points on the hull 
can be obtained from calculations, or measurements, 
of pressures at image points on the double-hull, 
with the double-hull and propeller deeply submerged. 

Turning now to the formula (B-2) for the vertical 
force, we obtain 


F(t) = - J [renee - k ds 
iS) 


> 


+ Pip (44 7t) 0 - k ds (B-7) 


s 


_> => => 
But since n(x) * k = —n (x; ) - k, (B-7) may be 
written as 


Pe (Ve) Ss Dip (Xr t) n (x) +k das 
Ss 
—2) — i => 
= Pryp (Xj ot) n(x, ) * k ds (B-8) 
Ss 


or, since the image hull S; is geometrically iden- 
tical to the hull proper, 


—>) > 
aoe) S = Pyp et) n> k ds 


S+S; 
and consequently the unsteady vertical force on the 
hull can be obtained from force calculations, or 
force measurements, using the double model and 
propeller deeply submerged. 


A Determination of the Free Air Content 

and Velocity in Front of the “Sydney-Express” 
Propeller in Connection with Pressure 
Fluctuation Measurements 


A. P. Keller 


Technical University Munich 


and 


E. A. Weitendorf 


University of Hamburg, 


Federal Republic of Germany 


ABSTRACT 


The Special Research Pool within the Institut ftir 
Schiffbau and the Hamburg Shipmodel Basin (HSVA) in 
collaboration with the Technical University Munich 
and Det norske Veritas executed extensive full-scale 
measurements on the Single-Screw Container Ship 
"Sydney-Express." The main task of the project was 
the determination of the free air content of the 
seawater in front of the propeller during the voyage 
from Australia to Europe. 

Simultaneously the velocity was measured at the 
control point within the Laser-beam, where the free 
air content was measured by the scattered light 
technique. Additional investigations were a deter- 
mination of the water-quality, high speed films and 
sterophotography of the cavitation at the blade, 
and pressure fluctuation measurements above the 
propeller. 


ils INTRODUCTION 


For several years the dynamic behaviour of small 

gas bubbles or nuclei in hydrodynamic pressure 
fields has been recognized as an important influence 
on cavitation inception and its extent. Besides 
other scale effects in the field of model propeller 
testing, the importance of this influence of nuclei, 
which also effects propeller excited pressure 
fluctuation measurements, was often underestimated 
and neglected. Thus, for instance, the results by 
van Oossanen and van der Kooy (1973) have shown 

that for equal non-dimensional flow conditions but 
different absolute revolutions (i.e. n = 20 and 

n = 30 Hz) the non-dimensional propeller excited 
pressure amplitudes were different. After the 
development by Keller (1973) of a practicable laser- 
scattered-light (LSL) method for measuring the 
undissolved air content, systematic cavitation and 
pressure fluctuation measurements were carried out 


300 


in the medium cavitation tunnel of the Hamburg Ship 
Model Basin (HSVA) with the model propeller of the 
"Sydney Express" [Keller and Weitendorf (1975) ]. 

The results were similar to those by van Oossanen 
and van der Kooy. Due to the additional application 
of the (LSL) technique, the differences of the 
nondimensional pressure amplitudes for different 
revolutions could be clearly explained by the 
influence of the free air content or nuclei on the 
cavitation. A further finding was that the non- 
dimensional pressure amplitudes and the cavitation 
for a revolution of n = 15 Hz were increasing with 
growing free air content, whereas the cavitation 
and those amplitudes for n = 30 Hz remained more 

or less constant. The different behaviour for 

n = 15 Hz and n = 30 Hz were explained by Isay and 
Lederer (1976, 1977). Using the theory of bubble 
dynamics they found that the reactions of the 
bubbles on the respective pressure gradient of the 
propeller blades at n = 15 or n = 30 Hz were differ- 
ent. Further, these investigations led to criteria 
of cavitation similarity of such a kind that the 
number of nuclei per unit volume of the model flow 
had to be increased compared with the number of 
nuclei of the full scale flow. By geosim tests 

with hydrofoils or propellers it should be determined 
to what extent these additional criteria for cavi- 
tation similarity are applicable. 

Keeping in mind these physical connections, the 
full scale trials on the container ship "Sydney 
Express" were planned. These investigations were 
the first attempt to measure the nuclei distribution 
in seawater around a ship by means of the LSL 
technique. The nuclei distribution could serve as 
a basic value for the geosim tests and perhaps as a 
comparative standard value of the water quality for 
model cavitation investigations. Furthermore, the 
experiences, made during the almost adventurous 
measurements on the "Sydney Express" with the LSL 
technique in front of a full scale propeller, could 
be of common interest because the introduction of 


optical laser methods is a promising tool in the 
research fields of boundary layers and propeller 
flows. 

The additional investigations on the "Sydney 
Express" help in full-scale model correlation only 
slightly; the main purpose of these measurements 
was the securing and better interpretation of the 
scattered light results. The following additional 
measurements were performed: 

1. Propeller-excited pressure fluctuation 
measurements with six pressure pick-ups above 
the propeller. 

2. Cavitation observations for determination 
of the thickness and extent of the cavity by 
means of stereo photography. 

3. Investigations of water-quality by means of 
a simple scattered light method (Aminco- 
colorimeter) for detecting suspended particles 
and total air content by means of a Van-Slyke- 
apparatus. For both measurements water 
samples were taken. 

4. Velocity measurements in the control volume 
of the scattered light measurement in order 
to estimate the bubble concentration. 

The "Sydney Express", as one of the fastest 
German single screw merchant ships, was chosen for 
the investigations because its propeller has an 
interesting cavitation extent. 


2. SHIP DATA AND PREPARATION OF THE MEASUREMENTS 


The single screw, turbine-driven ship "Sydney Express" 
has been built by Messrs. Blohm and Voss AG, Hamburg 
(No. 872) and belongs to the so-called second 
generation of container ships. 

The main data of the ship are given in Table 1: 


TABLE 1 - "Sydney Express" - Data 
Ship Data 

Length b.p. L = 210.00 m 
Breadth, moulded BPP = 30.50 m 
Design Draft D = 121.00 m 
Block coefficient c = 0.616 
Displacement (Design) V 43,457 m° 
Container about 1,600 
Max. Power Pp 23,870 kw 
Service Speed Ve = 22.0 kn 

Propeller Data 
Diameter Dp = 7.00 m 
Pitch (mean) Ph 6.550 m 
Blade-Area-Ratio Ae/Ag = 0.78 
Number of Blades Z = 5 


The necessary conversions of the ship construction 
for the installation of the measuring devices in 
the after peak of the ship were carried out at the 
Hapag-Lloyd ship yard at Bremerhaven during the 
latter part of September 1977. Figure 1 shows 
allusively to what an extent the narrow steel 
construction had to be cut free. The installation 
of three windows for the reception of the scattered 
laser light proved to be the most complicated of all 


301 


installations. For reasons of the ship's safety 
and also to enable proper cleaning these windows 
were pushed through 350 mm sluice valves together 
with their tubular guide pipes. The windows, of 
which only that opening was marked in Figure 1 
which had been used for measurements, were arranged 
between frames 12 and 13. Also, the fitting of the 
three 350 mm sluice valves required skillful impro- 
visation on the spot. The installations of the 
sluice valves for the pressure pick-ups, dimensioned 
in Figure 1, and of the cavitation observation 
windows were carried out without any difficulties. 
In addition, all electric lines were laid out 
from the measuring pick-ups to the measuring con- 
tainer during this period. The necessary amplifiers, 
digital magnetic tape recorders, and computer (HP 
2100 A) with its peripheral equipment were located 
in this measuring container. The measuring container 
was located in hole 6 directly on the tank deck of 
the after peak, in the last bay. For the determi- 
nation of the performance data of the ship, strain 
gauges were attached to the shaft. In addition, 
the shipborne electro-magnetic log (system Plath) 
for determination of the ship's speed was connected 
to the computer via an isolation amplifier. Thus, 
the ship's speed and power could also be recorded 
at each pressure fluctuation- and LSL-measurement. 
A recalibration of the log was made on the outward 
voyage in the North Sea by means of a speed measure- 
ment carried out by the Hamburg Ship Model Basin 
using their method with a resistance log. 


Bo PROPELLER EXCITED PRESSURE FLUCTUATIONS AND 
CAVITATION OBSERVATIONS 


The measurements of the propeller excited pressure 
fluctuations were started on the outward voyage 
when leaving the English Channel and continued until 
the arrival at Marseille (Tests No. 1-11). Further 


details on these measurements as well as for the 
pressure fluctuation measurements carried out in 
13-16) 


the Mediterranean (Tests No. 


Tables 2a and 2 b. 


are given in 


Laser-Beam 


Photo - 


Propeller— multiplier Laser 


Plane 


Cavitation - 
©) observation 


N 
© .U 
o—- 


Windows 


me 


815 + 815 


1255 a4 
Oo 


a nu 
Ov 


Pressure 
pick-ups 


FIGURE 1. Locations of test setups. 


., Laserbeam 


1) 
Frame 137 l 
12 cS 
iN 
Control - ) 
Volume 
1450 VX 
Propeller - 
circle 
rame 
6 8 10 13 ] 


| Photom. Laser 
Lo] 


Longitudinal 
cross section 


Venu -Tip 


FIGURE 2. Arrangement of test setups. - 


The results of the pressure fluctuation measure- 
ments for the Tests No. 1-4, 11 and 13-16 are given 
in Figure 3 showing the dimensionless pressure 
amplitudes of the blade frequency for the pressure 
pick-ups Pl, P3, P4, and P6. They have been 
harmonically analysed on the HP-computer in the 
measuring container. As usual with right-hand 
propellers the pressure pick-up on the starboard 
Side (here: P3) clearly shows higher values than 
that on the port side (P4). Figure 4 shows the 
amplitudes measured by these two pressure pick-ups 
up to the 15th harmonic. The harmonic analysis has 
been carried out for a "representative" revolution, 
resulting from the average of 60 propeller revolu- 
tions. 


Figure 5 shows the pressure fluctuations measure- 


ments versus propeller rpm for the pick-ups P3 and 
P4 for two drafts applied during the voyage in the 
Indian Ocean. At this point in time the propeller 


was already damaged. Further data of these measure- 


ment runs can be found in the Tables 3a to 3h. 
Examples of the results of harmonic. analyses up to 
the 15th harmonic order for the pressure pick-ups 
P3 and P4 are shown in Figure 6. In Figure 7 a 
comparison is given of the pressure amplitudes of 
these harmonic orders for the pick-up P4 (port) in 
shallow and deep water. In shallow water the 
pressure amplitudes are only slightly higher (5.8%) 
than that in deep water. With the pick-up P3 
(starboard) the difference was even smaller (1.0% 
increase) . 


TEST NO | 1-4|11 [13-16 
SYMBOL | = | 0 | 


is) 


TEST NO 
SYMBOL 


0.04; 


0.03 


0.02 


80 90 


100 —»n [RPM] 
Undamaged Propeller 
Mediterranean 


FIGURE 3. Pressure fluctuations. 


The lower pressure amplitudes of the blade 
frequency in the Indian Ocean (Figures 5 and 6) 
compared with that in the Mediterranean (Figures 
3 and 4) are to be attributed to a significantly 
stronger, but mainly stationary cavitation of the 
damaged blade (No. 3). A comparison between Figures 
4 and © shows that due to the damage the pressure 
amplitudes of the "not-blade-number" frequencies 
have been strongly increased in opposition to the 
blade frequency. It should be noted that the ship 
superstructure vibrated strongly after the propeller 
had been damaged. This damage resulted from a ground- 


Pressure Pick-Up P3 (St-B) 


Pressure Pick-Up P4 (Port) 


1 5 (0 15 


TEST NO 15; 17-10-1977 
n = 100.4 RPM; V, =21.3Kn 


Harmonic Order n 


FIGURE 4. Harmonic components of pressure fluctuations. 


303 


Table 2a 

Test No. iy. 2 3 4 1] 

Date ———— 9.10.1977 ————— 125 No 77 

Speed V, [kn] 17.2 18.6 19.8 21.4 20.7 

Revolution n_ [RPM] 89.1 94.3 98.8 104.7 100.4 

Power Pp [MW 11.4 13.0 14.6 16.8 14.9 

Draft aft [m 8.94 8.94 8.94 8.94 8.33 

Draft forward [m 6.35 6.35 6.35 6.35 6.96 

Course 209° 231° 2 Bie 230° 37° 

Sea region —— English Channel ——— Medit. 

Wind [Beauf | 4 4 4 4 (6) 

Wind direction 180° 180° 180° 180° ) 

Water Depth [m] 36 35 41 51 1040 

Table 2b Measurements in the Mediterranean 

Test No. 13 14 15 16 

Date ——— 17.10.1977 

Speed Vs [kn] 18.6 19.5 Dio) DD,,3) 

Revolution n_ [RPM| BONS) 98n2) MOOG 105.1 

Power Pp [MW Ook 12.2 15a Ws) 

Draft aft [m O99 Bold Dats 9.73 

Draft forward [nm] O63 9s08 9463) 9.63 

Course 11s? MA . 1ae 114° 

Sea region — 36°45'N; 18°49'E (Mediterr.) 

Wind [Beauf] = | 806 

Wind direction ——_———_ 90° 

Water Depth [m] = 3500 =— 

Table 3a Measurements in the Indian Ocean 

Test No. 47 59 60 61 62 65 

Date 30.11.77 1.12.77 —_—— Dol2ovT 

Speed V, [kn] Die? 21.4 2S Dilo8) DNS Di33 

Revolution n [RPM] 101.1 101.6 101.2 101.8 101.0 101.3 

Power Pp [Mw Nod = = —= = —= 

Draft aft [m 9.30 —————— _ 9.30 ——__=— 9.30 

Draft forward [m] 7.620 — 7.62 — 7.62 

Course 294° — 294° — 2gi” 

Sea region or 1a 1S Onende Oe 9°09'S; 

aasieien Hones | Ta Ne 20 88 Te San CATE 

Wind [Beauf | 6 = 3 — 2 = 3) 

Wind direction 100° ——__$_—_ 70°. —__—_=— 230° 

Water Depth [m| 3300 ——__—— 4900 —————= _ 2000 

Table 3b Measurements in the Indian Ocean 

Test No. 70 7\ 72 

Date ————— 4.12.77 ———— 

Speed V, [kn] 21.8 22.1 21.8 

Revolution n_ [RPM] 101.7 101.9 101.3 

Power Pp [MW = = =e 

Draft aft [m — Qo S7 — 

Draft forward |m —_ 8.08 — 

Course — 314° — 

Sea region or position = 2°58'N; 59°44'— ——=— 

Wind [Beauf] = 142 — 

Wind direction en BRO 

Water Depth [] ——__§_—— 3250 ———_=— 

Table 3c Measurements in the Indian Ocean 

Test No. 73 74 75 76 77 78 

Date — A V2o07 —_ 

speed V, [kn] Dito fl 20.9 — zee 16.9 

Revolution n [RP] OT 2 96.8 95).4 92.6 B2I9) 82.5 

Power P, [MW — 11.6 

Draft aft [m —- 9.37 — 

Draft forward [m) — 8.08 — 

Course =— 314° = 

Sea maglon Of POSa —————— 2°58'N; 59°44°—  ———————e— 

Wind [Beauf] = 32 —_ 

Wind direction = 235° — 
3250) — 


Water Depth {m] 


Table 3d 
Test No. 
Date 


Speed V. [kn] 
Revolution n [RPM] 
Power Pp [MW 
Draft aft [m] 
Draft forward [m] 
Course 

Sea region or pos. 
Wind [Beauf] 

Wind direction 
Water Depth [m] 


Table 3e 
Test No. 
Date 


Speed V [kn] 
Rexoilwe ton n [RPM] 
Power Pp [MW 
Draft aft [m 
Draft forward [m] 
Course 

Sea region or pos. 
Wind [Beauf] 

Wind direction 
Water Depth [m] 


Table 3f 
Test No. 
Date 


Speed V, [kn] 
Revolution n [RPM] 
Power Pp) [MW 

Draft aft [m} 
Draft forward {m] 
Course 


2°58'N;59°44'E; 


Measurements in the Indian Ocean 


79 80 81 
—_—_____—_ 4.12.77 
11.9 11.9 Nite’) 
5 OFZ 60.1 61.3 
7 Io8 = 

7 9.37 9.75 
08 8.08 7.82 
—_____—_——._ 314° 


——____—__ ] +2 


2252 


3250 3250 3250 


82 83 

Noe 12.0 
0.9 61.3 
ors) oS) 
7.82 7.82 


_———— 


—— 3°15'N359°27'E —— 


————————— ee 


a 


3250 3250 


Measurements in the Indian Ocean 


84 85 86 87 88 89 
as 1D 5.12.77 => 
17.5 17.5 17.5 19.7 20.4 20.4 
85.3 85.1 85.2 95.4 96.5 96.3 

13.9 — — 
= 9.75 = 
oe 7.82 _ 
a 8) eee 
SBIR SOP OTN — 
= te? = 
— 230° — 
— 3250 —— 


Sea region or position 


Wind [Beauf] 
Wind direction 
Water Depth {m] 


Table 3g 
Test No. 
Date 


Speed Vg [kn] 
Revolution n_ [RPM] 
Power Pp [MW 
Draft aft [m 
Draft forward [m] 
Course 

Sea region of pos. 
Wind [Beauf] 

Wind direction 
Water Depth [m] 


Table 3h 
Test No. 
Date 


Speed V, [kn] 
Revolution n [RPM] 
Power Pp oa 
Draft aft [m 
Draft forward [m] 
Course 

Sea region or pos. 
Wind [Beauf] 

Wind direction 
Water Depth [ml 


Measurements in the Indian Ocean 


90 91 92 
SIZ 
—_> 21.6 = 
100.8 101.3 101.3 
———_———- 9.75 eee 
= 7.82 — 

) 


SS SG 


8 


15 9Ng SOS. —— 


= Ori 
= 23009 —_—_—> 
— 3250 


Measurements in the Indian Ocean 


93 94 95 96 97 
ae 95>.12.77— 
Des 21.4 21.2 Die? Diath 
101.0 101.0 101.0 101.2 101.2 
== = = — DDD 
— 8.63 Som 
= 8.23 =— 
<i ie) Fee 
——————._ 8°28'N; 54°40'E ————— 


es 4 


at SS ES 


SS OOO 


Measurements in 


99 100 10] 
G26 
22.4 22.2 22.3 
103.1 102.3 102.8 
16.9 — 
= 8 
= 8. 
— 273° 

12°21'N; 47°03'E 
ee ee 


2. 


si 


the Gulf of Aden 


103 104 105 
——— Zoe 
22.3 DP <3} 225) 

101.8 101.9 101. 

17.0 = — 
2730, = 

Bab-el-Mandab 
37 38 40 


304 


{Kes ae Pressure Pick-Up P3 (St.-B) 
0.03 


TEST NO |70-80| 81-92 
. + 
0.02 


0.01 
0 
Pressure Pick-Up P4 (Port) 
0.03 
TEST NO 70-80) 81-92 
SYMBOL 
0.02 
0.0 
0 
60 70 80 90 100 Nn 
[RPM ] 


Damaged Propeller 
Indian Ocean, 4-12-1977 


FIGURE 5. Pressure fluctuations during laser-scattered- 
light (LSL)-measurement. 


ing due to a thunderstorm at the entrance of the 
Suez Channel. The cavitation of the damaged blade 
was so strong that it existed during the total pro- 
peller revolution. This could be seen through the 
cavitation observation windows. Unfortunately, no 
photographies were made because the measuring crew 
of Det Norske Veritas carrying out the cavitation 
observations left the ship in Port Said. 

In the Mediterranean, however, a large number 
(about 800) of black-white photographs of the 
undamaged propeller were made with the equipment 
of Det Norske Veritas with stroboscopic lighting. 
Since pictures were always taken with two Hasselblad 
cameras it might be possible to carry out stereo- 


Pressure Pick-Up P3 (St-B) 


Pressure Pick-Up P4 (Port) 


1 5 ae Re 15 
TEST NO 70 rmonic Order n 
Damaged Propeller V, = 21.8 Kn 
Indian - Ocean n =101.7RPM 


FIGURE 6. Harmonic components of pressure fluctuations 
during LSL-measurement. 


305 


Deep Water with Low Nuclei Content 
Test No 99: V, =22.4Kn; n= 103.1RPM 


Shallow Water with High Nuclei Content 
Test No 105: V.=22.5Kn;n = 101.7RPM 


1 5 10 —» 15 


Pressure Pick-Up P4 (Port) 
Damaged Propeller, 
Indian Ocean 


Harmonic Order n 


FIGURE 7. Harmonic components of pressure fluctuations. 


metric measurements of the cavitation layers in 
dependence of the blade positions. As an example 

for the cavitation extension of n = 105 rpma 
collection of photographs is shown in Figure 8. 

These pictures were made with a camera with a fisheye- 
objective. The photographed condition belongs to 

Test No. 16. 


4. INVESTIGATION OF THE WATER QUALITY 
Measurements of Suspended Particles 


In addition to nuclei measurements, which will be 
described later, the content of suspended particles 
was investigated as often as possible. This was 
necessary for two reasons: the LSL-method does 

not allow direct differentiation between solid and 
gaseous particles. Thus it became necessary to 
estimate the proportion of dirt or organic particles 
(probably contained in the water) in the measured 
nuclei sprectra. For these investigations a 
scattered-light instrument (nephelometer) was used; 
the J4-7439 fluoro-colorimeter of the American Instru- 
ment Company (Aminco). The Aminco-scattered-light 
instrument works on almost the same physical princi- 
ple as the LSL instrument. Water samples of 1 cm, 
investigated in the Aminco-colorimeter under a 
scattered light angle of 90° were exposed to a green 
light (514 nm) as in the laser control volume. The 
geographical positions where the Aminco scattered 
light measurements were carried out (as well as all 
the other measurements described in this report) 

are shown in Figure 9. 

The results of the Aminco scattered light investi- 
gations, given in Figure 10, were obtained in the 
following way: 

Water samples were taken from the condenser in- 
flow of the ship's turbine during the voyage. One 
part of this water was poured through a filter with 
a pore size of 0.4 um. Another part was used for 
unfiltered samples, which previously were roughly 
degassed by stirring and shaking. Subsequently, 
the unfiltered and filtered samples were investigated 


306 


KDrsesth 1 15°stb. 


Ta =9.75m 
Gy =0.22 


FIGURE 8. Cavitation "Sydney-Expess." 


in the Aminco-colorimeter. The deflection of the 
meter for the filtered sample was adjusted on the 
indicating scale to "0", which served as reference 
value. Measured values of unfiltered samples are 
shown in Figure 10; Relative Intensity is an 
arbitrary unit. 

The first measurements, at the end of October, 
were made with a one-hole-aperture in the beam 
path, the following ones with a four-hole-aperture 
due to a thereby increased intensity. 

In order to obtain a general idea of the sensi- 
tivity of the Aminco scattered light method, 
standard solutions were produced using the plastic 
spheres also used for the calibration of the LSL- 
instrument. It is apparent from this that five 
parts per cm? with a diameter of D = 25.7 um could 
still be measured. 

Many results from investigations of sea water 
did not show any difference between filtered and 
unfiltered samples. The content of suspended 
particles was thus very small in the Indian Ocean; 
it was below the response level of the Aminco-device. 
The samples taken on the 7th December 1977 contained, 
however, suspended particles. They descended from 
the shallow water region of the Bab-el-Mandab at 
the entrance to the Red Sea. 

The lack of knowledge about the back scattering 
qualities of the particles appears to be a problem 
when applying this scattered light method with the 


£A.\ 20°Stb. 


f 


HAL 25°stb. 


FAL 30°s tb. 


Veg =22.3kn; n=105RPM 
P,=17. IMW, Jkq=0.69 


Aminco-colorimeter. A more expanded and intensive 
investigations of suspended particles, for instance, 
with coulter counter, could not be carried out within 
the frame of this research work. 


¥ Aminco - Colorimeter 
T Total Air Content 


+ Scattered Laser Light M. 
® Velocity Measurements 


FIGURE 9. Positions of measurements. 


0 1,0 2,0 3,0 40 5,0 6,0 7,0 Relative Intensity 


1 n 4 


307 


iL 
Seawater, 27. 10.77 (One-hole aperture) 


Seawater 28. 10.77 

: 31. 10.77 

2.11.77 

2.11.77, 20 naut. miles before Fremantle 

Swan-River 3.11.77; Fremantle 

7.77 

11.11.77, Swanson Dock West 
7, White Ba’ 


212.77 
612.77 (Gulf of Aden) 

612.77 

[_]Seawater 712.77 (Shallow water, Bab-el-Mandab) 


71277 (Red Sea) 
Seawater 10.12.77 Great Bitter Lake 
11.12.77 (Mediterranean, 15 naut.miles behind Port Said) 
Seawater 12.12.77 (lonian Sea) 


Measurements of Total Air Content 


Although the water should always just be saturated 
at the surface, the gas concentration c_ of the 
sea water was also continuously determined from 
water samples with a Van-Slyke-apparatus. The 
results are given in Figure 11, in dependence of 
the temperature. 

For the calculation of the gas content ratio, 
Ee = c/e , the gas saturation capacity, c_, is 
necesSary for the specific salt content and temper- 
ature. Since the corresponding data were not known 
some water samples were left overnight in a basin 
with a large surface and the gas concentration c 
was determined on the following day, which in this 
case should indeed correspond to the saturation 
concentration c_. The two values obtained for the 
saturation concentration c_ are also plotted in 
Figure 11 (with the symbol -O- ). They are within 
the range of tolerance of the measured total gas 
content, Co, for the voyage leading southward. 
Subsequently the measured total gas content present 
values which correspond to the gas content ratio, 
€ = 1, i.e., to saturated water. Due to the 
dissolved salt the total gas content values, Co, 
for sea water should lie below the values for fresh 


water. This is, in fact, the case with the exception 


of some values of the voyage leading northward. It 
must be left to other investigations to find out 
whether the wind, seaway, and temperature "history" 
of the sea surface has an influence on the total 
gas content. 


5. MEASUREMENTS OF THE NUCLEI SPECTRA AND LOCAL 
VELOCITY 
Device for Nuclei Measurement 


The LSL method was applied to the measurement of 
the nuclei spectra in front of the "Sydney Express" 


Standard-Solution, D = 25,7 um; Cpart = 1,0 - 10 n/cm3 2810.77 
St=Sol D=25,7 um; Cpa = 54 nicm3 31.10.77 cea for particle size 
St-Sol D=25,7 um; Cport = 5 nicm? 1.1.77 D= 25,7 wm an 
St=Sol D= 1011 wm; Coon = 176-108 niem3 2311.77 WA7Ret Int 
aes 610 2311.77 
Cpart = 176-108 n/cm? 23.11.77 = a a vale a ah 
St.-Sol D = 1,011 pum; Cpart = 1,76 102 n/cm3 23.11.77 


ae} 39 
(ey Rel. Int. 


a DPV 7 Ny ; same sample after 6 hours) 


FIGURE 10. Measured suspended particles. 


propeller. This method was also applied to the 
model tests, described by Keller and Weitendorf 
(1975). Detailed information about the measure- 
ment principle has been supplied, for instance, by 
Keller (1970, 1973). Thus, it is not necessary 

to go into the details. 

Compared with previous measurements carried out 
in the laboratory the measuring distances were 
essentially larger at these full scale investigations. 
Thus, some new components for the measuring device 
were required. The distance between the measuring 
volume and the receiving lens amounted up to 2 m so 
that the laser power and the diameter of the 
receiving lens had to be markedly increased, in 
order to obtain usuable measuring signals. 

The arrangement of the measuring unit on board 
is shown in Figures 1 and 2. The path of the laser 
beam is bent three times and enters the water almost 
horizontally; the path of the beam of the receiving 
system is bent once and proceeds in the water 
vertically. With this arrangement the flow direction 


Co 
[foo] 


20 Se 


(O59) aso= Saturation of air in pure water (70 mm Hg ) 
Measured total air content in seawater southbound voyage 
Measured total air content in seawater northbound voyage 
Saturation of air in seawater 

2 Port Phillip Bay (Melbourne, 7. 11.1977 ) 


10 15 20 25 We 


FIGURE 11. Measured total air-content. 


308 


and the direction of the laser beam as well as the 
optical axis of the scattered light receiving system 


are standing vertically, one upon another. This is 
optimal for the measuring technique used. 
The homogenization of the laser beam, i.e. the 


conversion of the Gaussian intensity distribution 
over the beam cross section into a rectangular 
distribution, was made with a special filter. The 
homogenous intensity distribution as well as the 
shape of the laser beam (square or rectangular) 
were maintained quite well by the very long focal 
length of the laser system (about 6 m). 

The control volume, optically defined, was 
positioned in such a way that the stream line 
through the control volume came into the range of 
the propeller tip. The position of the control 
volume in front of the propeller was determined by 
the position of the reception window of the scattered 
light between frames 12 and 13, i.e., 4.2 m in front 
of the propeller plane. The additional geometrical 
fixing of the control volume in the vertical direction 
resulted from the laser window (located between 
frames 13 and 14) with its horizontal beam outlet 
into the water. Subsequently the positions for 
the control volume was fixed as follows: 90 cm of 
the ship's hull vertically downward and 145 cm from 
midship on the port side between frames 12 and 13 
(see Figure 2). 


The Calibration Device 


The relationship between the photomultiplier impulse 
amplitude and the size of nuclei was determined by 
a calibration with latex spheres. For this purpose 
a special device was put through an opening in the 
ship's hull when the ship stopped in calm water. 
With this device it was possible to maneuver a 
fine nozzle near to the control volume and to inject 
the latex spheres into the control volume. The 
apparatus was operated by means of small hydraulic 
elements from the inside of the ship (Figure 12). 
For the calibration latex spheres of 45 and 25 
Um were used. The corresponding photomultiplier 
impulse amplitudes fit excellently to the theoretical 
curve of the scattered light intensity. The measuring 
range was set to 8-117 Um for the nuclei diameter. 
In addition to the scattered light intensity, 
the dimensions of the control volume were important 
data for the determination of nuclei spectra and 
nuclei concentration. Since a direct measurement 
or calculation of the cross section of the laser 
beam in the control volume was not possible in this 
case, a new method had to be applied to determine 
the laser beam dimensions. By means of the above 
mentioned hydraulic device a small rotating wheel 
with thin platinum wires was adjusted in such a 
way that the wires cut the laser beam vertically 
at the location of the control volume. Thus the 
light in the direction of the photomultiplier was 
scattered. The dimensions could then be determined 
from the width of the photomultiplier impulses, the 
distance between the axis of the small wheel, and 
the light point on the small platinum wires 
(determined by crossed platinum wires) and the 
revolution number of the small wheel. The diameter 
(25 um) of the platinum wire had also to be con- 
sidered. The exact knowledge of the control volume 
dimensions was also important for the measurements 
of the local velocity, as described below. 
The dimension of the control volume in the longi- 


Sluice Valve 
Micro - Hydraulic Device 


6mm Pipes flexible 
6m long 


Pipe 10cm (Length 2.5m ) 


Seawater 


FIGURE 12. Calibration device and arrangement. 


tudinal direction of the laser beam was adjusted as 
usual through the measuring slit in front of the 
photomultiplier, after the enlargement factor of 
the reception optic was determined. This again was 
done by means of the hydraulic device with which 

an object, whose dimensions were known, was placed 
in the control volume; its picture was measured in 
the plane of the measuring slit. 

For the nuclei measurement the dimensions of the 
control volume were then fixed as follows: 0.86 mm 
x 0.86 mm X 1.33 mm = 0.98 mm®. The cross section 
of the control volume rectangular to the flow 
direction amounted to 0.86 mm X 1.33 mm = 1.14mm’. 
This detail was required for the determination of 
the nuclei concentration. 


Measurement of Local Velocity 


When the cross section of the control volume and 

the number of the nuclei measured per unit time, 
were known, it was necessary in addition to know 

the local flow velocity at the control volume in 
order to determine the nuclei concentration. Since 
the conversion of model test results from wake field 
measurements to full scale appeared to be too in- 
accurate for the determination of the local velocity 
and because the measurement of the velocity with a 
Prandtl tube, for instance, was not possible, a 

new method was applied to measure the velocity and 


flow direction. If the dimensions of the control 
volume in the flow direction are known, the velocity 
can be determined from the measured impulse width. 
In order to estimate the local flow direction at 

the control volume, an aperture is put into the 
beam path of the laser (Figure 13), which gives 

the laser ray a rectangular shape. This aperture 

is turned until the photomultiplier impulses have 
reached a maximum. Then it is possible to determine 
from the position of the aperture the position of 
the plane, formed by the flow direction and the 
laser beam. If now the measuring slit is turned 
until the half width of the distribution of the 
impulse width has reached a minimum, it is possible 
to read - from the position of the measuring slit - 
the plane which is formed by the flow direction 

and the optical axis of the reception system. The 
flow direction in the volume results from the inter- 
section of the two determined planes; the impulse 
width gives the flow velocity, and the impulse 
width spectra provides information on the degree 

of the turbulence flow. 


In this way flow characteristics can be determined, 


undisturbedly and locally, with one measurement; 
otherwise they could only be determined with a 
three-component measurement. Furthermore, the 
control volume simultaneously reaches the optimum 
inclination for the measurement of the nuclei size. 
Thus one signal provides data about the distribution 
of nuclei size and about the flow field. 

Large particles or bubbles require a longer 
period to completely cross the control volume than 
smaller particles at the same speed. This means 
that besides the larger impulse amplitude there is 
also a larger impulse width. These facts have to 
be considered in the measurement of the velocity. 
Therefore, a single-channel discriminator is inserted 
into the impulse processing electronics. The 
discriminator choses for the measurement only 
impulses of the amplitude or a strongly limited 
range of amplitudes. Thus it is possible to draw 
a clear conclusion from the measured impulse width 
on the speed of the particles in the control volume. 

The new technique to measure the velocity is 
illustrated in the Appendix. A rectangular beam 


cross section whose breadth is the vertical to the 
flow direction, has proved to be the optimum for 
the measurement of velocity and the determination 
of the flow direction. 


4 Spectral Filter 


309 


General Remarks 


Originally it was planned to shift the height of 
the measuring point on the optical axis of the 
reception system by different laser beam directions. 
In addition, this axis should be shifted laterally 
through two additional observation windows between 
the frames 12 and 13. This would make it possible 
to measure at several points in the plane between 
the frames 12 and 13. Unfortunately, this could 
not be realized due to lack of time, because the 
installation of the measuring equipment at the 
beginning of the voyage had taken too much time. 

It is not intended to describe all the diffi- 
culties which occurred at the installation of the 
equipment. The problem of vibration, however, must 
be mentioned. 

To protect the laser, vibration damping should 
be guaranteed as far as possible. It was, however, 
observed during the outward voyage that the pneu- 
matic vibration isolation, which had a resonant 
frequency of £, = 1.8 up to 3.0 Hz, could not be 
used, - even if the exciting blade frequency of the 
propeller was within the range of 8 and 9 Hz. 
Excitations occurred, of course, also at a propeller 
speed of f = 1.8 Hz and due to seaway frequencies. 
When it was obvious that different damper devices 
also did not help, the support, on which the laser 
and the photomultiplier were fixed, had to be 
stiffly connected with the steel construction of 
the after peak. This labor and the laser adjust- 
ments required more than half the time of the voyage 
to Australia during difficult climatic conditions. 
The laser adjustment was carried out mainly when 
the ship was stopped. The calibration of the 
nuclei impulses and the determination of the control 
volume, in which the nuclei were measured, were 
also carried out during these periods. These were 
kindly granted by the captain and his officers and 
had to be regarded as a special concession since 
the "Sydney Express" was on a fixed schedule. In 
this connection it must also be mentioned that the 
calibrations and later the measurements, made on 
the return voyage, could only be carried out after 
dark. For this reason, extra maneuvering watches 
had to be set in the engine control room, usually 
while the ship had a "16-hours-unattended-machinery- 
space". 

The above mentioned stiff support solved the 


1 Laser 5 Grey Wedge Filter 
2 Beam Expander 6 Rectangular Aperture 
3 Aperture 7 Lens 


Idealized Photomultiplier 
signal from scattering 


(ets objects. 
A Idealized signals with grey 
wedge filter(5) in rays. 


(eee Slope indicates direction. 


8 Microscope Lens 
9 Flow Section 

10 Control Volume 
11 Receiving Lenses 
12 Measuring Slit 
13 Photomultiplier 


FIGURE 13. Principle of LSL-measurements. 


310 


vibration problem almost completely. It provided, 
however, the risk that the laser might fail. 
Fortunately, this did not happen. The laser, a 
Coherent-Radiation (4 Watt) product, achieved the 
same performance (900 mW), to which it was adjusted 
at the beginning, up to the end of the voyage with- 
out any failure. The small vibration still observed 
at the measuring point had no significant effect. 


6. RESULTS OF MEASUREMENTS 
Local Velocity 


It was mentioned already that the number of measure- 
ments originally planned could not be carried out 
due to lack of time. Thus, for instance, the size 
and the direction of the local velocity could only 
be determined at one measuring point. This measure- 
ment took much time since there was no special 
electronic device available. It was the first 
measurement of this kind and it was included in the 
program at a late date, which made it impossible 

to establish a special measurement before the 
departure. The measurement was, therefore, partly 
performed with the electronic device which was also 
used for the scattered light measurements, and with 
some special interfaces. 

The velocity and one plane of the flow direction 
at the place of the control volume could be deter- 
mined for one velocity. At the ship's speed of 
22 kn the velocity at the measuring point amounted 
to 7.22 m/s and the direction was found at an angle 
of 5° downward. The corresponding result from the 
model test for the geometrically corresponding 
position of the "Sydney Express" amounted to 7.47 
m/s. This model test, however, was carried out for 
the propeller plane of the towed model, without a 
running propeller. - In full sale, on the other 
hand, the plane formed by the flow direction and 
by the optical axis of the reception device could 
not be determined due to lack of time. 


[eal 


Relative Abudance —= = 


0) 50 100 150 —» 200 
T [asec | 
FIGURE 14. Pulse width distribution and mean pulse 


width dependent on the inclination of the flow. 


NU Test 47 
cm? Measuring Time: 
100 try = 6-8 sec 
= 3 
1.0 €, = 37 N/cm 
0.1 n =101.1RPM 
0.01 Vs =21.2kn 
0.001 
0 20 40 60 80 100 ~m 
— Diameter 
30 -11 -1977; Wind Force : 6 Beauf. 
FIGURE 15. Nuclei distribution. 


The ratio, local velocity to ship's speed, 7.22 
m/s to 11.32 m/s, and which corresponds to the 
local wake in the control volume for the ship at 
22 kn, was applied for all nuclei concentration 
measurements. The nuclei concentration was then 
calculated from the recorded ship's velocity, the 
measuring period, and the measuring cross section. 

Figure 14 shows examples of the velocity measure- 
ment and also the change of the impulse width 
distribution for the rotation of the rectangular 
laser aperture. The value i = 0° corresponds to 
the horizontal plane. At 5° downward (A = -5°) the 
mean impulse width, evaluated on the HP-computer, 
reaches its minimum at 59.6 Usec. The large half- 
width of the distribution curve results from the 
turbulent flow. With a laminar flow the distribution 
curve would be smaller. (See Figures A 2.2 and 
IX AoS)o 

On the basis of these measurements a quantitative 
statement about the turbulent degree of the flow 
cannot yet be made. On the one hand we have no 
experience with this measuring technique, on the 
other hand the ratio, length to width of the laser 
beam cross-section, was too small at this measure- 
ment (2:1). At high turbulent flow the corners of 
the beam cross-section were dispersed by a relatively 
high amount of nuclei which resulted in shorter 
photomultiplier impulses than with nuclei running 
through the middle of the beam. A higher ratio, 
length to width, would be more favourable. 

The first practical experiences with this mea- 
suring technique are so promising that its further 
development is being promoted. The advantages which 
this measuring procedure offers in connection with 
the determination of the size of nuclei are quite 
remarkable. 


Nuclei Spectra 


About one third of the spectra obtained between 30 
November and 7 December 1977 are demonstrated in 
Figures 15 through 24. The spectra contain the 
respective sum of nuclei per cm? for the respective 
range of diameters. In the diagrams one range of 
diameter is marked by a horizontal line. The single 
ranges of diameters do not have the same width. 

The dissimilarity of these spectra, which obviously 
results from different conditions, will later be 
described in detail. 

First, it has to be noticed that for all spectra 
in the range of a bubble diameter from 20 to 40 um 
(micron) there is either a relative maximum or an 
absolute maximum of nuclei. The relative maximum 


tea Test 59 
10 Meas. Time : 
; tm = 19.3 sec 
0.1 8 = 13 N/cm3 
0.01 n =101.6RPM 
Vs =21.4kn 


0 20 40 60 80 100 Hm 
—~= Diameter 


1-12-1977 Wind Force 3 Beauf. 


Test 65 
Measuring Time : 
tm = 1.59 sec 
B= 159 N/cm?3 


n=101.3RPM 
V=21.3 kn 


0 20 40 60 80 100 Am 
—= Diameter 
2-12-1977; Wind Force: 2-3 Beauf., Heavy Swell 


FIGURE 16. Nuclei distributions in seaways. 


was detected in Test 47 - Figure 15, Tests 60 and 
62 - Figure 17 and Tests 90 to 92 - Figure 20. 

The absolute maximum was detected in Test 61 - 
Figure 17 and Test 65 - Figure 16. The strong 
fluctuation of the number of nuclei per em? (nuclei 
concentration €9) can be read from the diagrams. 
Figures 15 to 17 as well as 18 and 19 show spectra 
which have been measured in different seaways. 
During the performance of Tests 47 to 62 (Figures 
15 to 17) there was a seaway and swell from astern. 
The strong pitching motions of the "Sydney Express" 


ae Test 60 
Measuring Time: 
tm = 17.03 sec 


0.1 b= 15N/cm3 
0.01 n=101.2RPM 
Vez 21.3 kn 
0.001 
10.0 Test 61 
Measuring Time: 
10 t py = 1-16 sec 
01 5 =219 N/cem? 
n =101.8RPM 
Co Ve= 21-3kn 
0.001 
Test 62 
1.0 Measuring Time: 
tr = 12.33 sec 
041 i——| €)=21N/cm? 
n=101.0RPM 
a9) s= 21.3kn 
0.001 


0 20 40 60 80 100 ~§=©120 Bm 
1-12-1977; —= Diameter 
Wind Force: 3 Beaufort with Swell 


FIGURE 17. Nuclei distributions in a seaway. 


Shia 


€5=15N/cm3 
1 9 
Cell 2000 17 21 25 No of Rev 
Cell 0 — Time 
Test 60 ; Measuring Time: t,,=17.03 sec 
225 
Cell 2000 2933 37 No of Rev 
Cell 0 


Test 61; Measuring Time : tm= 1-16 sec 
€5=219N/cm3 


FIGURE 18. Nuclei distributions in a seaway. 


resulted in a strong fluctuation of the nuclei 
concentration, i.e. from Co = 21 N/em? 10) (Gi) = AILS) 
N/cm*? in the Tests 62 and 61 - Figure 17. Depending 
on the quality of the water, either swarms of 
bubbles or clear water, which hits the control 
volume of the laser beam, 2000 nuclei were counted 
more or less quickly. The process of counting the 
2000 bubbles is demonstrated in Figure 25 for two 
cases. There the analog-output voltage of the 
memory is plotted against time and propeller revo- 
lutions respectively. The output voltage of 1 Volt 
is reached by the memory when its 2000 cells are 
filled. In Test 61, 2000 bubbles were counted 
within 1.16 s and in Test 60 within 17.03 s at an 
almost linear processing of the output voltage. 
Figures 18 and 19 show also a series of nuclei 
spectra measured, one immediately after the other, 
at a seaway of Beaufort 4. At this series the 
direction of the seaway was, however, athwartships 
up to "slightly from fore". The seaway motions of 
the "Sydney Express" (length 210 m) were very 
small in this case. Subsequently the number Co of 
nuclei per cm? was higher than the smallest number 


N 
lon? Test 93 
10.0 Meas. Time: 
tm = 412 sec. 
Lo -| €5=62N/em3 
0.1 n= 101.0RPM 
Ve=21.3k 
0.01 s : 
0.001 
Test 94 
1.0 Meas. Time : 
tm = 5.63 sec 
0.1 |__| § = 45N/cm? 
n =101.0RPM 
0.01 Vg = 21-4 kn 
0.001 
10.0 | Test 95 
0 Meas. Time: 
1. tm =4.20sec 
01 b= 61N/cm3 
n =1010RPM 
0-01 Vs =21.2kn 
0.001 
0 20 440 60 80 100m 


—e Diameter 
5-12-1977; Wind Force: 4 Beaufort 


FIGURE 19. Nuclei distributions in a seaway. 


N 
ae Test 96 
10 Meas. Time: 
tripsisgoisec 
04 = 43N/cm? 
0.01 n =101.2RPM 
Ve =21.2 kn 
0.001 
Test 97 
1.0 Meas. Time : 
tm =6.5 sec 
0-1 by =39N/cm3 


n = 101.2RPM 
Ve = 21-4kn 


0 20 40 60 80 100 wm 
—e= Diameter 


5-12-1977 Wind Force: 4 Beaufort 


FIGURE 20. Nuclei distribution for calm water. 


with the pitching ship but still within the narrow 

range of To = 39 up to To = 62 N/com?. Considering 

the Tests 47 to 65 (Figures 15 to 17) and the Tests 
93 to 97 (Figures 18 and 19) it can be said that in 
a seaway the nuclei concentration G9 is higher than 
in smooth water, and further, that the influence of 
shipmotions on the concentration Cg superimpose on 

the influence of the seaway. 

The measurement series carried out with the 
Tests 90 and 92 (Figure 20) under ideal weather 
conditions, show on one hand the good repetitive 
accuracy of the results for a constant speed in 
good weather. In this case the number of nuclei 
per em? amounts to So = 18, 19, and 18 N/cem? at 
constant measuring periods of 13.6, 13.2, and 13.9 
for 2000 nuclei. This, however, shows - on the 
other hand - a clearly lower concentration, Co, 
than with the Tests 61 and 65 in a seaway (Figures 
17 and 16). 


lens Test 90 
Measuring Time: 
try = 13.6 sec 
b= 18 N/cm? 


0.01 n =100.8RPM 
Vg= 21.6 kn 
0.001 
Test 91 
1.0 + Meas. Time: 
try = 13.2 sec 
0.1 Se 19N/cm3 
0.01 n =101.3RPM 
V,= 21-6 kn 
0.001 
Test 92 
1.0 Meas. Time: 
tm = 13.9 sec 
0.1 “69 =18N/cm? 
0.01 n =101.3RPM 
Vo =21.6 kn 


0 20 40 60 80 100 §=9120 Am 
—» Diameter 


4-12-1977 


FIGURE 21. Nuclei distributions for different ship 
speeds. 


N 
(5 Test 70 
10 Measuring Time: 
tm=13.55sec 


ca = 18N/cm?3 


n =101.7RPM 
Ve =21.8 kn 


0.001 
Test 77 
1.0 t———; Meas. Time : 
tm =11.5 sec 
0.1 = bo =27 N/cm3 
0.01 n =82.9RPM 
Ve =17.1kn 
0.001 
10.0 Test 79 
1.0 Meas. Time: 
tr =3.3sec 
0.1 6, =137 N/cm? 
n =59.2RPM 
eo Vg =11-9kn 
0.001 


0 20 40 60 80 100 um 
—=Diameter 


4-12-1977, Draft aft Dg=9-37m; Draft forw. D¢ =8:08m 


FIGURE 22. Nuclei distributions for different ship 
speeds. 


A further difference, previously mentioned, has 


to be noticed when comparing measurements in a 


seaway and in calm weather. Whilst with typical 
spectra in a seaway (Tests 61 and 65) the absolute 
maximum is between the nuclei sizes 30 and 40 um, 
it can be detected for calm weather in the smallest 
measured, nuclei range. This phenomenon will be 


described later (in Section 7). The measuring 
series of different speeds for two drafts are 


as Test 90 
1.0 Meas. Time: 
tm =13.6 sec 
01 € 5 =18N/cm3 


n =100.8RPM 
Ve=21-6 kn 


0.001 
Test 85 
1.0 Meas. Time : 
ta = 11.3 sec 


bo = 28N/cm3 


n =85.1RPM 
Ve=17-5kn 


0.001 
10.0 Test 83 
| Meas. Time: 
1. tm= 3.2 sec 


b= 140 N/cem3 


n =61.3RPM 
V2 12.0kn 


(in 20) AOMGON MN GOMMNLIOO) rn 
—= Diameter 
4-12-1977; Draf aft Dq=9-75m; Draft forw.D¢= 782m 


FIGURE 23. Nuclei distributions on deep water. 


N 
ae Test 99 
1.0 Measuring Time: 
tm = 20.04sec 
01 9 =12N/cm3 
0.01 n =103.1RPM 
= 22.4kn 
0.001 Vs 
i Test 100 
0 Meas. Time : 
try = 16.90 sec 
01 ===) 
5 = 14 N/cm3 
0.01 t——-- 
n = 102.3RPM 
0.001 —| Vs = 22.2 kn 
Test 101 
velo Meas. Time : 
04 tm = 22.06sec 
C= tN /cm? 


n =102.8RPM 
Ve = 22.3kn 


0 20 40 60 80 100 um 
—e Diameter 


6-12-1977; Gulf of Aden 


FIGURE 24. Nuclei and particle distributions on 
shallow water. 


shown in Figures 21 and 22. In the second case the 
"Sydney Express" was ballasted with 5,160 tons of 
water. In both series it should be noted that with 
decreasing speed the number T 9 of nuclei per cm? 
increases. At the lowest speed of ca. Vg= 12 kn 
the bubble range of a diameter between 20 and 40 

Um contains the absolute maximum number of bubbles. 
The differences between the two cases are, however, 


{ N 
‘cms 
104 Test 103 
Meas. Time: 
1.0 tm = 1-56sec 
= 3 
O1 b= 155 N/cm 
n = 101.8RPM 
0.01 Vg = 22.3 kn 
0.001 
10. 
oo Test 104 
1.0 Meas. Time: 
tm = 1.77 sec 
0.1 bo = 137 N/cm3 
n =101.9RPM 
oo Vg = 22.3 kn 
0.001 
Test 105 
10.0 Meas. Time : 
10 tr = 0.98 sec 
Go =270 N/cm3 
o n =101.7RPM 
0.01 Vs =22.5 kn 
0.001 
9 2 2 WW Go >P) |[/Aun] 


7-12-1977 ; Bab-el-Mandab 


FIGURE 25. Analog output of memory. 


313 


small. The differences between the drafts were 
obviously not sufficient to provide stronger differ- 
ences between the nuclei spectra. 

The two measurement series shown in Figures 23 
and 24 were made under ideal weather conditions, 
the one 7 hours later than the other. The spectra 
from Figure 23 were obtained in deep water in the 
Gulf of Aden; the spectra shown in Figure 24 were 
obtained from shallow water at the entrance of the 
Red Sea at Bab-el-Mandab. With these two series 
it was intended to clarify the point that the 
propeller excited vibrations which occur on shallow 
water result (apart from the shallow water effect) 
to a higher extent from a stronger instationary 
cavitation, which arises on occount of an increased 
nuclei concentration in shallow water. It must be 
said that this question could not be answered. On 
the other hand a comparison of these two measurement 
series shows that the number of nuclei per cm? 
(nuclei concentration Co) increases from a fo of 
11 to 14 N/cm? in deep water to a Z) of 155 to 270 
N/cm*® in shallow water. This will be described in 
the following Section. The absolute maximum of Tp 
is here again in the range between the nuclei 
diameter of 20 to 40 um. In this connection it 
should be noted that hardly any nuclei with a 
diameter of above 60 Um were detected. 


7. DISCUSSION OF THE RESULTS OF NUCLEI SPECTRA 
AND COMPARISON WITH OTHER INVESTIGATIONS 


Simultaneously with the nuclei measurements in 
shallow water - Figure 24 - water samples have been 
taken. The results of the tests carried out with 
these water samples with the Aminco scattered light 
device appear in Figure 10. These samples from the 
shallow water region at Bab-el-Mandab showed a 
Relative Intensity of 0.4 for the difference between 
unfiltered and filtered water. Even after six 

hours the unfiltered sample still showed a Relative 
Intensity of 0.28. From this it can be concluded 
that the suspended particles, existing at this 
coastal strip, settled in the samples within six 
hours. From this Aminco scattered light measurement 
it can further be concluded that the high nuclei 
concentration shown by the LSL measurement - of the 
shallow water measurement series, Figure 24 - results 
mainly from suspended particles. There were probably 
also solid particles concerned (it is likely to be 
sand at the coast of Arabia) which show no inclusion 
of gas. This is assumed because the cavitation did 
not increase in the shallow water. The corresponding 
propeller excited pressure fluctuations in deep and 
in shallow water show practically no difference, 
Figure 7. 

In Figure 26 the results of the laser-scattered- 
light technique and the Aminco scattered light 
measurement for investigations in shallow water 
(Test 105) are shown together. Figures 27 and 28 
(in the diagrams marked with "Sydney Express") show 
further results of the Aminco-scattered-light 
measurements and the LSL measurements. In the 
Aminco scattered light investigations the differences 
between unfiltered and filtered water were equal 
to zero [A(Rel.-Int.) = 0] in these cases. This 
means that the concentrations of the suspended 
particles were imperceptibly small; they were in 
any case below the response level of the device. 

In each top diagram of Figures 27 and 28, results 
of the investigations of suspended particles from 


“ Sydn.- Expr.“ 
n=101.7RPM — Test No 105 


Latit. Longit. | Date 
12° 28'N | 43° 55'E | 7.12.77 


Aminco-Sc.-Light : 
A(Rel.- Int) = 0.4 


20 40 60 80 100 “m D+ 


Shallow Water 
Bab-el-Mandab 


FIGURE 26. LSL-technique compared with Aminco- 
Sc.-Light measurement. 


comparable locations are shown which were carried 
out on the occasion of the Indian Ocean expedition 
of the "Meteor". For the investigations, which 
have been made by Krey et al. (1971), the so-called 
inverted microscope and the Zeiss particle-counter 
were used. These results lie always one magnitude 
above the "Sydney Express" measurements for the 
operating revolution (n = 101 rpm). In case of the 
low revolution number of n = 60 rpm the nuclei 
concentration measured at the "Sydney Express" 
expedition (To = 15 N/cm?) reaches the values from 
the "Meteor" expedition in the range 20 - 35 um 

and exceeds in the range 35 - 92 um. Since, further- 
more, the water sample tests carried out with the 
Aminco scattered light device do not show any 
difference between filtered and unfiltered water 
(medium diagram - Figure 27) it is justified to 
state that with the LSL measurement mainly bubbles 
were recorded. The investigation of Keller et al. 
(1974) of the optical qualities of the latex spheres, 
applied for the calibration, supports this fact. 
According to his investigation the latex spheres 


“Meteor’’ Suspended 
1964/65 Particles 


Latit. | Longit. | Date 


| 2°06'N[57°53'E [2.2.65 | 


Station 179 


“Sydn-Expr."’ Test No 79 
Latit. | Longit. | Date 
2°58'N | 59°44'E | 4.12.77 

n=60RPM 


Aminco-Sc- Light: 
A(Rel-Int.) =0 


Test No70 
n=101.7 RPM 


Latit.| Longit. | Date 
Like Test No 79| 4.12.77 


Aminco -Sc-Light : 
A(Rel.-Int.) =0 


" Sydn.- Expr." 


0 20 40 60 80 100 “wm D—. 
Deep Water 
FIGURE 27. LSL-technique compared with other 


investigations. 


Meteor " 
1964/65 


Suspended 
Particles 


[ Latit. | Longit. | Date 
12°43'N | 48°32'E |17. 12.64 


13 Station 93 


" Sydn.- Expr.” 
n=103.1RPM _ Test No99 
Longit. 


12°21'N | 47°03'E | 6.12.77 
Aminco-Sc-Light 


A(Rel-Int.) =0 
Deep Water 
Gulf of Aden 
FIGURE 28. LSL-technique compared with other 


investigations. 


show scattering characteristics similar to the 
bubbles. Therefore, it can be said that the 
sensitivity of the LSL measurements is - toa 
certain extent - adjusted to the scattering 


behavior of bubbles via the calibration. With the 
LSL technique mainly bubbles are measured whose 
number is always smaller than that of all solid 

and gaseous nuclei. It is known, for instance, 

that silica algae are almost transparent. It is, 
therefore, understandable that there must exist 
differences between the LSL method on the one hand 
and the microscope method (with coloration perhaps) 
and the conductivity measurement with Coulter 
Counter on the other hand. The assumption that, 
with the LSL method, mainly bubbles are measured is 
supported by the good conformity of the LSL method 
with the holographic method of an ITTC-comparison 
measuring, Peterson et al. (1975). In this investi- 
gation a holographic method, the laser scattering 
light method, and a microscope method have been com- 
pared with each other. The first two methods agreed 


well with each other in the range of the bubble sizes 


20-40 um, whilst the microscope method also showed 
a nuclei concentration higher by one order of 
magnitude. The higher concentration of nuclei 
according to the microscope method apparently 
results from mistakes arising from the focusing of 
the nuclei. Similar difficulties might also occur 
with the inverted microscope applied at the "Meteor" 
expedition. This argument, however, does not say 
that the highest nuclei concentration of the "Sydney 
Express"-investigation, frequently occurring in 

the smallest ranges of size, results from bubbles 
only. (See, for instance, Test 70 - Figure 27 and 
Test 99 - Figure 28 or all diagrams of Figure 20). 
In the class of the smallest size nuclei solid 
particles which always exist in the sea water have 
certainly also been measured. 

Oceanographic studies with the Coulter Counter, 
for instance, carried out in the Gulf of California 
by Zeitzschel (1970) show a strong increase in the 
number of particles with a diameter of 14 to 4 um. 
In addition, Zeitzschel cites the size distribution 
of particulate carbon in the Indian Ocean by means 
of fractional filtration investigated by Mullin 


--e- Gordon (1970 ) 
Microsc., Organic matter 
Surface Atlantic 


—>— Carder et al.( 1971) 
Coulter counter 
Surface Pacific 


“Sydney - Express ” 


--- Test 79,n=60RPM 
= — et 10; n=1017RPM 


100 =um 
Diameter 


FIGURE 29. LSL-technique compared with other 
investigations. 


(1965). In Mullin's report the following average 
percentage in the different size categories for 

near surface samples (15 m) are given: 500 - 350 

um: 3%; 350 - 125 um: 5%; 125 - 95 um: 4%; 95 - 60 
um: 6%; 60 - 33 um: 6%; 33 - 10 um: 18%; and 10 - 1 
um: 58%. The content of organic carbon can amount 
to 4.5 - 34% of the particulate matter in the 
different regions of the oceans [see Zeitzschel 
(1970)]. Zeitzschel continues: "It can be concluded 
from the results obtained at the Gulf of California 
and the above mentioned references that small par- 
ticles, mainly in the range from 1 to 10 um in diam- 
eter, predominate in offshore surface waters of the 
oceans." Investigations by Gordon (1970) and Carder 
et al. (1971), which are compared with our results in 
Figure 29, revealed the same results. It is obvious 
that the "Sydney Express" results - ending at a 
diameter of 10 to 20 um for reasons of intensity - 
would probably show strongly increasing particle 
numbers below this range. This can be seen from 

the results of Gordon (1970) and Carder et al. 

(1971) which have been published by Jerlov and 
Nielsen (1974). 

The fact that a large number of small particles 
in sea water show every arbitrary geometrical shape 
(according to Zeitzschel) also reminds one of the 
shapes of particles from the water of a cavitation 
tunnel, shown by Peterson et al. (1975) - Figure 6. 

These sea water particles of different shapes 
(diameter 1 to 10 Um), which according to Figure 29 
are always available in a high concentration can 
easily nucleate cavitation, as we know from many 
investigations [(e.g., Peterson (1972) and Keller 
(1973) ]. 

The problem of the difference between real shapes 
of the nuclei, detected by the laser beam in the 
sea water and the diameters evaluated for the 
measuring results can only be mentioned here. In 
this connection one should remember that the cali- 
brations on the "Sydney Express" were performed 
with latex spheres, whereas the real shape of the 
nuclei in the seawater is unknown. This problem 
also arises with the Aminco-method and with the 
Coulter Counter measurements, the latter working, 
however, according to the conductivity principle. 


35 


A further uncertainty is probably included in 
the comparison of results obtained from oceanographic 
studies carried out with water samples from the 
open sea and those obtained from laser scattered 
light measurements carried out in the flow and in 
the boundary layer of the ship. The low-pressure 
area of the boundary layer with its vortices of 
different size most likely have a great influence 
on the conversion of pore nuclei into bubbles when 
they are moved from the calm free sea through the 
boundary layer of the ship and thereby increase. 
Due to the long running-time along the ship's hull 
diffusion will also have an effect. 

These physical processes accompanying the growing 
of the bubbles in the low-pressure areas of the 
boundary layers and the effect of diffusion could 
be the explanation for the fact that the lower 
speeds (12 kn, Tests 79 and 83) show a larger 
bubble concentration Gg (due to the long running- 
time along the ship's hull) than the higher speeds 
(21.6 - 21.8 kn, Tests 70 and 90) with a shorter 
running-time. (See measuring series with different 
speeds - Figures 21 and 22). Thus - at a ship's 
speed of about 60 rpm - a characteristic size of 
bubbles has been formed. The measurements in a 
seaway (Tests 61 and 65 - Figures 17 and 16) show 
similar characteristic sizes of bubbles between 20 
and 30 um. In a seaway the turbulence is larger 
due to wave and ship motions. According to Sevik 
and Park (1973) the turbulence can lead to character- 
istic bubble sizes in connection with the pressure 
history. 

All considerations concerning bubble sizes must 
finally lead to those bubbles participating in the 
cavitation process. According to the calculations 
by Isay and Lederer (1977), small bubbles, which 
can also arise from pore nuclei, will grow faster 
than big ones (Figure 30). The result of such 


Distance from 
suction side 
Y= 0-005 Greatest 


6-03 bubble Ro, 


ug = 95 zm) 


ad 
2 
k 
om 


| - 


Smallest Bubble = Ro1= 54m 
Py= 1 kp/cm?2 


Smallest Bubble Ro j=5 am 


=0.2 
gees, 


0.004 0.01 0.02 0.04 0.1 0.2 
Chordlenght c=2A 


FIGURE 30. Calculated growth of a single bubble in 
a hydrofoil flow [Isay and Lederer (1977)]. 


316 


calculations is valid for a hydrofoil of length 

c = 10 cm, wherein the pressure distribution was 
calculated by means of the profile theory for 
incompressible flow with the completion of shock 
pressures caused by the compressibility of the 
water. With these calculations one question re- 
mains unsolved: Up to which negative values can 
the local pressures on the profile really decrease 
in natural water? On the full scale propeller of 
the "Sydney Express" the local pressure gradients 
are probably steep and reach negative pressures, 
causing bubbles with a diameter of 10 Um, or less, 
to cavitate. Regarding the measurements, bubbles 
with diameters of about 10 lim to 20 um were still 
recorded in the results from Test 47 (Figure 15) up 
to Test 65 (Figure 16). For unknown reasons, how- 
ever, from Test No. 65 on nuclei with a diameter 
of less than 20 um frequently could not be measured. 
On the other hand one has to consider that, the 
smaller the nuclei concentration U9 becomes, the 
smaller the bubbles enlarged by cavitation. 

It is apparent from these remarks that it would 
have been desirable to record bubbles or nuclei 
with a diameter below 5 um. But this was impossible 
even with a 4 Watt laser which delivers 900 mW on 
the green line. Therefore, it has to be admitted 
that not all bubbles, which possibly are partici- 
pating in the cavitation process, could be detected. 
The question arises whether this will be possible 
without any doubt in the future and if it is 
necessary or not. Also the following aspects would 
have to be considered: the required laser intensity 
is limited; the exact local pressure distribution 
on the propeller blades is difficult to determine 
and on the other hand the tensile stress that can 
actually be supported by the sea-water is quanti- 
tatively unknown. : 

Before closing this paragraph a personal impres- 
sion in connection with the bubble sizes should be 
mentioned which is supported by the collection of 
photographs in Figure 8 and by numerous additional 
pictures and propeller observations on the "Sydney 
Express": The propeller will always find in the 
flow a sufficient number of small nuclei leading 
to cavitation. Therefore, the fullscale cavitation 
will always be more stable than the model cavitation 
with its smaller negative pressures and its different 
nuclei distribution. 

The white foam on the cavitation pictures of the 
full-scale propeller clearly indicates a large 
number of nuclei, which have led to cavitation and 
grown together. 


8. SUMMARY 


The comprehensive laser scattered light measure- 
ments on the "Sydney Express" showed the following: 
1. The nuclei spectra measured in a seaway in 
the Indian Ocean are quite different: In the range 

of the nuclei diameter of 20 - 40 um either a 
relative or an an absolute maximum of nuclei was 
measured. (Figure 16). The motions of the ship, 
especially the pitching motion, are in this con- 
nection as decisive as the wave motiens on the sea 
surface (Figures 18 and 19). The nuclei of this 
range (diameter: 20 to 40 um), consist of bubbles, 
since the scattered light method, carried out at 
the same time with the Aminco-colorimeter did not 
show any difference between unfiltered and filtered 
water. 


2. In good weather conditions the absolute 
maximum of the bubbles with a diameter between 20 
and 40 um (Figure 20) disappears. The nuclei of 
smallest diameter show the largest nuclei concen- 
tration. It probably consists of bubbles and 
suspended particles, as the comparison with micro- 
scope- and Coulter Counter measurements has shown. 

3. Measurements made at different speeds 
(Figures 21 and 22) have again resulted in an 
absolute maximum at a diameter of 20 to 40 um for 
the smallest ship speed at 12 kn. These nuclei 
certainly consist of bubbles, since the Aminco mea- 
surement in this case also did not show any dif- 
ferences. 

4. Measurements in shallow water show an 
absolute maximum at a diameter between 20 and 40 
um. The majority of these nuclei consists of 
suspended particles, as the Aminco scattered light 
measurement have shown. These suspended particles 
probably do not contribute to cavitation, since the 
comparison of propeller excited pressure fluctuation 
measurements between deep and shallow water shows 
practically no difference (Figure 7). 

5. The ship's vibrations caused by the propeller 
do pose a big problem for measurements of this type. 
The insensibility of the laser against vibrational 
stresses, however, after it was stiffly connected 
with the ship, was suprisingly good. Even the 
high loading caused by the temperature did not 
create any bad effects in the laser. 

6. Future laser measurements should possibly 
anticipate diameter ranges below 5 um. A more 
precise determination of suspended particles 
requires a greater effort than the present method. 

7. Further results of this trial will be 
published later. 


ACKNOWLEDGMENT 


The comprehensive measurements on the "Sydney Express" 
represented a project of the Sonderforschungsbereich 
98 "Schiffstechnik und Schiffbau" (Special Research 
Pool 98 "Marine Technology and Naval Architecture") 

to which Det Norske Veritas (propeller observation) 
contributed. The project was sponsored by the 
Deutsche Forschungsgemeinschaft. 

The authors wish to express their gratitude to 
Hapag-Lloyd who made the "Sydney Express" available 
for this investigation. Many thanks are expressed 
to Captain W. Scharrnbeck, the Chief Engineer H. 
Zwingmann, and the whole crew of the "Sydney Express", 
who, by their excellent co-operation, made possible 
the measurements and good results. 

The authors are grateful to: Ing.(grad) L. Hoffman 
(Hamburg Ship Model Basin - electronics, programming 
and evaluation); U. Steidlinger and W. Folkers, 
(Institute for Shipbuilding, Hamburg), and F. Meier 
(Technical University Munich) - all three provided 
for the mechanical construction and repair on board; 
Mrs. U. Schmidt (Institute for Shipbuilding, Hamburg ~ 
- for drawing the diagrams); Miss A. van Blericq 
(Hamburg Ship Model Basin - for the translation of 


the German original into English); and Mrs. I. Jurschek 


(Institute for Shipbuilding - for typing the manu- 
script). Dr. R. Doerffer (Institute for Hydrobiology 
of the Hamburg University, SFB 94) recommended the 
Aminco scattered light device and made suggestions 
concerning oceanography. 


REFERENCES 


Carder, K. L., G. F. Beardsley, and H. Pak (1971). 
J. Geophys. Res. 76 5070-5077. 
Gordon, D. C., Jr. (1970). Deep-Sea-Res. 17, 175- 


185. 
Isay, W.-H., and L. Lederer (1977). Kavitation an 
Flugelprofilen. (Cavitation on Hydrofoils). 


Schiffstechnik 24, 161. 

Jerlov, N. G., and E. S. Nielsen (1974). Optical 
Aspects of Oceanography. Academic Press, London 
and New York. 

Keller, A. P. (1970). Ein Streulicht-Zahlverfahren, 
angewandt zur Bestimmung des Kavitationskeims-— 
pektrums. (A Scattered-Light Counting Method 
used for the Determination of the Cavitation 
Nuclei Spectrum) Optics 32, 165. 

Keller, A. P. (1973). Experimentelle and theore- 
tische Untersuchungen zum Problem der modellma- 
Bigen Behandlung von Stromungskavitation. 
(Experimental and Theoretical Investigations on 
the Problem of Cavitation in a Flow with Models). 
Versuchsanstalt fur Wasserbau der Technischen 
Universitat Munchen. Rep. 26/1973. 

Keller, A. P., E. Yilmaz, and F. G. Hammit (1974). 
Comparative Investigations of the Scattered-Light 
Counting Method for the Registration of Cavitation 
Nuclei and the Coulter Counter. University of 
Michigan, Rep. UMICH 01357-36-T. 

Keller, A. P., and E.-A. Weitendorf. (1975). Der 
Einfluf des ungelosten Gasgehaltes auf die 
Kavitationserscheinungen an einem Propeller und 
auf die von ihm erregten Druckschwankungen. 
(Influence of Undissolved Air Content on 
Cavitation Phenomena at the Propeller Blades 
and on Induced Hull Pressure Amplitudes) . 
Institut fur Schiffbau, Universitat Hamburg. 

Rep. 321A. 


APPENDIX 


DESCRIPTION OF THE NOVEL TYPE OF VELOCITY 
MEASUREMENT 


When particles or bubbles pass through a light beam, 
they scatter a finite amount of light which is 
dependent principally on the object shape, size, 
index of refraction, and optical characteristics 

of the beam. For this technique a small, homoge- 
neously illuminated control volume (see No. 10 in 
Figure 13) is optically defined by the cross- 
sectional dimensions of the laser beam and the 
optics of the system detecting the scattered light 
(see No. 11 and 12 in Figure 13). 

The amplitude of the electrical output pulses 
from the photomultiplier (see No. 13 in Figure 13) 
is proportional to the "nucleus" size, and thus is 
the parameter used for "nucleus" spectrum determi- 
nation. 

The pulse width corresponds to the time in which 
the scatterer remains in the scattering volume, and 
therefore, by knowing the dimensions of the control 


Siby/ 


Krey, J., R. Boje, M. Gillbricht, and J. Lenz (1971). 
Planktologischchemische Daten der "Meteor"-Expe- 


dition in den Indischen Ozean 1964/65. (Plank- 
tological-Chemical Data of the "Meteor"- 
Expedition to the Indian Ocean 1964/65). "Meteor" 


Forschungsergebnisse, edited by Deutsche For- 
schungsgemeinschaft, Reihe D-No. 9; Borntraeger- 
Verlag, Berlin-Stuttgart. 

Lederer, L. (1976). Profilstr6mungen unter 
Beriicksichtigung der Dynamik von Kavitationsblasen. 
(Hydrofoil Flow with Regard to Bubble Dynamics) . 
Institut fiir Schiffbau, Universitat Hamburg. 
Rep. 341. 

Mullin, M. M. (1965). Size Fraction of Particulate 
Organic Carbon in the Surface Waters of the 
Western Indian Ocean. Limnol. Oceanogr. 10, 
453. 

Oossanen, P. van, and J. van der Kooy (1973). 
Vibratory hull forces induced by cavitating 
propellers. Transactions RINA 115, 111. 

Peterson, F. B. (1972). Hydrodynamic Cavitation 
and some Considerations of the Influence of 
Free Gas Content. 9th Symposium on Naval Hydro- 
dynamics, 2 1131, Paris. 

Peterson, F. B., F. Danel, A. P. Keller, and Y. 
Lecoffre (1975). Comparative Measurements of 
Bubble and Particulate Spectra by three Optical 
Methods. 14th ITTC, Ottawa. 2, 27. 

Sevik, M., and S. H. Park (1973). The Splitting of 
Drops and Bubbles by Turbulent Fluid Flow. 
Transaction ASME, Journ. of Fluids Engineering, 
95), Seriesh ay, Now 53. 

Zeitzschel, B. (1970). The Quantity, Composition 
and Distribution of Suspended Particulate Matter 
in the Gulf of California. Marine Biology, 7, 
4; 305. 


volume, the velocity of the "nuclei", i.e., the 
flow velocity, can be evaluated. 

The sketch in Figure A 2.1 shows the shapes of 
the optically bounded measuring volume for different 
positions of the rectangular laser aperture and 
the measuring slit in front of the photomultiplier. 
The time the "nuclei" need to cross the control 
volume is a function of the dimensions of the 
volume in the flow direction, and of the flow 
velocity. Therefore, the resulting photomultiplier 
pulse width is a measure of the flow velocity if 
the dimensions of the control volume are known. 

To get an accurate relation between pulse width and 
flow velocity, only nuclei of one known size, 
defined by their pulse height, should be selected. 

Example I in Figure A 2.2 displays an arbitrary 
position of the control volume relative to the flow 
direction. In that case, even for laminar flow one 
gets a certain fluctuation for the pulse widths, 
because the dimensions of the volume in the flow 
direction are not equal. 


318 


Rectangular 
aperture 


Measuring 
slit 


FIGURE A2.1. Principle of velocity measurement. 


FIGURE A2.2. Sketch of inclined control volume and 
received photomultiplier signals. 


FIGURE A2.3. Sketch of inclined control volume and 
flow direction. 


In example II Figure A 2.2 the main axis of the 
rectangular aperture is positioned parallel to the 
projection of the flow direction versus the plane 
vertical to the optical axis of the laser, and 
consequently the peak of the pulse width distribution 
is at a maximum value of t. 

In example III in Figure A 2.3 the direction of 
the measuring slit is also parallel to the projection 
of the flow direction versus the plane vertical to 
the optical axis of the photomultiplier, so that 
all dimensions of the measuring volume in the 
direction of the flow are the same, and the pulse 
width distribution therefore shows its most narrow 
shape. The peak of the distribution indicates the 
velocity in the main direction, whilst the shape of 
the curve is a measure of the turbulence level. 

The direction of flow can now be determined by 
the position of the rectangular aperture and the 
measuring slit. They each define a plane containing 
the corresponding optical axis, whereby the line 
of intersection represents the direction of the 
main flow in this region. 


Discussion 


319 


ORVAR BJORHEDEN and TORE DALVAG 


We congratulate the authors of this very 
interesting paper. For hull designers as well as 
propeller manufactures the problem of predicting 
the propeller induced vibration forces is a most 
essential task indeed. In this context we wish to 
inform you briefly about some recent developments 
at the KMW* Marine Laboratory related to the model 
testing technique applied in our cavitation tunnels. 

The first item concerns the method of hull 
wake simulation. For some time the well-known 
dummy technique, involving ship afterbody models 
and transverse net screens, has been used in our 
tunnels for the purpose of simulating model wake 
pattern. This is a rather time consuming process 
since the net screens have to be adjusted step by 
step until the correlation with the wake pattern 
obtained in the towing tank appears satisfactory. 
Moreover, the method has some technical drawbacks 
as regards the stability of the wake as well as 
the interaction between propeller and hull and the 
influence of the propeller on the wake pattern. 

In connection with hydro-acoustic tests, cavitation 
occurring on the nets may worsen the background 
noise level. 

In order to eliminate the above drawbacks a 
new technique involving longer afterbody hull 
dummies has been introduced. The method aims at 
simulating the full-scale ship wake pattern based 


upon the concept of equivalent relative boundary 
layer thickness, i.e., the frictional boundary 
layer thickness in relation to some characteristic 
length, e.g., the propeller diameter should be the 
same in the model and in full-scale. For ordinary 
cavitation testing purposes utilizing propeller 
model diameters around 250 mm and tunnels speeds 

of 4 to 8 m/sec this criterion results in hull 
dummy lengths of 2.5 to 3.5 m for most types of 
vessels. In principal, the model stern contour as 
well as the aftermost water-lines are made to scale, 
whilst the maximum breadth of the dummy is chosen 
on the basis of 2-dimensional potential flow cal- 
culations comparing the ship water-lines in unre- 
stricted water to the dummy lines within the bound- 
aries of the cavitation tunnel test section and 
aiming at similarity in the potential wake 
distribution. 

Figure 1 shows a picture of a 3 m hull dummy 
used for the testing of a 150 m, single screw, con- 
tainer ship. In Figures 2 and 3 the model wake 
distribution as obtained in the towing tank and 
then corrected for scale effect according to the 
so-called Sasajima method is given. In Figure 4, 
finally, a comparison between the corrected model 
wake and the wake distribution obtained in the cav- 
itation tunnel is shown for a few radii close to 
the propeller blade tip. As can be seen from the 
diagrams, the agreement is quite good, particularly 
as regards the wake peak in the 12 o'clock propel- 
ler blade position. 


*Karlstads Mekaniska Werkstad 


Figure 1. Hull dummy for wake simulation in cavita- 
tion tunnel. 


Apart from the advantage of a quicker and more 
direct simulation of the full-scale wake, the 
method with long afterbody dummies results in a far 
more stable wake distribution which in turn implies 
more consistent recordings of fluctuating propeller 
forces, propeller induced pressure pulses against 
the hull, etc. Probably, the interaction between 
propeller and hull is also more realistic with this 
method of wake simulation as compared to the method 
utilizing transverse nets. 

The second item refers to the instrumentation 
employed for recording of propeller forces and the 
propeller induced pressure pulses on a ship's hull. 
In both KMW tunnels a data collecting and evaluation 
system consisting of an on-line connected desk com- 
puter together with a printer and a plotter has 
been used for several years. For the measurement 
of propeller induced pressure pulses with the aid 
of pressure pickups fitted into the hull, a pulse 
sampling technique giving time averaged values from 
a number of propeller revolutions at each blade 
position has been the practice. With this method 
the pressure signals are given in analogue form and 
recordings can be obtained from only one pickup at 
atime. Recently, a new data collecting unit was 
put into service enabling simultaneous recording 
on 6 channels and storing test results from every 


r/R = 0.886 


r/R = 0.709 


r/R = 0.532 


r/R = 0.355 


r/R =0.177 


r/R = 0.177 
r/R = 0.355 
r/R = 0.532 
r/R = 0.709 
r/R = 0.886 
r/R = 1.063 == 
€ 
Figure 2. Model wake distribution as obtained in Figure 3. Wake distribution corrected for scale effect 
towing tank. according to the Sasajima method. 


W 
Wake Distribution Corrected 
1.0 Acc. to the Sasajima Method 
6 m/s 
r/R = 0.96 Simulated in Cavitation Tunnel 
SOSSSe 3 m/s 


r/R = 0.80 


0.5 


360 270 0 90 180 
) 
mparison between corrected model wake and wake simulated in cavitation tunnel. 


321 


Figure 5. Data collecting memory. 


second degree of a propeller revolution in digital 
form in a RAM semi-conductor memory controlled by 
the desk calculator. With this instrument, instan- 
taneous or time averaged test results can be stored 
and are readily available for printing, plotting, 


ERLING HUSE 


The authors in their presentation draw atten- 
tion to the problem of calculating cavity geometry 
and thus the excitation force due to cavitation. 

At the Norwegian Ship Model Tank in Trondheim we 

are at present developing a procedure to overcome 
this difficulty. In the cavitation tunnel we 
Measure the propeller-induced pressure at only 4 
positions on the hull model above the propeller. 

The measurements are made for non-cavitating as well 
as cavitating propellers. From the results of these 
Measurements we calculate an equivalent singularity 


O. RUTGERSSON 


First I would like to congratulate the authors 
on this interesting paper. The possibility of cal- 
culating hull forces and moments and their distri- 
butions directly on the body without the roundabout 
way over freestream pressures and solid boundary 
factors is especially elegant. Being somewhat in- 
volved in calculations and measurements of pressure 
fluctuations (with and without cavitation) at SSPA* 
I would like to ask if the authors intend to use 
this new method also to calculate solid boundary 


*Statens Skeppsprovningsanstalt, Goteborg, Sweden 


Figure 5. Desk calculator with printer and plotter. 


transformation to full scale, and harmonic analysis 
as well as integration of resulting hull surface 
forces and similar calculations with the aid of 

the desk calculator. 


distribution to represent the propeller. This is 
next combined with a theory similar to that of 
Dr. Vorus to obtain the excitation force on the hull 
referring to any given vibratory mode of the hull. 
As a second comment on the paper I notice in 
Figure 4 integration areas extending up to 30 pro- 
peller diameters upstream. This is, in my opinion, 
not very realistic because one is then passing one 
or more nodal points of practically occurring modes 
of vibration. 


factors for different afterbody shapes and propeller 
configurations? 

Unfortunately the authors' investigation is 
limited to non-cavitating propellers. This is a 
severe limitation as the contribution from the 
transient cavitation often is of a much higher mag- 
nitude than the contributions from blade loading 
and thickness. When discussing this subject the 
authors declare that methods "for predicting trans- 
ient blade cavity geometry and the attendant pres-— 
sure field" are not available. I would like to ask 
why the methods developed by Huse (1972), Johnsson 
and Sgndvedt (1972), and van Oossanen (1974) have 


322 


not been considered? These methods have been used 
in Europe for several years and the agreement with 
experiments is usually good. 

I agree that it is important that the integra- 
tion of hull forces and moments is carried out over 
a not too small part of the hull surface. This is 
even more important when the forces from a cavita- 
ting propeller are considered, as those pressures 
have a slower decay than those induced by a non- 
cavitating propeller [Lindgren and Johnsson (1977)]. 

Assuming that the hull forces should be used 
for an estimation of the vibration level for a 
certain ship project, I think that the problem is 
far more complicated than just a matter of integra- 
tion area. First, the described method is a near 
field theory where the influence of the propagation 
velocity of the pressure wave has been neglected. 
When calculating forces far from the propeller this 
could cause some difficulties. Secondly, the ship 
hull is not a rigid body. The vibration response 
will therefore be dependent not only on the hull 
forces but also on their location relative to the 
nodes of the vibration mode. Forces located close 
to the nodes will contribute very little and those 
located on different sides of a node will more or 
less cancel each other. Calculations with the Fi- 
nite Element Method have shown that hull forces aft 
of the aftermost node are particularly efficient 
in exciting high vibration levels. This could be 
the explanation for rather good results often being 
achieved in vibration calculations in spite of the 
fact that the excitation forces have been obtained 
by integration over a rather small area. 

The correct treatment of the problem will, of 
course, include vibration calculations, with a very 
detailed Finite Element model with the complete ex- 


citation forces and moments. Since this is very 
complicated and expensive it is seldom done. In- 
stead, different approximate procedures have been 
developed by different institutions. Referring to 
the integration problems the authors claim that 
"the current practice in European model basins is 
highly suspect." I very much doubt that this is 
current practice. At SSPA for example, we use the 
pressure fluctuations in a reference point above 
the propeller as a basis for estimation of the risk 
of vibration. On the basis of full-scale measure- 
ments we have established an approximate relation 
between excitation at this point and the vibrations 
at another reference point [(Lindgren and Johnsson 
(1977) 1]. 


REFERENCES 


Huse, E.,(1972). Pressure Fluctuation on the Hull 
Induced by Cavitating Propellers. Norwegian Ship 
Model Experiment Tank Publ. No. III. 

Johnsson, C. A.,.and T. Sgndvedt, (1972). Propel- 
ler Excitation and Response of 230,000 TDW Tankers. 
SSPA Publ. Wo. 70. 

Oossanen, P. van, (1974). Calculation of Perform- 
ance and Cavitation Characteristics of Propellers 
Including Effects of Non-Uniform Flow and Viscosity. 
NSMB Publ. No. 4657. 

Lindgren, H., and C. A. Johnsson, (1977). On the 
Influence of Cavitation on Propeller Excited Vibra- 
tory Forces and Some Means of Reducing its Effect. 
PRADS-Internattonal Symp. Tokyo. 


323 


Authors’ Reply 


BRUCE D. COX, WILLIAM S. VORUS, JOHN P. BRESLIN, 


and EDWIN P. ROOD 


Our thanks to the discussers for their interest 
and encouraging remarks. On Mr. Rutgersson's 
question of calculating solid boundary factors, we 
do believe it would be useful to perform computa- 
tions for a series of hull afterbody forms and pro- 
pellers. The results would illustrate sensitivity 
to the various physical parameters and could pro- 
vide guidance during the early stages of a ship 
design. However, for realistic predictions of pro- 
peller exciting forces, the complete calculation 
should be carried out using the actual wake, hull 
geometry, and propeller design under consideration. 

As noted in the paper and by Mr. Rutgersson, 
only the non-cavitating propeller case is consid- 
ered which is a severe limitation in many pratical 
applications. The principal purpose of the paper 
was to present analytical methods and simple form- 
ilae for predicting hull surface forces for a given 
representation of the propeller and show compar- 
isons with experiments. Future improvements in the 
propeller theory, in particular, the allowance for 
transient cavitation, can be incorporated quite 
readily into the surface force analysis. It can 
be shown [Breslin (1977)] that the time rate of 
change of the cavity volume plays a crucial role 
in generating the propeller pressure field. We 
are familiar with a number of proposed methods for 
predicting blade cavity geometry including those 
cited by Mr. Rutgersson. These approaches for the 
most part are empirical. An alternative procedure, 
described in Mr. Huse's discussion, consists of 
finding an "equivalent" singularity distribution 
so as to produce agreement between calculated and 
measured values of pressure at selected locations 
near the propeller. The problem of analytically 


predicting the proper singularity distribution to 
represent the cavity volume dynamics is now the 
subject of active research. 

We agree with Mr. Rutgersson that compress- 
ibility effects should be examined when considering 
the far field pressures generated by a propeller. 
A 5-bladed propeller operating at 100 rpm produces 
a blade rate frequency disturbance with a acoustic 
wavelength on the order of 600 feet. The relative 
phase of the distrubances generated far ahead of 
the propeller may be important in the integrated 
pressure force amplitude and phase. 

The theory presented in this paper assumes a 
rigid hull boundary, intended to provide a first 
estimate of propeller exciting forces acting on 
the hull girder. Certainly for detailed stress and 
vibration analyses, the interplay between fluid 
loading and hull structural deformation would have 
to be accounted for. In principle, the present 
theory can be extended to satisfy the boundary 
condition on a deformable body. The complete 
analysis would then involve coupled equations des- 
cribing the fluid loading and structural response, 
and could be solved by finite methods. 


REFERENCE 


Breslin, J. P., (1977). A Theory for the Vibra- 
tory Forces on a Flat Plate Arising from Inter- 
mittant Propeller Blade Cavitation. Sympostwm on 
Hydrodynamtes of Ship and Offshore Propulston 
Systems, Oslo, Norway 


' vr’ 
, € were £4 
d ‘ 
5 i 
s 
1 
4 
Pe 
’ “OL ay 4 
Pat yi 
<1 . - #1 « 
— + 
=. as Ah 
. - 
+ i 
’ i 4. 
r ~ = 
i ‘ 
tt t 
= 
1 ' 
i % \ 
= t ~ 
i 
> 2 
om 
E i 
4 ’ 
r + 
\ 
> y 
- l 
\ 
=I 
. 


Session V 


CAVITATION 


ERLING HUSE 

Session Chairman 

The Ship Research Institute of Norway 
Trondheim, Norway 


Ae) 


=({ 


Mi 


‘ 


ii 


Pressure Fields and Cavitation 
in Turbulent Shear Flows 


Roger E. A. Arnd 


University of Minnesota 


t 


Minneapolis, Minnesota 


William K. George 
State University of New York at Buffalo 


Buffalo, New York 


ABSTRACT 


Cavitation in turbulent shear flows is the result 
of a complex interaction between an unsteady 
pressure field and a distribution of free stream 
nuclei. Experimental evidence indicates that 
cavitation is incited by negative peaks in pressure 
that are as high as ten times the rms level. This 
paper reviews the current state of knowledge of 
turbulent pressure fields and presents new theory 
on spectra in a Lagrangian frame of reference. 
Cavitation data are analyzed in terms of the avail- 
able theory on the unsteady pressure field. It is 
postulated that one heretofore unconsidered factor 
in cavitation scaling is the highly intermittent 
pressure fluctuations which contribute to the high 
frequency end of the pressure spectrum. Because of 
limitations on the response time of cavitation 
nuclei, these pressure fluctuations play no role 

in the inception process in laboratory experiments. 
However, in large scale prototype flows, cavitation 
nuclei are relatively more responsive to a wider 
range of the pressure spectrum and this can lead to 
substantially higher values of the critical cavi- 
tation index. Unfortunately, this issue is clouded 
by the fact that higher cavitation indices can be 
found in prototype flows because of gas content 
effects. Some cavitation noise data are also 
examined within the context of available theory. 
The spectrum of cavitation noise in free shear 
flows has some similarity to the noise data found 
by Blake et al. (1977) with the exception that there 
appears to be a greater uncertainty in the scaling 
of the rate of cavitation events which leads to a 
substantial spread in the available data. 


1. INTRODUCTION 

The physical processes involved in cavitation 
inception have been studied for many years. Much 
of this research has been directed toward an under- 


327 


standing of the dynamics of bubble growth and the 
determination of the sources of cavitation nuclei 
and their size and number in a given flow situation. 
This research has led to a general understanding of 
some of the environmental factors involved in 
scaling experimental results from model to prototype. 
More recently, considerable attention has been 

paid to the details of the boundary layer flow over 
streamlined bodies and the role of viscous effects 
in the cavitation process. This research has shown 
that viscous effects such as laminar separation 

and transition to turbulence can have a major impact 
on the inception process and that there can be 
considerable variation between model and prototype 
in the critical conditions for cavitation. 

In the absence of viscous effects, the scaling 
problem reduces to an understanding of the size 
distribution of nuclei and the temporal response 
of these nuclei to pressure variations as viewed 
in a Lagrangian frame of reference. This was first 
treated in detail by Plesset (1949). As already 
mentioned, consideration of viscous effects shows 
that the cavitation inception process can be 
considerably altered by either laminar separation 
or transition to turbulent flow. Obviously these 
phenomena are interrelated and are strongly Reynolds 
number dependent. The recognition of the importance 
of these factors has had considerable impact on the 
direction of cavitation research in recent years. 
Several papers in this symposium deal directly with 
this aspect of the cavitation scaling problem. 

It is reasonably well understood that intense 
pressure fluctuations, either at the trailing edge 
of a laminar separation bubble or in the transition 
region, can have a major effect on the inception 
process on streamlined bodies. However, these 
phenomena will be excluded from this review. The 
focus of this paper will be on the relationship 
between the temporal pressure field and cavitation 
inception in free turbulent shear flows and fully 
developed boundary layer flows. Scant attention 
has been given to this problem, even though the 


328 


topic is of practical significance. Turbulent 
shear flows are very common in practice and what 
cavitation data are available for these flows 
indicate that there can be significant scale effects. 
For example, Lienhard and Goss (1971) present a 
collection of cavitation data for submerged jets. 
It is observed that the critical value of the 
cavitation index increases with an increase in 

jet diameter, with no upper bound on the cavitation 
index being defined by the available data. The 
cavitation index is observed to vary from 0.15 to 
3.0 over a size range of 0.1 cm to 13 cm. Arndt 
(1978) reviews the available data for cavitation 
in the wake of a sharp edged disk. These data 
increase monotonically with Reynolds number and 
again no upper limit on the critical cavitation 
index can be determined from the available data. 
At present, it can be said that laboratory experi- 
ments do not provide a reasonable estimate of the 
conditions that can be encountered under prototype 
conditions. From a practical point of view the 
situation is much more critical than the scaling 
problems associated with streamlined bodies since 
at present there is no definable upper limit on 


the cavitation index for these free shear flows. 
There are a myriad of factors that enter into 


the inception process in turbulent shear flows. 
As a minimum, we need information on the turbulent 
pressure field, such as spectra and probability 
density. We require an understanding of the diffu- 
sion of nuclei within the flow, and we need to 
know how these nuclei respond to temporal fluctu- 
ations in pressure. In taking into account the 
bubble dynamics inherent in the problem, consider- 
ation must also be given to gas in solution which 
can have an influence on both bubble growth and 
collapse. p 

The theory of bubble dynamics is well founded 
and reasonable estimates of critical pressure can 
be determined under flow conditions that are well 
defined. Needless to say, the flow conditions in 
a turbulent shear flow cannot be defined in 
sufficient detail. However, the problem of flow 
noise has led to a more comprehensive understanding 
of turbulence; in particular, recent aeroacoustic 
research has provided a wealth of data on turbulent 
pressure fluctuations. These data are a by-product 
of the need for understanding turbulence as a source 
of sound. At this point in time, it seems only 
logical to review the inception problem in terms 
of both classical bubble dynamics and the more 
recent results of the field of aeroacoustics. 


2. THEORETICAL CONSIDERATIONS FOR CAVITATION 


Cavitation Index 


The most fundamental parameter for cavitating flows 
is the cavitation index 


wherein p_ is a reference pressure, p_ the vapor 
pressure, U_ a reference velocity, and p the 
density of the liquid. The flow state of primary 
interest in this paper is characterized by a 
limited amount of cavitation in an otherwise Single 
phase flow. There is a specific value of 0 associ- 


ated with this flow condition, which for convenience 
will be defined as the critical index: 


If it is necessary to have completely cavitation 
free conditions, one design objective for various 
hydronautical vehicles is the minimization of 0 . 

Cavity flows are assumed identical in model 
and prototype for geometrically similar bodies 
when O is constant, irrespective of variations 
in physical size, velocity, temperature, type of 
fluid etc. In practice O0_ is found to vary over 
wide limits. Simply stated, these so-called scale 
effects are due to deviations in two basic assump- 
tions inherent in the cavitation scaling law; namely 
that the pressure scales with velocity squared and 
the critical pressure for inception is the vapor 
pressure, p_. As will be shown, the two factors 
can be interrelated, since in principle the critical 
pressure is a function of the time scale of the 
pressure field. 

In order to provide a foundation for the ensuing 
discussion, consider a steady uniform flow over a 
streamlined body devoid of any viscous effects. 

The following identity can be written: 


wherein C_ is a pressure coefficent defined in the 
usual manher. Generally speaking, C_ is defined 
by the pressure on the surface of a Given boedy- east 
is generally assumed that cavitation first occurs 
when the minimum pressure, p_, is equal to the 

: m™m 3 
vapor pressure, p This results in the well-known 
scaling law ys 


Consider next the case where the pressure in the 
cavitation zone is less than the minimum pressure 
measured on the surface of the body, then 


nt Pl E Py a Pl 
Oo = ——— -C fb ee 
Lou 2 1p me) W 2 

2 Oo ts (e) 


Here we have to distinguish between the pressure 
at the surface of the body p, and the pressure 
sensed by cavitating nuclei, p _. Assuming 
cavitation occurs when Pal = Py we have 


OL S=5= Cee (1) 


Equation (1) is one version of the superposition 
equation that is commonly referred to in the 
literature. 


Bubble Dynamics 


It is generally accepted that the process of 
cavitation inception is a consequence of the rapid 


or explosive growth of small bubbles or nuclei 
which become unstable due to a change in ambient 
pressure. These nuclei can be either imbedded in 
the flow or find their origins in small cracks 
and crevices at the surfaces bounding a given flow. 
The details of how these nuclei can exist have been 
considered by many investigators. A summary of 
this work is offered by Holl (1969, 1970). 
Theoretically, liquids are capable of sustaining 
large values of tension. However, the nuclei in 
the flow act as sites for cavitation inception 
and prevent the existence of significant tensions. 


The mechanics of the inception process are adequately 


described by the Rayleigh-Plesset equation, which 
considers the dynamic equilibrium of a spherical 
bubble containing vapor and non-condensable gas 
and subject to an external pressure Boy ft)? 


iN) 
n 
w\|we 


co qo il 
AER ——— 
RR 2 0 (2) 


+ - -—-4 
2. PG Ig SE) = Hl 


wherein R is the bubble radius and dots denote 
differentiation with respect to time. It should 
be emphasized here that even for the case of steady 
flow over a streamlined body, p_,(t) is a function 
of time since we are concerned with the pressure 
history sensed by a moving bubble. If the problem 
is simplified to consider the static equilibrium 
of a bubble, we find that there is a critical 

value of p - p below which static equilibrium 

is not possible. This is found to be 


(ee = Dae = 45/3R* (3) 
wherein R* is defined as the critical bubble radius. 
Substitution of Eq. (3) into Eq. (2) with dynamical 
terms identically zero will indicate that R* is a 
function of the partial pressure of noncondensable 
gas within the bubble. If p __(t) varies rapidly 

in comparison to the response time of the nuclei, 
then even greater values of tension are possible. 
Thus in general we can write 


PY y Pal Ss) 


AS /SR*in my c 3 


where 


ll 


o(o) =~, o(~) =1 


The function ¢ depends on the flow field. The 
argument of ¢ contains a characteristic time scale 
of the pressure field (t_) anda GREACEOENSIELS 
response time of the nuclei, (PR, 37s)* 7 En) the 
case of a streamlined body in une absence of viscous 
effects, t_ would be proportional to the quotient 
of body diameter and velocity. In the case of 
cavitation induced by turbulence, the characteristic 
time scale could be any of the turbulence time 


scales. For example, 
u' OD 
L4/ Vine 3 


ae factor (PR, js)? is 
derived from the asymptotic solution to Eq. (2) 
for the case of negligible gas diffusion. Under 
these conditions 


is often appropriate. 


329 


and the growth rate stabilizes at a value given by 


| 2) a (4) 
3 p 


Assuming a characteristic bubble response time 
given by R*/R, with 1 = ea = 4S/3R*, we obtain 


R* 
= —— O. 
qT, R 87 \ (5) 


A typical variation of ¢ based on the theoretical 
computations of Keller (1974) is given in Arndt 
(1974). 


The Influence of Dissolved and Free Gas 


The discussion in the previous section is based on 
the assumption of a healthy supply of free nuclei 
which is generally the case in recirculating water 
tunnels and in the field. Generally speaking, a 
reduction in O_ due to bubble dynamic effects 
usually only occurs on model scale. To some extent 
the level of dissolved gas and the number and size 
of free nuclei are interrelated. Some recent 
experimental results are documented in Arndt and 
Keller (1976). The level of dissolved gas can 
play an important direct role when the time of 
exposure to reduced pressure is relatively long. 
Under these circumstances Holl (1960) has shown 
that gaseous cavitation can occur at values of 6 
much greater than those for vaporous cavitation. 
Using an equilibrium theory, Holl (1960) deduced 
an upper limit on on given by 


wherein @ is the concentration of dissolved gas 
and 8 is Henry's constant. 

In summary, an overview of the effects of bubble 
dynamics and free and dissolved gas indicates that 
short exposure times such as are the case ina 
model implies that cavitation will occur at pressures 
lower than vapor pressure and OF is less than 
expected. Long exposure time, Such as can occur 
in vortical motion of all types, including large 
scale turbulence, implies the possibility of gaseous 
cavitation with © being greater than expected. 


3. PRESSURE FLUCTUATIONS IN TURBULENT SHEAR FLOWS 
Background 


Considerable progress has been made over the last 
five years in the understanding turbulent pressure 
fluctuations in free shear flows in an Eulerian 
frame of reference. Of particular importance is 
the development of pressure sensing techniques 
which under certain circumstances can lead to 
reliable measurements of pressure fluctuations. 


330 


The first theoretical arguments on the pressure 
fluctuations associated with turbulent flow appear 
to be due to Obukov and Heisenberg [Batchelor 
(1953) ]. Heisenberg argued that Kolmogorov scaling 
should be possible for small scale pressure fluc- 
tuations. Batchelor (1951) was able to calculate 
the mean square intensity of the pressure 
fluctuations as well as the mean square fluctuating 
pressure gradient in a homogeneous, isotropic 
turbulent flow. This work was extended by Kraichnan 
(1956) to the physically impossible but conceptually 
useful case of a shear flow having a constant mean 
velocity gradient and homogeneous and isotropic 
turbulence. 

Apparently there were no attempts made to extend 
this theoretical work until the 1970's when George 
(1974a), Beuther, George, and Arndt (1977a, b, c) 
and George and Beuther (1977) applied the concepts 
developed by Batchelor and Kraichnan to the calcu- 
lation of the turbulent pressure spectrum in 
honogeneous, isotropic turbulent flows with and 
without shear. When compared with experimental 
evidence gathered in turbulent mixing layers, the 
theory is found to be remarkably accurate. The 
predicted spectrum (with no adjustable constants) 
agrees with pressure measurements in turbulent jet 
mixing layers from several sources, including 
those of Fuchs (1972a), Jones and his co-workers 
(1977), and the authors themselves. As shown in 
Figure 1, the experimental data and the theory are 
remarkably consistent, especially in light of the 
fact that several different experimental techniques 
and different flow facilities are involved. 

The current state of knowledge of turbulent 
pressure fluctuations can be summarized as follows: 
1) Pressure fluctuations in a shear flow can 

arise from three sources. The first two involve 
interaction of the turbulence with the mean shear. 
These are second order and third order interactions, 
of which only the second order interactions are 
important at small scales. The last involves only 
interactions of the turbulence with itself. 

2) Kolmogorov similarity arguments can be 
applied to each of the spectra arising from these 


-400 Uy X/D 
198 15 
19.8 30 
30.5 1.5 
5 30.5 3.0 
po Michalke 
% |X Fuchs 
Ss (75) 
= 
ne -6.00 
mo 
2 


RT) 200 300 


log Kx 


FIGURE 1. Experimental confirmation of the theoretical 
pressure spectrum for a turbulent jet. 


terms. These arguments are valid for the small 
scale fluctuations. 

3) If the turbulent Reynolds number is high 
enough, there exists an inertial subrange in each 
of the three spectra in which 


us, D DPD gail 
Ted (k) iP Kk 
ses 2 =9/5 
Uap ie) Se xs 
= 2) WV} 7/3} 
Tp (K) = 4,0 € k 
wherein a . = 2, a. =0, a= 1.3, ¢ is the rate of 


dissipation of ene Sailene enérgy per unit volume, 
K is the mean shear, and k is the disturbance wave 
number. 

4) There is considerable evidence that coherent 
structures play an important role in determination 
of at least the large scale pressure fluctuations 
[Fuchs and Michalke (1975), Fuchs (1972a, b), Chan 
(1974a, b), and Chan (1976)]. 


Relation to Cavitation 


Since the above spectral results are expressed in 
Eulerian frames, they cannot be directly applied 
to the problem of cavitation inception which is a 
Lagrangian problem. Nonetheless, Kolomogorov scaling 
has been successful in an Eulerian frame of reference 
and therefore we can, with some confidence, infer 
that similar scaling will be valid for Lagrangian 
time spectra (i.e. the frequency spectra that would 
be seen by a moving material point). The results 
of such an exercise are as follows: 

1) The Lagrangian turbulent spectrum can be 
separated into interaction of the turbulence with 
the mean shear and the interaction of the turbulence 
with itself. 

2) The high frequency (analogous to small scale) 
will be well described by Kolmogorov scaling such 
that 


Apps (W) 


i 
A 
i) 
< 
ow 
SS 
ie) 
| 
vr 
Fh 
n 
—N 
|E 
Ss 


L 
es AppT (Ww) ye 6? fo Gal 


2 
p wr 
where 1 
2 
Rete (oe. 
cncy 
3) In the inertial subrange these reduce to 
-5 
1 
= 5 =i 
Fees () se we 2) 
fo) Ww 
a 
-3 
3/2) es Ww 
i AppT (W) = Vv / Se (=) 
Ww 
fe) d 


In summary it appears plausible to assume that 
the basic picture of pressure fluctuations arising 
from mean-shear turbulence interactions will be 
unchanged in a Lagrangian frame of reference, 
although the actual spectra are different. The 
postulated relations for Lagrangian spectra should 
be directly applicable to any Lagrangian phenomenon; 
in particular the relations should be applicable 
to the inception of nuclei in a fluctuating pressure 
field. 

In relating the information on the pressure field 
to the problem at hand, it is evident that two 
criteria must be satisfied for turbulence induced 
inception: 

1) The pressure must dip to the vapor pressure 
or lower. 

2) The pressure minimum must persist for a time 
that is long in comparison to the characteristic 
time scale of the bubble, say Tp (taken to be the 
time scale for growth at inception). 

Both factors lead to scale effects. Consider 
first the second factor. The preceding arguments 
for the pressure field in a Lagrangian frame of 
reference lead to the hypothetical spectrum shown 
in Figure 2. For convenience we have normalized 
the spectrum with respect to the mean square pressure 
and the Lagrangian time scale JY. (c.f. Tennekes 
and Lumley, Chapter 8). Requirement (2) for bubble 
growth is plotted at the frequency W = 1/T,- te 
is clear that as long as w << 1/Tp, any pressure 
flucuation persists for a time longer than the 
time scale of the bubble. Thus at frequencies less 
than » = 1/Tg cavitation inception can occur with 
minimal local tension. Moreover, by integrating 
the spectrum from wW = O to W = 1/T, 1 we can deter- 
mine that fraction of the mean square pressure 
which can contribute to bubble growth without 
appreciable tension (assuming a normal distribution 
of nuclei). 

Consider now the effect of maintaining Tg con- 
stant while varying the Reynolds number. Taking 
J~ &£/a' and noting that there are essentially no 
pressure fluctuations of interest above the 
Kolmogorov frequency, W = (e/v)2 we find that 
after 1/Tp, exceeds (e/v) 4, the entire spectrum 
can potentially contribute to bubble growth. This 
will occur when the Reynolds number is roughly 


3 
Q 
a 
< 
Ve Ve 0.2(%) "2 
E B 
FIGURE 2. Hypothetical pressure spectrum in a 


Lagrangian frame of reference. 


331 


Mean Square Pressure Fluctuation (Lagrangian) 
Sea eS Ses SS oo oS SS 


Vy 0.2 () 2 


wy —> 


FIGURE 3. 
spectrum. 


Integration of Lagrangian pressure 


ul/v ~ (2/uTR)?. By noting the spectral dependence 
on frequency and performing a running integral, a 
plot such as shown in Figure 3 can be generated. 
This graph illustrates how rapidly the asymptotic 
state is reached. This occurs when 7/T -ts J (€/v) 
> (ut £/v) 2 or when £/u'T_ > (u'&/v) 72 as previously 
stated. B 

As an example,* cavitation is observed to occur 
in submerged jets at an axial position, x, that is 
roughly one diameter from the nozzle. Assuming 
the dissipation rate to be approximately 0.05U;°/x, 
where Uz is the jet velocity, results in a criterion 
that the jet diameter must exceed the following 
before scale effects are absent: d > 0.05U;°TR7/v. 
Using typical values of Uz = 10 m/s and Tp = Om 
sec., we conclude that the asymptote is reached for 
d ~ 50 meters. Thus size effects could be important 
in many model experiments. 


1 
i) 


Effect of Intermittency at Small Scale 


In 1947, Batchelor and Townsend concluded from 
observations of the velocity derivatives in 
turbulent flow that the fine structure of the 
turbulence (small scales, high frequency) was 
spatially localized and highly intermittant in 

high Reynolds number flows. Subsequent work [c.f. 
Kuo and Corrsin (1971)] has confirmed that there 

is a decrease in the relative volume occupied by 

the fine structure as the Reynolds number is 
increased. Thus the spatial intermittancy increases 
with Reynolds number. The effect of this phenomenon 
on filtered hot wire signals is shown in Figure 4. 
These data are derived from Kuo and Corrsin (1971). 
It is obvious from these data that the signal is 
increasingly intermittant as the filter frequency 

is moved to higher and higher values. 

Since the dissipation of turbulent energy takes 
place at the smallest scales of motion, it is clear 
from these observations that the rate of dissipation 
of turbulent energy must vary widely with space 
and time. It was this consideration that led 


*Strictly speaking, these results are only applicable 
when the Lagrangian turbulent field is stationary. 

In most flows of interest this is seldom the case. 
However, the smallest scales of motion can often 

be considered to be in quasi-equilibrium. 


FIGURE 4. Filtered hot-wire signals in grid- 
generated turbulence [adapted from Kuo and Corrsin 
(1971)]. (i) £ = 200 Hz, £/£ = 0.52, 20 ms/division 
(horizontal scale); (ii) 1 kH2, OoSA, 4a (sists) Gp 
0.52, 1; (iv) high-pass signal, f = kHz, 1 ms/ 
division. 2 


Kolmogorov (1962) to reformulate his original 
similarity hypothesis in terms of the average rate 
of dissipation of turbulent energy <e> , and to 
assume that the logarithm of € was governed by a 
normal distribution. Later work by Gurvich and 
Yaglom (1967) showed that any non-negative quantity 
governed by fine scale components has a,log normal 
distribution with a variance given by » =A+B 

ln R, where A is a constant depending on the 
structure of the flow, B is a universal constant and 


Ro is the turbulence Reynolds number. 

These results have implications for the cavita- 
tion problem at hand. Beuther, George, and Arndt 
(1977a, b) have shown that Kolmogorov similarity 
scaling is applicable to the high wave number 
turbulent pressure spectrum. As a consequence of 
this and the observed intermittancy and spatial 
localization of small scale velocity fluctuations, 
it is reasonable to expect the same trend in the 
small scale pressure fluctuations. This could 
result in an important cavitation scale effect. 

To make this point clear, a set of hypothetical 
band passed pressure signals at high and low 
Reynolds number are presented in Figure 5. For the 
sake of argument, assume that the filter is set 
around a range of frequencies which will result in 
bubble growth (wTgp £1). Since the spectra of these 
two signals will be identified in terms of Kolmo- 
gorov variables and since the low Reynolds number 
signal is less intermittant, there is a greater 
probability that the high Reynolds number signal 


FIGURE 5. Hypothetical band-passed pressure signals: 
(i) low turbulent Reynolds numbers, (ii) high turbu- 
lent Reynolds number. 


will contain more intense deviations from the mean. 
In particular, with all other factors held equal 

it is more likely that the local pressure will fall 
below the critical pressure when the Reynolds number 


is high, even though the spectra are identical. 
This is shown in Figure 5. If the log normal 


arguments were applicable, then it can be expected 
that this will depend on the Reynolds number. 

The effect of intermittancy coupled with effects 
cited earlier could be of considerable importance 
to the problem of predicting cavitation inception 
in the prototype from small scale experiments in 
the laboratory. The Reynolds number in model and 
prototype can vary by many orders of magnitude. 
For example, experimental observations of boundary 
layer cavitation by Arndt and Ippen (1968) were 
carried out at Reynolds numbers, u'6/v, of the 
order 5000. On large ships, Reynolds numbers of 
10° and greater are not uncommon. 


Coherency of the Pressure Field 


An important factor related to cavitation in- 
ception in jets is the existence of coherent 
structure in the flow. Cavitation in highly turbu- 
lent jets is observed to occur in ring like bursts, 
smoke rings if you will. These bursts appear to 
have a Strouhal frequency fd/U_ of approximately 
0.5. This point is underscored by some recent work 
of Fuchs (1974). Fuchs made 2 and 3 probe pressure 
correlations as shown in Figure 6. His results are 
summarized in Table 1. Signals filtered at a 
Strouhal number of 0.45 were highly coherent. For 
comparison, velocity correlations are shown in 
parentheses indicating that the velocity field is 
much less coherent than the pressure field. 


The Turbulent Boundary Layer 


Because of the relative ease of measurement, there 
exists a considerable body of experimental data 


Jet Nozzle 


Probes (la2) 


“| 
Probe (0)—g\ A\ 
A \E 3d 


General Arrangement 
(a) (c) 


(b) 
a ae 
PoP, Py (P,*P5) Py P, P, Po 


FIGURE 6. Measurement of pressure coherency in a 
turbulent jet [adapted from Fuchs (1974)]. 


for wall pressure due to turbulent boundary layer 
flow. However, in many ways less is known about 
the turbulent pressure field for boundary layers 
than for free turbulent shear flows. Not only is 
the theoretical problem made more difficult 
(impossible to the present) by the presence of the 
wall, the experimental problem is considerably 
complicated by the dynamical significance of the 
small scales near the wall. 

Thus, in spite of over two decades of concentrated 
attention we cannot say with confidence even what 
the rms wall pressure level is, although recent 
evidence points to a value of [Willmarth (1975)]: 


mp = oo 


c 2 ico) 3} 

The basic problem is that the most interesting part 
of a turbulent boundary layer appears to the region 
near the wall where intense dynamical activity 
apparently gives rise to the overall boundary layer 
activity. While the details of the process are 
debatable, most investigators concur on the importance 
of the wall region on overall boundary layer 
development. Unfortunately, under most experimental 
conditions, the scales of primary activity are 
smaller than standard wall pressure probes can 
resolve [Willmarth (1975)]. Thus we have virtually 
no information concerning the contribution of the 
small scales to the pressure field, although we 
suspect that the small scales are significant or 
even dominant. 


Pressure Spectra in Boundary Layers 
Our knowledge of the pressure spectra may be 


summarized as follows: 
1) Pressure fluctuations arising from motions 


333 


in the main part of the boundary layer (y/é > 0.1) 
scale with the outer parameters iB, and 6. 

2) Pressure fluctuations arising from the inner 
part of the boundary layer scale with the inner 
parameters: 

a) hydraulically smooth, uy, Vv 
b) hydraulically rough, u h; where h is 
roughness height 

3) Pressure fluctuations arising from the 
inertial sublayer (logarithmic layer) scale only 
with u, and y, the distance from the wall. 

4) The wall pressure spectrum is a composite 
of all these factors and has a distinct region 
corresponding to each factor. 

A composite picture of the wall pressure spectrum 
is shown in Figures 7a and 7b. The 1/k range is 
evident in both the inner and outer scalings and 
arises from the inertial sublayer contribution 
[Bradshaw (1967)]. 

The pressure spectrum within the near wall region 
should closely resemble the wall spectrum (although 
this has never been confirmed). The spectrum in 
the main part of the boundary layer, should, however, 
resemble that obtained for a free shear flow at 
high Reynolds numbers. Again there is no information 
available to either prove or disprove this conjecture. 

The Lagrangian model developed in the preceding 
section depends in part on the assumption that a 
material point is in a stationary random field. 

As long as the Eulerian field is homogeneous, there 
is no problem. This is approximately true in many 
shear flows, but is never true in a turbulent 
boundary layer. Thus our Lagrangian spectral picture 
must be abandoned entirely (or used with great 
restraint). 

However, a number of features of the Lagrangian 
model can be applied to this problem. In particular, 
the "spectral peaks" in the outer flow can be 
identified with the Lagrangian integral scale, 

J ~ &/u'. The highest frequencies in the flow will 


*! 


Table 1. 


Normalized correlation functions with pressure probes 
arranged as shown in Figure 6 (corresponding velocity 
correlations in brackets). 


Signals 
Unfiltered 


Signals 
Filtered at St = 0.45 


Increasing 


(cm! 
m* 
3 . 
= Increasing 
ee ud 
a: 
za 
re 
[oo 
Do 
2 
log xv/u, 
FIGURE 7. Wall pressure spectra: (a) outer scaling, 


(b) inner scaling. 


be e/a) or u*/h, depending on whether the wall is 
hydraulically smooth or rough, and there will be 
increasing intermittency with increasing Reynolds 
number. The latter effect is most interesting and 
is quite evident in the many observations of dye 
streaks in the wall layer [cf Kim, Kline, and 
Reynolds (1971)]. 


Effect of the Pressure Field on Cavitation 


Whether or not the pressure fluctuations play a role 
in the cavitation inception process, depends on 
the previously cited criteria: 

1) The minimum pressure must fall below a 

critical level. 

2) The minimum pressure must persist below the 

critical level for a finite length of time. 

The first criterion depends greatly on the yet 
unresolved question of intermittency and its effect 
on the probability density of the pressure fluctua- 
tions. At this point in time we can say that the 
critical cavitation index will increase with 
Reynolds number because larger excursions from the 
mean pressure are more likely. Without justification, 
it is hypothesized that the effect on the pressure 
variance will be approximated by a log-normal 
dependence on the Reynolds number. ‘Detailed study 
of the wall pressure such as that proposed by 
George (1975) should aid considerably in resolving 
this question. 

The question of time scale is more easily con- 
fronted. Since most of the energy in the pressure 
spectrum scales with u, and 6 it is clear that the 
criteria for bubble growth without appreciable 
tension reduces to 


u,T,/5 <1 


In words, we again require a pressure fluctuation 
to persist for a time which is long in comparison 
to the response time of a typical nucleus. 

Since v/u,? is the shortest time scale ina 


smooth wall boundary layer, all of the pressure 
spectrum is sampled by the nuclei when 


2 

u, T,/Y < il 
This criterion is especially important in view of 
the highly intermittant process near the wall. 


For rough walls, the last criterion can be 
expressed in terms of the roughness height h by 


u,T,/n <a 


Since in fully 
that the small 


rough flow u,h/v Sl ale GUS} Giles 
scale criterion is more easily 
satisfied with rough wall experiments. 

In summary, the information we have on pressure 
fields in turbulent boundary layers and its 
relationship to cavitation inception can be 
summarized as follows: 

Significant scale effects can be expected when 
u'T,/5 > 1. As the ratio of T. to the smallest 
time scale in the flow decreases, the scale effect 
would be expected to level off i.e. when uxTp/V or 
u,T,/h <1. Further increase in the cavitation 
n er with Reynolds number will be due to the 
Reynolds number dependent effects on the probability 
density of the pressure fluctuations as a result of 
increased intermittancy of the small scale structure. 
The latter effect should produce a more gradual 
dependence of the cavitation index on Reynolds 
number than the former effect. 

The picture, as displayed above, is plausible 
and perhaps even appealing, but it must be viewed 
simply as conjecture until definitive experimental 
information is made available. An important hint 
of the relevance of these results can be found in 
the work of Arndt and Ippen (1967) where it was 
found that the region of maximum cavitation ina 
rough boundary layer shifted inward with a decrease 
in u,T /n. However, the change in this parameter 
varied only by a factor of 15 in their experiments. 
This will be discussed in more detail in subsequent 
sections. 


4. CAVITATION INCEPTION DATA 


A rather limited amount of experimental data have 
been collected under controlled conditions. The 
types of flows considered to date include the wake 
behind a sharp edged disk, submerged jets from 
nozzles and orifices, and smooth and rough boundary 
layers. There is a dearth of information relating 
the observed cavitation inception with the turbulence ~° 
parameters. Some of the earlier efforts in this 
direction are summarized in a paper by Arndt and 
Daily (1969) and by Arndt (1974b). A collation of 
available data is presented in Figure 8. Here the 
data are presented in the form of Eq. (1): 


o fC = 
Pp 


* f£ (Cp) 


Ke) T om Se T T Vapaly oe em Seer an au 
© Smooth, Daily & Johnson ('56) 


Boundary |@Sawteeth, Arndt & Ippen ('68) 
Layer © Sand, Messenger ('68) 
™ Sand, Huber ('69) 


eo: 


f 


Jet 4 Rouse, ('53) 
oo | Wake 4 Kermeen, Et Al ('55) | 
+ 
co 
OF a 
Best Fit Curve 1 
br ° >| 
| 10 100 
1000 C, 
FIGURE 8. Collation of cavitation inception data. 
wherein 
2t )/pU* Boundary Layer Flow 
Ce = 
W145 Free Shear Flows 
U 
fe} 


In this expression Cg is computed either from the 
measured wall shear jn the case of boundary layer 
flows or from turbulence measurements made in the 
air at comparable Reynolds numbers for the case of 
a free jet and a wake. The measured value of C, is 
only significant for the case of the disk wake and 
the pressure data was determined from the experi- 
mental work of Carmodi (1964). The available data 
seem tc be well approximated by the relation 


which was originally proposed for boundary layer 
flow by Arndt and Ippen (1968). These data would 
seem to imply that a relatively simple scaling law 
already exists and would further imply that the 
previous discussion in this paper on turbulence 
effects is superfluous. This is not the case. 
Arndt and Ippen (1968) made observations of the 
bubble growth in turbulent boundary layers. Some 
of their results are depicted in Figures 9 and 10. 
Figure 9 shows sample bubble growth data. The 
growth rate is observed to stabilize at a constant 
value during most of the growth phase. Using Eq. 
(4), the levels of local tension are found to be 
quite small, of the order 20 to 100 millibar. These 
data correspond to observations in a rough boundary 
layer. Of particular interest is the fact that, 

in all cases, the life time for bubble growth is 

a fraction of the Lagrangian time scale, J = d/u'. 
In fact growth times were observed to be of the 
order h/u,. Unfortunately there is not enough 


*Tp was estimated from Eq. (5) using observed values 
of R, reported in Arndt and Ippen (1967). For 
convenience, the results are normalized to equivalent 
Sand grain roughness, hg. 


335 


O16 
ig) Fol 
xo 
[= 
£ 
i) 
OG 008 
2 
we} 
Te} 
J) 
a " 
004 © Run P35 k-0400 
; © Run P47 k-0025 
CORIEOMSIRaIN SMG II 
Time (m sec) 
FIGURE 9. Sample bubble growth data [after Arndt 


and Ippen (1968)]. 


experimental evidence available to completely 
illuminate this point. As shown in Figure 10, 
cavitation occurs roughly in the center of the 
boundary layer with a tendency for the zone of 
maximum cavitation to shift inward as uxTp/h, 
decreases from about 1.5 to approximately 0.1*. 
In the cited boundary layer experiments, Cp is 
negligible. Thus 0, = 16 Cf. Noting that p' is 
approximately 2.5 pux* at the wall, we estimate 
that cavitation is incited by negative peaks in 
pressure of order 6 p'. This compares favorably 
with Rouse's (1953) data for jet cavitation which 
indicate that negative peaks of order 10 p' are 
responsible for cavitation. 

A strong dependence on Reynolds number can be 
observed even in free shear flows. Figure 11 
contains cavitation data for a sharp edged disk. 
These data were obtained in both water tunnels 
and a new depressurized tow tank facility located 
at the Netherlands Ship Model Basin. The water 
tunnel data are for cavitation desinence, whereas 
the tow tank data are for cavitation inception 
determined acoustically. The cross hatched data 
were determined in a water tunnel at high velocities 
by Keermeen and Parkin (1957). All the other data 
were obtained at relatively low velocities (2 - 10 
m/sec). There is considerable scatter in these 
data and this is traceable to gas content effects 


T T T T T T T T a reas |e 
all bubbles all bubbles 
30F ---- cavitating bubbles 30+ ---- Cavitating bubbles 4 


Relative Concentration (%) 
Relative Concentration (%) 


2 


FIGURE 10. Observation of cavitation in turbulent 
boundary layers [after Arndt and Ippen (1968) ]. 


336 


AL = 44 +0,0036R//¢ 


Od x 4 


aS 
Sur 


Onna «5 


Ow 


High Reynolds Number Asymptote 4 


ff Water Tunnel 
B O Smooth 


WY Kermeen & @ Rough 
Parkin O Vacutank 


(0) 2 4 6 8 10 l2 14 16 18 20 
Reynolds Number x 109 


Cavitation Index 


FIGURE 11. 
edged disk. 


Cavitation inception data for a sharp- 


which are dominate at low velocities as will be 
discussed later. At low Reynolds number the data 
appear to be satisfied by the empirical relationship 
discussed by Arndt (1976): 


o, = 0.44 + 0.0036 (Ud/v) 2 (7) 


It was found that the tow tank data agree with this 
relationship at relatively high Reynolds numbers. 
Equation (7) was developed from a model which 
assumes laminar boundary layer flow on the face of 
the disk. It would be expected that this condition 
would be satisfied at higher Reynolds numbers in 

a tow tank than in a highly turbulent water tunnel. 
At high Reynolds number (and also high velocity 
where gas content effects are negligible), there 

is a continuous upward trend in the data with 
increasing Reynolds number. This underscores the 
need for further work as suggested in the intro- 
duction to this paper. 

A systematic investigation of gas content effects 
in free shear flow was recently reported by Baker 
et al. (1976). Cavitation inception in confined 
jets, generated either by an orifice plate ora 
nozzle, was determined as a function of total gas 
content in the liquid. The results are shown in 
Figure 12. When the liquid was undersaturated at 
test section pressure, the critical cavitation 
index was independent of gas content and roughly 
equal to that observed by Rouse (1953) for an 
unconfined jet. When the flow is supersaturated, 
the cavitation index is found to vary linearly with 
gas content as predicted by the equilibrium theory, 
Eq. (6). This effect occurs even though the 
Lagrangian time scale is much shorter than typical 
times for bubble growth by gaseous diffusion. For 
example, in the cited cavitation data, a typical 
residence time for a nucleus within a large eddy 
is roughly 1/15 of a second. At a gas content of 
7ppm and a jet velocity of approximately 10 m/s, 
inception occurs at a mean pressure equivalent to 
a relative saturation level of 1.25. Epstein and 
Plesset (1950) show that for growth by gaseous 3 
diffusion alone, 567 seconds is required for a 10 
cm nucleus to increase its size by a factor of 10. 
One additional point should be kept in mind here. 
The local pressure within an eddy is much less than 
the mean pressure and highly supersaturated con- 
ditions can occur locally. Arndt and Keller (1976) 
also reported extreme gas content effects in their 


experiments with disks when the flow was super- 
saturated. The magnitude of the effect also depends 
on the number of nuclei in the flow. Gas content 
effects were noted only in their water tunnel 
experiments (where there is a healthy supply of 
nuclei). No gas content effects on inception were 
noted in the tow tank (where the flow is highly 
supersaturated but there is a dearth of nuclei). 
Thus the picture becomes more cloudy as the influence 
of dissolved, noncondensable gas is taken into 
consideration. 


5. SOME REMARKS ON CAVITATION NOISE 


A complete discussion on cavitation noise would be 
beyond the scope of this paper. Recognizing the 
unique features of cavitation inception in 
turbulent shear flows, it appears appropriate to 
review what is known about cavitation noise under 
the same circumstances. 

The general features of cavitation noise were 
reviewed by Fitzpatrick and Strasberg (1956), Baiter 
(1974), and Ross (1976). The spectrum of cavita- 
tion noise can in its simplest form be defined as 
the linear superposition of N cavitation events per 
unit time. Thus we can write 


S(£) = N G(£) (8) 


The function G(f) is the spectrum of a single 
cavitation event. If p, is the instantaneous 
acoustic pressure due to the growth and collapse of 
a single bubble, then by definition 


J ccrar - ee dt 
oO 


—co 
Fitzpatrick and Strasberg (1956) have shown that a 


characteristic bubble spectrum can be written in 
the form 


a 
b 
g 
= _047B(a-a,) 
5 ; 2 
5 1/2 pUG 
S 
S 
my (a) 
7] 
Cc 
Fy 
a 


(b) 


Contoured Nozzle 


FIGURE 12. Cavitation inception in confined jets. 


wherein Tt, is a characteristic bubble collapse 
time, Ry is the maximum bubble radius, and R is 
the distance to the observer. In addition, it 
appears reasonable to assume that N is related to 
the number of nuclei per unit volume, n, the 
velocity, the size of a given flow field, and the 
relative level of cavitation. Therefore we write 


N/nu a2 = £(c/fo_) 
0) c 


Thus a normalized version of Eq. (8) would be 


——_——F = £(o/o |) G(fr 7) (9) 


It is difficult to obtain appropriate scaling 
factors for R_ and T_ in a turbulent shear flow. 

The problem iS discussed briefly by Arndt and Keller 
(1976). Lacking more detailed information, the 
following assumptions can be used 


mR os Gl 
m 


aiee 
Be {UG 


If we interpret S(f) as the mean square acoustic 
pressure in a frequency band Af, Eq. (9) can be 
written in the form 


p 2/p2u “) 2 
eee, 


Afa/U 3 
( / By ond 


1 
a2) 


Se (a/o.) G(f£d/Uo ~) (10) 


Blake et al. (1977) circumvented the requirement 
of measuring n. They reasoned that 


2 —— 
= = 2 
i G(£) dt i i Che SES Ish, 


wherein Pp? is the time mean square of pp and y,, 

is the total lifetime of the bubble (including 
growth, initial collapse times and rebounding times) . 
Further, they simply reasoned that 


or that 


S(e 32) = We eee _)) 
() (0) (o) 


This results in the normalized spectrum 


Di 2 
pi” (£,4£) yt or 

S(t, £) SSS (11) 
A£N R 4op 


Making the same assumptions as before, we would 
expect that 


a ee G! (£a/vo®) (12) 
(hea/o,) 53/2” 2 


Blake et al. were able to determine S(tgf) for the 
case of noise due to cavitation on a hydrofoil 
using measured values of Ry. They assumed N equal 
to unity and found that Eq. (11) resulted in 
excellent collapse of the data. 

Arndt (1978) used Eq. (12) to normalize cavitation 
data previously reported by Arndt and Keller (1976). 
These data correspond to noise from cavitation in 
the wake of a disk and were collected under a 
variety of conditions in both a water tunnel and in 
a depressurized towing tank. Both the level of 
dissolved gas and the number of free nuclei were 
monitored. As shown in Figure 13, the normalization 
is not very successful. It would appear that Eq. (10) 
would be more effective in taking all of the 
variables into account. However, n could only be 
measured in unison with acoustic observations in the 
water tunnel. Because of the nature of the laser 
scattering measurements used to determine n in the 
depressurized towing tank, these measurements had 
to be made separately from the acoustic measure- 
ments. The assumed form for S(£T)) in Eqs. (10) and 
(11) varies by a factor na3/o's. As an example, n in 
the depressurized towing tank appeared to be rela- 
tively constant and equal to about 15/cm?. Therefore 
the factor nd3/o% was found to have a maximum varia- 
tion of 23 dB. This does not account for the scatter 
shown and one can only assume that there are other 
complicating factors. It should be emphasized that 
these data were collected under carefully controlled 
conditions. This underscores the fact that the 
current state of knowledge in this area is poor. 


6. CONCLUSIONS 


Cavitation inception in turbulent shear flows is 
the result of a complex interaction between an 
unsteady pressure field and a distribution of free 
stream nuclei. There is a dearth of data relating 
cavitation inception and the turbulent pressure 
field. What little information that is available 
indicates that negative peaks in pressure having a 
magnitude as high as ten times the root mean square 


4 

(dB) 
foe} 
(eo) 


joerc ere re are) 


ae 2|5 100F 
| Water Tunnel Dao U %/ed 1 4 Gas Content nd 
Bcm 6M/sec 0.76 I5S/cem, 39ppm 

I20F 2cm  4M/sec 0.80 4I/m> 45ppm 

2cm  4MYsec 0.63 219/m> 7.5ppm 4 


140 Bem 4/sec 064 75/em> 6 %ppm J 
Vacu -Tank 


ie 16cm 30™/sec 0.66 — 10 ppm 4 
160 fem 3.eM/sec O66 — : 10 ppm ; ca 
10 100 1000 10000 100000 
fel 1 
Uo ola 


FIGURE 13. Normalized cavitation noise spectra. 


338 


pressure can excite cavitation inception. This fact 
alone indicates that consideration should be given 
to the details of the turbulent pressure field. 
The available evidence indicates that two basic 
factors related to the pressure field enter into 
the scale effects. First, as the scale of the 
flow increases, cavitation nuclei are relatively 
more responsive to a wider range of pressure 
fluctuations. Secondly, the available evidence 
indicates that large deviations from the mean 
pressure are more probable with increasing Reynolds 
number. This would explain some of the observed 
increases in cavitation index with physical scale. 
In view of the almost total lack of information on 
the statistics of turbulent pressure field (aside 
from some correlation and spectral data) and the 
potential importance of this knowledge to under- 
standing cavitation, it is strongly recommended 
that careful experiments be initiated to remedy the 
situation. Such experiments have been proposed by 
George (1974b, 1975). 
Direct application 
tion to cavitation is 


of the pressure field informa- 
unfortunately clouded by gas 
content effects which also increase the cavitation 
index with increasing exposure time. The fact that 
a reasonably precise scaling law for cavitation 
noise has not yet been found (perhaps a consequence 
of the lack of knowledge about the pressure field) 
further complicates interpretation of experiments 
and theory. Therefore it is also strongly recom- 
mended that the problem of the response of cavita- 
tion nuclei to turbulence receive particular attention. 
Such experiments have been proposed by Arndt (1978). 


ACKNOWLEDGMENTS 


R. E. A. Arndt gratefully acknowledges the support 
of the Air Force Office of Scientific Research and 
the Seed Research Fund of the St. Anthony Falls 
Hydraulic Laboratory. W. K. George gratefully 
acknowledges the support of the National Science 
Foundation under grants from the Engineering (Fluid 
Dynamics) and Atmospheric Sciences (Meterology) 
Programs and the Air Force Office of Scientific 
Research. Both authors are grateful to Mrs. Sandra 
Peterson who typed the manuscript. 


REFERENCES 


Arndt, R. E. A. (1974a). Cavitation inception and 
how it scales: a review of the problem with a 
summary of recent research. Proc. Symposium on 
High Powered Propulsion of Ships, Wageningen, 
The Netherlands, December (available from the 
Netherlands Ship Model Basin). 


Arndt, R. E. A. (1974b). Pressure fields and 
Cavitation. Trans. 7th IAHR Symposium, Vienna, 
September. 

Arndt, R. E. A. (1976). Semi-empirical analysis of 


cavitation in the wake of a sharp edged disk. 
J. of Fluids Engr., September. 
Arndt, R. E. A. (1977). Cavitation and erosion: 
an overview. Proc. Corrosion/77, Natl. Assoc. 
of Corrosion Engineers, San Francisco, Calif. 
Arndt, R. E. A. (1978). Investigation of the effects 
of dissolved gas and free nuclei on cavitation 
and noise in the wake of a sharp edged disk. 
Proc. Joint IAHR/ASME/ASCE Symposium on Fluid 
Machinery, Ft. Collings, Colorado. 


Arndt, R. E. A., and J. W. Daily (1969). Cavitation 
in turbulent boundary layers. Proc. of Symposium 
on Cavitation State of Knowledge, ASME, 64-86. 

Arndt, R. E. A., and A. T. Ippen (1967). Cavitation 
near surfaces of distributed roughness. MIT 
Hydrodynamics Laboratory Report No. 104. 

Arndt, R. E. A., and A. T. Ippen (1968). Rough 
surface effects on cavitation inception, J. of 
Basic Engr., Trans. ASME, Series D, 249-261. 

Arndt, R. E. A., and A. Keller (1976). Free gas 
content effects on cavitation inception and 
noise in a free shear flow. Proc. IAHR Symposium 
on Two Phase Flow and Cavitation in Power 
Generation Systems. 

Baker, ©. B., J. W. Holl’, and R.\E. Aj Arndt (@o7o)e 
The influence of gas content and polyethelene 
oxide additive upon confined jet cavitation in 
water. ASME Polyphase Forum, March. 

Baiter, J. J. (1974). Aspects of cavitation noise. 
Proc. Symposium on High Powered Propulsion of 
Ships, Wageningen, The Netherlands, Dec. 


Batchelor, G. K. (1951). Pressure fluctuations in 
isotropic turbulence. Proc. Camb. Phil. Soc., 
4727) 359-314. 

Batchelor, G. K. (1953). The Theory of Homogeneous 
Turbulence, Cambridge University Press. 

Batchelor, G. K., and A. A. Townsend (1949). The 
nature of turbulence at large wavenumbers. Proc. 


Royal Soc. (London), A199, 1057; 238-255. 

Beuther, P., W. K. George, and R. E. A. Arndt (1977a). 
Modelling of pressure spectra in a turbulent 
shear flow. 93rd Meeting Acoustical Soc. of 
Amer., State College, June. 

Beuther, P., W. K. George, and R. E. A. Arndt (1977b). 
Pressure spectra in a turbulent shear flow. 
Bulletin Amer. Phys. Soc., 22, 12, December. 

Beuther, P., W. K. George, and R. E. A. Arndt (1977c) . 
Pressure spectra in homogeneous isotropic 
turbulent flow. Bulletin Amer. Phys. Soc., 22, 

12, December. 

Blake, W. K., M. J. Wolpert, and F. E. Geib (1977). 
Cavitation noise and inception as influenced by 
boundary layer development on a hydrofoil. J. 
Fluid Mech., 80, Part 4, 617-640. 

Bradshaw, P., (1967). Inactive motion and pressure 
fluctuations in turbulent boundary layers. J. 


Fluid Mech., 30, 241. 

Chan, Y. Y. (1974a). Pressure sources for a wave 
model of jet noise. ATIAA J., 12, 2. 

Chan, Y. Y. (1974b). Spatial waves in turbulent 
Jets eRUST CSO Ger laTaS el 7a el. 

Chan, Y. Y. (1976). Noise generated wavelike eddies 


in a turbulent jet. JICAS paper, No. 76-42. 

Epstein, P. S., and M. S. Plesset (1950). On the 
stability of gas bubbles in liquid gas solutions. 
J. Chem. Phys., 18, 11, November, 1505-1509. 

Fitzpatrick, H. M., and M. Strasberg (1956). Hydro- 
dynamic sources of sound. Proc. lst Symposium 
on Naval Hydrodynamics, Natl. Acad. of Sciences, 
Natl. Res. Council, Publication 515, September. 

Fuchs, H. V. (1972a). Measurements of pressure 
fluctuations within subsonic turbulent jets. 
Ui, Wloletere! Etetol WAllaa np 2An So 

Fuchs, H. V. (1972b). Space correlations of the 
fluctuating pressure in subsonic turbulent jets, 
Wn Clotdel Eveel WAllon 5» 2p Ip Tis 

Fuchs, H. V. (1974). Analysis of circumferentially 
coherent pressure fluctuations relevant to jet 
noise. Colloquium on Coherent Structure in 
Turbulence, Southampton. 


Fuchs, H. V., and A. Michalke (1975). On turbulence 


and noise of an axisymmetric shear flow. J. 
Fluid Mech., 70, 1. 

George, W. K. (1974a). The equilibrium range of 
turbulent pressure spectra, Bull. of the Amer. 
PAYS}. SOC PLO eS 8) 

George, W. K. (1974b). 
dynamics Research. 

George, W. K. (1975). Proposal to ONR, Fluid 
Dynamics Branch and presentation at David H. 
Taylor Naval Ship Research and Development Center, 
February. 

Gurvich, A. A., and A. M. Yaglem (1967). Breakdown 
of eddies and probability distribution for small 
scale turbulence, Phys Fluids, 10, Supplement, 
S59-S65.- 

Heisinberg, W. (1948). Zur statistichen theorie 
den turbulenz. F. Phisik, 124, 7-12; 359-374. 

Holl, J. W. (1960). An effect of air content on 
the occurrence of cavitation. J. Basic Engr., 
Trans. ASME, Series D, 82, 941-946. 

Holl, J. W. (1969). Limited cavitation. Proc. of 
the Symposium on Cavitation State of Knowledge, 
ASME, June, 26-63. 

Holl, J. W. (1970). Nuclei and cavitation. fran. 
ASME, J. of Basic Engr., December, 681-688. 

Jones, B. G., R. J. Adrian, C. K. Nithianandan, 
and H. P. Planchen, Jr. (1977). Spectra of 
turbulent static pressure fluctuations in jet 
mixing layers, ATAA Paper No. 77-1370. 

Keller, A. P. (1974). Investigations concerning 
scale effects of the inception of cavitation. 
Proc. Conf. on Cavitation, Inst. of M. E., 
Heriot-Watt University, September. 

Kermeen, R. W., and B. R. Parkin (1957). Incipient 
cavitation and wake flow behind sharp-edged 


Proposal to General Hydro- 


339 


disks. Report 85-4, Hydrodynamics Laboratory, 
California Institute of Tech., August. 

Kim, H. T., S. J. Kline, and W. C. Reynolds (1971). 
The production of turbulence near a smooth wall 
in a turbulent boundary layer. J. of Fluid 
Mech., 50, 133. 

Kolmogorov, A. N. (1962). A refinement of previous 
hypotheses concerning the local structure of 
turbulence in a viscous incompressible fluid at 
high Reynolds number. J. Fluid Mech., 13, 1; 
82-85. 

Kraichnan, R. H. (1956). Pressure fluctuations in 
turbulent flow over a flat plate. J. Acoust. 
Soc. Amer., 28, 3. 

Lienhard, J. H., and C. D. Goss (1971). Influence 
of size and configuration on cavitation in sub- 
merged orifice flows. ASME Paper No. 71-FE-39. 

Monin, A. S. and A. M. Yaglem (1975). Statistical 
Fluid Mechanics: Mechanics of Turbulence, Vol. 
II, MIT Press, Cambridge. 

Obukov, A. M. (1949). Pressure fluctuaticns in a 
turbulent flow. Dokl. Acad. Nauk., SSSR, 66, 


1S el Os 

Plesset, M. S. (1949). The dynamics of cavitation 
bubbles. J. of Applied Mech., September, 277- 
282. 

Ross, D. (1976). Mechanics of Underwater Noise, 
Pergamon Press. 

Rouse, H. (1953). Cavitation in the mixing zone of 


a submerged jet. La Houille Blanche, February. 
Tennekes, H., and J. L. Lumley (1972). A First 
Course in Turbulence. MIT Press. 
Willmarth, W. W. (1975). Pressure fluctuations 
beneath turbulent boundary layers. Annual 
Review of Fluid Mech., 7, Palo Alto, Calif. 


Secondary Flow Generated 
Vortex Cavitation 


Michael L. Billet 


The Pennsylvania State University 
State College, Pennsylvania 


ABSTRACT 


Secondary flow theories are employed to calculate 

the secondary vorticity near the inner wall of a rotor 
for several flow conditions. This calculated vortic-— 
ity is used in a simple vortex model to calculate the 
minimum pressure coefficient of the resulting vortex 
behind the rotor. The influence of inflow velocity 
distributions on the generation of secondary vortic-— 
ity is discussed. Comparisons are given between the 
calculated pressure coefficients and the measured 
cavitation indices of the vortex. 


1. INTRODUCTION 


Secondary flows generate additional streamwise vor- 
ticity when a boundary layer flow is turned by a 
rotor. The apparent effect of this additional vor- 
ticity is evidenced by the high cavitation numbers 
of the vortex formed downstream of the rotor plane. 
One example of the cavitation associated with a 
vortex can be found in the draft tube of a Francis 
turbine operating in the part load range. The 
cavitation depends directly on the square of the 
streamwise vorticity associated with the vortex. In 
most cases, the critical cavitation numbers typical 
of this vortex are often higher than those associ- 
ated with any other type of rotor cavitation. 
Previous experimental results have shown that a 
cavitation inception prediction of this vortex is a 
very difficult problem. All rotors operating with 
a wall boundary layer have a vortex ‘along the inner 
wall. The appearance of this cavitating vortex varies 
from rotor to rotor. The critical cavitation number 
can vary aS much as an order of magnitude. Small 
variations in the wall boundary layer can cause a 
significant change in the critical cavitation number. 
Some confusion in cavitation inception data asso- 
ciated with this vortex is due to a confusion of 
types of cavitation, i.e., vaporous versus nonva- 
porous cavitation. Vortex flows tend to be good 


340 


collectors of gas bubbles which can cause non- 
vaporous cavitation. This often leads to confusing 
nonvaporous for vaporous cavitation giving high 
cavitation numbers. In general, results indicate 
for vaporous- limited cavitation that 


< 
Tg = SC oee (1) 


Thus, the minimum pressure coefficient is of partic- 
ular importance in a study of vortex cavitation in- 
ception. 

It is appropriate then to find a simple descrip- 
tion of the vortex in order to calculate its minimum 
pressure coefficient. Unfortunately, the vortex is 
composed of a finite number of vortex filaments 
and a difficulty arises in specifying this number. 
This is particularly difficult when the vortex exists 
in the low pressure region near the inner wall of the 
complicated flow behind a rotor. In this region, 
there are vortex filaments in the primary flow in 
addition to the secondary vortex filaments which can 
influence this vortex. The combined effect of these 
filaments is to induce a swirl velocity distribution, 
Vg, which can be easily measured. 

Some preliminary tests show that in many cases 
small changes in the incoming velocity profile near 
the inner wall cause large differences in the crit- 
ical cavitation number of the vortex. Measurements 
of the primary flow field show only a change in down- 
stream velocity profile near the inner wall. This 
is especially true if the rotor was designed to be 
unloaded near the inner wall. For these cases, 
changes in the critical cavitation number can be 
directly related to changes in the secondary vortic- 
ity near the rotor inner wall. 

The secondary vorticity can roll-up into a vortex 
like flow in the blade passage or it can simply com- 
bine with other vortex filaments aft of the rotor to 
form a larger vortex flow. In either case, there 
will be a circulation and a characteristic dimension 
of the passage vorticity which will determine the 
critical cavitation number of the resulting vortex. 


In this paper, a brief summary is given of the 
method for calculating the secondary vorticity in 
the blade passage with comparisons to flow field 
measurements. Initially, the primary flow field 
through the rotor had to be determined in order to 
calculate the passage secondary vorticity. This was 
accomplished by using a streamline curvature method. 
Flow field results are given in detail for one basic 
flow configuration so named Basic Flow No. 1. Com- 
parisons between the calculated minimum pressure co- 
efficients and measured critical cavitation indices 
are given for several basic flow configurations or 
inflow velocity distributions. 


2. CALCULATION OF FLOW FIELD 
Primary Flow Field 


A schematic of the calculation procedure for the 
flow through a rotor is given in Figure 1. This 
outlines the iterative procedure for the calcula- 
tions and indicates the point at which refinements 
to the deviation angle are necessary and where 
secondary flow calculations are employed. 

It is important to realize that in this discus- 
sion the flow field is being solved for a given 
rotor configuration. For this case, the boundary 
conditions are (1) the geometric or metal angles 
of the blades, (2) the rpm of the rotor, (3) the 
velocity profile far upstream of the rotor plane, 
and (4) the bounding streamlines of the flow. 

After solving for the bounding streamlines, the 
iterative calculation procedure is started by 
establishing the velocity profile far upstream of 
the rotor. The initial conditions (Step 1) to the 
solution for this boundary condition are (1) bounding 
streamtube and (2) velocity profile in rotor plane 
without rotor. With this information, the initial 
streamlines without rotor can be calculated using 
the streamline curvature equations (Step 2). The 
result of this calculation is the boundary condition 
of an initial velocity or energy profile at a station 
far upstream of the rotor plane. 


CALCULATION OF PRIMARY FLOW FIELD 


STEP 1 INITIAL CONDITIONS 
STEP 2 CALCULATION OF FLOW WITHOUT RoToR | 
STEP 3 FIRST ESTIMATE OF ROTOR OUTLET ANGLE 


Si? J—= CALCULATION OF FLOW FIELD WITH ROTOR 


Sup 7 SECONDARY FLOW CALCULATION 
STEP 8 THIRD ESTIMATE OF ROTOR OUTLET ANGLE 


STEP 9—— | FINAL CALCULATION OF FLOW FIELD WITH ROTOR 


FIGURE 1. Schematic of calculation procedure for 
primary flow field. 


341 


AG 


VELOCITY PROFILE 
(B.C. #3) 
y— BOUNDING STREAMLINES (B.C, #4) 


METAL ANGLES OF — 
BLADES (B.C. #1) 


IER eT = 


ROTOR RPM (B.C. #2) 


FIGURE 2. Schematic of boundary conditions. 


Knowing the blade metal angles, the first estimate 
of the flow outlet angles (Step 3) can be calculated. 
These flow outlet angles depend on the blade metal 
angles and on a deviation angle. The deviation 
angle correlation developed by Howell as discussed 
in Horlock (1973) is initially applied. This 
relationship considers only thin blade sections and 
assumes that each blade secticn operates near design 
incidence. As shown in Figure 2, all of the boundary 
conditions are now known and the flow field can be 
solved with the rotor included (Step 4) by using 
the streamline curvature equations [McBride (1977)]. 

Once a converged solution is obtained for the 
flow field using Howell's deviation angles (Step 4), 
the axial velocity distribution is known whereby the 
inlet angles can be estimated in addition to the 
acceleration through the rotor. Now a second 
estimate of the rotor outlet angles (Step 5) can be 
made. For this deviation angle, the effects of 
acceleration, Aéd', blade camber, 69, and blade 
thickness, Aé*, are calculated separately. For the 
calculation of the deviation term due to axial 
acceleration through the rotor, an equation developed 
by Lakshminarayana (1974) is applied. For the 
calculation of deviation terms due to camber and 
thickness effects, the data obtained by the National 


‘Aeronautics and Space Administration [Lieblein 


(1965)] are used. The result is an improved outlet 
flow angle profile which can be used to again calcu- 
late the flow field (Step 6). 

The converged solution of the flow field (Step 6) 
is then used to solve the secondary vorticity 
equations (Step 7) and to determine a deviation term, 
Aéds, which is due to nonsymmetric flow effects. The 
details of the secondary flow calculations will be 
discussed later in this paper. An improved outlet 
flow angle profile (Step 8) is obtained by adding 
this secondary flow term to the deviation terms 
thus far calculated to obtain 


Bo* = Bo — AS' + AS* + bo + AS. (2) 


where 85* is the outlet flow angle and 8) is the 
blade metal outlet angle. This outlet flow angle 
distribution is then used as a boundary condition 
in the calculation of the flow field (Step 9). 

Finally, all of the deviation angle calculations 
are checked based on the flow field calcvlated in 
Step 9. If the angles did not change significantly 
then the result obtained in Step 9 is used as the 
final flow field. 

In all, twenty-eight streamlines were calculated 


342 


© WITHOUT UPSTREAM STRUTS 
WITHOUT SCREEN 
WITHOUT ROTOR 


© WITHOUT UPSTREAM STRUTS 
WITHOUT SCREEN 
DESIGN FLOW COEFFICIENT 
(BASIC FLOW NO. 1) 


DISTANCE FROM 
SURFACE, Chie 


R' (inches) CALCULATED PROFILE 
inc 


—_—— 


r or 
02 OC HS WE a ke 


AXIAL VELOCITY RATIO, Vien 


FIGURE 3. Comparison between velocity profile with/ 
without rotor. 


through the rotor with the first streamline being 

at the inner wall and the last streamline going 
through the rotor tip. The streamlines were spaced 
more closely near the inner wall because the second- 
ary flow calculations are most important near the 
wall. Also, the streamline curvature equations are 
inviscid so that there is a finite velocity at the 
inner wall streamline. 

A sample of the calculations for the flow field 
is given in Figures 3,4, and 5 for the flow config- 
uration called Basic Flow No. 1. For.Basic Flow 
No. 1, the boundary layer entering the rotor is 
axisymmetric with no upstream distribution such as 
screens or struts forward of the rotor which is 
operating at its design flow coefficient. In Figure 
3, the calculated axial velocity profile in the 
plane of the rotor without the rotor and the calcu- 
lated axial velocity profile in front of the rotor 
with the rotor operating on design is shown. In 
addition, experimental data measured in the 48-inch 


(a) WITHOUT UPSTREAM STRUTS 
WITHOUT SCREEN 
DESIGN FLOW COEFFICIENT 
3 (BASIC FLOW NO. 1) 
DISTANCE 
FROM 
SURFACE, 2 DATA 
R' (inches) 
ake 
VE 
l ined 
CALCULATED 
PROFILE 
° 
0 rt A (PAL al 
0 0.2 #O4 0.6 0.8 1e Ome Zed 
VELOCITY RATIOES, V5 NES AND Vy 
FIGURE 4. Rotor outlet velocity profiles for basic 


flow no. l. 


WITHOUT UPSTREAM STRUTS 
WITHOUT SCREEN 

DESIGN FLOW COEFFICIENT 
(BASIC FLOW NO. 1) 


DISTANCE 
FROM 
CENTERLINE, 5 max 


R’ (inches) 


CALCULATED PROFILES 


n°) 

(0) 

O° OF O41 G6 O86, RO be a 
8 


x 
VELOCITY RATIOS, —— AND) — 
Ves co 


4 


FIGURE 5. Tangential and axial velocity profiles 
at cap. 


water tunnel by a LDA system are given for a com- 
parison. In Figure 4, the calculated outlet velocity 
profiles are shown with comparison to measured data. 
Finally, Figure 5 shows the calculated and measured 
tangential velocity, component, Vg, downstream of the 
rotor plane where cavitation occurs under certain 
flow conditions. In general, the flow field calcu- 
lations show very good agreement with the experi- 
mental data. 


Secondary Flow Field 


The major equations used in the streamline curvature 
method for calculation of the flow field were derived 
from the principles of conservation of mass, momentum, 
and energy. The fluid was assumed to be incompress-— 
ible, inviscid, and steady. In addition, the flow 
field was assumed to be axisymmetric. 

The resultant equations allow for streamline 
curvature and for vorticity in the flow. However, 
it is important to realize that the solution to the 
flow field does not contain all of the vorticity. 

In particular, only the circumferential vorticity 
is totally included. The other components of 
vorticity contain derivatives with respect to the 
circumferential direction which are assumed to Ee 
zero. As discussed by Hawthorne and Novak (1969), 
the neglected vorticity terms can be related to the 
secondary flows that occur in the blade passage 
along the inner wall. 

Using the generalized vorticity equations, Lak- 
shminarayana and Horlock (1973) derived a set of 
incompressible vorticity equations valid for a 
rotor operating with an incoming velocity gradient. 
Their expressions for the absolute vorticities, 

Ws', Wn', defined along relative streamlines, s', 
n', were modified for the boundary conditions imposed 
by this problem and were integrated. The resulting 
equations are 
W) ap} 
Oe Sapte ead (3) 
2 abo 


and 


2 2 
AAW OY 22. Hue 
Oo = Wo WR" ds' + Wo 3 ds! 
1 1 
2 
1 ' 
W 
-wW ecla ds' + w_! sts (4) 
2 2 S$] Wy 


where the primes refer to a rotating frame of 
reference and the subscripts, 1, 2, refer to com- 
puting stations along a streamline within the rotor. 
As shown in Figure 6, s', n', b' represent the 
natural coordinates for the relative flow, W is the 
relative velocity, we' and W,' are absolute vorticity 
resolved along the relative streamline, s', and the 
principal normal direction, n', 2 is the rotor 
rotation vector, and R' is the radius of curvature 
of the relative streamline. 

The means by which the streamwise component of 
vorticity is produced in this relative flow are 
similar to those discussed by many investigators 
for a stationary system. However, it is important 
to note that additional secondary vorticity is 
generated when x W has a component in the relative 
streamwise direction. Rotation has no effect when 
the absolute vorticity vector lies in the s'-n' 
plane and the rotation, 9%, has no component in the 
binormal direction, B'. 

These equations were employed to calculate the 
secondary vorticity along a relative streamline 
through the rotor. All of the quantities in the 
equations were calculated by an iterative procedure 
using the primary flow calculations. The initial 
normal component of absolute vorticity, Wn , for a 
streamline was calculated from the incoming axial 
velocity profile to the rotor. In all, the vorticity 
along twenty-eight streamlines was calculated. 

As an example, Figure 7 shows the importance of 
each term in Eq. 4 in the rotor exit plane for Basic 
Flow No. 1. The sum of these terms is given in 
Figure 8. The secondary passage vorticity is the 
difference between the exit vorticity, Ws5, and the 
inlet vorticity, Ws}, along a streamline. 


CALCULATION OF FLOW FIELD THROUGH ROTOR IN RELATIVE COORDINATE SYSTEM 


VELOCITY COMPONENTS 


Ww 
Vx, 
U 


BLADE ROW 


VORTICITY COMPONENTS ROTATION COMPONENTS 


STREAMLINE 
STREAMLINE 


Wr, Ty 
| a2) a 


1 
/B2 


s. 


FIGURE 6. Description of relative coordinate system. 


343 


3 T T T Ta r t~— 
WITHOUT UPSTREAM STRUTS 
Q @ WITHOUT SCREEN 
o 2u, ds 20,4, ds! DESIGN FLOW COEFFICIENT 
3 = (BASIC FI i 
FE - Wo liwRumenis C FLOW NO. 1) 
€ 
z 2 22_,u, ds! a 
=f | 
2 qe W 
z : 
a 
> O 
= 
Oo 
& 
wot = 
r= 
< c) 
nn 
a 
CALCULATED DATA 
0 n L | 
6 5 “4 3 -2 -1 0 1 2 3 
uy RR 


RELATIVE ABSOLUTE STREAMWISE VORTICITY, G, = 


FIGURE 7. 
ge} dhe 


Streamwise passage vorticity for basic flow 


The effect of this additional vorticity, Ws - 
Ws}, is to induce secondary velocities which are 
assumed to occur at the exit plane of the rotor. It 
is important to note that the normal component of 
vorticity, Wndr is accounted for in the axisymmetric 
flow analysis. Thus, only streamwise secondary 
vorticity calculated as a function of radius influ- 
ences the flow field. 

The effect of the streamwise component of vorti- 
city within the blade passage is similar to that 
obtained in the flow through a curved duct [Hawthorne, 
(1961), Eichenberger, (1953)]; however, there is 
the difficulty of devising a reasonable approximate 
method of satisfying the Kutta-Joukowski condition 
at the exit of the rotor. The method used in this 
investigation assumes that the flow is contained in 
a duct defined by the blades and streamlines of the 
primary flow leaving the exit of each blade. In 
this exit plane, a flow solution devised by Hawthorne 
and Novak (1969) was applied. The secondary stream- 


Sire 
WITHOUT UPSTREAM 
STRUTS 
WITHOUT SCREEN 
re DESIGN FLOW COEFFICIENT 
c (BASIC FLOW NO. 1) 
fe C= =| 
wwf 
oO 
fhe 
[4 
—) 
wn 
= 
is) 
& 
seals | 
= EXIT STREAMWISE 
=< VORTICITY 
= pe 
QB a, 
a ) 
1 i L 
-4 3 =a = 0 1 2 3 
Wo Rp 
RELATIVE STREAMWISE VORTICITY, @, = 
Sy We 
FIGURE 8. Relative passage streamwise vorticity at 


rotor exit plane for basic flow no. 1. 


344 


wise vorticity was divided into tangential and axial 
components whereby the former, (We5-Ws 1) sinBo, 
causes a radial gradient of axial velocity and the 
latter leads to an equation for a stream function 
describing the radial and tangential velocities in 
the exit plane, r,6. 

The form of the secondary stream function equation 
is 


2 


ay ley 1 02y xd * 
y2y = we OE i ue a) ee CE 
Sed Soe Ste Da 1 pl ae (ete) 
2 C)s) Yr 
* 
- Ge ")secBo = I5((Fe)) 5 (5) 


where Vx is the secondary axial velocity and is 

obtained from the solution of the tangential com- 

ponent of streamwise vorticity. The solution to 

Eq. (5) was found by applying standard differential 

techniques. The solution and the necessary boundary 

conditions will not be discussed in this brief paper. 
The deviation angle due to the secondary flow 

can be calculated using 


N cos? B> UA 
N\ = ou dr 
s 271V Cha (S) 
a 0 


where N is the number of blades and ¥ is obtained 
from the solution of Eq. (5). The axial velocity, 
Vx, and outlet angle, 85, are determined in the 
calculation of the primary flow field. 

The results of the secondary flow calculations 
for various basic flows indicate that the effects 
are significant only near the inner wall where the 
incoming vorticity is the largest. The deviation 
angles calculated for Basic Flow No. 1 are shown 
in Table 1. 


3. CAVITATION EXPERIMENTS 


The cavitation experiments were conducted in the 
48-inch diameter water tunnel located in the Garfield 
Thomas Water Tunnel Building of the Applied Research 
Laboratory at The Pennsylvania State University. In 


Correlation with cavitation data for basic 
1 and 4. 


FIGURE 9. 
flow nos. 


CAVITATION NUMBER 


TABLE 1. Deviation Angles for Basic Flow No. 1 
Normalized Distance 


from Surface Deviation Angles 


R/R AS 
s 
0.00 -5.4 
0.04 =239) 
0.14 -1.0 
0.24 
0.34 
fo} 
0.44 <0.2 
0.54 
0.64 


all cases, desinent cavitation was employed as the 
experimental measure of the critical cavitation 
number. The cavitation in the vortex system occurred 
on the rotor cap. Also, the occurrence of the 
cavitation was very sporadic. 

The air content of 3.1 ppm was chosen for all of 
the cavitation experiments because gas effects are 
reduced and the relative saturation level was always 
much less than unity. Desinent cavitation number 
data were obtained for different incoming velocity 
profiles to the rotor. The incoming velocity profile 
was varied by changes in the configuration of the 
upstream surface in addition to varying the rotor 
flow coefficient. Results were obtained with/without 
upstream struts, with/without a screen on the upstrea 
surface, and on/off design rotor flow coefficients. 
In all, there were sixteen different flow configura- 
tions or Basic Flow Nos. tested. 

Figures 9-11 display the effects on the desinent 
cavitation number over a range of velocities due to 
variations in the inflow velocity distribution. In 
general, the cavitation number increased for in- 
creasing free stream velocity for all flow config- 
urations shown. As shown in Figure 9, the addition 
of upstream struts which consisted of four struts 
placed at the 0°, 90°, 180°, 270° points on the 
upstream surface caused the cavitation number to 


8 T == T T 
WITHOUT SCREEN 
DESIGN FLOW COEFFICIENT 

Le © WITHOUT UPSTREAA) STRUTS 


GQ WITH UPSTREAM STRUTS 
AIR CONTENT - 3.1ppm 


(BASIC FLOW NO. 1) 
(BASIC FLOW NO. 4) 


st 5 Qg— 

oad 
A CALCULATED 

fo} 

T ec agent : 

REFERENCE POINT 
2 
10 15 20 25 x0 35 4 45 


VELOCITY ~ ft/sec 


8 ior aT] Saree Teor 
DESIGN FLOW COEFFICIENT 

WITHOUT UPSTREAM STRUTS 

© WITHOUT SCREEN (BASIC FLOW NO. 1) 
7} 4WITH SCREEN (BASIC FLOW NO, 3) 
WITH UPSTREAM STRUTS 

G WITHOUT SCREEN (BASIC FLOW NO, 4) 
LL. © WITH SCREEN (BASIC FLOW NO. 5) 


wpsahe 
l2pve 
a 


CALCULATED 


CAVITATION NUMBER ~ 04 
° 


10 15 20 25 30 35 4 
VELOCITY ~ ft/sec 


increase. In contrast to this result, the addition 
of upstream screens causes the cavitation number to 
decrease as shown in Figure 10. Data in Figure 11 
show that a decrease in the flow coefficient by 10% 
causes a dramatic increase in the cavitation number, 
whereas a 10% increase in the flow coefficient 
causes the opposite trend which is not shown in the 
figures. Additional cavitation results are given in 
iBslililene, {(lS)745)) o 


4. CORRELATION OF SECONDARY FLOWS WITH THE CRITICAL 
CAVITATION NUMBER 


Because of the complicated flow field where the 
vortex exists, an absolute calculation of Cppiy of 
the cavitating region would be very difficult. The 
minimum pressure associated with the cavitation 
occurs within the vortex which is located along the 
inner wall. This minimum pressure is not only 


© DESIGN FLOW COEFFICIENT (BASIC FLOW NO. 1) 


aah © 10% LOW IN FLOW COEFFICIENT  g 
Malis (BASIC FLOW NO. 2) 
a8|S 
iT] > 
© 3 
2 
a 
S 4 
5 8 
Sool eae 
S ° ae 
= 3 fo} .e os) 8 
S eS 
= 
Co 
al WITHOUT SCREEN 
WITHOUT UPSTREAM STRUTS 
1 
10 15 2 25 0 35 ry) 


VELOCITY ~ ft/sec 


FIGURE 11. Correlation with cavitation data for basic 
flow nos. 1 and 2. 


345 


45 FIGURE 10. Correlation with cavitation data for 
basic flow nos. 1, 3, 4, and 5. 


determined by the vorticity associated with the 
vortex but also by the location of the vortex in 
the primary flow field. 

Considering only the vortex, there are many fac- 
tors which can influence the minimum pressure coef- 
ficient. If one models a vortex by a simple 
rotational core combined with an irrotational outer 
flow, the Cpmin is found to be 


2 


IP 
atm 7S a m=} V7) 


where T is the circulation and r, is the radius of 
the core. Thus, the factors which influence Cpmin 
are those which influence the circulation or core 
size. 

Assuming that secondary flows control the vortex, 
Eq. (7) can be used to predict changes in critical 
cavitation number due to changes in the secondary 
vorticity produced along the inner wall. Therefore, 
Eq. (7) can be arranged into the form 


T 
Bef re Wf 
A Ss CeOUA (8) 
Ga. - 2 
see) iP 
we Von B 


where [T is now the integrated component of stream- 
wise passage vorticity and ro is approximated by the 
characteristic dimension of the resulting passage 
vorticity. The letters A and B refer to different 
flow states. 

The passage streamwise vorticity was calculated 
along several mean streamlines in the blade passage 
by the method outlined in this paper for four basic 
flow configurations which are described in the left 
hand column of Table 2. For all flow configurations 
considered, the results show a large amount of 
streamwise vorticity at the rotor exit plane near 
the inner wall. An example of the exit streamwise 
passage vorticity is shown in Figure 8 for Basic 
Flow No. 1. 

As can be seen in Figure 8, the vorticity near 


346 


TABLE 2 - Vortex Circulation and Core Size Calculated Flow Vorticity Data 


Circulation Characteristic Nondimensional Planar Momentum 
Basic Flows Th Dimension Ratio Thickness 
(Gacsee) Ro (inch) Ue Wi) 6 (inch) 
Basic Flow No. l 
without upstream struts - 11.64 0.81 -0.080 0.85 
without screen 
design flow coefficient 
Basic Flow No. 2 
without upstream struts - 8.23 0.57 -0.091 O)g7/al 
without screen 
0.9 design flow coefficient 
Basic Flow No. 3 
without upstream struts =) 10599 0.20 -0.076 0.94 
with screen 
design flow coefficient 
Basic Flow No. 4 
with upstream struts - 8.29 0.45 -0.102 Gol 


without screen 
design flow coefficient 


the inner wall has a characteristic dimension 
associated with it. A measure of the circulation 
associated with this vorticity can be found by 
integrating the vorticity from the inner wall to 
the radius where the vorticity changes sign. In 
addition, the characteristic dimension of the 
passage streamwise vorticity must be related to the 
difference between the radius where the vorticity 
changes sign and the inner wall radius. The results 
for several basic flow configurations are shown in 
Table 2. Also, the nondimensional ratio, T'/x Voor 
which is a measure of the minimum pressure coef- 
ficient of the vortex is given in addition to the 
planar momentum thickness of the mean boundary 
layer profile entering the rotor for each flow 
configuration. 

In order to make absolute comparisons between 
calculated minimum pressure coefficients and 
cavitation data, a reference point is necessary 
and the effect of Reynolds number must be calculated. 
A reference point for Basic Flow No. 1 of 6 = 2.8 at 
a velocity of 15 ft/sec was chosen. The influence 
of Reynolds number was determined by solving for the 
relative streamwise vorticity at two different free 
stream velocities. For these calculations, a bound- 
ary layer profile at the reference Reynolds number 
was used in one calculation and the boundary layer 
profile at three times the reference number was 
used in the other calculation. 

Now using Eq. (8) with Basic Flow No. 1 as the 
reference point, comparisons between cavitation 
data and Cpmiyn calculated using the passage stream— 
wise vorticity can be made. Some of the results 
are shown in Figures 9, 10, and 11. As can be noted, 
the changes in Cpmin or 06 for the vortex as calcu- 
lated, using secondary flow theory, correlate well 
with the cavitation results. Only the correlation 
with the rotor operating off-design (Basic Flow No. 
2) is poor at the higher velocities. It is felt 
that this is due to primary flow problems. 


5. SUMMARY 


A secondary flow analysis has been developed which 
can be employed to assess the effect of inflow 
velocity distribution on the strength and core 
size of a vortex. This analysis has been success— 
fully applied to a rotor where the secondary flows 
dominate the flow field near the inner wall. 


NOMENCLATURE 
ap' - streamline spacing in bi-normal direction 
Rg - radius of rotor 
WwW - relative velocity 
Bo - relative outlet metal angle 
g% - relative outlet air angle 
Aé':- deviation angle due to axial velocity accel- 


eration 
AS, - deviation angle due to secondary flows 


09 - deviation angle due to blade camber 

og - cavitation number = (Pw - Py) /(1/2pVa~) 

dg - limited cavitation number 

dq - desinent cavitation number 

Wg' - component of absolute vorticity vector in 
relative streamwise direction 

Wn' - component of absolute vorticity vector in 
relative normal direction 

Wp' - component of absolute vorticity vector in 
relative bi-normal direction 

Qn' - component of rotation vector in relative 
normal direction 

%!' - component of rotation vector in relative bi- 
normal direction 

ACKNOWLEDGMENT 


This research was carried out under the Naval Sea 
Systems Command General Hydromechanics Research 


Program, Subproject SR 023 01 01, administered by 
the David W. Taylor Naval Ship Research and 
Development Center, Contract NO001773-C-1418. 


REFERENCES 


Billet, M. L. (1976). Cavitation results for a 
secondary flow generated trailing vortex. 
Applied Research Laboratory TM 76-234. 


Eichenberger, H. (1953). J. Math. and Phys. 32; 34. 


Hawthorne, W. R. (1961). Proc. Seminar Aero. Sci., 
Bangalore, India, 305. 

Hawthorne, W. R., and R. A. Novak (1969). ‘The 
aerodynamics of turbo-machinery. Ann. Rev. 
Fluid Mechanics 1; 341. 


347 


Horlock, J. H. (1973). Axial Flow Compressors, 

R. E. Krieger Company, New York, 55-60. 
Lakshminarayana, B. (1974). Discussion of Wilson, 
Mani, and Acosta - A note on the influence of 
axial velocity ratios on cascade performance. 

NASA SP-304, 127. 

Lakshminarayana, B., and J. H. Horlock (1973). 
Generalized expressions for secondary vorticity 
using intrinsic coordinates. J. Fluid Mech. 59; 
97. 

Lieblein, S. (1965). Experimental flow in two- 
dimensional cascades. NASA SP-36, 209. 

McBride, M. W. (1977). A streamline curvature 
method of analyzing axisymmetrical axial, mixed, 
and radial flow turbomachinery. Applied Research 
Laboratory TM 77-219. 


On the Linearized Theory of 
Hub Cavity with Swirl 


G. H. Schmidt 


Technical University of Delft 


and 


J. A. Sparenberg 


University of Groningen 


The Netherlands 


ABSTRACT 


In general, there is a cavity astern of the hub of 

a ship screw. This cavity is rather stable and is 
roughly in the shape of a long circular cylinder. 
There is circulation about it, which occurs in the 
case of a real screw propeller, when the circulation 
around the blades at their roots is nonzero. 

Because the divergence of the vorticity field is 
zero, this circulation at the roots "flows" down- 
stream in the form of circulation about the hub. 

At the end of the hub the flow contracts and the 
swirl velocity increases. The pressure becomes 
lower and a cavity forms where the pressure decreases 
to the vapor pressure. 

We introduce the following simplifications: 
First, we neglect the influence of the finite number 
of blades and consider a half infinite axially 
symmetric hub immersed in an inviscid and incom- 
pressible fluid. The incoming flow consists of a 
homogeneous part, parallel to the axis of the hub 
in the direction of the endpoint, and of a swirl 
which represents the circulation around the hub. 

In the upstream direction the hub tends to a 
circular cylinder while its radius tends to zero 
towards the end point. Second, our theory will be 
linear: The difference between the radius of the 
hub and the radius of the cavity is assumed to be 
small and quantities which are quadratic in this 
difference will, in general, be neglected. 

Using these simplifications we determine the 
shape of the cavity for given values of, for 
instance, the swirl, the incoming velocity, the 
ambient pressure, and the vapor pressure. The 
surface tension is also included in the general 
formulation of the problem. The more detailed 
considerations, as well as the numerical calculations, 
will be confined to zero surface tension. 

One of the unknowns of the problem is the 
position of the point of separation. This position 
can be determined by demanding that the pressure 
exceeds the vapor pressure everywhere on the wetted 


348 


surface of the hub and by demanding that the flow 
cannot penetrate the surface of the hub. 

The shape of the cavity is roughly a circular 
cylinder. There are waves on the surface of this 
cylinder which are, within the limitations of our 
theory, steady with respect to the hub, and their 
crests and throughs are perpendicular to the axis 
of the hub. We will give numerical results for 
the wavelengths and amplitudes of the waves as 
functions of, for instance, the incoming velocity 
and of the shape of the hub. 


1. INTRODUCTION 


A long cavity generally begins somewhere at the 
end of the hub of a ship screw. This cavity, which 
has circulation around it, does not close or widen, 
it has a rather stable mean value to its radius. 
The circulation or swirl occurs in the case of a 
real screw propeller when the circulation around 
the blades at their roots is not zero. Because 

the divergence of the vorticity field is zero, this 
circulation at the roots "flows" downstream in the 
form of circulation about the hub and then about 
the cavity. 

In order to gain some insight in this phenomenon 
we introduce some simplifications. We neglect the 
influence of the finite number of blades and con- 
sider a half infinite axially symmetric hub immersed 
in an inviscid and incompressible fluid. The 
incoming flow consists of a homogeneous part paral- 
lel to the axis of the hub in the direction of the 
endpoint and of a swirl which represents the 
circulation around the hub. In the upstream 
direction the hub tends to a circular cylinder 
while its radius tends to zero towards the end 
point. Hence, near the endpoint the flow contracts 
and the swirl velocity increases proportional to 
the inverse of the radius. This means that the 
pressure becomes lower and a cavity starts where 
the pressure decreases to the vapor pressure of 


the fluid. Another approximation is that our theory 
will be linear. In order for this theory to be 

valid it is necessary that there be no abrupt changes 
in radius of the hub and cavity. In real fluids the 
viscosity can have an important influence on the 
point of separation [Wu (1972)], however, this 

effect is too complicated to be treated by our 
method. We will not take into account the dependence 
of the local vapor pressure on the curvature of the 
interface between vapor and liquid. Surface tension 
is included in the general formulation of the prob- 
lem. The more detailed considerations, as well as 
the numerical calculations, will be confined to 

zero surface tension. 

One of the unknowns of the problem is the value 
of the axial coordinate of the point of separation. 
This value can be determined by demanding that there 
is no place at the wetted area where the pressure 
is lower than the prescribed pressure in the cavity 
and by demanding that the flow cannot penetrate the 
surface of the hub. 

The problem is very similar to the shrink fit 
problem, in the theory of elasticity, of an unbounded 
elastic medium with a circular two-sided infinite 
hole [Sparenberg (1958)]. This hole is occupied by 
a half infinite axially symmetric rigid body and 
the problem is to calculate the contact pressure 
between the body and surrounding medium when for 
instance shear stresses are supposed to be zero. 
Also, in this case, the edge of the region of con- 
tact has to be determined. 

The way in which we solve our problem is 
analogous to the way in which the aforementioned 
elastic problem can be solved. First we determine 
a Green function. This is, in our case, the 
deformation of the two-sided infinite cavity with 
swirl when a rotationally symmetric pressure of a 
Dirac 6 function type is applied at the circular 
cylindrical wall. By using this Green function as 
a kernel we can write down a Wiener-Hopf integral 
equation for the unknown contact pressure causing 
the fluid flow along the hub. This integral 
equation is solved numerically by the finite element 
method. 


2. EQUATIONS OF MOTION AND BOUNDARY CONDITIONS 
First we consider a two-sided infinite circular 


undisturbed cavity of radius Yor with swirl in an 
inviscid and imcompressible fluid of density ~p. 


FIGURE 1. 


Undisturbed cavity flow. 


The undisturbed velocity field and pressure field are 


~ ~ ~ 1p ~ 
u=U, v=0, w==, p= Po(r), 12 Baap ((aL)) 


349 


where u, Vv, and w are the velocity components in the 
x, xr, and 9 direction, p is the pressure, and T is 


27 times the circulation around the axis. From 
Bernoulli's equation it follows that 
po (x) = ppl? /2xr? (2) 


Po is the ambient pressure in the fluid and p (r) 
> p, for r+. On the wall of the caviity for 
if oP ra we have 


(x) = - pr? 2 = = 
ID We Ie, OE fae a = eye. (3) 
where p_ is the pressure inside the cavity and po 
is the surface tension of the fluid. In the 
following we assume 
= > 0 4 
5 Te (4) 


hence, the ambient pressure at infinity is larger 
than the pressure in the cavity. From (3) it 
follows 


- po + Vp2o2 + 2 pl? (p - P,)- 

Sn et eo ee OY 
c 2 - 5 
(pela Pe) (5) 


We had to choose the positive root under the 
assumption (4). For (p, - Pp.) < 0 we would have 
chosen the negative root, however, this would 
yield an unstable situation. In the case of zero 
surface tension (5) simplifies to 


x, = 1¥p/2(p,-Pe) (6) 


(o} 


The equations of motion for a time dependent fluid 
flow are 


St, = Ot. . ot 1 ap 
ae oP Wee 
De Oe VY Be p dx ! 2) 
= ©. = OF we 1 op 
—+0—4+7V—- —-=- = = 
at : ox M or Yr @ Or 9 (8) 
ow  . dw. ow. ww 
—F+t1—+7—+—= 
DE” “Oe” Y Oe 12 e f 2) 
Also, we have to satisfy 
diy (hy Yo te Be 2a Veo, (10) 
x ag 1G) 


For a disturbed motion which satisfies (7)... (10) 
it remains true that (1) 


Tj 
Wheat (11) 


otherwise a circular contour floating with the 
fluid would change its circulation which is im- 
possible when external force fields inside the 
fluid are absent. This follows also from (9) which 
is satisfied by (11). Hence substituting (11) into 
CAD eseeverene (10) we are left with the following three 
equations for the three unknown functions u,v, and 
Pp, 


350 


Am = Bel ei 1 ap 
oy Wa peseses , (12) 
AVSuh a Ges ta fore eax 
~ ~ ~ ~ 2 
OU Rie On a SRO Cole ay (13) 
at x or (0) he ~ 
dn Oy 
— + — + — = 0 (14) 
ax or ie 


We now linearize these equations with respect to 
the undisturbed swirl flow, 


a=U+u, V=v,p=p,+ Dd, (15) 


where the perturbation quantities widens nie) 6. WA(Ssp3e7 
t), and p(x,r,t,) are supposed to be of OE) FSub- 
stituting (15) into (12)...(14), neglecting terms 
of O(c2) and using (2) we find 


We Geog a Se, (16) 
at ox p dx 
ov av 1 ap 

+ =a A aL7/ 
at uv Ox po or nT) 


<¥+V%;42%=0 (18) 


Because the (u, v) velocity field is without rota- 
tion we can write 


(a, v) = Ge, 2%) (19) 


where ¢ = $(x, r, t) is a scaler potential function 
satisfied by (18) 


(20) 


We now suppose the disturbed cavity wall to be at 
fa Sac ab wep (21) 
c c 


where 6r (x, t) is O(€). On this axial symmetric 
boundary we demand the difference between the 
pressures inside the cavity and in the fluid to be 
in equilibrium with the effect of the surface ten- 
sion and with some still unspecified external 
normal loading 0U*E£ (x, t) of the cavity wall, 


i 
R 


= =\ 0} 
12) 19 p ( 


where R,and Ry are the principle radii of curvature 
of the boundary, reckoned positive when the centers 
of curvature are at the side of the cavity. 

Within the accuracy of our linearized theory we 
can put 


2) 4 Meee eee ef Oe, (a). 
1 Ro Cc c 


Substituting (23) into (22) and using (2) and (15) 
we find 


il 
Pye (OR? /2r0) + P=) Re a. OPyS — sen On.) cua 


[-) 


16 SB ae Org. (24) 


Expanding the functions of r in (24) with respect 
to 6r,, neglecting second order quantities, and 
using (3) the boundary condition (24) changes into 


2 2 en 
er op a 2 
= (- ——+ = + = 
p(x, roe t) ( im + 5 )ox. 0 a5 ox. + pUCE. 
(co) Cc 


From (16) we find, because p + 0 and $¢ > O for x 


+ - om, 


yp So? OW Seo Ss (26) 


which is Bernoulli's law for the unstationary lin- 
earized flow. Herewith the dynamical boundary 
condition (25) becomes 


32 > 
+—= (—-- — -o——_ - 
U5 5 (3 = ) br. oe 5 ox, Use (27) 


The kinematical condition at the boundary of the 
cavity is 


a 3 ad 
— + — 6 = — 
at ore y ox ae Gre = We) 


Hence, we must solve (20) under the conditions (27) 
and (28) while » > 0 for r > © and for x > - ~. 


3. THE GREEN FUNCTION 


We suppose the dimensionless loading of the boundary 
(22) to have the form 


(GS 12) SY 229) Bere (29) 


where € is a "Small" positive parameter which has 
no connection with the linearization parameter ec. 
Because our problem is linear we assume 


ét =t 
Di Crate) = (3705) Ny dx (x,t) = éx_(x)e~ Bn (3.0)) 
Then equation (20) and the boundary conditions (27) 
and (28) change into 


Gaz + goo t+ = gp) PORE) = 0, ay 


ee Org (x) = U*E(x), (32) 
To* 


SURES ole elCxee) = Ole (33) 
c or 


a 
edr (x) a Ux 


We introduce the Fourier transform g (1) of a function 


g(x) by 
p00 WE 


= ipx 1 
= dx, =— 
g(u) Von g(x)e x g (x) V2, 


2 
(= reals u? ¢(u,r) =0 . (35) 


Hence, for real yu 


¢ (ur) = Ay (u) K (ule) + Ag (wT (lulz), (36) 


where Ky and Ip are modified Bessel functions. 
Because ¢ > 0 for r > © we have 


Aj (uv) = 0. (37) 


Substitution of (36) with (37) into (32) and (33) 
yields 


2 — 
(iw) x (lulz) aay - Go- 5+ woe ww 
@ ta XG © 
= -U*F(u), (38) 
Ju] Ky {ule ) ay) + (é-inv) 6x, (u) = 0. (39) 


Solving (38) and (39) for 6x, (u) and applying the 
inverse Fourier transformation we obtain 


6x, (x) = 
a = ans 
i £(u) |ulK, (Julzye "du 
a 2 
Jon [(é-ino) 2x, (Lule) + - Sy +u?0) |ulx, (ulze)] 
co c 
(40) 
We now choose 
f(x) = 6(x) , (41) 


where 6(x) is the delta function of Dirac, hence 
£(u) = 1/V20. Next we split the range of integra- 
tion into two parts namely - ~ < uw < O and O < 4u 

< © and neglect terms of 0(&2) in the denominator, 
then we find 


def ee] 
k (x) == 82 C2) | een ee) 


ix 


Ri(Be =" ale 


1 fos) 
Sis J 
(E 


2480 OK (E)-(a+BE~) Ky (E) 
U {e) 


351 


~iEx 
Ae{o] 
ne Ky (E) Le dé 
grate = 
an (E+ “**¥o ) K (6) - (otBE2)K, (E) , (42) 


U 


where a and 8 are dimensionless quantities given by 


T2 
-< Bees. . (43) 


ie Cc r U2 
c 


It can be easily proved that under the assumption 
(4), 


a/B = pf + [2r? w, - p,)/007| > Ale (44) 


In order to find the Green function for the 
stationary case we have to take the limit € > 0 in 
(42). 

We now make some remarks for the case o # O and 
hence 8 # 0. 

First, the integrals in (42) are absolutely con- 
vergent for 0 < x < ». This means that when sur- 
face tension is present Green's function k(x) is 
finite even at the point of application of the 
Singular loading (41). This could be expected 
because the surface tension can be represented by 
a membrane placed at the boundary of the cavity 
and a membrane has the possibility to locally 
sustend such a loading by a jump in its first 
derivative while its deformation is still a contin- 
uous function of x. 

Second, we consider the denominators in (42) for 
€ = 0 and look for positive real roots of 


a K, (é) 
(¢ ap {8 s) = Ki (a) . (45) 


The left hand side of (45) is curved upwards for 

— > 0, while the right hand side is curved down- 
wards. The proof of the latter statement is rather 
complicated and will not be given here. However, 
taking this for granted, it means that there are 
none or two real positive roots, which is analogous 
to the case of ordinary gravity waves with surface 
tension. One of the roots corresponds to a wave 
primarily due to the swirl, the other one to 
capillarity. [Whitham (1973), p.446] 


4. THE CASE OF ZERO SURFACE TENSION 


Green's function (42) in the stationary case for 
zero surface tension, when we take a different 
positive value for € which of course is irrelevant, 


ILS} 
, x 
co 1g 
A ig () © © ae 
GS) dena) ee eee: 
E>0 [(E-i E)K, (E)- a Ky (E)I 


Ky (jive 


[ (+i E)K,(&) = @ Kj, (&)] 


352 


First we investigate the number of poles of the 
integrands for € = 0, hence, the number of positive 
real roots of 


K (&) 
° 


RG, 2 ae 


(47) 


where now (43) a We fire = 2(p, - p)/pu". 

From the well known expansions of K,(&) and 
K,(&) it follows that the right hand side of (47) 
is zero for — = 0 and tends to infinity for § >>. 
We prove that this function increases monotonically 
with &, hence, we have to consider 


2 
re (U3) 
= 65 == . 
K, (&) 


au poe alen 


a— > Kj () 


K, (8) 
K, (6) 


(48) 


Instead of proving that the right hand side of (48) 
is positive we will show that 

Ais (3) iG) =e Sole) | BR) 20. (49) 
This is easily shown to be true for § >. 


when the derivative of (49) is negative the 
function itself has to be positive. 


Hence, 


2K, (6) (K, (E)+ & KE) I/E - Ky>(&) + Ky? (E), (50) 


is negative, since K,(&) for 0 < §€ < ~. This means 
that the right hand side of (47) increases mono- 
tonically, hence, there is one and only one root 
Ss Cre WAY) alin. SS C5 

We will estimate the value of bee 
show that 


Therefore we 


(€+1) K (8) SS LS) 2) Oe (51) 
From well-known expansions for K 
inequality holds for §€ + >. 

left hand side of (51) being 


and Kj, this 
The derivative of the 


(E+1) [Kp (&) - Ki (E)] , (52) 


is clearly negative, and hence (51) holds in 0 < 
— <e, From K,(&) > Kj(&) and (51) it follows 
that the root BS of (47) satisfies 


oe 
24a)? 


(Ns tae try OH DN CAM) (53) 

Second we have to determine at which side of the 
real axis this root is situated when € is small 
but not zero. Consider the denominator of the 
first integral of (46), hence a root of 

(E-i €) Ee MS) 75 K, (€) = 0. (54) | 

The zero in the neighborhood of the real axis of 
(54) is assumed as 
Ts) tn 


ES ae (55) 


where — satisfies (47) or (54) with € = 0. Sub- 
stituting (55) into (54), expanding the modified 


This derivative, 


Bessel functions, and using the definition of Se 
we find 


ii Kee) 


6 &= 5 5 (56) 
(2K) (E,)K (60) -E Ki (E+E K (E,)} 
Hence by (49) we find that 
Im(é +6 5) 22 Oy, (57) 


or the pole of the integrand of the first integral 
in (46) is slightly above the real axis for € small 
and € > 0. In the same way the pole of the inte- 
grand of the second integral in (46) is slightly 
below the real axis. 

Now we want to give a different representation 
of (46). We distinguish between two cases x > 0 
and x < 0. In the case of x > 0 we rotate the 
direction of integration of I, and Ip as follows. 


1, =e M@ paris) pp By = (pmsl) p (58) 
and in the case of x < 0 
at =e (Oped) ip) ath == (Opa) 5 (59) 


From the foregoing it follows, that for x > 0, a 
pole has to be added to TI, as well as to In. The 
question arises: are there still other poles in 
the complex half plane Re — > O which are passed 
by rotating the lines of integration? We now 

shall give a proof that this does not happen. This 
proof was kindly given to us by our colleague Prof. 


Dr. B. L. J. Braaksma. 
Consider the function 
def 
ENS) —— ee K, (6) = oh eS) = = [NY (iE) 

+ (o+1)K, (€)], (60) 
which is real for real values of €. Suppose sj 
with Re sj > 0 and Im s; # 0 is a zero of F(&), 
then also s» = 8; (complex conjugated value) is 
such a zero. The functions K)(s.&), j = 1,2, 
satisfy J 

2. 42 
‘da d : 
a 2S Se Se eee Sy reG.s) = ©, J = 1,2 
ae2 dé 5) J 
(61) 


Multiplying (61) by K,(s_&) with k = 2 for j =1 


and k = 1 for j = 2, we find by subtracting the 
results 

2 2 a d 
(s] - So) EK, (S16) Ky (so&) = ag £11 (S28) ae K, (s1&) 


= K1 (818) $1 (S28) (62) 


Hence = 
2 2 a 
(s] = So) ite K; ($7 &) Ky (sg&) dé = EL (Ky (S28) gpk (S18) 5 
at 
d (s &)] 
= Lene) ae a | 


(63) 


It is easily seen that the right hand side 
vanishes for — > © and because s) and Sp are zeros 


1 


of F(&) (57) this right hand side also vanishes 
for § > 1. Because the integral is positive we 
have found that the assumption Im s; = -Im so # 0 
yields a contradiction. It follows that no other 
residues have to be added to the resulting integrals 
after the rotations as denoted in (58) and (59) 
besides the two we mentioned for x > 0. 

Using some formulae from Watson (1922) we find 


f TL : 
OG at 3) ar Ue J, 68) ak ween 0 & P Op (64) 


iG & G) > len) 2 ea) 75 = Op GS) 


Adding poles to the integrals in the case x > 0 we 
can transform the Green function (46) into 


Avsin, bx. yak > 0 


k(x) = h(x) + , (66) 


where 


ro) _, Ie 


2 e S dé 
ind): SY ers | eee angen ee SES Qe 

7 [ET (E) +05) (6) ) +[EY, (€)+a¥)(E)] (67) 

oO 
and 
2 
AS 28 (Murs ) 5. SB fs (68) 
(o) (o) Omc 
The function h(x) is symmetric, h(x) = h(-x). 


For x > 0 it has a logarithmic singularity because 
for x = 0 the integrand as a function of &, behaves 
as 1/2€, hence 


h (x) = cs Si Sop <a Or (69) 
For x > © the behavior of h(x) depends on the behav- 
ior of the integrand in (67). For & > 0, this 

turns out to be as 


(402/n2E2)+0(E32ne) . (70) 


Then it follows from Doetsch (1943) p. 233 that 
h(x) = Sin cies oc |x| +. (71) 


Now suppose that for x < 0 the shape of the 
cavity is prescribed. 


FIGURE 2. 


Flow with swirl along hub. 


im = ie dp Oe (3) ,» Ow see © (729) 
c c 


and that the unknown pressure between hub and fluid 
US} ie) oF pU2£ (x). 


353 


Then we have to solve the following integral 
equation 


) 
[ k(x-x') £(x')dx'= Ox | (x) 7 $8 < O} (73) 


which is of the Wiener-Hopf type. 


5. THE EXPLICIT SOLUTION OF a K,(&) -€ K (&) =0 
(0) 


In order to find an explicit solution of Eq. (47) 

we first have to make some preliminary considerations. 
Assume the following loading of the otherwise undis- 
turbed cavity boundary 


f(x) = €,6(x), (74) 


where €,; is a small parameter. By (66) we find for 


the deformation of the cavity 
€, A sin bx 7 2820, (75) 
Sx (x) = €, h(x) + 


(0) pm 3% S Oe 


Next we consider 
e(e3) 3 Sep , 1S O 9 GEG) S O7 3 2 Os (76) 


The loading given in (74) is the derivative with 
respect to x of the loading given in (76). Hence 
the derivative of the deformation 6*r,(x) caused 
by (76) has to be equal to (75), we take 


BWeos lope 5 2. Op (77) 


-e] 5 


d*x (x) = =€) hi(Eyidercr 


* 
[> 


, x <0, 


o 


where we have chosen the constant of integration 
in such a way that for x > +” we have a harmonic 
wave with mean value zero. 

Finally consider 


2G) SO »,*s<@ 3 5 82 Op (78) 


£ (x) = €j 
The loading given in (74) is also, in this case, 
the derivative with respect to x of the loading 
given in (78). Hence the same argument applies as 
before. However, now the constant of integration 
has to be chosen so that the disturbance tends to 
zero for x > -~ , we find 


(79) 


A 
oa p (i-cos lop) 5 38 E25 


(0) pn 28S, Os 


Subtraction of the disturbances (77) and (79) yields 


354 


+00 
A 
bar (x) - O*#r (x) = =e | Mae a= 


Cole Xu LCON (80) 
which is constant as could be expected because 
belongs to a constant loading of magnitude -€1pU* 
of the whole cavity. 

This displacement however can be calculated in 
another way by using (6) where we have to replace 
Diy, 


Py + pU*f = Die €)pU*. (81) 


Expanding (6) with respect to €), we find 


ip 
2 ee 
c Cc (p,-P te 1pU?) 2 
T €,0U 
= \[& ules ). (82) 
Oe Sac 2p -p ) 
Po Po C6 
Combining (80) and (82) yields 
+0 
u*r /2(p.-p_) = h(x) dx + A/b (83) 
(o} 2 fe} 


Substituting h(&), A, and b from (67) and (68) into 
(83) and carrying out the integration with respect 
to x we find after some reductions 


-1 2 1 


a = 2(2ata 5.) =L (84) 
where 
a dé 
a 2 2 
EL LET, (€) tad) (E)] FEY (E)+a0¥, (E)] } 
(85) 
Solution of (84) with respect to 5 yields 
a 2L ys 
bj(a) =a. {1 SF ene (86) 


by which we have found the unique solution of (47) 
for € real and & > 0. This derivation rests on 
some mechanical considerations such as uniqueness 
of the solutions in relation to radiation conditions. 
The result however, which is interesting from the 
point of view of zeros of transcedental equations 
connected with Bessel functions, has been verified 
by others in a more straight forward way and found 
to be correct. 

By (86) it follows that an axial symmetric wave 
moving along the cavity with velocity U has a wave- 
length A(U) given by 


d(U) am/b =[mT/E (a)] [20/(p.-Pe) ] 


oO 


2 (P-Po) /pu* (s7) 


’ 


Equation (87) describes the dispersion of these 
waves when surface tension is neglected. 


6. NUMERICAL SOLUTION OF THE INTEGRAL EQUATION 
In the left hand side of (73) the function k(x) is 


given by (66-68) and the dimensionless quantity 
f£(x') is unknown. For x < O the right hand side 


is determined by the geometry of the hub. Let this 
geometry be described by 
= = + 6 me) 9 88 
r ry x) ae x ) (88) 
where r(x) is a given function. Then the right 


hand Sigs is known up to an unknown shift, s, of 
the hub along the x-axis, since the position of 


the point of separation is a priori unknown. Hence 
for x < 0 we can write (73) as 
) 
{ kK (GcR Sexi") (Exe) xa or (Ge a> Sip x <= O 
(89) 
where the function f(x') and s are unknown. First 


we will describe how f£(x') is computed numerically 
from (89) for arbitrary values of s. Then s will 
be determined by a condition to be satisfied by f 
at x = 0. 

We make some remarks concerning the behavior of 
£(x') for x'tO and for x' > -©. As will be shown 
in the Appendix, the behavior of f(x') near the 
origin is, for arbitrary values of s, 


B Ieee US H(0) (90) 


eG) 2 Pi 
where B is some constant which will be discussed 
later. 

The hub has a constant radius far upstream, 
hence 6r (x) tends to a finite value for x > -~, 
Since the kernel k(x) vanishes for x > -~, the 
perturbation due to the end part of the hub van- 
ishes far upstream. Hence, the pressure distribu- 
tion there is the same as that of a two-sided 
infinitely long hub with constant radius. This 
case was also considered in the preceding section. 
We' find £(-~©) from (82), with -e, = £(-~), and (6) 
and (87); 


£(-~) =a 6r (-~)/r . (91) 
h c 


In order to transform (89) into a discrete 
function we choose n + 1 points on the negative 
x-axis: 


oO < a < x1 S boo S wy) < x) = 0, (92) 
and construct n coordinate functions, f (x), ..., 
f,(x), defined on -~ < x < 0 as follows: For 

WS By .-, n-l the function f,(x) vanishes out- 
side the interval (x41, Xm-1), and inside this 
interval its value is 


£ (G25 x 


) Baal (xtc ) paex 
™m ™m 


ae, 9) (x = xy 


) Ws S38. ayo 
m m- 


The function £ (x) vanishes for x SE5 a and: 


1/2 1/2 


fi, (x| = |x| 7 |x| 5 $y SR SO; (94a) 


27) SB Reo sp) 7 Geo oc oS 8 Sear co (94b) 
1 2 
Finally £ (x) vanishes for X41 < x < 0 and: 


= = Bi (95a) 
a) ak a / ( Soe d * ee el 


£ (x) (95b) 
n 


i] 
a 
* 
1A 
* 


These functions are plotted in Figure 3. We approx- 
imate the function f(x') in (89) by a linear com- 
bination of the coordinate functions: 

n 


S@)S 6 2.62) 4 (96) 


where the C_ are unknown coefficients. In order to 
approximate f(x') well near the origin, we have 
chosen f; in a special way and, besides, the points 
x are more densily distributed near the origin. 
Since f£(x') is almost constant for large negative 
values of x', we have chosen f to be constant in 
(© Qp 'F3_)o ie 

Next we have to determine the coefficients C), 
oetain Ch We substitute (96) into (89) and then 
the C_ must be chosen so that the difference 
between the right hand side and the left hand side 
of (89) is as small as possible, in the sense of 
some norm. The computed values of the C_ appeared 
to depend strongly on which norm was chosen for this 
difference; many of these norms give unreliable 
results. We obtained reliable values of the C as 
follows: m 


Equation (89) with x = Kor 2=0,1,..., n-2 yields 

n 
M = + = Reni a 

= on Cc ox, (x) s), R= O, n-2 (97) 

m=1 
where: 
(0) 
= = ' ' O 
My ih ss, = a0) BG) Gh? (98) 


—0oo 


At the points x,-}; and x, we minimize the difference 
between the right hand side and the left hand side 


of (89). The expression 
n n 
z (eM EG ote be 2 S\iZ (99) 
fe gen eek 


is a quadrative, non-negative function of the C . 
Now the Cj, -, Cy are determined so that they 
minimize (99) with the constraints (97). 


FIGURE 3. The coordinate functions 


eG o >. £.. 


355 


We have checked these numerically computed values 
of the C, as follows. First, the computed approxi- 
mation has the square root character (90) even in 
the interval (x3,0). Second, the value of C, equals 
the right hand side of (91) within an error of 0.5%. 
Third, if we replace the kernel k(x) in (89) by a 
kernel k(x), which has the same behavior (69) at 
the origin, and which for x > ~ is also given by a 
term A sin bx as in (66), then (89) can be solved 
effectively by the Wiener-Hopf method. If we, apply 
our numerical method to (89) with the kernel k, 
then the numerically computed function, f, equals 
the analytically computed solution within an error 
of 1s. 

We have tried to compute the Cj, ..., C_ in 
different ways; for instance: 
i) By collocating the points x), ..., x, with 

the exception of one point xj, so that the number 
of equations equals the number of unknowns. This 
method had to be rejected because the computed 
approximation for f(x) appeared to have oscillations 
near Xj. 

ii) By minimizing the sum of squares of the 
differences between the right hand and the left 
hand side of (88) at the points x, ..-., X,- We 
have also rejected this method, because oscilla- 
tion occurred in f(x) near the origin. 

We make some remarks concerning the computation 
of the matrix elements. M Inya(98) Seek orems— a5, 
---, n - 1 the integrand iS non-zero only ina 
bounded region. The kernel k(x) is written as the 
sum of a logarithm and a function which is bounded 
at x - 0. The integral over the logarithm is 
evaluated analytically; the remaining term is 
integrated numerically. For m =n the integrand is 
non-zero in an unbounded region. For x < x, we 
have f,(x) = 1 and we must evaluate 


x 
n 


J k(x] = 220) s@bst s (100) 


—o 


Note, that the integrand does not tend to zero for 
x' > -~, as follows from (66). However, the express- 
ion (100) represents the deformation of the cavity 
due to a loading which equals a step-function. This 
deformation has been computed in (76,77). 

We now come to the determination of the shift, s. 
The pressure in x < 0 at r = r, must exceed the 
vapor pressure. Hence, by (22) with o = 0, we 
must have f 2 0, and by (90), B 2 0. As will be 
shown in the Appendix, the shape of the cavity for 
small values of x is given C 


' L 
6x (x) = 6x (0) + br (0)x - 4B(x3/m) 7/3, x + 0 
c h h 


(101) 


This implies that the radius of curvature tends to 
ZOO) EO exXiy) 0. 


Since the fluid may not penetrate 


356 


the hub, we must have B < O for a hub with a smooth 
surface. We found above that B > O and hence B 
must vanish. For our numerical approximation this 
implies that the coefficient, C,, must vanish. Now 


the value of the shift, s, is determined by iteration 


so that C, vanishes. 

When f(x) has been computed we can compute the 
shape of the cavity in x > 0 with a numerical 
integration of (73) for x > 0. Using (40) with 
o = 0, we can derive an expression for dr,(x) for 
x > ©. The derivation is similar to the derivation 
of (66) and, therefore we give the result only: 


2E (102) 
Oo j 
OS (0) ee TASH (ERX) Wet A COS| (SX/.-m) he 
c Beene ) ) 0) G 
to) 
where: 
(0) 
ex 
A, = lim e cos (E x/r) oid (3x3) Gea (103a) 
e+O -2 
to) 
IN = nin f e Sen (Ee) sm (5%) obren (103b) 
e+O -2 


7. NUMERICAL RESULTS 


In this section we give computermade plots of the 
shape of the cavity dr,(x) for a number of shapes 
of the hub dr;},(x) and for a number of values of the 
dimensionless parameter a. We consider the case of 
zero surface tension, hence a is given by (87) or 
by (43) with o = 0: 


2(p, - P_) Tr 
os = 5 (104) 


It follows from (88) that 6r},(x) depends on rq for 
a fixed hub. The value of r, is given by (6): 


1 
mae 


= neue 
Bese Gy) NG). raise) (105) 


However we can vary 4 without changing drp,(x) by 
varying U and keeping p, IT and p, - Poe constant. 

In the Figure 4 the function dédrp(x) is plotted; 
it consists of a straight horizontal line and part 
of a parabola. The x-axis is chosen so that x = 0 
at the point of separation. No scale-unit is given 


in the vertical direction, since ér,(x) and 6r¢(x) 
are the linearized perturbations of the undisturbed 


FIGURE 4. The functions 6rp [(x+s)/r,] 
(hatched curve), Sr¢(x/r,) and the asymptotic 
expression (102) (a.e.). The values of a are 
a)4, b)2, c)l, d)0.5, e)0.25. The point of 
separation is at x=0. bry is given by 


Sry (x/r,) = 1 for x < 0 and = 1- (x/xQ)? for 
x > 0. The values of s/r, are a)1.092, 
b)1.017, c)0.558, d)0.070, e)0.014. 


357 


cavity (1); see Figure 2. The dimensionless quantity the fluid particles leave the hub. This effect is 
x/Yc is on the horizontal axis. In x > 0 the important in the case of a low speed. 

numerically computed function érg(x) is plotted In Figure 5 we have plotted the same functions 
and also the asymptotic expression (102) is given. for a different shape of the hub. Here bry, (x) 

It appears that the asymptotic expression is a good consists of a straight line, a part of a parabola, 
approximation for 6r,(x) also for rather small and another straight line. It appears that quali- 
values of x/rc. tatively the same effects occur. 

The Figures 4(a) through (e) correspond to In Figure 6 we have only one value of the param- 
decreasing values of a. This is equivalent to eter, a, (a4 = 1), but we have plotted a family of 
increasing values of the speed U with constant p, functions Oxy, (x). The plot of 6r,(x) is omitted, and 
T, and p. - Pe- The length of the waves on the we have indicated the point of separation with a 
cavity is an increasing function of a, as was stated dot. The amplitudes of the waves in dr,(x) at x = — 
in Section 4. Further we observe from these figures are denoted in a table underneath Figure 6. These 
the following: numbers are the amplitudes divided by dr} (-~). 

i) An increase of the speed U induces an increase From this figure we observe the following: 

of the amplitude of the waves on the cavity. iii) If dr, (x) decreases abruptly as a function of 
ii) When U is relatively large, the point of x, then the amplitude of the waves on the cavity is 
separation is near the point where dr,(x) attains relatively large. 

the value 6r;(-”). When U is small, the point of iv) The sign of ér,(x) at the point of separation 

separation is near the point where érp(x) = 0. can be positive or negative, depending on the 

The latter phenomenon is easily understood, since function dry,(x). If é6r,(x) decreases more and more 
we can imagine two reasons for which the fluid may slowly as a function of x, then the value of 6rp, (x) 
separate from the hub: First, the radius of curva- at the point of separation approaches zero from 
ture of the hub may be so small that the fluid below. 
particles are unable to keep contact with the hub. We will compare the effects on the cavity by 
This effect dominates in the case of a relatively changing a, or U with constant p and p_ - p_,, and 
high speed U. Second, the value of dr},(x) may of the function 6ér, (x). In order to give a rough 
become negative. Then the centrifugal force makes description of the dependence on Sr (x), we use the 


T r El of a af Do Ot 
-4.0C -2.0c o> 2.00 y 700 8.ac 10.a0 


D> ae. 


7? 
=r T 
-4. 0c -2.00 ON CO 


CUUUUULU EU UE 


“4.00 73.00 


d 


FIGURE 5. The functions 6r}[(x+s)/r,] 

(hatched curve), 6x (x/r¢) and the asymptotic 
expression (102) (a.e.). The values of a are 
a)4, b)2, c)1, d)0O.5, e)0.25. The point of 
separation is at x=0. dr, is given by 

Srp (x/rQ)=1 for x < 0, = 1-(x/r,)*/2 for 

0 < x < r,, and =1.5-x/r for x > r_. The values 
of s/Xo are a)1.683, b)1.759, c)1.805, d)0.361, 
e)0.052. 


FIGURE 6. 


at ile 

2 al 

3 1m 

4 Alo 

5 ING 

6 0. 

i QO. 

8 Q. 

fc) 0.115 0.607 
10 0.0807 0.516 
alal 0.0565 0.437 
12 0.0395 0.370 
13 0.0277 0.311 
14 


pany 
oO 


A family of functions Sr, (x/rc) with a = 1. They are given by Srp (x/r,)=1 for x < 0 and 


=1-)(x/r¢) 2 for x > 0. The value of \ is given in the table. The point of separation is denoted by a dot. 
The amplitude of the waves at x =“ divided by r,(-‘), denoted by A, is also given in the table. 


DCU UL TU UE CU 
7 
7. 


LD LLL EA MMM UA 


Sot ee SCORE ema BI 5 a7 ayes 
Sine tub -5 00 -4, 0 


UU UU UU UU IE UY 


xr 
2 00 14. @¢ 
aa 
15 
7 c/f 
: ES ae 
2 (yt 77108 B25 (iy V08 5S. uv 
Ly, x /r, 
pa eee 
ua OF 


x/P, 


1 
6.00 


VU in TTT. 
T T T insert De 
-3.00 =5. 00 -4 a0 -2. 00 a. 00 


d 


FIGURE 7. The functions 6r,(x/r,) (hatched vl 
curve) and 6x4 (x/rQ) for vanishing waves at (HUA TVARINTE 
x = ©, The values of a are a)4, b)2, c)l, 


d)0.5, e)0.25. The point of separation is at aaa “6 00 4 00 


x= 0. 
e 


-2.00 70’ 00 


x/T 
| 
2.00 yaa 8 ud 

x/r 
TI 
2.00 4.00 8.00 


curvature kK of the hub, which must be interpreted 
as some mean value of the curvature of 6ér, (x) in 
the wetted region not too far upstream. 

First we consider the amplitude of the waves on 
the cavity. This amplitude is an increasing function 
of both U and k, as follows from i) and iii). Next 
we consider the point of separation. An increase 
of U or k tends to shift this point from the point 
where Sr}(x) = 0 to the point where éry,(x) = drp(-2) 
as follows from ii) and iv). 
an increase of U has roughly the same effect as an 
increase of kK. However, the length of the waves on 
the cavity is, as follows from the previous theory 
(87), independent of k but is a decreasing function 
of U. 

Finally we consider a shape of the hub which 
induces, for some value of a, no waves on the cavity 
at x = ©. The existence of nontrivial shapes can 
be shown as follows. Let dér,)(x) and 8x49 (x) be 
two shapes of the hub and let 6X oe, (x) and 6X G5 (x) 
be the corresponding shapes of the cavity for some 
value of a. In each case the point of separation 
is at x = 0. We choose A > 0 so that the amplitude 
of ASXe, (x) at x = © equals the amplitude of OY Gy (x)- 
(Notice that their wavelength is already the same.) 
Next we choose a shift, s > 0, so that 


lim AGL] (x+s) + 8X Qo (x) = 0% (106) 
xo 

Now we construct a shape, 6r, (x), which induces no 

waves at infinity, as follows: 


xcs (107) 


Sx, (x) Adxy, , (xts) + Sry 2 (x), 


Sx, (x) = ASX, (xts) + OY 5 (x) - of S sg <7) 


The dimensionless load f was introduced in (22). 
By virtue of the linearity of our equations we 
obtain the load corresponding to the shape (107) by 
shifting the load due to 8rpy over a distance s, 
multiplying it by » and summing the load due to 
6X49: This load is nonnegative in x < O and 
vanishes in x > 0. The condition for the point of 
separation described in the preceding section is 
satisfied at x = 0. The shape of the cavity is 
obtained by a similar construction as for the load 
f. The value of ér (x) tends to zero for x + © by 
virtue of (106). 

Using this method we have constructed five 
functions dr},(x) which induce no waves at x = © for 
five different values of a respectively. The 
functions 6r,) and 6r,2 are the functions plotted 
in Figures 4 and 5 respectively. The results are 
plotted in Figure 7. The point of separation is at 
x = 0. The value of Sx (x) in x > O is unessential. 


APPENDIX 


SOME RESULTS DERIVED BY USING WIENER-HOPF-TECHNIQUE 


In equation (73) we let the path of integration 
be from -~ to ~ and we assume the load f(x) to 
vanish for x > 0. Then applying Fourier transform 
(34) to both sides of this equation, we obtain: 


Hence in these respects, 


859) 


bx (E) = k(E) £(E), (A,1) 
where: 
‘i = 
K(E) = Ky, ((E]) (a Ky, ({E]) - [é] Kole} (A,2) 
Here we have chosen the unit of length to be equal 
to ¥_, so that r. = 1. We have to solve (A,1) with 
f(x) = 0 for x > O and with dr,(x) being prescribed 
agoye og S Wr 
In order to apply Wiener-Hopf-Technique, we need 

a multiplicative decomposition of K(E). We define: 


My ee 


H(E) = -k(E) (E2 - cae) (E2 + : (A,3) 


where € is the root of (47). This function is 
continuous and positive in -~ < — < ~. By virtue 
of well-known asymptotic expressions for the 
Bessel functions, K, and K,, we have: 

BS) S Ika © Gye)o Esra © (A,4) 
Hence, we can decompose H(&) in the usual way, see 
for instance Noble (1958); we find: 


me) on) /m) 4 (A,5) 


where: 


2 i S4 
H (&) = exp == | 7 in {H(Z) ] ae (A,6) 
Cc 


This represents two equations; the upper or the 
lower of the + signs must be read. The contour of 
Ci (resp. C-) is the real axis, indented into the 
upper (resp. lower) half of the complex C-plane at 

t = &. The function H*(E) [resp. H~(&)] is analytic 
in the upper (resp. lower) halfplane. Using (A,2-3) 
and (A,5) we can write (A,1l) as 


1/2 


Gag (5) (2 = 2) (BA) > ae) — 
{o) 


peli) = 
AC) d: sai” (3). 28S) (A,7) 
The function (E+i)% has a cut from -i to -i» and 
(E-i) 1/2 has a cut from ito i~. They are both 
chosen so that they are positive for §>°%. 

The function Sr (x) is prescribed for x < 0. 


First we assume: 


Ax 


bx (x) =e xI<0) (A,8) 


for some positive \; later we discuss the general 


case. Fourier transformation gives: 
= = - + 
Sale) =a On (Ey * + 8x (6), (A,9) 
where 
6x (E) = (Cay i! a Sx (x) dx (A,10) 


O 


360 


is unknown. We substitute (A,9) into (A,7) and 
separate functions which are analytic in the upper 
half plane and respectively in the lower one. There- 
fore we write: 


@a)ne = (Gaas (sD eG) = 
ne (&)) + he(é) , (A,11) 
where: 
h (&) = Seaer tits (CANFEIE2}) 
(27) (iA+i) g-id 


is analytic in the lower half of the complex €-plane, 
and: 


+ “ H (in) 
h (&) = —— [—*- 


Hie) a 
Qn) 4 (iA4+1) 


5 (A,13) 
(esi) 7? E-id 


is analytic in the upper half plane. 
and (A, 11) we can write (A,7) as: 


Using (A,9) 


SY sy + 
[6x _*(E) (E+i) / H (é)]+h (&)] (E2 = Ea) = 
- . 1 2 = =- 
=Is\ (0S) (E - e) = (( > al) / Ist (15) 2e{(S)) (A,14) 
+ 

The function 6r_ (&€) is analytic in the upper 
half plane by virttie of (A,10). Hence the left 
hand side of (A,14) is analytic in the upper half 
plane. Since £(x) vanishes for x > 0, £(&) is 


analytic in the lower half plane and hence the 
right hand side of (A,14) is also analytic there. 
Hence, both sides of (A,14) represent an entire 
function. The H7™(&) tend to 1 and the nt (E) are 
O(1/E) for — > ©. We assume ér (&) gl/2 and £(&) 
gl/2 to be bounded for — + ~. Then the entire 
function must be a first order polynomial C_ + Cj 
—, where the values of the constants C and°C, will 
be given later. We can now solve for the unknowns 
6x0 and f: 


pele (e Hens 
ore) = Xe, tay 
H (é) Be 
= (Es - 2 
£(E)) =) IG, en BP) EV Sey cus 
H (E) 


The value of Cl] is chosen so that £(£) is o(e 1/4 


for —& +> ™. Hence, by (A,12): 


) 


er SHUG (net 5 EN) (A,17) 


We choose Ce so that £(E) is an order smaller, i.e., 
0(£73/2), for — + ~; hence: : 


c= - Opt NMEA aS BON) (A, 18) 


The meaning of this choice for f(x) will be discussed 
later. 


We obtain f(x) with the inverse Fourier transform: 


eo Léx (ei) 1/2 


H (é) 


£(x) = 


Esa 
(2m) 1/7 


IG, Cie He (G) (l= Bees - (A,19) 


Since the integrand is analytic in the lower half 
of the E-plane, the right hand side of (A,19) 
vanishes for x > 0, as follows from the calculus 

of residues. In order to obtain an expression for 
f(x) for x + 0 we investigate the integrand in 
(A,19) for — ++. The function f(x) is continuous 
at x = 0, since the integrand is 0(E73/2). Hence 
£(07) = 0. For real values of € we have, by (A,6): 


: a Pe at 
+ Sy (A 
H(E) = (H(E))-/? exp Neat ahi = a 


where the integral is a Cauchy principal-value. 
Hence for real — we have 

ie (2) ~ Sn 2 Eee), (A,21) 
where H(&) is a continuous function, which is 0(1/&) 
for — > + ™ by (A,4). Therefore, if the factor 
H7~(€) in the denominator in the integrand in (A,19) 
is omitted, the value of the integral changes by a 
term which is O(x) for x + O. The other factors 
in the integrand in (A,19) are an exponential and 
a rational function of €. Using the calculus of 
residues and Tauberian theorems we can obtain an 
asymptotic expression for the value of this integral 
for x t 0. We do not go into details and give the 
result only: 


=P |x| 1/2 


ie (5) SS Qur Ape ee 0) (A,22) 


where: 


1/2 


(AE ey QUST Ae GR) - (a,23) 


In a similar way we investigate dr (x) for x ¥ 0. 
Substitution of (A,17) and A,18) into (A,15) and 
then into (A,9) gives: 


or, (8) = 

i Hit (GA) (En) a NES 
eee (A,24) 
Ci) (ayy Fe) (Ee >) (=a) 


This expression is OES) for §€ + © and hence 

6r (x) and its first derivative are continuous at 
x = 0. An expression for x ¥ 0 is obtained in the 
same way as for f(x): 


Ges (GF) Ss Gre (()) GE -WreY (@)) se = a B sel? x10} 
c c (o} a2 

(A,25) 

At this point we return to our choice (A,18) for 


Go. If (A,18) does not hold, then it can be shown 
that: 


£i(x)) 2) BS |x| 3 se P Oy (A, 26) 


1/2 


(Gre (9) wes _(@)) — rests 6% xe Ol (A,27) 
c c 


where the constants B* and B** have the same sign. 
The condition f> 0 implies B* > O and the condition 
that the fluid does not penetrate the hub implies, 
in the case of a smooth hub, that B** < 0. Hence 
they must vanish both, which is achieved only by 
giving C_ the value (A,18). 

Finally we consider a hub of arbitrary shape. 
By virtue of the Laplace transform we can write: 


oe Ax 
Sr (x) Some g(A)e GA 5 SOG (A,28) 
C-ic 
where c is a positive number, and: 
) 
g(X ) = if mes bx (x) dx (A,29) 


oo 


First we assume that 6dr (x) is such that the integral 
aay (VAp2S))) ats absolutely convergent iors JN SI Cp Joyihe, 
our results will appear to hold for a more general 
case. By virtue of the linearity of our questions, 
the expressions (A,22) and (A,25) hold with B given 
by: 

Ctic 


1/2 


H (id) an. 
(A,30) 


g () 2 + es) (Remy) 


GRales 


Substitution of (A,29) into (A,30), interchanging 
the order of integration, and applying partial 
integration with respect to x twice, gives: 


(o) 


B= | L(x) {E7 Sx, (x) ae Oig UU) Ir Cbs p (A,31) 
ic 
where 
Cctjio 
1 Se EY aR 
L(x) = Daa e Paks, Gy pp 32S Os. Hypa?) 
(A+1) 
c-ics 


361 
In this expression for L(x) we substitute ay SS lly, 


take the limit c + 0 and use some symmetry-properties 
of Ht (1). Then we find: 


ete 
il i H (u 
L(x) = al Re e ae HSL, Chil, se = O, (H\,55))) 
(1-in) 


—ily/) 
Since the integrand in this expression is 0(H / ) 
we can derive for L(x): 
~ ly SLY 
aeeeae* (sl 47 xt 0 (a, 34) 


As stated in Section 6, the position of the 
point of separation is determined by the condition 
B=0. By (A,31) this condition becomes: 


(0) 


ff L(x) [e2 Sy (x) + Sr 0 | dx = 0. 
= ° c c 


We can give an interpretation to the two terms in 
this integrand. There are two reasons for which 
the fluid may separate from the hub. First the 
value of é6r, may become negative, so that the 
centrifugal force makes the fluid particles leave 
the hub. This corresponds to the first term in the 
integrand. Second, the radius of curvature of the 
hub may be so small that the fluid particles are 
unable to keep contact with the hub. This corres- 
ponds to the second term in the integrand. 


(A, 35) 


REFERENCES 

Doetsch, G. (1943). Laplace Transformation. Dover 
Publications. 

Noble, B. (1958). Methods Based on the Wiener-Hopf 
Technique. Pergamon Press 


Sparenberg, J. A. (1958). On a shrinkfit problem. 
Applied Scientific Research, Section A, Vol. 7. 

Watson, G. N. (1922). Theory of Bessel functions. 
Cambridge University Press. 

Whitham, G. B. (1973). Linear and Non Linear waves. 
J. Wiley and Sons. 

Wu, Th. Y. (1972). Cavity and Wake flows. 
Review of Fluid Mechanics, 4. 


Annual 


Unsteady Cavitation on an 
Oscillating Hydrofoil 


Young T. 


Shen and Frank B. 


Peterson 


David W. Taylor Naval Ship Research and Development 
Center, Bethesda, Maryland 


ABSTRACT 


Bent trailing edges and erosion are often observed 
on marine propellers and are attributed mainly to 
unsteady cavitation caused by the nonuniformity 

of the flow field behind a ship's hull. In order 
to improve the physical understanding of the 
cavitation inception and the formation. of cloud 
cavitation on marine propellers, a large two 
dimensional hydrofoil was tested in the DTNSRDC 
36-inch water tunnel under pitching motion. Fully 
wetted, time dependent, experimental pressure 
distributions were compared with Giesing's unsteady 
wing theory. The influence of reduced frequency 
and pressure distribution on inception was determined. 
A simplified mathematical model to predict unsteady 
cavitation inception, was formulated. Good corre- 
lation between theoretical prediction and experi- 
mental measurements on cavitation inception was 
observed. The reduced frequency, maximum cavity 
length, foil surface pressure variation, and time 
sequential photographs were correlated with the 
formation of cloud cavitation. A physical model 
based on the instability of a free shear layer 
defining a near-wake region provides a reasonable 
explanation of the observed results. 


1. INTRODUCTION 


Hydrofoil craft are typically designed to operate 
both in calm water and waves; and marine propellers 
normally operate in the nonuniform flow field 
behind a ship. Unfortunately, due to the complexity 
of the experiments, only a few experiments have 
been specifically concerned with unsteady leading 
edge sheet cavitation on hydrofoils and propellers, 
Morgan and Peterson (1977). It is the intent of 
this paper to report the results of experiments 
concerned with leading edge sheet cavitation on an 
oscillating two dimensional hydrofoil. Following 

a brief review of the most pertinent experimental 


362 


data available in the literature, an analytical 
method for the prediction of inception will be 
developed and compared with the experimental data. 
Once the cavity is present on the foil, cavity 
instabilities develop due to the foil oscillation 
and also due to the inherent instability of the 
cavitation process. This general process of 
instability in the leading edge sheet cavity is the 
subject of this paper. 

It has been observed by innumerable investigators 
that a leading edge sheet cavity can, under certain 
circumstances, be quasi steady with relatively few 
collapsing vapor bubbles to produce erosion. How- 
ever, if a propeller blade enters a wake field, the 
inception angle of attack at the leading edge may 
not agree with the uniform flow inception angle. 

In addition, the developed cavity may exhibit 
instabilities not produced in uniform flow fields. 

One form of cavity instability is manifest by the 
shedding of a significant portion of the sheet 

cavity. This shed portion appears to be composed 

of microscopic bubbles and is commonly referred to 

as "cloud" cavitation, van Manen (1962). Cloud 
cavitation is now considered to be one of the main 
causes of erosion and bent trailing edges, Tanibayashi 
(1973). 

Model experiments have been performed by many 
organizations in an attempt to simulate full-scale 
wake fields in which propellers operate. One of 
the first detailed experiments concerned with 
unsteady cavitation was reported by Ito (1962). 
These experiments were with pitching three dimen- 
sional hydrofoils and propellers in a wake field. 

A principle result directly applicable to the work P 
to be reported here was that the reduced frequency 

had an important influence on the cavitation. He 

also concluded that the "critical" reduced frequency 
at which a leading edge sheet cavity broke up into 
cloud cavitation was 0.3 to 0.4. His latter con- 
clusion will be considered in more detail in the 
context of the results to be reported here. 

A recent discussion of this subject was given 


by Tanibayashi (1962). He concluded that the 
occurrence of cloud cavitation in nonuniform flow 
cannot be predicted on the basis of uniform flow 
experiments. In earlier work by Tanibayashi and 
Chiba (1968), it was concluded from experiments 
with an oscillating two dimensional foil that an 
unsteady flow was required for the formation of 
cloud cavitation. However, unlike the earlier 
results of Ito, no distinct critical reduced 
frequency was found. Since these latter results 
were for nominally hemispherical travelling bubbles, 
instead of a leading edge sheet, it remains to be 
established whether the type of cavitation in the 
growth phase is of importance to cloud cavitation 
formation. 

Chiba and Hoshino (1976) carried out extensive 
measurements of induced pressures on a flat plate 
above a propeller. On the basis of comparing 
results with and without a wake field and with and 
without cavitation, they determined that strong 
pressure impulses were detected on the flat plate 
and these correlated with the presence of cloud 
cavitation. 

Strong pressure fluctuations of very short 
duration have also been detected by Meijer (1959) 
on the surface of a cavitating two dimensional foil. 
He attributed these pressure fluctuations to a 
stagnation point at the rear of the sheet cavity 
passing over a pressure gage. Chiba (1975) has 
attempted to correlate cavity collapse on a two 
dimensional oscillating foil with the response from 
a pressure gage mounted in the foil. He concluded 
that, as expected, when the shed vapor collapses 
large pressure impulses occur. The essential 
points for both of these experiments are that foil 
mounted pressure gages can be used in the presence 
of cavitation and when correlated in time with 
photographs can assist in the interpretation of the 
physical processes involved. This technique was 
also used in interpreting the results to be reported 
here. 

Two other oscillating foil experiments have also 
been reported, Miyata (1972), Miyata et al. (1972), 
and Radhi (1975), that demonstrate the importance 
of the reduced frequency on the whole cavity 
inception, growth, and collapse process. Both have 
shown that for the particular conditions of their 
experiments, inception could be delayed. The 
greatest suppression occurred for reduced frequencies 
in the range of 0.4 to 0.5. Both of these experi- 
ments will be discussed later in more detail within 
the context of the results to be reported in this 
paper. 

All of the experiments reviewed above describe 
various aspects of cavitation instabilities that 
are associated with the cavitation performance of 
oscillating foils and propellers in a wake. This 
cavitation performance appears to be uniquely 
related to the unsteady flow field that exists 


Pe (*/e = 0.033) 


P, (*/e = 0.10) 
P, (¥/e = 0.25) 


PITCH AXIS LOCATION 


alles ba C = 241 m 


KULITE PRESSURE GAGE 


363 


over the cavitating surface. In the sections that 
follow analytical and experimental results will be 
presented in an effort to provide a better under- 
standing of how these various results are related 
and of the associated physical processes involved. 


2. EXPERIMENTAL APPARATUS AND TEST PROCEDURE 
Foil and Instrumentation 


The foil was machined from 17-4 PH stainless steel 
to a rectangular wing of Joukowski section with the 
trailing edge modified to eliminate the cusp. To 
simulate the viscous effects at the leading edge 

as close to a prototype as possible, the model was 
designed with a chord length of 24.1 cm and a span 
of 77.5 cm. The maximum thickness to chord ratio 

is 10.5 percent. The foil surface was hand finished 
within 0.38 wm RMS surface smoothness. 

Pressure transducers were installed at a distance 
of 7.96, 24.1, and 60.3 mm from the leading edge. 
These locations correspond to 3.3, 10, and 25 
percent of chord length from the leading edge. 
Kulite semiconductor pressure gages of the diaphram 
type were mounted within a Helmholtz chamber con- 
nected to the foil surface by a pinhole. With this 
arrangement one could measure the unsteady surface 
pressures due to foil oscillation and high frequency 
pressure fluctuations inside the boundary layer 
over a pressure range of +207 KPa (+30 PSI) anda 
calibrated frequency range of 0 to 2 kHz. In order 
to increase the spatial resolution in measuring 
the local pressure fluctuations inside the boundary 
layer, the diameter of the pinholes installed on 
the foil surface were kept at 0.31 mm (0.012 inches), 
(see Figure 1). This arrangement also reduces the 
danger of cavitation damage to the pressure 
transducers. Extreme care was taken to fill the 
Helmholtz-type chamber through the pinhole under 
vacuum with deaerated water to minimize the possible 
occurrence of an air bubble trapped inside the 
chamber. If a gas bubble was present within the 
gage chamber, the resonant frequency of the chamber 
would be reduced below its 3880 Hz value. For 
example, with the above procedures for filling the 
gage chamber at a pressure of 3.4 KPa, a bubble of 
0.6 mm diameter at atmospheric pressure is produced. 
This bubble will lower the chamber's resonant 
frequency to 1100 Hz. The danger of becoming a 
Helmholtz resonator was not observed in our dynamic 
calibration tests up to 2000 Hz. The calibration 
procedure used here was developed by the National 
Bureau of Standards, Hilten (1972), modified to 
the extent that water rather than silicone oil was 
the fluid medium. Since it was very important to 
determine the relative phase difference between the 
foil angle and the pressure gage signals, all 
amplication and recording equipment was selected to 
minimize the introduction of unwanted phase shifts. 


PINHOLE 


HELMHOLTZ TYPE 
CAVITY 


FIGURE 1. A sketch of the foil 
and three pressure gage locations. 


364 
Photographic Instrumentation 


All photographs used to document the inception and 
cavity instability processes were taken with two 
35 mm cameras. Illumination was provided by strobe 
lights having a light duration of 10 microseconds. 
With the camera shutter open, the first frame of a 
sequence was taken when a foil position indicator 
triggered the strobe lights. Each succeeding 
exposure was taken 10 and 1/25 foil oscillations 
after the preceeding exposure. An electrical pulse 
from a light detector was recorded on a channel of 
the same magnetic tape that was used to record the 
foil position, pressure gage responses, and a time 
code. Oscillograph records then allowed a direct 
correlation between these events. Both top and 
spanwise photographs were taken simultaneously by 
exposing the film with one set of flash lamps. In 
order to focus the camera lens in the same region 
as the location of the pressure gages when viewing 
in the spanwise direction, the camera was elevated 
at an angle of 4° and directed slightly downstream 
by an angle of 10°. 

High-speed 16 mm movies were taken at a rate of 
9,300 frames per second to assist in the interpre- 
tation of the 35 mm pulse camera sequential 
photographs. Adequate exposure for these photographs 
was achieved by using high intensity tungsten 
filament flash bulbs of 25 millisecond duration. 


Test Section 


The closed jet, test section of the 36-inch water 
tunnel was modified by the insertion of sidewall 
liners to provide two flat sides as shown in Figure 
2. On each end of the foil a disc was, attached. 
This disc rotated in a sidewall recess. Thus the 
foil could be rotated without gap cavitation 
occurring between the end of the foil and the 
sidewall of the tunnel. One sidewall assembly was 
fitted with clear plastic windows to permit side 
view photography. 

The foil was oscillated by a mechanism whose 
conceptual design is shown in Figure 3. With this 
type of design the foil mean angle (a _) can be 
adjusted statically and the amplitude of foil 
oscillation (a]) can be continuously adjusted 
between 0° < a; < 4° while in operation. The 
oscillation frequency is continuously variable 
between 4 Hz < £ < 25 Hz. Air bags, shown in 
Figure 3, were installed to reduce the fluctuating 
torque requirements on the motor drive system. 


VIEWING PORTS 


FOIL DISC 


FIGURE 2. 


Schematic of closed jet test section. 


PNEUMATIC 
AIR BAGS 


FOIL 
OSCILLATOR 
ARM 


ADJUSTABLE 
PIVOT POINT 


— FOIL SHAFT 


SLIDE 


CONNECTING 


ROD —— ECCENTRIC CRANK 


~ DRIVEN BY VARIABLE 
SPEED D.C. MOTOR 


FIGURE 3. 
mechanism. 


Conceptual design of foil oscillation 


Water Tunnel Resonant Frequencies 


The study of cavity dynamics in a water tunnel gives 
rise to a fundamental question, namely, the effect 
of tunnel compliance on transient cavity flows. If 
the tunnel was perfectly rigid and if there were 

no free surfaces other than that of the cavity 
itself, then an infinite pressure difference in an 
incompressible medium would be required to create 

a changing cavity volume. To make sure that this 
kind of tunnel effect would not be present in our 
model tests, a hydraulically operated piston having 
a frequency range of 0 to 45 Hz was initially 
oscillated in a test section opening to simulate 
the maximum expected change of cavity volume. A 
sharp peak of fundamental tunnel resonance was 
observed at 4.7 Hz. Consequently, all of the foil 
oscillation experiments reported here were carried 
out at frequencies either above or below this 
resonant frequency. 


Data Reduction 


Due to the installation of two sidewall liners in 
the test section, the tunnel velocity was corrected 
according to the area-ratio rule. The tape recorded 
time histories of foil angle and pressures were 
digitized using a Raytheon 704 minicomputer and 
reduced using algorithms implemented on the DTNSRDC 
CDC-6000 series digital computers. The time histo- 
ries were recorded on one inch magnetic tape at 

15 inches per second (38 cm/s) using IRIG standard 
intermediate band, frequency modulation techniques. 
During digitization, these data were filtered using 
eight-pole Butterworth low pass filters that have 

a -3 db signal attenuation frequency at 40 Hz. 

They were then sampled at 125 hertz. The run 
lengths used in the data reduction were nominally 
40 seconds. For the oscillating foil data the 
computer output consists of values of mean and 
standard deviations, sine wave amplitudes and 
frequencies, and transfer function magnitudes and 
phases. Mean and standard deviation values were 
obtained from the stationary foil data. For the 
transfer functions, the system input was foil angle, 
where the pressures were responses to this input. 
For the dynamic runs, foil angle was sinusoidal; 


nominally one percent of this channel's signal 
energy consisted of harmonics or noise. 

The methods used in data reduction are now 
described. The mean value, wt, and the standard 
deviation, o_, were calculated in the usual manner. 
The sine wave amplitudes and frequencies, and the 
transfer functions were obtained using operations 
on measured autospectra and cross spectra. These 
spectra were obtained using overlapped fast Fourier 
transform (FFT) processing of windowed data segments, 
Nuttall (1971), where the following reduction 
parameters were used: FFT size of 1024, 50 percent 
overlap ratio, and full cosine data window. The 
true autospectrum of a sine wave is an impulse, 
0.5A°6(£ - £_); the measured autospectrum is this 
true spectrum convolved with the spectral window. 
The spectral window associated with the cosine has 
the form: 


sin mf 


£1 (1-£7) 


The wave frequency is, in general, not sampled at 
a rate which is an integral multiple of the sampled 
frequency. Thus, the measured spectrum consists 

of this spectral window sampled at evenly spaced 
frequencies where the location of the samples 
relative to the sine wave frequency or spectral 
window maximum is unknown. The sine wave frequency 
and amplitude are found by fitting the spectral 
window shape to the three largest samples that are 
closest to where the sine wave is expected. The 
transfer functions are given by the cross spectra 
between the input and output data channels divided 
by the autospectra of the input channel. The 
transfer functions were evaluated at the frequency 
of oscillation of the foil. Quadratic interpolation 
between spectral samples was used to obtain the 
cross and autospectrum values. Once evaluated, the 
complex transfer functions were converted to magni- 
tudes and phases. The transfer function magnitude 
is then the output sine wave amplitude, and the 
transfer function phase is the phase angle of this 
output sine wave. Except for data runs when cavi- 
tation was present, the cross spectra coherency 
was always greater than 0.98; this high coherency 
implies low noise and high linearity at the foil 
oscillation frequency. 


3. UNSTEADY HYDRODYNAMICS IN FULLY WETTED FLOW 


Basic knowledge in the general field of unsteady 
aerodynamics has been compiled, condensed, and 
presented by several authors [for example see 
Abramson (1967)]. Available experimental hydro- 
dynamics information for oscillating wings and 
foils is very limited, especially at high values 
of Reynolds number. Most of the available experi- 
mental data concern lift, drag, and moment 
coefficients from flutter and craft control 
investigations. For cavitation inception studies, 
accurate determination of the pressure distribution, 
especially around the leading edge, is of major 
importance. In the present investigation, three 
pressure gage transducers were installed on the 
foil to measure the unsteady surface pressures. 
Experimental data were then correlated with an 
available unsteady flow theory with the intent 


365 


of providing adequate information to analyze unsteady 
cavitation inception. 

The foil was pitched about an axis at %% chord 
length from the leading edge. The instantaneous 
foil angle a is given by 


= 1 


6 + a) sin wt (1) 


where & _, 4), and W™ are the mean foil angle, pitch 
amplitude, and circular frequency of pitch oscil- 
lation. Let Cp(t), Cys, and Cpy(t) denote the 
total pressure coefficient, the magnitude of the 
steady pressure coefficient at the foil mean 
angle, and the magnitude of the dynamic pressure 
coefficient, respectively. At a given location on 
the foil, it is assumed that: 


De ap 2 (Ge), oe 
P oO 15 oo 
c(t) = BE = Fe = 8 4 
4p v2 A (e) W 
(2) 
= © a © (ie) 
ps pu 
where 
B = EAS 
ces Ey V2 (3) 
and 
P (t) 
4 
é re) = wl (4) 
a 5p v2 


where P(t), Ps, Py(t), and p are the local total 
pressure on the foil, static pressure on the foil, 
dynamic pressure on the foil, and the fluid density, 
respectively; P, and V, denote the freestream 
pressure and freestream velocity. We have: 


sin(wt + 6) (5) 


where |Acpu| and ¢ are the amplitude of dynamic 
pressure response and phase angle, respectively. 
A positive value of > means that the pressure 
response leads the foil angle. 

Let the Reynolds number, Rn, and the reduced 
frequency, K, be defined by 


Ww, € 
nm = (6) 
n v 
and 
wC 
= 7 
K 2 Voo 7) 


where C, Vv, and w are the chord length, kinematic 
viscosity of the fluid, and the circular frequency 
of the oscillating foil, respectively. Fully wetted 
experiments covered the range of Reynolds number 
Rn = 1.2 to 3.7 x 10© and reduced frequency K = 0.23 
to 2.30. The test results are given in Tables la 
to lc. The phase angles and the amplitude of 
dynamic pressure response per radian of pitch 
oscillation are given in Figures 4 and 5 at values 
Oe Ch = Os55 WsO, Etre BoOPs 

An unsteady potential flow theory for small- 
amplitude motion recently developed, Giesing (1968), 
is used here to correlate the experimental results. 
The unsteady part of the pressure coefficient is 


(DEG) 


PHASE ANGLES, ¢ 


-50 


PHASE ANGLES, ¢ (DEG) 


100 


PHASE ANGLES, # (DEG) 


-50 


@ = 3.25 + 0.5 Sin wt 


X/C EXP THEORY 


@= 3.25 + 1.0 Sin wt 


X/C EXP. THEORY 
0.033 O --- 
0.10 A ---— 


0.25 [e) 


a@ = 3.25 + 2.0 Sin wt 


X/C EXP. THEORY 


REDUCED FREQUENCY, K 


REDUCED FREQUENCY, K 


REDUCED FREQUENCY, K 


FIGURE 4a. Phase angles of dynamic pressure 
response at pitch amplitude a; = 0.5 deg. 


FIGURE 4b. Phase angles of dynamic pressure 
response at pitch amplitude a, = 1.0 deg. 


FIGURE 4c. Phase angles of dynamic pressure 
response at 4] = 2.0 deg. 


obtained as the difference between the total pressure 
coefficient minus the steady part. The steady 
solution is based on an exact nonlinear theory. 

The theoretical values obtained from Giesing's 
program are plotted on Figures 4 and 5 along with 

the experimental data. 

The phase angles obtained from experiments and 
calculations will be discussed first. As seen in 
Figures 4a to 4c, the agreement between experimental 
measurements and theoretical calculations of pressure 
and phase angles is quite good for all three pressure 
locations. The agreement is good between experi- 
mental measurements and theoretical predictions of 
Magnitudes of dynamic pressure for the cases of 
X/C = 0.25 and 0.10, as seen in Figures 5a to 5c. 

At low values of K the measured pressure coefficients 
are seen to be slightly lower than the values 
calculated for the case of X/C = 0.033. The exact 
cause of this small discrepancy between measurements 
and theoretical calculations has not been determined. 

The cause of small discrepancies between the 


theory and experiments requires further investigation. 


Nevertheless, the overall good agreement observed 
between our experimental measurements and Giesing's 
method is extremely encouraging. It is noted that 
Giesing's method is based on unsteady potential 

flow theory. The combined theoretical and experi- 
mental results by McCroskey (1975, 1977) indicate 
that unsteady viscous effects on oscillating airfoils 
are much less important than the unsteady potential 
flow effects, if the boundary layer does not interact 
significantly with the main flow. The present study 
appears to agree with his conclusion for the case 

of a fully wetted foil. On the basis of this 
relatively good agreement between Giesing's method 
and the experimental data, this method will be used 
in the next section to predict cavitation inception 
as a function of the reduced frequency, K. 


4. UNSTEADY EFFECTS ON CAVITATION INCEPTION 


The major objective of this section is to examine 
what effect unsteadiness has on cavitation inception. 
The question of the occurrence of cavitation is of 
particular importance when comparing model test 
results for marine propellers or hydrofoils with 
the full-scale prototype data. We would like to 
know whether a noncavitating model is also free 
from cavitation in the prototype. When calculating 
the flow about propeller blades or hydrofoils, it 
is important to know whether the cavitation bubbles 
form on the blades, and if so, under what circum- 
stances. The cavitation number o, defined by 


has proved useful as a coefficient for describing 

the cavitation process. Here, p and P_ denote the 
density and vapor pressure of the fluid and P_ and 
V,, denote the freestream static pressure and the 

freestream velocity, respectively. 

In addition to the incoming flow properties such 
as freestream turbulence and nuclei content, the 
surface finish and boundary layer characteristics 
on the body surface are also of paramount importance 
to the cavitation inception process Acosta and 
Parkin (1975). To limit the scope of the test 


367 


program, air content of the water was not varied. 
The air content was measured with 70% saturation 

in reference to atmospheric pressure at a water 
temperature of 22.2° C and tunnel pressure of 103.6 
kPa. 

The foil was pitched sinusoidally around an axis 
at the quarter chord location aft of the foil leading 
edge. The cavitation tests were carried out by 
lowering the ambient pressure from the previous 
fully wetted tests. The determination of cavitation 
inception was based on visual observations. For 
every test condition, 30 pictures were taken to 
record the cavitation process on the foil. A 
picture was taken every ten oscillations plus 1/25 
of the time period of the foil oscillation. Thus, 

a series of high quality short duration photos 

were taken that together simulate one and 1/5 cycles 
of the foil oscillation. A pulse signal was 
simultaneously recorded on magnetic tape when a 
picture was taken. In this way, each cavity pattern 
observed on the foil could be related directly to 
the instantaneous angle of attack of the foil. 


Analytical Prediction 


A simplified mathematical model will be formulated 
first to explore the possible effect of unsteadiness 
on cavitation inception. A significant delay in 
dynamic stall was observed experimentally and 
discussed in a recent review paper by McCroskey 
(1975), who showed that the pressure gradient AC, /dx 
around the leading edge was of paramount importance 
in dynamic stall. The studies by Carta (1971) 
indicate that the mechanism involved in the delay 
of dynamic stall is the large reduction of unfavor- 
able pressure gradient dC /dx during any unsteady 
motion. P 

The mechanism involved in cavitation inception 
is different from the mechanism of aerodynamic 
stall. It is generally assumed that cavitation 
occurs on a body when the local pressure, including 
the unsteady pressure fluctuations within the 
boundary layer, falls to or below the vapor pressure 
of the surrounding fluid, Huang and Peterson (1976). 
Aside from the effect of nuclei content of the 
water, it is the value of the local pressure 
coefficient that governs the occurrence of cavita- 
tion. Prior to the occurrence of cavitation on 
an oscillating foil, the foil is in a fully wetted 
condition. Thus, the knowledge of pressure distribu- 
tion on the foil in the fully wetted condition 
can be expected to provide useful information for 
unsteady cavitation inception prediction. 

As previously mentioned, the combined theoretical 
and experimental results reviewed and summarized 
by McCroskey (1977) indicate that unsteady viscous 
effects on oscillating airfoils are much less 
important than unsteady potential flow effects, if 
the boundary layer does not interact significantly 
with the main flow. In the present study, as 
discussed in the previous section, the three 
pressure coefficients measured at three points 
around the leading edge are predicted reasonably 
well by Giesing's method both in amplitude and 
phase within the range of reduced frequencies 
examined. This unsteady potential flow theory will 
now be used to investigate cavitation inception. 

In the tests, the foil was oscillated about a 


mean angle of 3.25°. The mean values of dynamic 
foil loadings determined from measurements are 


368 


MAGNITUDE OF DYNAMIC PRESSURE RESPONSE, | py! fey (PER RADIAN) 


FIGURE 5a. 


oy 


MAGNITUDE OF DYNAMIC PRESSURE RESPONSE, locpyl jen (PER RADIAN) 


FIGURE 5b. 


Oy 


o 


ow 


@ = 3.25 + 0.5 Sin wt 
THEORY 


o) 1.0 1.5 2.0 2.5 
REDUCED FREQUENCY, K 


Magnitude of dynamic pressure response at 
0.5 deg. 


@= 3.25 + 1.0 Sin wt 


x/C EXP. THEORY 
0.033 a SoS 
0.10 A - - 
0.25 [e)} — 


+ oa -, 
1.0 125: 2.0 2.5 
REDUCED FREQUENCY, K 


Magnitude of dynamic pressure response at 
1.0 deg. 


@ = 3.25 + 2.0 Sin wt 


x/C EXP. THEORY 
15: 
0.033 «3 --- 
0.10 A a 
= 0.25 Oo = 
3 
= 
« 
fre 
= 
Ss 
a QO 
3 
_ 10 
a 
fS 
= 
a 
is 4 2 
& \ 
2 | \ BAS 
ale Z 
Co 
ray F & ge 
z Ne ee 
> K A - 4 
a en Te a 
5 iene xe) 
3° A é p25 
2 - 
cook eee) 
= Ss oO” 
S= 9-9 
o | 
0 
5 1.0 1.5 2.0 2.5 
REDUCED FREQUENCY, kK 
FIGURE 5c. Magnitude of dynamic pressure response at 
a) = 2.0 deg. 


plotted in Figure 6. Aside from some scatter in 

the data, they are seen to be independent of 
frequency (or reduced frequency). The steady 
pressure distribution calculated theoretically at 
3.25° is given in Figure 7. A suction peak appears 
at around 1.8 percent of the chord length aft of 

the leading edge. Reasonably good agreement between 
the theoretical prediction and the three experimental 
measurements should be noted. Experimental data 
confirm the basic assumption made in Eq. (2) that 
the total pressure coefficient Cp(t) is the sum of 
the dynamic pressure coefficient Cpu (t) plus the 
static pressure coefficient Cys at the mean foil 
angle, i.e. 


CAGE) =n€ stn Can Gt) 
1) ps pu 


We will now proceed to examine the possible 
relationship between the dynamic pressure coeffi- 
cient Cpu(t) and the static pressure coefficient 
Cps- Let the instantaneous foil angle be expressed 
as in Eq. (1). The dynamic pressure response is 
then given by 


u | : 


sin (wt + 6) (5) 


Here |ACpul| and $ are the amplitude and phase angles. 
They are functions of reduced frequency K and 
location X/C. They can be obtained either from 
experimental measurements or theoretical calculations. 
In the following study, Giesing's program will be 
used to compute these variables. In our oscillating 
tests, the mean foil angle was always maintained at 

a = 3.25°. The type of cavitation observed in our 


369 


TABLE la - MEASURED DYNAMIC PRESSURE RESPONSE AT PITCH AMPLITUDE %= 0.5 DEG. 


x/c = 0.25 x/c = 0.10 x/c = 0.033 
ue v f K = é [SC nul/ oy é JOC lay, é [SCou| /oy 
m/s HZ x10 Deg Per Radian Deg Per Radian Deg Per Radian 
7002 4.88 5.5 846 1.2 43.4 3.48 7.1 5.53 =3).1 9.92 
7006 4.88 10.0 1.539 1.2 77.7 5.40 31.2 5.60 7.0 9.62 
7010 4.88 15.0 2.305 1.2 95.7 8.51 48.3 6.77 15.5 10.10 
7014 6.71 5.5 635 1.7 35.8 3.07 1.2 5.69 -6.4 10.15 
7019 6.71 10.0 1.154 1.7 58.4 4.13 13.6 5.72 ol 9.72 
7024 6.71 15.0 1.730 1.7 81.6 6.59 32.5 6.21 7.6 10.32 
7029 9.75 5.5 +423 2.4 18.3 2.69 -2.0 4.74 =11.5 10.43 
7034 9.75 10.0 .769 2.4 44.9 3.34 12.6 4.79 -7.1 9.93 
7039 9.75 15.0 1.153 2.4 62.2 4.45 24.8 5.32 =2.3 10.21 
8002 13.11 5.5 -317 3.3 5.7 2.61 =11'53) 5.89 -15.1 10.90 
8006 13.11 10.0 -576 3.3 29.5 2.76 =3.2 5.33 -12.4 10.10 
8010 13.11 15.0 -865 3.3 48.5 3.50 6.4 5.36 -8.2 10.19 
8029 6.71 4.0 +462 1.7 24.9 2.65 -4.3 5.60 =9.2 10.09 
8041 9.75 4.0 308 2.4 7.1 2.58 -7.6 4.78 =13 73) 10.64 
8045 13.11 4.0 -231 3.3 =2.5 2.77 -13.3 5.86 -15.8 11.68 
8057 14.94 5.5 «282 3.7 7 2.88 =1'5).1 6.62 -16.6 11.80 
1121 11.58 4.0 +264 2.8 = 39) 3.05 -9.9 4.54 -15.0 11.85 
1122 11.58 5.5 +363 2.8 10.5 2.86 -1.0 4.44 =13.3 11.12 
1123 11.58 7.5 495 2.8 20.3 2.91 4.3 4.44 -11.6 10.77 
1124 11.58 10.0 +660 2.8 32.1 3.03 10.0 4.53 =9.5 10.72 
1125 11.58 15.0 -990 2.8 52.3 3.93 24.0 5.00 -5.4 10.69 
TABLE 1b - MEASURED DYNAMIC PRESSURE RESPONSE AT PITCH AMPLITUDE = 1.0 DEG. 
x/c = 0.25 x/c = 0.10 x/c = 0.033 
Run v f K Rn é | ac. Jo, | bel /er, é [oC pul /oa 
Rinses m/s HZ x10~° Deg Per Radian Deg Per Radian Deg Per Radian 
7003 4.88 5.5 -846 1.2 42.4 3.39 10.5 5.46 -1.5 9.66 
7007 4.88 10.0 1.539 V2 71.1 5.13 28.5 6.09 7.4 9.77 
7011 4.88 15.0 2.305 1.2 93.6 8.51 47.0 7.40 16.0 10.73 
7015 6.71 5.5 +635 1.7 37.4 3.04 5.8 5.28 -5.3 10.03 
7020 6.71 10.0 1.154 1.7 62.8 4.57 23.6 5.67 2.0 10.05 
7025 6.71 15.0 1.730 1.7 79.5 6.71 38.8 6.75 8.6 10.70 
7030 9.75 5.5 +423 2.4 20.7 2.67 -2.6 5.26 -10.4 10.36 
7035 9.75 10.0 +769 2.4 44.9 3.28 9.5 5.19 -5.3 9.91 
7040 9.75 15.0 1.153 2.4 62.6 4.64 20.9 5.76 =.6 10.41 
8003 13.11 5.5 -317 3.3 7.6 2.64 -9.5 5.91 -13.6 11.03 
8007 13.11 10.0 576 3.3 30.9 2.81 clo) 5.45 -10.2 10.22 
8011 13.11 15.0 -865 3.3 49.0 3.63 6.8 5.65 =659) 10.40 
8030 6.71 4.0 - 462 57) 23.5 2.69 =1.1 5.13 -8.6 10.07 
8042 9.75 4.0 +308 2.4 8.2 2.59 =7..9 5.34 12.4 10.72 
8046 13.11 4.0 231 3.3 =.8 2.78 -11.8 6.13 -14.2 11.76 
8058 14.94 5.5 +282 3.7 3.5 2.90 -11.8 6.46 -14.8 11.96 
1017 11.58 4.0 -264 2.8 208} 3.08 -11.3 5.76 -13.9 12.09 
1018 11.58 5.5 +362 2.8 9.4 3.03 -6.3 5.58 =12'57, 11.83 
1019 11.58 7.5 +493 2.8 21.7 2.97 -6 5.22 =10.5 10.99 
1020 11.58 10.0 +660 2.8 34.2 3.33 6.1 5.40 -9.0 10.83 
1021 11.58 15.0 +988 2.8 52.2 4.15 15.5 5.69 -4.9 11.10 
TABLE 1c - MEASURED DYNAMIC PRESSURE RESPONSE AT PITCH AMPLITUDE 
@% = 1.5 DEG 
x/ce = 0.25 x/c = 0.10 x/c = 0.033 
aay f K a 3 [Ac I/, [c,ui/a, 6 14Cp ul /on 
m/s HZ x10 Deg Per Radian Deg Per Radian Deg Per Radian 
7031 9.75 5.5 +423 2.4 21.0 2.65 =3.2 5.50 -10.0 10.26 
7036 9.75 10.0 +769 2.4 46.5 3.22 8.3 5.30 =5.1 9.76 
7041 ous 5% Opel 53) 2.4 62.3 4.62 19.0 5.98 92 10.40 
8004 13.11 5.5 -317 3.3 9.0 2.57 -8.6 5.72 -13.0 10.79 
8008 13.11 10.0 -576 3.3 31.5 2.78 -1.0 5.39 -9.7 10.19 
8012 13.11 15.0 -865 3.3 50.2 3.63 8.4 5.62 Bez 10.35 
= 2.0 DEG 
7004 4.88 5.5 - 846 1.2 46.7 3.36 16.9 4.93 Sai/ 9.34 
7008 4.88 10.0 1.539 1.2 71.3 5.14 33.7 6.01 7.2 9.76 
7012 4.88 15.0 2.305 1.2 90.7 8.47 49.0 7.96 16.0 10.98 
7016 6.71 9o5 -635 1.7 33.0 3.01 7.4 5.06 -4.9 9.75 
7021 6.71 10.0 1.154 1.7 58.5 4.16 23.2 5.59 2.0 9.83 
7026 6.71 15.0 1.730 7, 77.8 6.30 36.4 6.91 8.7 10.87 
7032 9.75 5/a15) =423 2.4 19.4 2.63 -2.8 5.38 =9.7 9.99 
7037 9.75 10.0 -769 2.4 44.7 3.18 8.3 5.30 -5.0 9.68 
7042 9.75 15.0 1.153 2.4 60.9 4.42 19.2 5.92 0.0 10.26 
8031 6.71 4.0 -462 7, 18.8 2.70 =-1 4.86 -8.3 9.67 
8047 13.11 4.0 +231 353) 2.0 2.50 -13.0 8.35 -16.5 10.16 
O, = 2.5 DEG 
7017 6.71 515 +635 1.7 32.8 2.89 6.9 4.91 -4.8 9.34 
7022 6.71 10.0 1.154 aloy 57.7 4.01 21.6 5.54 1.6 972 
7027 6.71 1550 1-730 slog 77.9 6.00 36.4 6.59 8.8 10.33 


370 


NEGATIVE PRESSURE COEFFICIENT, 


~Cp 


NEGATIVE PRESSURE COEFFICIENTS, 


wo 


FIGURE 6. 
oscillating tests. 


Seuie| ——+y— at 
a z eo ol x 


EXP x/C a@ = 3.25 + 1.0 Sin wt 


Vx = 13.1 m/s 


5 10 15 20 25 
FREQUENCY, HZ 


Mean pressure coefficients deduced from 


- THEORY AT a@, = 3.5 DEG 


THEORY AT a, = 3.25 DEG 
EXP. AT aw, = 3.25 DEG 


-05 -10 =) -20 
x/C 


test program always initiated near the foil leading 
edge. 

Let (dC,/da); denote the static pressure gradient 
with respect to foil angle at a given location on 
the foil. Similarly, let (dcp /da) 4 denote the 
dynamic pressure gradient with respect to foil angle 
at the same location on the foil with the reduced 
frequency, K, as the parameter. To simplify the 
writing, they will be referred to as the "Static" 


and "dynamic" angular pressure gradients respectively. 


Let &(k) be the ratio of dynamic angular pressure 
gradient versus static angular pressure gradient 
at a given location on the foil, namely 


&(K) = (ac_/da) / (dc_/da) (9) 
P u P s 


This ratio §(K) and the phase angle $ for several 
locations and reduced frequencies have been calcu- 
lated and are given in Table 2. The static angular 
pressure gradient (dC _/da)_ at a given location is 
approximated for mean? foil angles of 3.3 to 4.3° 
since leading edge cavitation inception typically 
occurred within this range. As seen in Table 2, 
for a given reduced frequency, the amount of 
reduction in dynamic pressure ratio (&€) remains 
almost a constant value in the range of 0.004 < 

X/C < 0.06 which covers the foil region over which 
leading-edge cavitation occurs. Consequently, if 
the foil is oscillated around the mean foil angle 
G54, the shape of the pressure distribution in the 
neighborhood of the suction peak and the peak 
location are essentially the same for both zero 


25 FIGURE 7. Static pressure distributions at 
foil angles of 3.5 and 3.25 deg. 


TABLE 2 - THEORETICALLY CALCULATED DYNAMIC PRESSURE 
RESPONSE AT VARIOUS (x/c) LOCATIONS 


REDUCED FREQUENCY, K 
0.05 0.1 0.3 0.5 0.75 1.0 1.5 2.0 


de 
At x/c = 0.0046, Gana = 33.52 


é -7-47, 10.53 711.25 -9.53 -7.91 6.97 -6.26 -5.96 
10 30.18 28.10 22.41 21.87 20.97 20.57 20.27 20.19 
Ls «90 84 67 65 +63 +61 -61 -60 


de 
At x/e = 0.0073, (F#) = 30.25 
8 


4 -7.46 -10.51 -ll.11 9.25 -7.45 6.32 -5.26 -4.61 
28.03 26.09 21.72 20.28 19.44 19.06 18.77 18.68 


g +93 -86 +72 67 +65 -63 +62 -62 


dc 
At x/c = 0.0117, G®). = 26.59 


da’s 
é -7.44 -10.46 -10.88 -8.78 -6.66 -5.22 -3.56 -2.31 
| apa! joy, 24.28 22.59 18.78 17.51 16.77 16.44 16.19 16.10 
& -91 -85 -70 -66 -63 +62 -61 -61 


dc 
At x/e = 0.018, (;4) = 23.0 
8 


é -7.41 -10.38 -10.53  -8.08  -5.48 -3.58  -1.02 41.13 
Joc} 20.72 19.27 15.99 14.89 14.25 13.96 13.75 13.67 

pu /% 

e 91 83 -70 +65 62 61 -60 -60 


de 
At x/e = 0.026,(7$), = 19.65 


é -7.37 -10.28 -10.08  -7.19 -4.01 -1.52 +2.15 +5.40 

Jac. | 17.72 16.47 13.63 12.67 12.12 1188s ella 71 ed. 68 
pu’ / 

E -90 84 69 65 62 61 60 -60 


dc 
At x/e = 0.035, (z£), = 16.79 


é  -7.31 -10.16 -9.53 6a LON = 2422 +.96 45.96  +10.51 

Jac. | TGAY  WAgiig)shlaehA 10.87 * 10.40 10.20 10.09 10.11 
pu /Q% 

z 91 -85 .70 65 62 -61 -60 -60 


d 


ic 
ia os 


At x/c = 0.058) Gab), = 12-78 
6 -7.18 -9.84 =g510)-=3633' +2533. 47-23 415.41) | +22.84 
, : 8.89 8.2 ; 7.78 7.83 8.0 
Vou joy 11.68 10.84 3 7.88 5 
g 92 85 .69 65 62 61 .61 63 


and nonzero reduced frequencies. This is an 

important conclusion which will be utilized later 

in the analytical prediction of cavitation inception. 
We will now proceed to develop a criterion to 


define the unsteady leading edge cavitation inception. 


Let 0;, denote the cavitation inception angle 
measured in a stationary test for given values of 

oO and Rn. As an example, at a cavitation number of 
Oo = 1.15 and Rn = 3 X 10°, cavitation inception 
occurred experimentally at ajo = 3.5°. The corre- 
sponding pressure distribution calculated using 
potential flow theory is given in Figure 7 with a 
suction peak appearing at around 1.6 percent chord 
aft of the leading edge. Let Cpsmin (jg) denote 
the minimum value of the static pressure coefficient 
Cops, at the foil angle 4 = aj,. It has been 
generally assumed that cavitation inception occurs 
when -Cpsmin (%is) = 9. Obviously, this simple 


Sh7fal 


relationship is not realized in the present test 
results (See Figure 7). This kind of discrepancy 
in applying the above scaling law for cavitation 
inception is a classic problem and has been exten- 
sively discussed in the literature [for example see 
Morgan and Peterson (1977) and Acosta and Parkin 
(1975) ]. 

One of the possible reasons for this discrepancy 
is that a finite amount of time is required for 
nuclei to grow. Thus, cavitation inception will 
depend not only on the magnitude of the suction 
pressure peak, but it will also depend on the shape 
of the pressure distribution in the neighborhood 
of the suction peak and the peak location. Since, 
as shown previously, these two features of the 
pressure distribution are essentially the same for 
zero and nonzero reduced frequencies of interest 
here, it will be assumed that the amount of time 
required for nuclei to grow is approximately the 
same for both a stationary and oscillating foil. 
Consequently, it is assumed that cavitation incep- 
tion occurs on the foil at nonzero reduced 
frequencies when the magnitude of ~Cpsmin (O;5) is 
encountered during the foil oscillation, for given 
values of oO and Rn. 

An analytical method will now be developed to 
predict leading edge cavitation inception on a 
oscillating foil based on inception measurements 
made on a stationary foil. Let AC, be given by 


Ke = |e. @ Jee, @). | (10) 
p Psmin 1s psmin fo) 


where Cpsmin (%) denotes the minimum value of the 
static pressure coefficient at 4 = 49 and Cysmin (dis) 
is the minimum static pressure coefficient at the 
cavitation inception angle 4;,. According to our 
assumption, unsteady cavitation occurs when the 
difference in the static loading AC, between ig 
and 4, is produced by the dynamic loading at some 
instant of time tj. Thus, unsteady cavitation 
occurs if 


= ACp (11) 


where Gan ea) | is the magnitude of the dynamic 
pressure response at time t = tj. If the value of 
Aig - % is small it follows from Eqs. (5) and (11) 
that 


a, (aC_/de). sin (Ob, + o)/=(@, = &))) (der /de) 
p u al is p s 
(12) 


where t:. corresponds to that instant of time at 
which Eq. (11) is satisfied. Small-amplitude 

motion has been assumed. The static angular 
pressure gradient is to be evaluated at the location 
of the suction peak corresponding to the steady 
condition a = a.. The unsteady inception angle 4;,, 
for a given reduced frequency K is obtained from 
Eqs. (1) and (12). 


cos 
Oo. = 6 ar (Gi, = Gh.) COSY) 
s ) 


= a] sing \j/l - ( (a, # 0) (13) 


As a consequence of Eq. (12), no singularity is 
expected inside the square root. Due to the 


372 


unsteady effect the inception angle ajy is generally 
different from AEC Let Aa be 


(14) 


which can be used to measure the magnitude of the 


unsteady effect. From Eq. (13), it follows that 
Ney S (SG) SOE, 1) 
is fo) & 
2 
Sagres (15) 
= Ci sino 1- a 
a 76 
(a, # 0) 


For the case where the phase angle > is small at 
the location of inception, we have 


Qa. 


1 
SO ae es oa) a 7 


a,& 
(16) 


Although a small phase angle, $¢, approximation is 
not required, it is useful to make this approxi- 
mation for the sake of discussing the implications 
of Eq. (15). The first term on the right-hand side 
represents the effect of the ratio of dynamic to 
static angular pressure gradients &(K) on unsteady 
cavitation inception. The second term represents 
the effect of phase angle, amplitude of oscillation, 
and the ratio of pressure gradients on cavitation 


6.0 


Sins a FROM EQ (15) 
<%= 6.0 DEG 


FROM EQ (15) 
m = 2.8 DEG 


CAVITATION-INCEPTION ANGLES a (DEG) 


RUNS 1205 TO 1208 
3.25 + 2.8 Sin wt 
9.75 m/s,o = 1.35 


4.0 
ab) 1.0 1.5 2.0 
REDUCED FREQUENCY, K 
FIGURE 8. Measured cavitation-inception angles for 


test runs 1205 to 1208, a) 2.8 deg. 


inception. For example, as seen 
phase angles of dynamic pressure 
leading edge lag behind the foil 
for values of K less than 1.0 at 
to this phase lag, 
tation inception is further delayed. 


in Figure 4 the 
response at the 
angle (negative 6) 
X/C 0.033. Due 
the occurrence of unsteady cavi- 
Contributing 


TABLE 3 - EXPERIMENTAL RESULTS ON UNSTEADY 


CAVITATION-INCEPTION ANGLES, & 


Run Vv f K 
No n/s HZ x10 

1205 9.75 4.0 2.4 - 307 
1206 9.75 B55) 2.4 ~423 
1207 C675) U5) 2.4 Belt 
1208 9.75 10.0 2.4 768 
1301 11.49 4.0 2.8 264 
1302 11.49 EGE) 2.8 - 362 
1303 11.49 is) 2.8 494 
1304 11.49 10.0 2.8 659 
1305 11.49 15.0 2.8 - 988 
1306 11.49 25.0 2.8 1.646 
1307 14.78 4.0 Sjo7/ 205 
1308 14.78 Dyed) 7) - 282 
1309 14.78 7.5 3.7 384 
1310 14.78 10.0 S\6i7/ 512 
1401 11.49 4.0 2.8 264 
1402 11.49 5.5 2.8 -362 
1403 11.49 7.5 2.8 ~494 
1404 11.49 10.0 2.8 659 
1405 11.49 15.0 2.8 987 
1406 11.49 25.0 2.8 1.646 
1407 14.78 4.0 3.7 205 
1408 14.78 5398) 3.7 282 
1409 14.78 708) 3.7 384 
1410 14.78 10.0 S}57/ 513 
1501 16.42 4.0 4.1 185 
1502 16.42 DoS) 4.1 255 
1503 16.42 Hoe) 4.1 347 
1504 16.42 10.0 4.1 - 462 
1505 16.42 15.0 4.1 694 
1506 16.42 25.9 4.1 al galey/ 


iu 
a a 

e ea is iu 

Deg Deg Deg 
1235) 2.8 4.3 5.28 
35 2.8 4.3 5.28 
35 AGEy 8) 5.28 
S5 2.8 4.3 5.28 
alga ks} “95 35 3.94 
aq ks} 395 315 3.94 
abe als} 395. 3.5 3.94 
LS 7995) Sop) 3.94 
3 195) S\o5) 3.94 
ib abs} 195 35 3.71 ~3.94 
Algal7d 1.00 35 Sey 19.4 
alent? 1.00 3.5 Sorat 
ial 1.00 S}G5) 3.94 
ye 1.00 365 3.94 
iL Gals} HI55 35) 3.90 ~4.20 
bats} 1.55 355 3.90 ~4.20 
13 Te55 35) 4.20 
alGals) 1.55 S}o5) 4.20 
1.13 1.55 375 4.20 
algal} 1.55 35) 3.60 ~ 3.90 
1.14 55) 3.5 3.90 
1.14 155, 3.5 4.20 
1.14 LG 5s) S355) 4.20 
1.14 IG S35) 35 3.90 
LS +95 S}58) 3.71 
1.15 095 355) So 7/4 
ilo als) +95 S}55) SJov/l 
algal) -95 e}55) 3.94 
1.15 295 Si5) Slozal 
nals) 95 &}55) Sef 394 


TABLE 4- THEORETICAL CALCULATION OF Aw AND & 
1205 TO 1208 AT x/c 


K é E Aa 
DEG DEG 

0.05 -7.41 91 42 
0.1 -10.38 .83 64 
0.3 -10.53 .70 .86 
0.5 -8.08 65 .87 
(57/5) -5.48 62 85 
1.0 -3.58 61 83 
1,5} -1.02 .60 .74 
2.0 +1513 .60 66 


to the inception delay is the oscillation amplitude 
a]. It is noted that the effect of oscillation 
amplitude on inception angle is strongly coupled 
with the phase angle. Thus, there will be no effect 
of q,; on inception if there is no phase shift. This 
is a consequence of the small oscillation amplitude 
assumption. As the reduced frequency K approaches 
zero, &>1 and 9-0, and the steady-state inception 
angle (Aa*0) is recovered. 


Experimental Results 


The range of Reynolds numbers covered in the cavita- 
tion tests was 2.4 to 4.1 x 10®. Because it is 
shown in Acosta and Parkin (1975) and Huang and 
Peterson (1976) that the existence of laminar 
separation may trigger premature cavitation in 
model tests, the boundary layer characteristics on 
the foil under stationary conditions were calculated. 
Within the Reynolds number range of the test program, 
the occurrence of laminar separation around the 
leading edge was not predicted. Flow visualization 
with dye injection supported this conclusion. The 
unsteady effect of foil oscillations on the boundary 
layer characteristics was not included in the 
calculation. 

In order to simulate prototype viscous effects 
as closely as possible, the model was tested at 
high tunnel speeds (11.5 to 16.4 m/s). Fora 
given body shape the laminar boundary layer thick- 
ness based on chord length decreases approximately 
as (Rn)-%. The effect of surface roughness on flow 
characteristics becomes more important at higher 
Reynolds number. This roughness effect was found 
in the present model tests with cavitation appearing 
prematurely in a few "weak" spots even though the 
surface was highly polished. This caused some 
difficulty in determining accurate values of 


373 


4y FOR TEST SERIES 


= 0.018 


NOTE: 


R 
fT 


4.30 DEG. 


R 
a 


3.25 DEG. 


2.80 DEG. 


R 
a 


cavitation inception angle. The relative importance 
of this uncertainty was minimized by applying the 
same cavitation inception criteria to both the 
steady and unsteady test results. 

Six series of oscillating foil tests were carried 
out. The test conditions and the test results are 
given in Table 3. Only 30 pictures were taken to 
cover one and 1/5 cycles of oscillating motion, 
and thus the angle at which inception occurred can 
only be related to two successive pictures. There- 
fore, in some cases, the inception angle is given 
in terms of a small range of angles instead of a 
single value. 

The test results from runs 1205 to 1208 are 
shown in Figure 8. In these cases, the foil was 
oscillated around a mean angle G) = 3.25° with a 


4.5 o = 3.25 + 1.55 Sin wt 


RANGE OF FOIL ANGLES IN 


4-0 TWO SUCCESSIVE PICTURES 


EQ(15), % = 1.55 DEG 


ZX RUNS 1401 TO 1406, o= 1.13 


CAVITATION-INCEPTION ANGLES, a |, (DEG) 


O RUNS 1407 TO 1410, w= 1.14 


T r 


oy 1.0 Wo 2.0 


REDUCED FREQUENCY, K 


FIGURE 9. Measured cavitation-inception angles for 
runs 1401 to 1410. 


TABLE 5 - THEORETICAL CALCULATION OF Aw AND @ FOR TEST SERIES 
1401 TO 1410 AT x/c = 0.018" 
K 6 rd Ao wy NOTE: 

DEG. DEG. DEG. 
0.05 =F) .91 .23 Be O,, = 3-5 DEG. 
0.1 -10.38 .83 134 3.84 
0.3 -10.53 .70 142 3.92 @ = 3,25 Wz8, 
0.5 -8.08 .65 .40 3.90 
0.75 -5.48 62 .36 3.86 CA =nEINSSEDEGE 
1.0 -3.58 61 .33 3.83 
a5 =1.02 -60 .26 3.76 
2.0 +1.13 .60 .20 3.70 


374 


TABLE 6 - THEORETICAL CALCULATION OF Aw AND ae 
FOR TEST SERIES 1301 TO 1306 
Iso Tomis06 nt x! Gio Ole 
NOTE: 
K é g Aa Mu 
DEG DEG DEG 
0.05 -7.41 ht 14 3.64 a, , = 3.5 DEG 
0.1 -10.38 - 83 oil 3.71 
0.3 -10.53 -70 +26 3.76 &% = 3.25 DEG 
0.5 -8.08 -65 -25 3.75 
0.75 -5.48 -62 23, SJoV/s) &% = .95 DEG 
1.0 -3.58 - 61 22 3.72 
ale) -1.02 - 60 .18 3.68 
2.0 Cra bo ale} - 60 ols} 3.65 
4.5 ‘ @ = Q,5. As seen in Figure 7, the steady suction 
@ = 3.25 + .95 Sin wt 5 
S peak occurs at a location near X/C = 0.018. The 
a predicted results based on Eq. (15) for the unsteady 
2 RANGE OF FOIL ANGLES IN cavitation inception are given in Table 4 and plotted 
J TWO SUCCESSIVE PICTURES in Figure 8 along with the experimental data. The 
a 4.0 phase angle ¢, and the ratio of dynamic to static 
2 angular pressure gradients, €, used in the predic- 
z tion were calculated with Geising's computer program. 
E Reasonably good agreement between theoretical calcu- 
S BUSI) o Gh Sh0 Use lations and experimental measurements is observed. 
2 6 ate Ree The test results from runs 1401 to 1406 and runs 
= Gite ees ere 1407 to 1410 are given in Table 3 and plotted in 
= El seen one nas Figure 9. In these cases, the foil was oscillated 
Ss around a, = 3.25° with a pitch amplitude of a) = 
1.55° and cavitation number of o = 1.14. The 
3 1.0 1.5 2.0 measured cavitation inception angle at the stationary 


REDUCED FREQUENCY, K 


FIGURE 10. Measured cavitation-inception angles for 
runs 1301 to 1310 and 1501 to 1506. 


pitch amplitude of a, = 2.8° and cavitation number 
fo} 1.35. The measured cavitation inception angle 
at the stationary condition was aj, = 4.3°. Within 
the range of reduced frequency 0.3 < K s 0.77, the 
measured unsteady cavitation inception angles were 
5, = 5-28°. That is, a significant delay of 
cavitation inception was observed at nonzero 
reduced frequencies. The unsteady inception angles 
computed. from Eq. (15) will now be examined. A 
previous discussion indicates that the suction 
pressure peak with the foil in oscillation is 
located at essentially the same X/C position as the 
suction peak corresponding to the steady condition 


TABLE 7 - THEORETICAL CALCULATION OF Aw AND @& 


AT x/c = 
K é g 
DEG 

0.05 -7.41 91 
0.1 -10.38 83 
0.3 -10.53 .70 
0.5 -8.08 .65 
0.75 -5.48 62 
1.0 -3.58 61 
ilo) -1.02 . 60 
2.0 casa li} .60 


0.018 


DEG 


1.06 


condition is ajg = 3.5°. The measured unsteady 
inception angles vary from ajy = 3.9 to 4.2° between 


K = 0.2 to 1.0 and oj = 3.6 to 3.9° at K = 1.65. 
The theoretical results obtained from Eq. (15) are 
given in Table 5 and plotted in Figure 9. Once 


again, a Significant delay in cavitation inception 
is observed experimentally and predicted theoret- 
ically at nonzero reduced frequencies. The agree- 
ment is fair. Part of the discrepancy between theory 
and experiment may be due to the lack of accurate 
resolution in measuring foil angles, since only 
30 pictures were taken to simulate 1 and 1/5 cycles 
of foil oscillation. The phase angle ¢$ is seen to 
change the sign from negative to positive values 
at K above 1.5. Consequently, at high values of 
reduced frequencies the amount of reduction in 
cavitation inception delay is reduced. This trend 
is observed experimentally and predicted theoret- 
ically. 

The test results from runs 1301 to 1306, 1307 


Lu 


NOTE: 


a, = 4.3 DEG 
s 


1 


a 


> = 3.25 DEG 


(ogy = 6.0 DEG 


Peres = 5 m/s 


8 fe) o = 2.5 


(DEG) 
O 


lu 


a =3+6 Sin wt 


CAVITATION-INCEPTION ANGLES, a. 


oy 1.0 1.5 2.0 
REDUCED FREQUENCY, K 


FIGURE 11. Measured cavitation-inception angles by 
Miyata (1972). 


to 1310 and 1501 to 1506 are given in Table 3 and 
plotted in Figure 10. The foil was oscillated 
around a, = 3.25° with a pitch amplitude of a) = 
0.95° and cavitation number o = 1.12 to 1.15. The 
measured cavitation inception angle at the stationary 
condition is aj, = 3.5°. The measured maximum 
steady inception angles are aj, = 3.70 to 3.93°. 

Once again, a Significant delay in cavitation 
inception at nonzero reduced frequencies is mea- 
sured. The theoretical calculations based on Eq. (15) 
are given in Table 6 and plotted in Figure 10. The 
agreement is reasonably good. 

In order to provide an insight into the effect 
of a] on cavitation delay, a theoretical example 
is computed in Table 7 and plotted in Figure 8. 

The foil is assumed to pitch around a, = 3.25° with 
an amplitude of a] = 6.0°. The stationary cavita- 
tion inception angle is assumed to be dj, = 4.3°. 

It is seen in Figure 8 that a significant delay in 
cavitation inception can be expected if the pitch 
amplitude is increased. This trend is also observed 
experimentally by comparing Figures 9 and 10. 

A two-dimensional foil undergoing pitch oscil- 
lations around an axis located at mid-chord was 
tested by Miyata et al. (1972). Two of the typical 
test results are produced in Figure 11 for com- 
parison. For the data shown the foil was oscillated 
with a pitch amplitude of a] = 6.0°. As expected 
(See Figure 8) a significant increase in the angle 
of cavitation inception is noticed for 0 < K < 1.2. 
For the second set of data shown in Figure 11, the 
foil was oscillated with a pitch amplitude of 
a] = 3.0°. A similarity between Figure 8 and 
Figure 11 is noticed. Although the foil shapes and 
the locations of pitch axes are different between 
Miyata's experiments and ours, the effect of 
unsteadiness on cavitation inception is similar for 
two model tests. A similar trend is also noticed 
in Radhi's experiments (1975). 

In the review papers by Acosta and Parkin (1975) 
and Huang and Peterson (1976), one is clearly 
reminded that even under steady conditions the 
cavitation inception process is extremely complex. 


375 


The theoretical prediction of cavitation inception 
angle under steady conditions is still very difficult. 
However, if the steady-state inception angle Ais is 
known from model tests, the effect of unsteadiness 

on cavitation inception may be estimated reasonably 
well by Eq. (15). Further investigations are 

needed to explore discrepancies between theory and 
experiment and the applicability of Eq. (15) to 
different foil shapes and for pitch axis different 
from the ones examined here. 


5. LEADING EDGE SHEET CAVITY INSTABILITY 


Wu (1972) has provided a very useful review of the 
physics of cavity and wake flows which may help to 
explain the observations of the present experiment. 
The essence of his description, applicable to the 
partial cavity condition, is that the free shear 
layer enveloping the cavity is unstable. The cavity 
occupies a portion of what can be referred to as 

the wake bubble or near wake, physically delineated 
in steady flow by a dividing streamline that is 
characterized by a constant or nearly constant 
pressure. For the condition where the cavity within 
the near wake is unsteady, the region is, strictly 
speaking, not defined by a streamline but by a 
material line which is difficult to observe experi- 
mentally. Because of this difficulty, we will 
initially assume that a quasisteady approximation 

is valid. When the cavity is just beyond the 
inception condition, its surface should be smooth 

as would be expected with a laminar shear layer. 

As the cavity grows in length the free shear layer 
would tend to become unstable. Transition from a 
laminar to turbulent shear layer initially takes 
place at the downstream end of the near-wake. A 
further extension of the cavity length causes 
transition to gradually move upstream along the 

free shear layer and the far-wake becomes irregular. 
This is comparable to the bursting of a short laminar 
Separation bubble in a single phase fluid. With a 
continued increase in cavity length, transition 

can begin at the leading edge of the cavity. 

In applying here the general features of the 
near-wake outlined by Wu (1972) no assumption is 
made as to whether the cavity occupies all of the 
near-wake region since the detailed physics of the 
region downstream of the cavity trailing edge are 
uncertain. One possibility is that the roll-up of 
the shear layer into vortices is completed at the 
near-wake closure where the vortices break away. 

If this occurs, it is reasonable to expect a 
periodicity in this shedding process. 

The variation in foil pressure at the P; location 
(see Figure 1) can give a useful insight into what 
is happening both downstream of the cavity and 
within the cavity when the foil is oscillaing with 
a pitch amplitude (a1) of 1.55°. Figure 12 shows 
an oscillograph record of the pressure variation 
P, for a cavity that reaches its maximum length 
downstream of the gage location. (A) is the 
region where the foil surface is fully wetted and 
the pressure appears to follow the variation 
expected as the angle of attack, a, is varied. At 
point (B), the cavity begins to cover the gage and 
in this example the pressure drops from the fully 
wetted pressure of 31.7 kPa to the cavity pressure 
in 0.003 seconds. The cavity pressure remains 
constant, except for several pressure spikes (C) of 


376 


FIGURE 12. Sample oscillograph record for 
the variation in foil surface pressure with 
foil angle at K = 0.26, Van) RS am/S\, 
Dey =i/6%2) kPa) =) 325°) Pal oo san Ot. 


millisecond duration, until the trailing edge of 
the cavity recedes past the P, gage (point D). 

The absolute magnitude of the cavity pressure 
could not be accurately determined from these 
experiments since the in situ pressure gages were 
not calibrated for the condition of a gas/liquid 
interface at the entrance to Helmholtz-type chamber 
over each gage. As shown in Figure 12, point B, 
the growing cavity does not appear to produce large 
foil surface pressure fluctuations at its downstream 
edge. However, when the cavity recedes, (ie., point 
D) then the foil surface pressure fluctuations can 
be comparable to the magnitude of the dynamic 
pressure. 

Based on photographic records it appears that 
when the cavity is expanding, its trailing edge is 
disturbed as one would expect if the shear layer 
were unstable at that location. Beginning at the 
cavity trailing edge and then moving forward, the 
cavity surface becomes highly disturbed, irregular, 
and bubbles are introduced into the shear layer, 
just as one would expect when transition in the 
shear layer moves forward. The cavity pressure, 
as measured by the gages Pj, P2 and P3, remains 
constant throughout this change in the surface of 
the cavity. 

During the early stages of sheet cavity growth, 
when only the cavity trailing edge appears disturbed, 
small regions of bubbles are shed from the sheet 
cavity trailing edge. This shedding process becomes 
more accentuated as the sheet cavity length increases 
and more of its surface becomes disturbed. High 
speed movies taken at 9,300 frames per second 
clearly show the highly turbulent characteristics 
originating at the trailing edge of the sheet 
cavity and progressively moving upstream. 

The sequence of vapor shedding from the cavity 
trailing edge, as determined by high speed movies 
taken at 9,300 frames per second, is as in the 
sketches of Figure 13. The photographs of Figure 
14 demonstrate a phase in the vapor bubble shedding 
process from the sheet cavity as sketched in 13c 
with two regions of shed vapor downstream. It 
should be noted that since the foil surface is 
very smooth, a reflection of the shed vapor is 
seen in the side views. Therefore a dashed line 


PRESSURE GAGE Ps 


PRESSURE GAGE Po 


PRESSURE GAGE P) 


FOIL ANGLE 


" FOIL ANGLE MAX. 


ai neeanimttamenen CAMERA PULSE TRACE 


has been added to Figure 14 to indicate the 
separation of the vapor and its reflected image. 
This shedding process is periodic and for the 
example shown in Figure 14 the shedding frequency 
at a given spanwise location is nominally 700 hertz. 
The view shown in Figure 14 covers nominally the 
center third of the foil span. Visual observations 
with strobescopic lighting indicate that the leading 
edge sheet cavitation, for nonzero values of K, 
typically consists of a series of 3 dimensional 
cavities across the span. 

In Figure 15 the top view shows a depression in 
the cavity surface (a) just above P); and a rise in 
cavity height (b) just downstream of the depression. 
At this instant a pressure "Spike" is detected by 
P, (see for example C in Figure 13). This condition 
precedes the shedding of a small region of vapor 
bubbles upstream of the sheet cavity trailing edge 
and significantly deforms the cavity trailing edge 
shape. It is the forerunner of the condition that 
will be referred to in this paper as "cloud" cavita- 
tion. It is interesting to note that after 
correlation: of over 600 photographs of the leading 
edge sheet cavitation with the pressure gage signal, 
the pressure "spike" always occurs when a depression 
in the cavity surface exists over the pressure gage. 
The converse, however, was not observed, ie., the 
"spike" can occur when no depression was discernable 
in the photographs. These "spikes" can occur without 
any significant gross change in the observed 
character of the sheet cavity surface in the general 


— 


Lo 


CAVITY FLOW 
+—— 


(a) FOIL LEADING (b) 
EDGE 
(c) x (4) 
FIGURE 13. Sequence of vapor shedding from the cavity 


trailing edge. 


<—TOP VIEW 


<— TOP VIEW 


vicinity of a pressure gage. 
"spikes" can occur during the life of the sheet 

cavity it appears improbable that they are due to 
the interaction of a postulated reentrant jet with 


Since numerous pressure 


the sheet cavity surfaces [Knapp et al. (1970)]. 
These "spikes" frequently have amplitudes which are 
comparable to the dynamic pressure and certainly 
exceed the estimated static pressure in the free 
shear layer over the pressure gage location. Quite 
possibly, these pressure "spikes" are due to the 
free shear layer itself since they only occur when 
the cavity surface indicates a turbulent shear 

layer is present. When the reduced frequency is 
high, for example at K = 1.65, the fully wetted 
pressure variation leads the foil angle by 68° and 
then no pressure "spikes" are produced at the pressure 
gage location as can be seen in Figure 16. At 

these high reduced frequencies the periodic shedding 


t SIDE VIEW 


4 ste VIEW 


7/7 


ROW 


FIGURE 14. Progressive shed- 
ding of vapor from sheet cavity 
trailing edge, K = 0.26, 

V_= 11.5 m/s, P_ = 76.2 kPa, 

© (o) come 

a = 3.25 + 1.55 sin wt. 


—— FLOW 


FIGURE 15. Cavity surface de- 
pression producing pressure 
"SPIKE" P, gage location, 

1S = Oo dain Ni, > aiboS m/s, 

Po =, 76-2 ry 6 = SoOSo, + 
1.55 sin wt. 


from the sheet cavity trailing edge downstream of 
the pressure gage is still observed. 

The last aspect of the leading edge sheet cavity 
instability to be described in this paper is that 
which will be called cloud cavitation. The three 
principle features of cloud cavitation for K 2 0 
are as follows: 


(1) A large surface area of the sheet cavity 
becomes highly distorted and undergoes a 
significant increase of overall cavity 
height in the distorted region, (Figure 17). 

(2) Once this distorted region begins to 
separate from the main part of the sheet 
cavity, the upstream portion of the sheet 
cavity develops a smooth surface and a 
reduced thickness (Figures 18 and 19). 

(3) The trailing edge of the smooth surfaced 


WwW 
~~ 
ss) 


FIGURE 16. Sample oscillo- 
graph record for the variation 
in foil surface pressure with 
foil angle at K = 1.65, 

Vina ellis> m/s, P_ = 76.2 kPa, 


a) colt 
ON= 35255 1 S5iisiniwwts. ERT 


FIGURE 17. Initial stage in 
the process of cloud cavitation 
formation, K = 0.51, woe 14.8 
m/s, mS 124.1 kPa, a = 3.25 

+ 0.95 sin wt. 


region then moves downstream, becomes 
unstable at its trailing edge, and quickly 
develops the characteristic appearance of 

the leading edge sheet cavity elsewhere 

along the span (see feature a in Figure 20); 
or, the trailing edge of the smooth portion 
of the sheet cavity moves upstream to the 
foil leading edge and the cavity disappears 
(Figure 21). In Figure 21 a dye trace 
injected at the foil leading edge can be seen. 


When the foil is stationary (K=0) cloud cavita- 
tion shedding can be very periodic as can be seen 
in Figure 22 which shows the oscillograph trace of 
the pressure gage response. The frequency of 
shedding for the condition illustrated in Figure 


ponvonslisa cbintasnsitauanoanwanshonanatbia 


PRESSURE GAGE P3 


PRESSURE GAGE Po 


PRESSURE GAGE Py 


FOIL ANGLE 


~=— FOIL ANGLE MAX. 


CAMERA PULSE TRACE 


se cecbetaonetta 


{ SIDE VIEW 


*—TOP VIEW 


22 is 42 Hz based on the response of the pressure 
gage P}. Figure 23 shows a photograph of the type 

of cavitation that produced the time pressure history 
of Figure 22. In Figure 23, (a) is a cloud just in 
the process of being shed, (b) is a cloud previously 
shed at a nearby spanwise location, and (c) is a 
cloud shed earlier at the same location as (a). 

The cavities did not shed in the manner of the two- - 
dimensional separation which typically occurs in 
sharp leading edge foils [Song (1969),Besch (1969), 
Wade and Acosta (1965)]. Instead, cavity shedding 
was highly three-dimensional and more or less 
independent of the sheet cavity instability occur- 
ring several cavity lengths away along the foil 

Span. However, it appears that for the trailing 

edge shedding and the cloud cavitation (at least 


| FLOW 


TOP VIEW 


—=TOP VIEW 


sa 


for K = 0) shedding occurrence alternated between 
several spanwise locations. This is clearly seen 
in Figure 23. 

Several other aspects of the cavity shedding 
process were apparent. The shed vapor had an 
initial gross rotation with the same direction as 
occurs in the free shear layer. This was evident 
from the high speed movies viewing the cavitation 
along the span (ie., a side view), and can also be 
inferred from the pulse camera photographs taken 
from the same view. The gross volume of the shed 
vapor had relatively little dispersion prior to its 
collapse but frequently developed within it regions 
of apparent bubble coalesence prior to collapse, 
as can be seen in Figures 14 and 24. 

On the basis of the previously described defini- 
tion of cloud cavitation, its occurrence was 
determined from available photographs. The presence 
of cloud cavitation as a function of the ratio of 


SIDE VIEW 


{ SIDE VIEW 


379 


———— [PIL 


FIGURE 18. Cloud cavity sepa- 
ration from leading edge sheet 
cavity (example 1), K = 0.99, 
We S dled m/s, P= 76.2 kPa, 
a = 3.25 + 1.55 sin wt. 


—— FLOW 


FIGURE 19. Cloud cavity sepa- 
ration from leading edge sheet 
cavity (example 2), K = 0, 

WV = deo8 m/s, P= 124.1 kPa, 
Oe—isey25eR 


maximum sheet cavity length, %m, to chord length, 

C, and reduced frequency K is shown in Figure 25. 

The data used to define the condition for the 

occurrence of cloud cavitation were all taken at 

nominally the same value of o. Figure 25 shows 

that for a given (2m/c) value, cloud cavitation 

can occur at nonzero K values whereas none would 

be apparent for K = 0. For example, if test 

conditions were adjusted such that 2m/e = O26, Ete 

0.3 < K < 0.4, then one could conclude as did Ito 

(1976) that there was a "critical" reduced frequency 

associated with the onset of cloud cavitation. 
Figure 25 also shows two curves representative 

of the influence of the value of a, on cloud cavita- 

tion. It is readily apparent from the data in 

Figure 25 that the conditions for cloud cavitation 

cannot be simulated by quasi-steady experiments. 

As shown in Figure 25, cavity length is strongly 

dependent on K. If the angle of a stationary foil 


380 


FIGURE 20. Final stage in cloud 
shedding process, K = 0.21, 

VEX = 14.8 m/s, Pi = 124.1 kPa, 

a = 3.25 + 0.95 sin wt. 


FIGURE 21. Desinent condition 
for leading edge sheet cavity; 
K = 0.49, V| = 11.5 m/s, 

P_ = 76.2 kPa, 0 = 3.25 + 1.55 
sin wt. 


was set to the maximum angle the oscillating foil 
attained (4.2° for a; = 0.95 in Figure 25), the 
maximum cavity length could be as much as a factor 
of two larger than for finite values of K (eg., 

K = 1.2). 

The data plotted in Figure 26 show that within 
the accuracy of the experiments, a variation in 
velocity from 11.5 to 16.4 m/s produced no signifi- 
cant change in the results shown in Figure 25 other 
than that expected for the small variation in o 
that occurred between tests. It appears that the 
parameters of K, 0, and aj, are sufficient to 
correlate all of the present data with the presence 
of cloud cavitation. 


6. CONCLUSIONS 


In order to improve the physical understanding of 
the cavitation inception process and the formation 


—— FLOW 


eee ee ee ee 
La tne 


i} SIDE VIEW 


LOCATION OF DYE INJECTION 


<—TOP VIEW f s10e VIEW 


of cloud cavitation on marine propellers, a large 
two-dimensional hydrofoil was tested in the DINSRDC 
36-inch Water Tunnel under pitching motion. The 
foil was instrumented with pressure transducers to 
measure the unsteady surface pressure due to foil 
oscillation, and photos were taken to correlate 
cavitation inception and cavity patterns. 

Prior to the occurrence of cavitation on an 
oscillating foil, the foil is in a fully wetted 
condition. Knowledge of the pressure distribution 
on a fully wetted foil can be expected to provide 
useful information for prediction of unsteady cavi- 
tation. Fully wetted, time dependent, experimental 
pressure distributions were compared with results 
from Giesing's method for calculating unsteady 
potential flow. Good correlation between the 
prediction and the experimental measurements was 
obtained for both dynamic pressure amplitudes and 
phase angles within the range of reduced frequencies 
investigated (K = 0.23 to 2.30). This good corre- 


Nera Ni eaten td Nt 


ee eee 


? 


i \| i | \ \ 
} | \ \ i\ i f | i\ j\ r é % A h\ 
Ls Ne Ned Nal Newel Sot Nga Singer mene! “enamel Neto Nin OES emf Site etna Sana 


Besa | 
} 


PAAR arte nnengnatn th RAFAL GR gee ac nN ty ANETTA NG 


| FLOW 


TOP VIEW 


lation supports McCroskey's conclusion that unsteady 
viscous effects on fully wetted oscillating airfoils 
are less important than unsteady potential flow 
effects, if the boundary layer does not interact 
significantly with the main flow. 

Six series of oscillating foil experiments were 
carried out in this test program to study the 
leading edge sheet cavity growth and collapse. 

A simplified mathematical model was developed to 
explain experimental results for leading edge sheet 
cavitation inception. The mathematical model 
utilizes Giesing's method for calculating the 
unsteady potential flow. A significant delay in 
unsteady cavitation inception was both predicted 
and measured. A further delay in cavitation 
inception was also observed and predicted with 
increasing pitch amplitude. It is shown that 
unsteady cavitation inception is a function of: 


ett aiat 


i cathnpinyyremrea FOIL ANGLE (K=0) 


SIDE VIEW 


381 


PRESSURE GAGE Ps 


“| PRESSURE GAGE P, 


\eotem.| PRESSURE GAGE Py 


FIGURE 22. Surface pressure 
fluctuations for K = 0, 
We datos m/s, P= 76.2 kPa, 


| CAMERA PULSE TRACE GS 3,250, 


a— FLOW 


FIGURE 23. Alternate spanwise 
cloud cavitation shedding for 
i= Op Wy = UitsS m/s, i = 1962 
kPa’, O) — 3/25). 


(1) the ratio of dynamic to static angular 
pressure gradients 


(dc_/da) / (dc _/da) 
iS) u Pp s 


and, 


(2) the phase shift between the foil angle 
and the dynamic pressure response. 


Due to the phase lag in pressure response a signifi- 
cant delay in unsteady cavitation inception is 
predicted theoretically and observed experimentally. 
Additionally, the angle at which cavitation inception 
occurs increases with increasing pitch amplitude. 

This effect results from a change in the phase angle. 
It is well known tha even in a steady condition 
the cavitation inception process is extremely complex. 

The theoretical prediction is still very difficult. 


FIGURE 24. Apparent coales- 
cence of vapor bubbles within 
cloud cavity; K = 0.28, 

We = Wot m/s, Po = 124.1 kPa, 
a = 3.259 + 1.55° sin wt. 


a = 3.25 +a Sin wt 
Vq = 11.49 m/s 6= 1.13 
(°) = FOIL ANGLE AT K = 0 


SEVERE CLOUD CAVITATION 
DURING CAVITY LIFE 


O O 


CAVITY LENGTH 2m/, 


MARGINAL OR NO 
CLOUD CAVITATION 


FIGURE 25. 


<—e—T0P VIEW 4 SIDE VIEW 


= 1.55 DEG 


CLOUD CAVITATION 
AT og ONLY 


= 0.95 DEG 


1.0 1.5 2.0 25) 


REDUCED FREQUENCY, K 


Variation in cloud cavitation with reduced frequency K and pitch amplitude Os 


@ (DEG) Px (kPa) Vo (m/s) 


@ 0.95 76.3 11.49 

Oo 0.95 124.3 14.78 

ro) 0.95 158.8 16.42 

\v] 1.00 165.7 16.42 

O 1.55 76.3 11.49 

a 1.55 127.7 14.78 
5 0 


FOIL ANGLE AT K = 0 


CAVITY LENGTH {m/¢ 


75 1.0 
REDUCED FREQUENCY, K 


Nevertheless, if the inception angle djg is known 
from the steady model tests, the unsteady effect 
on cavitation inception, to the first order, may 
be estimated by the present method. Since the 
present tests were carried out with only one foil 
shape and only one pitch axis location, further 
experiments are required, and in particular, the 
range of variables should be extended. 

Based on photographic observations of the leading 
edge sheet cavitation instabilities, it appears 
that the free shear layer and near-wake stability 
concepts reviewed by Wu (1972) give a reasonable 
qualitative description of the physical process. 
The inherent instability of the free shear layer 
and associated vortex shedding appear to provide 
a reasonable model for the breakup of a sheet 
cavity. However, the detailed hydrodynamics 
associated with the near-wake closure region can 
still only be postulated. The commonly held concept 
of a reentrant jet, Wu (1972), may provide a reason- 
able description applicable to the closure of the 
near-wake region during the actual shedding of 
vapor. For sheet cavitation extending over only a 
portion of the foil chord this reentrant jet may 
not actually penetrate the cavity itself but pene- 
trate only a locally separated region just down- 
stream of the sheet cavity trailing edge. In any 
event, the presence of a reentrant jet is not 
required to explain the inherent instability and 
breakup of the sheet cavity. 

For the conditions of the experiments reported 
here, where the gross flow is nominally two dimen- 
sional, the cavity instability is not coherent to a 
significant extent along the foil span. In other 


383 


FIGURE 26. Influence of Vie 
1.5 Por Gyr and K on cavity length 
(2m/c) . 


words, the cavity instability is highly three- 
dimensional and appears to be principally dependent 
on conditions in the immediate upstream free shear 
layer flow. The most extreme form of cavity insta- 
bility is manifest as a large shed cloud of vapor 
and thus referred to in the literature as "cloud" 
cavitation. 

Within the context of the experimental results 
reported here, the principle parameters controlling 
the formation of cloud cavitation are reduced fre- 
quency, K, cavitation number, o and foil oscillation 
amplitude, 4}. The maximum cavity length, (2m/c), 
is a function of these three parameters. However, 
it has been shown that predictions of lm/e at finite 
reduced frequencies cannot be based on the cavitation 
observations at zero reduced frequency. With o con- 
stant, the results show that it is possible to have 
no cloud cavitation at finite reduced frequencies - 
even though it was present on a stationary foil set 
to the maximum unsteady angle. However, if the 
steady foil is set to the mean angle of oscillation, 
@, and no cloud cavitation is present, then it is 
easily shown that at finite reduced frequencies 
cloud cavitation will be present. Thus, Ito's con- 
clusion that there exists a "critical" reduced fre- 
quency for the onset of cloud cavitation appears to 
be the result of the specific chosen values of the 
parameters, K, 0, and 4). 

The implication of the above results is that the 
prediction of the occurrence of cloud cavitation 
for hydrofoils in waves and propellers in wakes can- 
not be based solely on the performance in calm 
water or uniform flow. 


384 


ACKNOWLEDGMENTS 


Grateful appreciation is expressed to Mr. G. Kuiper 
for his skill in the pressure gage dynamic calibra- 
tion, assistance in the test set-up and his helpful 
discussions. Grateful appreciation is also due to 
Mr. R. Pierce for his excellent work in performing 
the data reduction. Finally, the reviews and 
constructive comments by Mr. J. McCarthy and 
Dr. W. Morgan are greatly appreciated. 

The work described in this paper was sponsored 
by Naval Sea Systems Command and the General Hydro- 
dynamic Research Program at DTNSRDC. 


REFERENCES 


Morgan, W. B., and F. B. Peterson (1977). Cavita- 
tion Inception, A Review-Progress Since 17th 
ATTC. Proceedings 18th ATTC, Annapolis, MD. 

van Manen, J. D. (1962). Bent Trailing Edges of 
Propeller Blades of High Powered Single Screw 
Ships. Proceedings of IAHR Symposium, Sendai, 
Japan 

Tanibayashi, H. (1973). Practical Approach to 
Unsteady Problems of Propellers. International 
Shipbuilding Progress, 20, No. 226. 

Ito, T. (1962). An Experimental Investigation into 
the Unsteady Cavitation of Marine Propellers. 
Proceedings of IAHR - Symposium on Cavitation 
and Hydraulic Machinery, Sendai, Japan. 

Tanibayashi, H., and N. Chiba (1968). Unsteady 
Cavitation of Oscillating Hydrofoil. Mitsubishi 
Heavy Industries Technical Report (in Japanese), 
Be WO Ao 

Chiba, N., and T. Hoshino (1976). Effect of Un- 
steady Cavity on Propeller Induced Hydrodynamic 
Pressure, Journal of the Society of Naval 
Architects of Japan, 139. 

Meijer, M. C. (1959). Some Experiments on Partly 
Cavitating Hydrofoils. International Shipbuilding 
Progress, 6, No. 60. 

Chiba, N. (1975). Behavior of Cavity Collapse as 
a Cause of Cavitation Damage of Propeller Blades. 
Cavity Flows, ASME Symposium Proceedings, 1975, 
ial als} 

Miyata, H. (1972). Pressure Characteristics and 
Cavitation. M. S. Thesis, Department of Naval 
Architecture, Tokyo University. 


Miyata, H. et al. (1972). Pressure Characteristics 
and Cavitation on an Oscillating Hydrofoil. 
Journal of the Society of Naval Architects of 
Japan, 132. 10; 107-115. 

Radhi, M. H. (1975). Theoretische und Experimen- 
telle Untersuchung uber den Kavitationseinsatz 
an Schwingenden Tragflugelprofilen. PhD Thesis, 
Technischen Universitat Berlin, D83. 

Hilten, J. S., et al. (1972). A Simple Sinusoidal 
Hydraulic Pressure Calibration. National Bureau 
of Standards, Technical Note 720. 

Nuttall, A. H. (1971). Spectral Estimation by 
Means of Overlapped Fast Fourier Transform Pro- 
ceeding of Windowed Data. WNUSC Report 4169. 

Abramson, H. N. (1967). Hydroelasticity with 
Special Reference to Hydrofoil Craft. NSRDC 
Report 2557. 

Giesing, J. P. (1968). Two-Dimensional Potential 
Flow Theory for Multiple Bodies in Small - 
Amplitude Motion. Douglas Aircraft Company, 
Report No. DAC-67028. 

McCroskey, W. J. (1977). Some Current Research in 
Unsteady Fluid Dynamics - the 1976 Freeman 
Scholar Lecture. Trans. ASME Journal of Fluid 
Engineers, 99, Series 1, No. 1, 8-38. 

McCroskey, W. J. (1975). Recent Review in Dynamic 
Stall. Proceedings, Unsteady Aerodynamics 
Symposium, The University of Arizona, 1-34. 

Acosta, A. J., and B. R. Parkin (1975). Cavitation- 
Inception - A Selective Review. Journal of Ship 
Research, 19, 4; 193-205. 

Carta, F. O. (1971). Effect of Unsteady Pressure 
Gradient Reduction on Dynamic Stall Delay. 
Journal of Aircraft, 8, 10; 839-841. 

Huang, T., and F. B. Peterson (1976). Influence of 
Viscous Effects on Model/Full Scale Cavitation 


Scaling. Journal of Ship Research 20, 215-223. 
Wu, T. Y. (1972). Cavity and Wake Flows. Annual 
Review of Fluid Mechanics, 4. 
Knapp, T. R., et al. (1970). Cavitation. McGraw- 


Hill Book Company. 

Song, C. S. (1969). Vibration of Cavitating Hydro- 
foils. St. Anthony Falls Hydraulic Laboratory, 
University of Minnesota, Project Report No. lll. 

Besch, P. K. (1969). Flutter and Cavity-Induced 
Oscillation of a Two-Degree-of-Freedom Hydrofoil 
in Two-Dimensional Cavitating Flow. NSRDC Report 
3000. 

Wade, R. B., and Acosta, A. J. (1965). Experimental 
Observations on the Flow Past a Plano-Convex 
Hydrofoil, J. Eng. for Power, ASME, No. 65-FE3, 
TOK 


Cavitation on Hydrofoils in 
Turbulent Shear Flow 


H. Murai, A. Ihara, 


and Y. Tsurumi 


Tohoku University, Sendai, Japan 


ABSTRACT 


Conditions and positions of inception, locations of 
zones, and aspects and behaviors of bubbles and 
cavities of cavitations occurring on two hydrofoils 
with the profiles of Clark Y 11.7 and 08 in shear 
flows and a uniform flow have been observed and 
measured, and correlated with measured pressure 
distributions on the hydrofoils and turbulence 
levels and size distributions of cavitation nuclei 
in free streams. 

At attack angles small for the profile, traveling 
cavitations begin near positions of minimum pressure 
and at cavitation numbers about the same as absolute 
values of minimum pressure coefficients, irrespective 
of flow shears in free streams provided local values 
are used. Discrepancies between conditions and 
positions of inceptions and pressure coefficients 
and their distributions, and sizes of traveling 
bubbles depend on qualities of free streams. 

On the hydrofoil with the Clark Y 11.7 profile, 

a traveling bubble in a zone of rising pressure, 
deforms, creating a projection in shear flow, or 
two projections in uniform flow, leaves only the 
projection and then collapses. On the hydrofoil 
with 08 profile, a traveling bubble collapses after 
the deformation caused by the instability of the 
bubble surface. On both hydrofils, bubbles collaps- 
ing symmetrically and asymmetrically, looking like 
micro jets forming, can be found. 

At attack angles large for the profile, fixed 
cavitations occur. Conditions and positions of 
inception are similar to those of traveling cavita- 
tions. In the boundary layers on both side walls, 
fixed cavitations occur at relatively large 
cavitation numbers, possibly equal to the absolute 
values of local minimum pressure coefficients, and 
even develop beyond the boundary layers. Cavitation 
zones on the low-speed side are larger than those 
on the other side, and those occurring in the 
boundary layers of uniform free streams are of an 
intermediate size. 


385 


At attack angles intermediate for the profile, 
fixed and traveling cavitations occur at the same 
time and tend to become fixed only on the Clark Y 
11.7 profile. On the 08 profile, fixed cavitations 
at the leading edge and traveling cavitations at 
about the mid-chord appear at the same time in shear 


flows, but only fixed cavitations occur and develop at 


the leading edge in uniform flows. 


1. INTRODUCTION 


Many researches on the cavitation characteristics 
of hydrofoil profiles have been published, and the 
appearance, the degree, and the effects on the 
hydrodynamic behavior of hydrofoil of the incipient 
and developed cavitations occurring on hydrofoils 
have been discussed by Numachi (1939, 1954), Daily 
(1944, 1949), and Kermeen (1956a, 1956b). Recently, 
the effects of the behavior of boundary layers and 
the turbulence in the free stream on the inception 
and development of cavitations on hydrofoils were 
reported by Casey (1974), Numachi (1975), and Blake 
et al. (1977). Although they have been concerned 
with cavitation occurring on hydrofoils in a free 
stream of uniform velocity, actual blades of 
hydraulic machines, including ships' propellers, 
work mostly in nonuniform flow, and the effect of 
nonuniformity might have to be examined as well. 

Investigations on cavitation occurring in shear 
layers have been made by Daily and Johnson (1956) 
in a zone of wall shear turbulence, by Kermeen and 
Parkin (1957) in a wake behind a circular plate 
and by Rouse et al. (1950) and Rouse (1953) in 
submerged jets. But research concerning the cavita- 
tion occurring on hydrofoils laid in a free stream 
with a shear is not available as far as the authors 
are aware. 

The present report is intended to clarify the 
influence of the spanwise shear, uniform in the 
core and the accompaning boundary layers on both 
sides of the free stream and its turbulence on the 


386 


> 


ice 
2 ALS, 
3 Lrgéte ep 
o 

he eee ae ee ea 
a MMM ha hdd Lhd hhh hhd hhhdE 


10096 


o 
Oo 
mM 


FIGURE 1. High speed water tunnel. 


inception and development of cavitation and the 
aspect and behavior of cavitation cavities occurring 
on two hydrofoil profiles with different cavitation 
characteristics. 


2. EXPERIMENTAL APPARATUS AND METHODS 


High Speed Water Tunnel 


The water tunnel used for the experiment is shown 
schematically in Figure 1. The tunnel contains 
180m? of water. The water is circulated by the 


centrifugal pump, P, whose revolution is controllable. 


Bubbles generated in the measuring section, the 
duct, and the pump mainly disappear in the reservior 
T. In the reservoir the water first flows upward 

to the free surface at the top of the reservoir, 

and then down very slowly through an area of 20m? 

to the bottom. Two spaces, one at the entrance 
corner of T and the other at the top of the tunnel, 
separate bubbles from the water and continuously 
remove the separated air. The water sucked up from 
the bottom of T turns to the horizontal direction 
through corner vanes, and enters the measuring 
section through the honey comb, S, made of synthetic- 
resin pipes of 26mm diameter, 6mm thick, and 450mm 
long. Then it flows through two nozzles, Nl and 


N2, which contract the cross section from 2100x1400mm? 


to 1500x1000mm2 and to 1200x200mm? , the room for 
installing the shear grid, and the nozzle for 
contracting the cross section from 1200x200mm? to 
610x200mm2. The contraction ratio is 24:1 in all. 
The water flowing out of the measuring section flows 
through the diffuser and back to the circulating 
pump P. 

The tunnel pressure is controlled by introducing 
compressed air to the top of the reservoir or by 


U/Uc 


3 
0,02 &, 
Uc 0.01 = 
oa 
© 8.23 m/s Q A 
| 4 5.60 m/s = 
07 | SIL 1 ati 


00 601 0.2 03 #04 O05 0.6 Or . OG 09 1.0 
y/h 


FIGURE 2. Velocity distribution at no grid condition. 


lowering the free surface led from the top of the 
tunnel, the maximum and minimum pressures being 
48x10° Pa and -0.8x10° Pa. The flow velocity at 

the measuring section is controlled from the measur- 
ing station by controlling the speed of the 
circulating pump P. 


Measuring Section 


The measuring section has a cross section 200mm 
wide and 610mm high and its total length is 3000mm. 
The first upstream 1000mm has two plexiglass windows 
in each side, and upper and lower wall. In this 
experiment, the hydrofoil is installed through two 
downstream-Side windows in both side walls. Figure 
2 shows the spanwise distributions of the velocity 
and the static pressure at the position of the 
mid-chord of the hydrofoil in the case of no grid. 
The velocity profile is almost uniform except in 
the 10% the boundary layers on both side walls. 

The static pressure, expressed as the difference 
from that at the side wall, is constant within the 
accuracy of this experiment. 


Hydrofoils 


Two hydrofoils have been prepared for the experiment, 
each of which has 100mm chord and 700mm span. ‘Two 
profiles have been selected; one is Clark Y 11.7 
and the other 08, dimensions of which are shown in 
Table 1. The former is selected for the purpose of 
examining the influence of the behavior of the 
boundary layer on the hydrofoil surfaces on the 
inception and development of cavitation and the 
aspects of cavitation bubbles or cavities, because 
it has a round nose and a surface pressure distri- 
bution rising toward the trailing edge. The latter 
is selected as a typical profile among ones designed 
by Numachi (1952) for high-speed flows, and has a 
sharp leading edge and comparatively good cavitation 
characteristics for its simple shape. 

The hydrofoil of the Clark Y 11.7 profile has 
14 and 13 piezometer holes of 0.4mm diameter on the 
suction and pressure surfaces respectively, and one 
of the 08 profile has 13 and 13 piezometer holes, 


“Table 1 Profile Forms of Hydrofoils 
@ilaals Ye Ayal g 7/ Og 
x Y x Y 

Upper Lower Upper Lower 

0.0 3590) | S550 0.0 OQclS O56 
Le2S (Sed 598 LS5A5 O35 1 OO 
2S GsSQ  do@7/ Bo) O57 O50 
by 50) T3990 O93 5) 50) 1247) O20 
ted Bois) © Moss} 10.0 Aol OO 
10.0 Veo 0) 5c7) WSs 35695 O.0 
USO) above Gite} (0) Gal) 20.0 S502 0.0 
2050 alt 3G 0.03 30.0 502 Oo) 
SJO6@! dhalo7o Oo 50.0 3400 OO 
Ayo ei;  alak 6A@) | OS @ 70.0 G502 60 
SOcoO)  dO6S2 . 0.) 80.0 S507, Wo 
60.0 Vols O50 90.0 259 OO 
70.0 3) DO lO) O70 oty oO) 
80.0 5222. ©. 7 > O57 Oo 
TORO 27.80) ORO OWI57S Oo35 OO 
100.0 OZ O10 100.0 OclG OO. 16 


Table 2 Positions of Piezometer Holes 
Culaels Ye Ibi 47 Og 
Upper Lower Upper Lower 
X & X % X & X % 
1 (0) al SON 4 2} 50) 
2 S}50)  alS 350 2 G60 as 6.0 
3 Go@) ike 6.1 3 i@,@ Le wO>o 
4 WO ca aly —al@ Gal 4 S50 aby abs 5 
5 IL Ey 5 ab iL} | LA Og} 5 BOO Ae 20.60 
6 PX0) Gal 19 2) Gal 6 30.0 19 30.0 
7 30.0 20 $3) Gal 7 40.0 20 40.0 
8 40.5 21 39) 59)  Soo0 2 5050 
©) S§@oal AA 30.10) J 690.0 22° O60 50 
EOS GlO\=2 2s) (50), 10 70.0 23 70.0 
dat 70.4 24 69°58 iil 8050 24 s050 
12 80539 25 V9 58 2 8550 25 85.0 
13 B60 26 84.8 13 GOO AS SO, O 


{ 
6789 
2345 


10 
Wi2i3 
aS 
ribo 
14116 118 19 20 21 22 23 24/26 
17 2) 


1S 


as are shown in Table 2. The holes are inclined to 
the direction of the free streams as to have no 


influence on the pressures measurements of each other. 


Pressures are measured by using a mercury-water 
manometer. 

For measurements of pressure distributions, the 
hydrofoil is shifted spanwise so as to allow the 
piezometer holes to cover the whole 200mm span. 

For observations of cavitations, the part of hydro- 
foil having no piezometer hole is used. 


Shear Grids 


In order to examine the influence of shear flow, 
the free stream at the measuring section has been 
made to have the simplest shear, that is, uniform 
shear. The grids for creating uniform shear flow 
are composed of straight rods arranged perpendicular 
to the free stream and the hydrofoil span with non- 
uniform spacings calculated by using the theory of 
Owen and Zienkiewicz (1957). The spaces near both 
side walls were modified according to Liverey and 
Turner, (1964) and Adachi and Kato (1973) and are 
shown in Table 3. In order to make two different 
free streams having the same shear but different 
turbulence, two grids were made, composed of rods 
with different diameters, 20mm for No. 1 and 15mm 
for No. 2. 


TABLE 3 Rod Spacings of Shear Grids 


Grid No. l 
Rod Number it 2 3 4 
distance from low-speed 
side wall (mm) AWoil SQLS) alosjo%4 al Sjs}3} 
Grid No. 2 
Rod Number 1 2 3 4 5 


distance from low-speed 
side wall (mm) 16 47.4 81.5 118.4 161.2 


387 


The shear grid is installed at a position 1500mm 
upstream from the mid-chord of hydrofoil, where 
the cross section of the duct is about twice as 
great as that of the measuring section so as to 
keep the grid free from cavitation. 


Measurement of Velocity and Static Pressure at the 
Measuring Section 


Spanwise distributions of the velocity and the static 
pressure are measured at the position of the mid- 
chord of the hydrofoil in the absence of the 
hydrofoil, by using a Prandtl-type Pitot tube of 
3mm diameter. They corresponded to the difference 
of static pressures at the inlet and exit of the 
second nozzle, N2, and the static pressures at the 
exit of the nozzle and the position 530mm upstream 
and 170mm below the position of the mid-chord of 
hydrofoil. 

It has been pointed out by Lighthill (1957) that 
total pressures measured by using a Pitot-tube in 
a shear flow exhibit larger values than real ones 
due to displacement effects of a Pitot-tube. The 
displacement thickness of the boundary layer on 
the Pitot-tube used in this experiment, having a 
ratio of outer to inner diameters of 0.6, is 
calculated as about 0.54mm by use of the empirical 
equation presented by Yound and Mass (1936) and 
Macmillan (1956). The error in this experiment 
caused by the displacement thickness is the order 
of 0.08mm/s for a shear factor of 0.15 in the core 
of the shear flow so that it can be neglected, 
except in the boundary layers. There the shear 
factor, on which the error is proportional, is 
considerably large, especially near both side walls. 

The static pressure at the measuring section is 
limited due to the following two reasons: at the 
upper limit, by the strength of the differential 
piezometer used for detecting the velocity at the 
measuring section; and at the lower limit by the 
need to prevent the shear grid from cavitating. 
The prescribed velocities at the measuring section 
are determined so as to keep the static pressure 
at the measuring section within the above-written 
limits for obtaining the inception and development 
of cavitation corresponding to the angles of attack 
of the hydrofoils, as shown in Table 4. 


Measurement of Turbulence 


Spanwise distributions of the components of turbu- 
lent velocity in the directions parallel to the 
free stream and perpendicular to the free stream 
and the hydrofoil span are measured at the position 
of the mid-chord of hydrofoil (in the absence of it) 
by using the Laser-Doppler velocimeter, DISA 55L 
Mark II. Each component of turbulent velocity is 


TABLE 4 Velocity and Pressure at the Test Section 
on Cavitation Experiments 


a (rad) Velocity (m/s) Pressure (105 Pa) 
0.0 iil ©) -0.64 ~ -0.45 
0.052 10.0 HOS  —0)53)5) 
0.105 9.0 —=0)5 (60) ~ 0), ALab 
(0) 5 15) 7/ 8.0 =0).33' + +0)-40 


388 


Pulse Modulator Pulse Power Amplifier 


Transmitter 


Sweep 
Oscillator 


X-Y Recorder 


Standard 
Vessel 


Stream 
Passage 


Receiver 


Frequency Analyser’ 
Esai cir ears Ang gl come 


ee Impedance Converter 
land Amplifier pe 


FIGURE 3. Schematic diagram for nuclei measurements. 


detected as an absolute value of the root mean 
square. 


Observations and Measurements of Cavitation 


Cavitation inceptions are seen by the naked eye 
under 50Hz, stroboscopic 3yus flash illumination. 
An incipient cavitation number is defined by using 
the static pressure at which the inception is 
detected while reducing the static pressure at a 
low rate and the local free stream velocity. How- 
ever, in the boundary layers the velocities at 
outside edges are taken while the free stream 
velocity is kept at the prescribed value. Desin- 
ences are too intermittent and indefinite to be 
detected definitely in the course of raising the 
static pressure. 

For the measurements of positions of inception 
and the observations of appearances of cavitation 
bubbles or cavities, photographs of 3 Us exposure 
and high-speed motion pictures of 3000 frames per 
second and 2 us exposure for each frame were taken. 
For the high-speed photography, the high-speed 
camera, FASTAX, was used synchronized with the 
high-speed stroboscope made by E. G. and E Co. Ltd. 


For the measurements of average locations and shapes 


of cavitation regions, photographs of 1/60 s 
exposure were used. 


U/Uc 


2 
PUc 


2 i eres mt aR 002 &, 
Wee Se eV ee = es 
ye ea ri 3 ¢ i -2—_¢—_2—¢ iS + — = ool x 
in ij | Uc —|0™ & 
0 583 m/s = 
4946ms 
07 + 4 ae i | Brn f 


(b) Grid No.2 


FIGURE 4. Velocity and static pressure distribution 
for shear grids. 


Relative Measurement of Cavitation Nuclei 


Size distributions of gas nuclei are measured by 
using the sound-attenuation method of Schiebe 1969. 
The measuring system is shown in Figure 3. The 
frequency range of swept pulses was 20kHz~1000 kHz. 
Both probes for emission and reception were 25mm 
diameters, made of a crystal, and exposed directly 
to water. The measurements were relative ones for 
comparison between the three cases of no grid and 
grids No. 1 and No. 2 because the system has not 
yet been calibrated for bubbles with prescribed 
definite diameters. 

Measurements were carried out at four positions 
in the spanwise direction at the mid-chord of 
hydrofoil perpendicular to the free stream and the 
hydrofoil span. 


3. RESULTS OF EXPERIMENT AND DISCUSSIONS 
Shear Flow at Measuring Section 
The velocity profiles normalized by each velocity 


at the mid-span and the distributions of the static 
pressure expressed as the difference from one at 


the side wall and normalized by each dynamic pressure 


at the mid-span for the grids No. 1 and No. 2 are 
shown in Figure 4. The flow shear for grid No. 1 
is uniform in the free stream core and the non- 
dimensional shear factor is 0.15. That for grid 
No. 2 is about the same as for grid No. 1 at half 
the core of the free stream on the high-speed side 
but smaller at the other half. The non-dimensional 
shear factor is 0.06. Both have boundary layers of 
10% thickness span on both sides. The static 
pressure is higher in the free stream core than on 
the side walls by about 1% or a little more of the 
dynamic pressure at mid-span. Scatters of plots 
are within the accuracy of this experiment. 


Spanwise Distribution of Turbulence 


Root mean squares of two components of turbulent 
velocity, one stream-wise and the other perpendic-— 
ular to it ahd the hydrofoil span, are measured 

in every free stream, and shown in Figure 5 
normalized by Uc. The velocity at the mid-span was 
kept at 9.86 m/s. When both are expressed as the 
turbulence levels based on the local velocity of 
free stream, U, for the cases of the two shear grids, 
both u'/U and w'/U vary so little in the spanwise 
direction that they can be regarded as constant 


10.0 =I T T 
i No.Grid| O | | Uc=9.86 
jo .Gri c=9. m/s 
NO. An) oan] 
8.0 |— [no.2 [ofan] A 
* a 
Bea arti 


u/Uc , w/Uc (%) 


FIGURE 5. Spanwise 
distribution of 
turbulence. 


104 103 104 1o% 
lam T Ir = = Tet oT T \aeeenal 
10S F ,aa a + 
- e} 4 
é 5o° & #8 6 ayo 
ra 
5 4 a 
oo F s a 4 a 
a 
lo* 08 8 Oo aj 
Y/h = 0.125 0.375 0.625 0.875 4 
kd = 0.65 ray 
a 
103 Ir Uc = 11.Om/s S © Bo 
© No.l ray ry o4 
& No.2 
ea el yp ss 1 I ! 
10-4 10% 104 1o3 


Ro (cm) 


FIGURE 6. Spanwise variation of size distribution of 
cavitation nuclei. 


within the accuracy of this measurement. u'/U was 
6.8 and 6.2% in the case of grids No. 1 and No. 2, 
respectively, and w'/U was 3.6% in the case of both 
two shear grids. It has been reported by Harris 
et al. (1977) that in a shear flow generated by a 
shear grid, w' and the other lateral component of 
turbulent velocity, say v', are almost the same. 
If it is also assumed that v' = w' in this experi- 
ment, the resultant turbulence levels were 8.5 and 
8.0% in the case of the grids No. 1 and No. 2, 
respectively, in the core of the free stream. In 
the case of no grid both u'/U and w'/U were 0.1%, 
and the turbulence can be regarded as isotropic 

at a level of 0.17%, in the core of free stream. 


Spanwise Variation of Size Distribution of Cavita- 
tion Nuclei 


Attenuations of sound pressures were measured at 
four positions in the spanwise direction ( 12.5, 
37.5, 62.5, and 87.5% span) from the low-speed side 
at the position of mid-chord in the absence of the 
hydrofoil, and at the cavitation numbers of 2.75 
and 0.65. Because the levels of attenuated sound 
pressures were not calibrated for micro bubbles 

of known sizes, sound pressure levels in the shear 
flows at each measuring position were compared with 
one in the uniform flow in which any spanwise 
variation was not noticed. Frequencies and 
differences of sound pressure levels were related 
to equivalent radii, and to differences of the 
numbers of cavitation nuclei from those in the 
uniform flow by using the formulae presented by 
Richardson (1947) and Gavrilov (1964). 

At a cavitation number of 2.75, any noticeable 
difference of size distributions between the shear 
flows and the uniform flow was not found. Ata 
cavitation number of 0.65, however, remarkable 
differences were ncticed as can be seen in Figure 
6. Numbers of nuclei with radii smaller than 24m 
in both shear flows are considerably larger than 
those in uniform flow, and the larger the numbers 
of nuclei the smaller the nuclei radii are. Size 
distributions in the two shear flows were not so 
different from each other in the high-speed sides 
of free streams, but in the low-speed sides, the 
shear flow made by the grid No. 2 is richer in 
nuclei, especially in the range of small radii, than 
the other. 


389 


Cavitation Inception 


Spanwise variations of local incipient cavitation 
numbers are plotted in Figure 7 for the Clark Y 
11.7 profile, and in Figure 8 for the Og profile. 
Also, spanwise variations of positions of minimum 
pressure for the case of no grid, grid No. 1, and 
grid No. 2, are shown. 


Clark Y 11.7 Profile 


In the case of no grid incipient cavitation 
numbers, kdi, are a little smaller than absolute 
values of minimum pressure coefficients, |cpmin|'s 
over the whole span at the attack angles, a, of 
0 and 0.052 rad, and in the core of free stream at 
a's of 0.105 and 0.157 rad. Differences between 
kdi's and |Cpmin|'s increase as a increases until 


06 | ae ! 
0) 0.2 0.4 0.6 0.8 
aS T T 
mo] 
= i Poe, (b) @= 0 052rad 
Ce ae ee ee 
ee a ae SS = = 
2 e \ ¥N 
By Numachi (1947) Sy 
0 L SiG | 
0 02 04 06 08 1.0 


(c) @=0O 105rad =I 


— | 


= - 
2.0 eS] 
oS 
18 alk eee | at 
(0) 0.2 04 0.6 0.8 1.0 


kdi 
on 
[o) 
S 


a= 0 157rad 


Gy 
PlOIP\O 

| 

| 


with tip clearance | 


Sor | 

z 

2 
z 
i] 
is 
B 

eens eee aee,| 


FIGURE 7. Spanwise variation of incipient cavitation 
numbers for the Clark Y 11.7 profile. 


390 


= + + + * 
go oe eG 


(a) a@=0O rad 


i pel et ec le a Th 
0 0.2 04 06 0.8 1.0 
20 
z xe 
15 A 
—— 
1.0 — =) 
05 : : ~| 
0 02 04 06 08 1.0 
35 a a el 


+ (c) a= 0.105 rad 
Ae 


0) 0.2 0.4 06 08 1.0 
Y/h 
FIGURE 8. Spanwise variation of incipient cavitation 
numbers for the Og profile. 


it reaches 0.105 rad, but become smaller at a = 
0.157 rad. 

At a's not smaller than 0.105 rad, fixed cavita- 
tions occur in the boundary layers at positions 
very close to both side walls at kd's much greater 
than local |Cpmin|'s. At the same time a zone of 
cavitation widens spanwise beyond each boundary 
layer with the inception so that detection of 
inception becomes difficult in the region neighbor- 
ing both boundary layers on the side walls. This 
is the reason the lack of points between y/h = 0.025 
~ 0.3 and 0.7 ~ 0.975. Frequency distributions of 
cavitation occurrences analyzed by using high speed 
motion pictures for 1 second illustrate those facts, 
as can be seen in Figure 9. 

In free streams with shears made by the grids 
No. 1 and No. 2, kdi's almost equal or are a little 
larger than local |Cpmin|'s. They vary spanwise 
under the influences of the flow shears in the 
core and the boundary layers on both side walls, 
and the accompanied secondary flows, except at 
a = 0.105 rad, which indicates that these free 
streams are rich in cavitation nuclei. At a = 0.105 
rad, kdi is a little smaller than |cpmin|, which 
can be assumed to be due to cavitations changing 
from traveling to fixed, as mentioned in the next 
section. 

Differences between kdi's and |Cpmin|'s in the 
boundary layers are larger than those in the case 
of no grid on the low-speed side, but are the 
contrary on the high-speed side, due to the 
secondary flows induced by the flow shears in the 
cores. The above-mentioned effect is most remark- 
able at a = 0.105 rad: kdi's in the boundary layer 
on the low-speed side in cases of the shear grids 
are larger than those not only in the case of no 
grid but also |cpmin|'s in the boundary layer, 
though only by a little. The mechanism causing 
the effect has been examined by measuring spanwise 
variations of static pressures on three points near 
the leading edge in the boundary layer on the low 


speed side at the attack angle of 0.105 rad in the 
case of the grid No. 2. It was confirmed that the 
detected incipient cavitation number, 2.53, in the 
boundary layer lies near the largest absolute value 
of the pressure coefficient based on the local 
velocities in the zone between 3 and 5 mm from the 
side wall. However, measured velocities in the 
zone are not very reliable. Symbols A in Figure 

7 show kdi's when the hydrofoil has a tip clearance 
of about 0.1mm on the high-speed side in the case 
of grid No. 1. It was found that effects of a 
boundary layer are weakened by tip clearances, 
especially at large angles of attack, although 
another cavitation occurs at the tip clearance. 


08 Profile 


At 0 angle of attack, traveling cavitations 

occurred and kdi aimost coincide with | Cpmin| in 

the case of no grid, but were larger than the latter 
in the case of grid No. 1. The difference decreases 
spanwise toward the high-speed side, in correspon- 
dence with the size distribution of cavitation 
nuclei. At angles of attack larger than O rad, 
however, fixed cavitations occurred and kdi's were 
much larger than measured |Cpmin|'s because of the 
lack of a piezometer hole at the position of the 
largest |Cpmin|, which is closer to the leading 

edge than the closest hole at 3% chord. At a = 
0.052 rad, in the case of the grid No. 1, another 
cavitation of the traveling type appears around 

the position of the measured second lc min | and the 
kd almost coincided with the measured Cpmin| . kdi's 
in the case of grid No. 1 were smaller than those 

in the other case on the high-speed side. The 
discrepancey can be surmised as due to the discrep- 
ancy between structures of laminar separation 
bubbles just behind the leading edge in the two 
cases because of the difference between turbulence 
levels. At a = 0.105 rad, kdi in the case of the 
grid No. 1 was larger than in the case of no grid 

in the core of free stream, but was the opposite 

in the boundary layer on the high-speed side wall. 
In the case of no grid, cavitations with long and 
wide zones occurred in the boundary layers on both 
sides close to the side walls and the leading edge 
as with the Clark Y 11.7 profile. 


Location of Incipient and Developed Cavitations 


Spanwise variations of positions of cavitation 
inception and front and rear edges\of (time) average 
zones of developed cavitation are shown in Figures 
10 and 11 for the profiles Clark Y 11.7 and 08 
respectively. Also are shown spanwise variations 


FIGURE 9. Frequency 
distribution of cavi- 
tation occurrence. 


pee I 
ee | 
(0) ray 
| SPiepree nf 8)eukeiee 
a 
08 + 0 g 4 $4 
a 4 a=Orad 
06 + 
i oT oo a 03 : 
= /By Numachi (1947) _| oo i Fae) Aut 
aaa aS Se on a= 0 rad 
| ze l al 
) 02 04 06 08 10 
Y/h 
a I T T 
S 
« ° e 5 =) @ 
os; 8 g EP 


04 
Re ko oe % 
Clark Y 11.7 
RO! a = 0.052 rad 
| i | | 
io) 0.2 0.4 0.6 0.8 1.0 
y/h 
FIGURE 10 (a) (b). Spanwise variations of 
position of inception and front and rear 
edges of cavitation zones for the Clark Y 
11.7 profile. 
1.0 aan Uhm 
° = 8 6 
_ 08 br 5 z 
ae 
eS a g a 3 al 
Q § 9 o g 
04 rn a= 0.105 rad a 
—_— kee os LO. =Cpmin 02 
No. | a at = = 
No.2 2 oF Be (c) 
[Ma lo. | @ > with Tip clearance] _| 0! (lees ab delay) 
i 6 s # —] a = 0.105 rad 
mI ! | | 0 
(0) 2 04 06 08 10 
y/h 
O5 T T 3 T 
<< 8 ) a 
i 04 b & a =) o a S) 
E a 
03 mal 
a 
; i) 
a g o 
o2 g 4 as g z 4 
a=0.157rad (d) 
kai [25 | 2.0 | -CPmin Clark Y 11.7 
ST RAE i a = 0.157 rad 
No 2 s a [o[—--— 0.1 
No! eo! | with fip clearance i 
AL i a 3 
| eS es Sem Shah's 
Ce) 02 04 06 08 1.0 


y/h 


FIGURE 10 (c)(d). Spanwise variations of 
position of inception and front and rear 
edges of cavitation zones for the Clark Y 
11.7 profile. 


391 


of positions of minimum pressure, in each case 
indicated in the figures. The bottom and the 
second (at a = 0, 0.052 rad) groups show the 
positions of inception or front edges of cavitation 
zones and refer to the scales written on the right- 
hand side, and the other groups show rear edges of 
cavitation zones and refer to the scales written on 
the left-hand side. Open symbols correspond to 
traveling cavitations and closed and semi-closed 


hey T | 
a 
HOM = 
S A 
Az a RK 
3 e 
08 6 ® “4 
Q 
in ) 
0.6 ° O =| 
° a= Orad 
|_kdi [056 [053 |-Comin | 
iegid TO Loke. 5 
No.1 [AT [AlAlAla|—-— 
= =| OB. 
ah RN Es Ree ae SL 3 
| | | 04 
(0) 0.2 04 06 0.8 10 
Y/h 
(a) 
08 
a= O rad 
1.0 
= 4 A 
= A 4& S) 
es re ) =) e g 
0.5 cal 
a = A a a 
0 2 R a o a i) 4 
a= 0.052 rad 
[ [_kdi [0.9 0.7 [=Cpmin 
[NoGrid | |@| {2 o 
[Not | fal lalaAtala|—-—= ae 
uae es a nee | os 
Za 
eg e | e ese e—20 
(0) 02 0.4 0.6 0.8 Ke) 
Y/h 
(b) 
O08 
a = 0.052 rad 
| I | 
a= 0.105 rad 
0.15 [ kdi [2.5 | 2.0 — 
[No Grid e| jo! jo o 
(No. 1 al [al [a g 
0.10 4 
me Hw 4 g e 
Ze A a 2 
0.05 fr z a 4 
0 ) ) a) ) 
9 he 
(0) 0.2 0.4 0.6 08 1.0 
Y/h 
(c) 
08 


a = 0.105 rad 


FIGURE 11. Spanwise variation of the 
position of inception and the front and rear 
edges of cavitation zones for the Og profile. 


392 


symbols to fixed. kd's indicated in the figure are 
based on the velocity at the mid-span. 


Clark Y 11.7 Profile 


In the case of no grid, traveling cavitations oc- 
curred a little downstream from positions of minimum 
pressure at a = 0 and 0.052 rad. Front edges of 
average zones of cavitation move forward beyond posi- 
tions of minimum pressure as kd is reduced, uni- 
formly in the core of the free stream. At a = 0.105 
rad, in the core of the free stream, cavitations, 
mainly traveling mixed with fixed, occurred just 
downstream from positions of minimum pressure. How- 
ever, with a small decrease of kd from the incipient, 
the type of cavitation changes to fixed and the 
front edges of cavitation zones move backward from 
positions of inception and forward with a further 
decrease of kd. In the boundary layers on both 

side walls, fixed cavitations occurred very close 

to the leading edge of the profile and to the side 
walls, and front edges of cavitation zones move 
little as kd is reduced. At a = 0.157 rad, fixed 
cavitations occurred just downstream from positions 
of minimum pressure and front edges of cavitation 
zones moved forward just a little and never ex- 
ceeded positions of minimum pressure, in the core 

of free stream. In the boundary layers on both 

side walls, fixed cavitations occurred just down- 
stream from the leading edge of the profile and al- 
most attached to the side walls, and front edges of 
cavitation zones moved little as kd was reduced. 

At all attack angles, lines of rear edges of 
cavitation zones have shapes similar to the velocity 
profile at kd's a little smaller than the incipient. 
But rear edges move backward with a further decrease 
of kd to be almost uniform in the spanwise direction. 

In cases of grids No. 1 and No. 2, positions of 
inception are closer to positions of minimum 
pressure than in the case of no grid, in correspon- 
dence with size distributions of cavitation nuclei: 
Front edges of cavitation zones move forward beyond 
positions of minimum pressure in the cores of free 
streams at a = 0, 0.052, and 0.105 rad. Ata=0, 
0.052, and 0.105 rad, incipient cavitations are of 
the traveling type, but at a = 0.105 rad, in the 
cores of the free stream, cavitations sometimes 
change their type from traveling to fixed as kd is 
reduced, and in those cases front edges of zones 
of fixed cavitations move backward from the inception 
position. In the boundary layer on the low-speed 
side wall a fixed cavitation occurred very close 
to the leading edge of the profile and to the side 
wall, but no inception of cavitation of any type 
can be detected in the boundary layer on the other 
side wall, in the range of kd in this experiment. 

At a = 0.157 rad, fixed cavitations occurred at 
positions of minimum pressure, including the boundary 
layers on both side walls, and front edges of 
cavitation zones move little. 

At kd's a little smaller than the incipient, 
lengths of cavitations are larger on the high-speed 
Side than on the other side at a = O and 0.052 rad. 
At a = 0.105 and 0.157 rad, however, they are larger 
near the wall on the low-speed side than on zones 
more distant from the wall. Rear edges of cavita- 
tion zones have a tendency to be uniform in the 
spanwise direction at all attack angles as cavita- 
tions develop. 


Much difference between the two grids in the loca- 


tions and movements of cavitation zones cannot be 
found. 


08 Profile 


At O angle of attack, positions of cavitation 
inception and movements of front and rear edges of 
cavitation zones with a decrease of kd, compared 
with positions of minimum pressure, are quite 
similar to those of the Clark Y 11.7 profile in 

the cases of no grid and grid No. 1. However, at 
a's larger than 0, fixed cavitations always occurred 
at the leading edge over the whole span, irrespective 
of the existence of the shear grid. Front edges of 
cavitation zones never moved from the leading edge 
as kd's were reduced. Lengths of cavitation zones 
do not grow much, owing to the steep negative- 
pressure zones just behind the leading edge, until 
kd's are reduced to about the second Cpmin|'s. 

But in the case of no grid, once kd's increase, 
they develop suddenly beyond positions of minimum 
pressure and tend to be uniform in the spanwise 
direction as can be seen in Figure 13(b) at a = 
0.052 and kd = 0.7. In the case of grid No. l, 
however, lengths of the fixed cavitation do not 
grow enough to reach positions of minimum pressure. 
Instead cavitations of the traveling type appear 
around positions of the second minimum pressure, as 
can be seen in Figure 13(b) at a = 0.052 and kd = 
0.7 and as shown in Figure 11(b) by the symbols A. 
The length of the cavitation zone is about the same 
as that in the case of no grid in the free stream 
core but smaller than that in the boundary layers 
on both sides, at the beginning of development. 

At a = 0.105 rad, the length of the cavitation zone 
is much larger than that in the case of no grid at 
the beginning of development, but becomes about the 
same as the others with a further decrease of kd. 


Aspect and Behavior of Cavitation Bubbles and 
Cavities 


Figures 12 and 13 show several examples among the 
3us-exposure photographs and an example of high- 
speed motion pictures of cavitations taken at the 
inception and each stage of development occurring 
on the Clark Y 11.7 and 0g profiles, respectively. 
Cavitation numbers indicated in the figure on the 
left hand side are based on the velocity at the 
mid-span. 


Clark Y 11.7 Profile 


At a = O and 0.052 rad, incipient cavitations are 
of the traveling type in all cases, and in general, 
the bubble radius and number of bubbles in the case 
of no grid were the largest and the smallest, 
respectively, of the three cases, followed by the 
case of grid No. 1, which agrees with the size 
distributions of cavitation nuclei given previously. 
Each bubble is circular when observed perpendicular 
to the hydrofoil surface, but as the cavitation 
number is reduced, two, in the case of no grid, or 
one, in both cases of two shear grids, horn-like 
projections are projected behind each bubble from 
the downstream or both sides. The groups of plots 
lying second from the bottom in Figures 10 (a) (b) 
show positions of the upstream tips of the projec— 


393 


No Grid Grid No.4 Grid No.2 
. (a) Oo Oar 


No. Grid 


FIGURE 12 (a) (b). Cavitation on the hydrofoil of the Clark Y 11.7 profile. 


Ww 


© 


Kd 


v/s) 


Kd 


2.5 


Z.25 


2.0 


FIGURE 12 (c) (d). 


No. Grid Grid No.1 
(c) @= 0.105 rad 


(d) = (Sz) Aol! 


Cavitation on the hydrofoil of the Clark Y 11.7 profile. 


200mm 


Kidii=n2 ai 


tions, which seem to be little affected by either 
kd or the shear of the free stream. The projections, 
in the case of no grid, are supposed to be generated 
in cores of trailing vortices and adhere to the 
hydrofoil surface, because velocities of the bubbles 
exceed those of surrounding water in regions down- 
stream from positions of minimum pressure. It can 
be seen in high speed motion pictures shown in 
Figure 12(f) that the main body of bubbles, having 
generated projections, decay, leave behind them 
projections of two string-like bubbles, and then 
collapse. In cases of the two shear grids, bubbles 
are inclined upward toward the high-speed side due 
to the secondary flow caused by the flow shears. 
Trailing vortices on the low-speed side reach the 
hydrofoil surface easier than those on the high- 
speed side. Bubbles which generate projections be- 
come fewer as kd is reduced in the case of the 
shear grids. Several bubbles can be found which 
seem to collapse and generate micro jets. 

At a = 0.105 rad, cavitations of both types, 
traveling and fixed, appear, though the former are 


395 


FIGURE 12 (e). Behavior of 
fixed cavitation on the hydro- 
foil of the Clark Y 11.7 pro- 
file, a = 0.157 rad, flow up 
to down, 12 ms between frames, 
2us exposure. 


fewer than the latter. Front edges of fixed cavi- 
tation zones are round compared with tips of the 
above-mentioned projections. At a = 0.157 rad, only 
fixed cavitations occur. A cycle of formation of 
the break off of a fixed cavity is shown in high 
speed motion pictures in Figure 12(e). At IEILASKE p 

a clear bubble is generated, like those observed in 
our laboratory on the surface of an axisymmetrical 
body with a hemispherical nose. The bubble develops 
in both streamwise and spanwise directions. The 
middle part of the spanwise breadth of the bubble 
becomes bubbly, then wavy, and after the development 
of the middle part breaks off in pieces of micro- 
bubble clouds which are transported downstream al- 
though a few remaining small parts grow and 
disappear. 


08 Profile 


At O angle of attack tiny bubbles of traveling 
cavitation can be found at a kd a little smaller 


396 


than kdi, in both cases of no grid and the grid No. 
1. Little difference between sizes of the bubbles 
can be noticed, although bubbles can hardly be 

found on the low-speed side in the case of the grid 
No. 1. As cavitation numbers are reduced, however, 
the bubbles grow larger and are fewer in the case 

of no grid, due to the difference in size of 
cavitation nuclei as stated above. Bubbles deforming 
to generate projections like those on the Clark Y 
11.7 profile can barely be found. Instead, cavities 
collapsing to clusters of small bubbles appear. 

The discrepancy of the collapse aspect between the 
two hydrofoils can be considered to be caused by 

the difference of pressure distributions. Cavita- 
tions of the above type become more than traveling 
bubbles with the decrease of kd, in the case of 

the grid No. l. 

At a = 0.052 rad, only fixed cavitations occurred 
at the leading edge in cases of both no grid and 
grid No. 1. The fixed cavitations grown without 
changing the front edges of cavitation zones from 
the leading edge and develop their lengths slowly 
until the kd's are reduced to about the second 

Cpmin|'s, being about equal to each other and 
existing at the mid-chord in both cases. Nonuniform- 
ity of lengths can be found in the case of grid 
No. 1. When kd's reach the second |Cpmin|'s, how- 


FIGURE 12 (f). Behavior of 
traveling cavitation on the 
hydrofoil of the Clark Y 11.7 
profile, a = 0 rad, flow up to 
down, 0.3 ms between frames, 
2us exposure. 


ever, a remarkable difference in the aspects of 
cavitations between the two cases occurs in spite of 
only a small difference in the measured pressure dis- 
tribution. In the case of no grid, fixed cavitation 
develops beyond the position of minimum pressure, 
whereas in the case of grid No. 1, the rear edge 

of the zone of fixed cavitation does not reach the 
position of minimum pressure. Instead, another 
cavitation of the traveling type appears around the 
position of minimum pressure, and bubbles of the 
traveling cavitation are found more on the high- 
speed side. The mechanism of this difference can 

be surmised as follows: a free shear layer on an 
interface between cavity and water may be laminar 
near the point of inception in either case, but 

the distance necessary for its transition in the 
case of no grid is larger than in the case of grid 
No. 1 because of the difference of the turbulence 
level in the free stream between the two cases, 

and the distance necessary for a cavity surface to 
reattach the hydrofoil surface might be the same. 
The fact that the cavity surfaces in Figure 13(b) 

at kd = 0.7 are clear in the case of no grid but 
wavy in the other case may show this. Furthermore, 


the effect of rolling up the cavity surface caused 
by the secondary flow may be expected in shear flow. 
At a = 0.105 rad, only fixed cavitations can be 


Kd 


0.6 


0.56 


054% 


(a) @ = 0 rad 


found in both cases. Even in the case of the grid 
No. 1, much uniformity of cavitation zones can be 
found, although some tail wisps of cavitation can 
be found in the case of no grid, e. g., ones 
gathered in cores of streamwise vortices. 


4. CONCLUDING REMARKS 


Conditions and positions of inception, locations of 
zones, and the aspect and behavior of bubbles and 
cavities of cavitations occurring on two hydrofoils 
with the profiles of Clark Y 11.7 and 0g in shear 
flows made by shear grids and a uniform flow have 
been observed and measured. They have been corre- 
lated with measured pressure distributions on the 
hydrofoils and the qualities of free streams, i.e. 
turbulence levels and size distributions of cavita- 
tion nuclei in free streams. The main conclusions 
deduced from the results may be summarized as 
follows. 

At attack angles small for the profile, when 
pressure distributions have gradual chordwise 
changes, traveling cavitations incept near positions 
of minimum pressure and at cavitation numbers about 
equal to absolute values of minimum pressure coeffi- 
cients, irrespective of flow shears in free streams, 
provided local values influenced by flow shears are 
-used. Discrepancies between conditions and posi- 
tions of inceptions, and pressure coefficients and 
their distributions depend on the free stream quali- 
ties. The sizes of traveling bubbles depends on the 
size distribution of cavitation nuclei. 

On the hydrofoil with the Clark Y 11.7 profile, 
having a relatively large positive pressure gradient, 
a traveling bubble in a zone of rising pressure 
deforms, creating a projection in shear flow, or 
two projections in uniform flow, leaves only the 


397 


FIGURE 13 (a). Cavitation on the 
hydrofoil of the Og profile. 


projection and then collapses. On the hydrofoil 
with the 08 profile having gradual pressure gradient, 
a traveling bubble collapses after the deformation 
caused by the instability of bubble surface. On 
both hydrofoils, bubbles collapsing symmetrically 
and asymmetrically, looking like micro jets forming 
can be found. 

At attack angles larger for the profile, when 
the pressure distribution declines steeply followed 
by a relatively large positive pressure gradient, 
fixed cavitations occur. Conditions and positions 
of inception are similar to those of traveling 
cavitations, although discrepancies of them from 
pressure coefficients and their distributions are 
less than those of traveling cavitations. In the 
boundary layers on both side walls, fixed cavitations 
occur at relatively large cavitation numbers, 
possibly equal to absolute values of local minimum 
pressure coefficients. They develop in both stream- 
wise and spanwise directions even far enough beyond 
the boundary layers to affect cavitation inceptions 
in zones neighboring the boundary layers. Cavita- 
tion zones on the low-speed side are larger than 
those on the high-speed side. Fixed cavitations 
of this kind occur in the boundary layers on both 
sides of uniform free streams also. 

At attack angles intermediate for the profile, 
fixed and traveling cavitations occur at the same 
time and tend to become fixed only on the Clark Y 
11.7 profile. On the 08 profile, fixed cavitations 
at the leading edge and traveling cavitations at 
about the mid-chord appear at the same time in shear 
flows, but only fixed cavitations occur and develop 
at the leading edge in uniform flows. Discrepancies 
of conditions and positions of inception from 
pressure coefficients and their distributions are 
the largest of the three cases mentioned on the 
Clark Y 11.7 profile, but about the same as above 
mentioned two cases, on the O08 profile. 


FIGURE 13 (b) (c). Cavitation on 
the hydrofoil of the Og profile. 


(c) 


Q>= 0.057 rad 


a=0.105 rad 


ACKNOWLEDGMENT 
The authors wish to express their thanks to Mr. S. 


Onuma, the technician of Institute of High Speed 
Mechanics for his assistance in the experiment. 


NOMENCLATURE 


Cp: pressure coefficient 


|Cpmin | absolute value of minimum pressure 
coefficient 
fp: number of total occurrences of 
cavitation 
fy: number of local occurrences of 
cavitation at position y 
h: width of measuring section 
kd: cavitation number 
kdi: incipient cavitation number 
1: chord length of hydrofoil 
n(R,): number of bubbles at radius R, 


p: static pressure at hydrofoil surface 
or in free stream 
Py: static pressure at side wall of 
measuring section 
Rg: bubble radius 
U: local free stream velocity 
Uc: velocity at mid span of hydrofoil 
installed in measuring section 
RMS values of turbulence velocity 
components parallel to free stream, 
parallel to hydrofoil span and 
perpendicular to u' and v', respec- 


tively 
X, Y; X, y : co-ordinate system fixed in hydrofoil; 

the X(x) axis is parallel and the 
Y(y) axis is perpendicular to the 
chord of the hydrofoil 

a attack angle in radian 

fe) water density 

r chordwise distance from leading edge 
of hydrofoil to rear edge of cavita- 
tion zone 

r 


o ? Chordwise distance from leading edge 
of hydrofoil to inception point or 
front edge of cavitation zone 


REFERENCES 


Adachi, T., and E. Kato (1973). An experimental 
study on the turbulent linear shear flow. J. 
Japan Soc. Aeronautical & Space Sci. 21, 573. 

Blake, W. K., M. J. Wolpert, and F. E. Gieb (1977). 
Cavitation noise and inception as influenced by 
boundary-layer development on a hydrofoil. J. 
Fluid Mech. 80, 617. 

Casey, M. Y. (1974). The inception of attached 
cavitation from laminar separation bubbles on 
hydrofoils, Proc. Confe. Cavitation held at 
Edinburgh, 1. 

Daily, J. W. (1944). Force and cavitation charac- 
teristics of the NACA 4412 Hydrofoil, Calif. 
Inst. of Tech. Hydrody. Lab. ND 19. 


399 


Daily, J. W. (1949). Cavitation characteristics 
and infinite-aspect ratio characteristics of a 
hydrofoil Section. Trans. ASME 71, 269. 

Daily, J. W., and Johnson (1956). Turbulence and 
boundary-layer effects on cavitation inception 
from gas nuclei. Trans. ASME, 78, 1695. 

Gavrilov, L. €. (1964). On the size distribution 
of gas bubbles in water. Soviet Phy.-Acoust. 
US, Ake 

Harris, V. G., J. A. H. Graham, and S. Corrsin (1977). 
Further experiments in nearly homogeneous tur- 


bulent shear flow. J. Fluid Mech. 81, 657. 
Kermeen, P. W. (1956). Water tunnel tests of NACA 


4412 and Walchner profile 7 hydrofoils in non- 
cavitating and cavitating flow. Calif. Inst. of 
Tech. Hydrodyn. Lab. 47-5. 

Kermeen, P. W. (1956). Water tunnel tests of NACA 
66-012 hydrofoil in non-cavitating and cavitating 
flows. Calif. Inst. of Tech. Hydrodyn. Lab. 
47-7. 

Kermeen, R. W., and B. R. Parkin (1957). Incipient 
cavitation and wake flow behind sharp-edged disks. 
Calif. Inst. Tech. Engr. Div. 85-4. 

Lighthill, M. G. (1957). Contribution to the theory 
of the Pitot-tube displacement effect. J. Fluid 
Mech. 2, 493. 

Liversy, J. L., and J. T. Turner (1964). The 
generation of symmetrical duct velocity profiles 
of high uniform shear. J. Fluid Mech. 20, 201. 

Macmillan, F. A. (1956). Experiments on pitot tubes 
in shear flow. Aero. Res. Counc., London 18235. 

Numachi, F., and T. Kurokawa (1939). Uber den 
Einfluss des Luftgehaltes auf die Kavitationsent- 
stehung am Tragfltigel. Tech. Rep. Tohoku Imp. 
Univ. 13, 236. Werft-Reederei-Hafen, Bd.XX. 

Numachi, F. (1954). Summary Report on the research 
of cavitation phenomena obtained hitherto by our 
Institute. Rep. Inst. High Speed Mech. Tohoku 
inulin C7 ALES). 

Numachi, F. (1975). Effect of turbulence in free 
stream on cavitation incipience of hydrofoil. 

J. Fluid Engr. Trans. ASME 97, 180. 

Numachi, F., M. Nakamura, and I. Chida (1952). 
Cavitation tests on hydrofoil profiles of simple 
form for blade elements. Memoirs Inst. High 
Speed Mech. Tohoku Univ. 6, 113. 

Owen, P. R., and H. K. Zienkiewiez (1957). The 
production of uniform shear flow in a wind 
tunnel. J. Fluid Mech. 2, 521. 

Rechardson, J. M. (1947). J. Acoust. Soc. Amer. 

ID, G66. 

Rouse, H. and A. H. Abdul-Fetouh (1950). Charac- 
teristics of irrotational flow through axially 
symmetric orifices. J. Applied Mech., Trans. 
ASME 17, 421. 


Rouse, H. (1953). Cavitation in the mixing zone 
of a submerged jet. La Houille Blanche, Janu- 
Feb, 9. 


Schiebe, F. R. (1969). The influence of gas nuclei 
size distribution on transient cavitation near 
inception. Univ. Minnesota Proj. Rep. 107. 

Young, A. D., and J. N. Maas (1936). Behavior of 
Pitot tube in a transverse total-pressure 
gradient. Aero. Res. Counc., London, Rep. and 
Mem. 1770. 


Scale Effects on Propeller 
Cavitation Inception 


G. Kuiper 


Netherlands Ship Model Basin 
Wageningen, The Netherlands 


ABSTRACT 


The boundary layer of four propeller models in 
uniform flow is investigated and related with cavita- 
tion inception. Laminar separation is found to be 
an important phenomenon on model propellers. The 
radius where laminar separation starts is found to 
be a limit for the radial extent of cavitation. 

No inception takes place in regions of laminar flow. 
The effect of nuclei in the flow is investigated 
using electrolysis. Nuclei seem to be important 

for cavitation inception when laminar separation 
occurs, but they do not initiate sheet cavitation, 
when the boundary layer flow is laminar. When the 
boundary layer on the blades is tripped to turbu- 
lence by roughness at the leading edge it is shown 
that this changes the cavitation by restoring cavita- 
tion inception at the vapour pressure. The effect 
of electrolysis on cavitation becomes very small 
when the propeller blades are roughened. Calcu- 
lations of the pressure distribution and the laminar 
boundary layer were made and related with test 
results. 


1. INTRODUCTION 


When cavitation patters, observed on full scale 
ship propellers, are compared with observations on 
model scale, differences are often found [e.g., 
Bindel (1969), Okamoto et al. (1975)]. These 
differences are caused by two main factors: 
correct scaling of the incoming flow of the 
propeller, including propeller-hull interaction, 

and incorrect scaling of cavitation. 

Considerable efforts have been made to improve 
the simulation of the incoming flow by testing the 
cavitating propeller model behind the ship model 
in a large cavitation tunnel or in a depressurized 
towing tank, or by correcting the measured model 
wake to simulate the full scale wake in a cavitation 
tunnel [Sasajima and Tanaka (1966), Hoekstra (1975)]. 


in- 


400 


In this paper the problem of proper scaling of 
cavitation will be investigated. 

Scaling rules for cavitating propellers can be 
formulated using dimensional analysis when the 


relevant parameters are known. 


This results in the 


following well-known dimensionless quantities: 


the advance ratio 


the cavitation index 


the Froude number By 


the Reynolds 


V = 


where 
A 


h = 


Vv = 


ee 1 
= oD (1) 
-p + 

ager ogh 
Fe (2) 
on2D2 
2 
ans D (3) 
g 
2 
number Re. = Das (4) 
N v 


advance velocity of the propeller 
number of propeller revolutions 
propeller diameter 

pressure at some reference level 
vapour pressure 

density of water 

acceleration due to gravity 

vertical distance from reference level 


kinematic viscosity 


When these dimensionless parameters are kept the 


same for model and prototype, 


the cavitation 


behaviour of a propeller is independent of size, 
provided that no additional parameters play a role 
in the cavitation process. 

The choice of the cavitation index as a parameter 
implies the assumption that inception occurs when 
the local pressure is equal to the vapour pressure. 
When the inception pressure deviates from the vapour 
pressure these deviations are called "Scale effects 
on cavitation inception". 

Two scaling problems do arise now. First it is 
impossible to maintain the Froude number and the 
Reynolds number at the same time. The Reynolds 
number is abandoned and is lowered on model scale 
by a factor of 3/2, where A is the scale ratio. 
Even if the Froude number is not maintained it is 
practically impossible to obtain the full scale 
Reynolds number on model scale. The second scaling 
problem is that nuclei play a role in cavitation 
inception. Both problems manifest themselves as 
scale effects. 

Pure water can withstand very high tensions and 
nuclei are necessary to generate inception of 
cavitation. Nuclei are mostly considered to be gas 
pockets in the fluid, possibly trapped in small 
crevices of hydrophobic particles. For a review 
see Holl (1970). In a cavitation tunnel, however, 
the flow will also contain free air bubbles which 
come out of solution at the pump, at sharp corners, 
or at the cavitating propeller in the test section. 
Resorbers are used to bring the free gas back into 
solution, or the tunnel can be prepressurized. 

When no large nuclei are present, however, scale 
effects on cavitation become larger [Hill and 
Wislicenus (1961)]. Inception of cavitation becomes 
related to the pressure at which the largest gas 
bubbles become unstable and start to expand, and 
this pressure is lower than the vapour pressure 

when the nuclei are small [Daily and Johnson (1956)]. 
In a towing tank there are very few nuclei since 
they will rise to the surface or to go into solution. 
Therefore Noordzij (1976) created additional nuclei 
in the NSMB Depressurized Towing Tank by electrolysis 
and showed the "stabilizing" influence of nuclei on 
propeller cavitation behind a ship model. A similar 
effect was reached by Albrecht and Bjorheden (1975) 
who injected additional nuclei into the water of 
their free surface cavitation tunnel after the low 
pressure in the test section had deaerated the 

water so much that nuclei were no longer formed in 
the tunnel. 

It is very difficult to control the nuclei content 
of the incoming flow [Schiebe (1969)]. When the 
nuclei are large enough, the inception pressure 
will be close to the vapour pressure. However, 
when the nuclei are too large they can lead to 
"gaseous cavitation" [Holl (1970)] with inception 
above the vapour pressure, or they can be removed 
from the region of lowest pressures by the pressure 
gradient in the flow, as was theoretically shown by 
Johnson and Hsieh (1966). 

Variation of the Reynolds number leads to viscous 
effects on cavitation inception. Arakeri and Acosta 
(1973) and Casey (1974) showed the effect of the 
boundary layer on cavitation inception. Laminar 
separation was shown to be especially important. 
Arakeri and Acosta (1973) visualized the boundary 
layer by a schlieren technique and they tentatively 
related the cavitation index at inception and the 
pressure coefficient at laminar separation or at 
transition. Increased pressure fluctuations in 
the reattachment region of a laminar separation 


401 


bubble and in the transition region were measured 
by Arakeri (1975) and by Huang and Hannan (1975). 
Van der Meulen (1976) also observed the inception 
process on headforms by means of holography. He 
showed that suppression of laminar separation by 
polymers also could suppress cavitation inception. 
The relation between the inception pressure and 
the pressure at laminar separation or transition 
was not always confirmed. In a recent case study 
[Kuiper (1978)], it was shown that viscous effects 
were responsible for a delay in cavitation inception 
on a propeller model. Additional nuclei had no 
effect in this case, but it was not yet clear if 
nuclei did interact with the boundary layer to 
create cavitation inception. 

In this study, scale effects on cavitation on 
three propellers with different characteristics 
were investigated. When a propeller operates in 
a wake, scaling problems of the incoming flow and 
of cavitation cannot be separated. Therefore the 
propellers were tested in uniform axial flow. The 
tests were carried out mainly in the Depressurized 
Towing Tank. A description of this facility is 
given by Kuiper (1974). The advantages of this 
tank for the research on scale effects on cavitation 
inception are the, supposedly, very low and constant 
turbulence level and nuclei content, the uniform 
inflow of the propeller, and the absence of wall 
effects. Both advance speed and propeller revolu- 
tions can be controlled very accurately. The range 
of Reynolds numbers which can be tested is lower 
than in a cavitation tunnel (maximum carriage speed 
is 4 m/sec.) but is not smaller. 

The aim of the present study is to gain insight 
into the occurrence of scale effects on cavitating 
propellers and to develop means to improve the 
correlation with full scale observations. Paint 
tests were carried out to visualize the boundary 
layer flow on the propeller blades. Methods to 
calculate the pressure distribution on the blades 
are discussed and the calculated pressure distri- 
butions are used for the interpretation of the 
results of the paint tests and the cavitation 
observations. The nuclei content is varied by 
using electrolysis, and roughness at the leading 
edge of the propeller blades is applied to make the 
boundary layer on the blades turbulent, thus simu- 
lating a higher Reynolds number. The relation 
between the boundary layer on the blades and 
cavitation inception is shown and the effect of 
leading edge roughness and electrolysis is investi- 
gated. 


2. TEST PROGRAM 
Propellers and Test Conditions 


Four propellers were investigated in uniform flow. 
Propeller A is the propeller which was investigated 
behind a model in a case study by Kuiper (1978). 
This propeller showed viscous scale effects on 
Cavitation inception but was insensitive for 
electrolysis (Figure 1). Behind the model, this 
propeller operated in a nozzle. In this study it 
was tested without a nozzle. 

Propeller B is the propeller which was tested by 
Noordzij (1976) behind a model. This propeller 
was very strongly influenced by electrolysis. 
Without electrolysis the sheet cavitation varied 
per revolution, (Figure 2). With electrolysis the 


propeller an electrolysis grid was mounted, as 
shown in Figure 5. The wires had a diameter of 
0.2 mm and a current of 0.2A was used to generate 
nuclei. The propeller shaft was at 0.4 meter below 
the water level and the lowest wire at 0.5 meter. 
Therefore the effect of electrolysis could only 
be observed in the upper half of the propeller disk. 
-The propeller boundary layer. Two ways of 
affecting the boundary layer were used. First, 
sandroughness at the leading edge was used to trip 
the boundary layer to turbulence. Second, the 


9 FULL SCALE OBSERVATIONS 


FIGURE 1. Viscous effects on cavitation inception on 
propeller A behing the model. 


cavitation pattern was present and identical at 
every revolution. This "stabilizing" effect of 
nuclei is important because it affects the induced 
pressure fluctuations on the hull. 

Propeller C had a very distinct collapse of the 
cavity when the blades left the wake peak, as can WITHOUT ELECTROLYSIS 
be seen in Figure 3. This irregular collapse of 
the cavity was thought to be caused by viscous 
effects and it can also strongly influence the 
pressure fluctuations on the hull. 

Propeller D was not tested in cavitating con- 
ditions. It was used only for boundary layer 
visualization. This propeller is an example of a 
smaller propeller model used behind models with a 
maximum length of 7 meters. This propeller was 
made of a copper-nickel-aluminium alloy (CUNIAL). 
Propellers A, B and C were of aluminium. 

The most important geometrical characteristics 
of the four propellers are given in Figure 4. The 
complete description, necessary for the calculations, 
is given in the Appendix. Most tests were done in 
the NSMB Depressurized Towing Tank. To obtain 
uniform inflow the propellers were mounted on a 
right-angle drive unit, which was kept afloat by a 
catamaran-type vessel, as shown in Figure 5. Only 
a few comparative tests were done in a cavitation 
tunnel. 

The following parameters were varied: 

-The propeller loading. Two advance ratio's were 
used, namely 70% and 40% of the pitch ratio at 
r/R=0.7. (Slip ratio's of 30% and 60% respectively). 
The slip ratio of 30% corresponds to a loading which 
is about normal behind the ship, the slip ratio 
of 60% corresponds to an overloaded condition, as 
occurs when the blades are in a wake peak. Propeller 
A was also investigated at an intermediate loading WITH ELECTROLYSIS 


: : of 
with a slip of 40% FIGURE 2. Effect of electrolysis on propeller B 
-The nuclei content. At 1 meter in front of the behind the model. 


FIGURE 3. 
C behind the model. 


Irregular collapse of cavitation on propeller 


propeller Reynolds number was varied with a factor 
of about three. 

-The cavitation index. Three values of the 
cavitation index were used: Oyq=1.5, 2.0, and 2.5. 
The reference level of the cavitation index was 
always taken at the propeller tip in the top position. 
In this paper most cavitation observations will be 
shown at Oyp=l.5. At higher revolutions a lower 
cavitation index was possible: Oymp=0.5 in the 
towing tank and oyp=1.0 in the cavitation tunnel. 


Paint Observations 


To visualize the character of the boundary layer 
at the propeller blades a surface oil flow technique 
was used [Maltby, ed. (1962)]. This technique was 
adapted for use in water on propellers by Meyne 
(1972) and Sasajima (1975). It is particularly 
useful on rotating bodies because the difference 
in friction coefficient between laminar and turbu- 
lent boundary layer flow, in combination with the 
centrifugal force acting on the paint, creates a 
clear difference in the direction of the paint- 
streaks in laminar and turbulent regions. 

The paint, used in our paint tests, consisted 
of lead-oxide, diluted with linseed oil and coloured 
with red "Dayglo" pigment. This mixture produced 
a finely detailed pattern of streaks on the metal 
surface of the propeller. When the propeller 
blades were painted yellow with a thin layer of 
zinc-chromate primer, as is done with the cavita- 
tion observations to improve contrast and to avoid 
reflections, no streaks were formed. Consequently 
the flow visualization tests were done with the 
propellers not painted. 

The viscosity of the paint was controlled by 
the amount of linseed oil and was chosen such that 
the formation of the pattern took about one full 
run in the towing tank. At least 500 revolutions 
were always available to form the patter. To 


403 


1,047 


1.0 
—07 
o5 
bo -0.2 
5 BLADES 
D = 0.3268m 
Ag/Ao = 0.820 
Co7/D = 0.368 


PROPELLER A tic ©) = 0.042 


TR 
1.0 
0.72 
053 
0.24 


TR 


4 BLADES 
D = 0300m 
Ac /Ag = 0.630 


Co7/D = 0.430 
t/c(07) = 0.022 


Ac lAg = 0.824 
Co7/D = 0.307 


PROPELLER D t/c (07) = 0.050 


FIGURE 4. Geometry of propellers. 


reach the desired condition took about 100 
revolutions, most of them very close to the final 
condition. Paint tests were also done in the 
cavitation tunnel. The pictures obtained there 
were more profuse, especially at high tunnel veloc- 
ities, because of the relatively long time it took 
to reach a stable condition. For runs longer than 


404 


ELECTROLYSIS GRID 


-CAMERA 


a 


DETAIL ELECTROLYSIS GRID 


~. 


0.15 | 


0.35m 


050m 


FIGURE 5. Test equipment for open-water tests. 


a few minutes the viscosity of the lead-oxide is 
too low and the blades are cleaned by the flow. 

The paint is put on the propeller blades at the 
leading edge to about 10% of the chord. The layer 
must be rather thick to provide enough paint to 
cover the whole blade. Some pictures were taken 
with UV light using the fluoriscent properties of 
the pigment. The bulk of the pictures of the paint 
tests was taken in colour photography with natural 
light. This gave good colour prints, but unfortu- 
nately the contrast in monochrome paper turned out 
to be rather poor. 


Roughness at the Leading Edge 


To trip the boundary layer to turbulence the leading 
edge of the propeller blades was covered with 
carborundum. The leading edge of the propeller 
blade is wetted with watery thin varnish to about 
0.5 mm from the leading edge. This is done by 
touching the leading edge with a pad wetted with 
varnish. The softness of the pad determines the 
length of the wetted area from the leading edge. 
Then carborundum is put on the wetted area by 
spreading the grains on a felt cloth and by wiping 
the wetted leading edge with that cloth. Two grains 
sizes were used: 30 wm (31-37) and 60 um (53-62). 
Microscopic inspection afterwards is necessary. An 
example is given in Figure 6. 


3. CALCULATION OF PRESSURE DISTRIBUTION 


The analysis of boundary layer phenomena and of 
cavitation on propeller blades becomes very specu- 
lative when the pressure distribution is not known. 
No firm experimental verification of calculations 
of the pressure distribution is available yet, only 
the total thrust and torque give some evidence of 
the value of calculations. The calculations are 
always potential flow calculations and the effect 
of viscosity on the propeller sections cannot yet 


be derived with suitable accuracy. Since the 
propeller thrust is least sensitive to viscous 
effects this quantity gives the most reliable 
verification of calculations. When the propeller 
geometry and the nominal inflow are known two 
approaches are available to obtain the distribution 
of propeller loading, viz. the lifting line theory 
and the lifting surface theory. Hereafter, both 
approaches will be considered with models going 
back to the work of Lerbs (1952) for the former 
and Sparenberg (1960) for the latter theory. 


Lifting Line Calculations 


The lifting line theory concentrates the loading of 
a propeller section at one point. Using the induc- 
tion factor method [Wrench (1957)], a relation 
between the hydrodynamic pitch angle, 8;, and the 
circulation, I, at each section is found. 


B. (i) = ae] ie (a) I (5) 


When a given propeller is analysed 8; and I are 
unknown. To find them a second relation is necessary, 
which is derived from two-dimensional profile 
characteristics. The lift coefficient 


de" 
Cy -( =) (ata,) (6) 


where a, is the zero lift angle of the propeller 


section. Since the angle of attack a is taken from 
oS (BS (2) (7) 
Pp i 
where 8, is the known geometrical pitch angle, a 


second relation is formulated between Cy, (or IT) 
and §;, in which dC;/da, is assumed to be known. 
When the two-dimensional value for dc; /da, based 
on the geometry of the propeller section, is used 
the results are rather drastically wrong. This is 
caused mainly by the finite length of the propeller 
section, which creates a distribution of induced 
velocities affecting camber and angle of attack. 


1mm 
60 Lm 30 Lm 
CARBORUNDUM CARBORUNDUM 


FIGURE 6. Microscopic picture of leading edge 
roughness. 


MEASURED 6 
Rey = 2.310 


CALCULATED L. SURF 
—— — — — (AUeVIYATH) (b, (HIN: 


No 


fo} 01 02 03 0.4 loks) 06 07 08 o9 10 
PROPELLER A 


MEASURED 
Rey = 2.9x10 


——-—-— CALCULATED L. SURF. 
———— CALCULATED L. LINE 


f°) 01 0.2 0.3 04 Os 06 07 08 09 1.0 
PROPELLER C 


FIGURE 7. 


So Eq. 6 has to be corrected to obtain a three- 
dimensional lift curve. At one point of the lift 
curve, at the ideal angle of attack, results of 
systematic lifting surface calculations are avail- 
able [Morgan et al. (1968)] and they can be expressed 
as correction factors on camber, Ko, and the angle 

of attack, Ky. Van Oossanen (1974) used these 
correction factors to define the three-dimensional 
lift curve over the whole range of angles of attack 
instead of at the ideal angle of attack only. He 


wrote 
fol 
(<2) Ry Saue ea 
da Zl Gls Ko (8) 


3a ‘ (9) 


MEASURED 
Rey = 1.6210) 


————— ——— CALCULATED EL SURE: 


Nu) Ka 


fo) 02 04 06 08 10 1.2 14 1 
PROPELLER B 


MEASURED 6 
Rey = 15x10 


CALCULATED L. SURF. 
—— —— CALCULATED LLINE 


fe) 0.1 02 03 04 o5 06 O07 08 09 1.0 
PROPELLER D 


Propeller open-water characteristics. 


where a; is the ideal angle of attack of the 
propeller section. Substitution of these three- 
dimensional values in Eq. 6 makes it possible to 
solve the set of Eqs. 5-7, resulting in a radial 
distribution of 8, Qos and Cy,. 

In Figure 7 the calculated open-water character- 
istics using this approach are compared with experi- 
ments. The agreement between measurements and 
calculations is acceptable. Propeller B could not 
be calculated since the regression formula's for 
K, and K, in the program were restricted to a 
maximum pitch ratio of 1.4. 

Viscosity is taken into account by assuming a 
viscous lift slope 


ac 3 
L t 
(<2)yan = 0.947-0.76 (£) 


where t= max. thickness of propeller section 
c= chord length of propeller section 


The drag is calculated using the characteristics 
of the equivalent profiles of the NSMB B-series 
propellers. 


Lifting Surface Calculations 


The lifting surface theory calculates the induced 
velocities over the propeller blades, in chordwise 
and radial direction, thus including the effects 

of finite aspect ratio of the blades. The draw- 
back is that the theory is linearized, which 
restricts the validity to lightly loaded propellers. 

Van Gent (1977) has shown in his thesis how 
heavily loaded propellers can be treated with a 
linearized theory since the vorticity in the wake 
induces an additional axial velocity component in 
the propeller plane, keeping the angles of attack 
of the propeller sections small. 

The boundary conditions on the propeller blades 
are fulfilled at a number of chordwise and spanwise 
points. In our calculations four chordwise and 
ten radial points per blade were chosen. The pitch 
of the vortex sheet in the wake was taken rather 
arbitrarily as the pitch at 0.7D. 

A very approximate description of the viscous 
effects is used. The drag force of the propeller 
sections is split into two parts: a drag force 
as a result of losses in the suction peak at the 
leading edge and a drag force due to friction. The 
latter is calculated using a friction coefficient 
of 0.0080, irrespective of the Reynolds number. 

The first drag force is taken as half the theoretical 
suction force. The same correction is also applied 
to the sectional lift, which is obtained from chord- 
wise integration of the lift distribution. In the 
calculation of the induced velocities the geometrical 
pitch angle is reduced by 3/4 degree to simulate 
viscous effects on the zero lift angle. 

The open-water diagrams as calculated with the 
lifting surface theory as described by Van Gent 
(1977) are shown in Figure 7 together with experi- 
mental results and lifting line calculations. The 
general agreement with measurements is as good as 
the lifting line calculations. This makes clear 
that the linearized lifting surface theory can 
indeed produce reliable open-water characteristics 
up to high propeller loadings. At very low advance 
ratio's the calculations deviate from the measure- 
ments but this might well be caused by an erroneous 
estimate of the viscous effects. 


Calculation of the Pressure Distribution 


Lifting line as well as lifting surface calculations 
give the radial distribution of the lift coefficient, 
of the angle of attack, and of the induced camber 
(or camber distribution) which can be translated 
into a zero lift angle. In Figure 8 these results 
are compared for propeller A at 40% slip. The 
lifting line calculation gives a higher loading at 
the tip and a lower loading at inner radii, compared 
with the lifting surface calculation. This is 
characteristic for all four propellers in all 
conditions. The total thrust does not differ very 
much. Large differences, however, are found for 

the angle of attack and for the zero lift angle. 


LIFT. SURFACE 
ET NES 


AroT= AincineNce * %o 


Q, =ZERO LIFT ANGLE 


FIGURE 8. Radial distribution of lift coefficient and 
angle of attack on propeller A at 40% slip. 


Since these values will be used in the calculation 
of the pressure distribution this discrepancy needs 
further attention. 

The source of the discrepancy is the choice of 
Eqs. 8 and 9, used in the lifting line calculation. 
The reduction of the slope of the lift curve with 
the lifting surface correction factor for the camber, 
Ko (Eq. 8), is an empirical one, first suggested 
by Lerbs (1951) when he analyzed the lift slopes 
of his "equivalent profiles". The physical meaning 
of this correction is not clear, but it still can 
lead to correct results for thrust and torque, 
since the lift slope for the equivalent profiles 
was derived using a lifting line theory and experi- 
mental values of thrust and torque. Therefore, this 
correction for the lift slope, used in combination 
with the same lifting line theory, should give 
results for thrust and torque not too far from the 
experimental results. The definition of the three 
dimensional zero lift angle (Eq. 9) is another 
empirical relation, bringing the calculated open 
water characteristics in line with experiments. 
However, this does not necessarily mean that the 
three dimensional angle of incidence and zero lift 
angle have a physical meaning and can be used for 
the calculation of the pressure distribution. 
Therefore, the results of the lifting surface cal- 
culations are used in the following to calculate 
the pressure distribution. 

To calculate the pressure distribution on the 
blades, the effect of propeller thickness has to 
be calculated and the leading edge singularity of 
the lift distribution has to be dealt with. Tsakonas 
et al. (1976) calculated the pressure distribution 
on the propeller blades using a singularity distri- 
bution for the thickness, in combination with a 
linearized lifting surface theory. These calcula- 
tions, however, remain linearized, producing an 
infinite velocity at the leading edge, which was 
removed by the Lighthill correction for thin air- 
foils [Lighthill (1951)]. In our study, three- 
dimensional effects on the pressure distribution 
are neglected. Interaction effects between thickness 


PROPELLER A 


PROPELLER C 


Tp =0.95 


and loading, which occur due to the non-planar 
surface of the propeller blades are taken into 


account by a correction factor [Morgan et al. (1968)]. 


This makes it possible to apply conformal mapping 

to calculate the pressure distribution. An approx- 

imation of the original theory of Theodorsen (1932), 
known as Goldstein's third approximation [Goldstein 

(1948)] was used. The determination of the "effec- 

tive geometry" was done using a camber line, derived 


from the calculated induced velocities of the lifting 


surface calculation. This can be done because the 
problem is linearized. The calculated induced 
camberline and the geometrical thickness distribution 
were combined in the NACA-manner to obtain the 
geometry of the effective profile. The pressure 
distribution on the propeller section was then cal- 
culated using the induced angle of attack from the 


407 


PROPELLER B 


FIGURE 9. Calculated pressure distribution on the 
suction side at 30% slip. 


lifting surface calculation. The lift coefficient, 
which is found from the lifting surface calculation, 
is maintained using the method of Pinkerton (1934). 
This is necessary because the potential flow lift 
coefficient of the effective profile is slightly 
lower at inner radii, where the sections become 
thicker. The differences are of the order of 0.02. 
In Figures 9 and 10 the calculated pressure 
distributions at the suction side are given for 
propellers A, B, and C. 


4. RESULTS OF PAINT TESTS 


In Figure 11 the paint patterns are shown for pro- 
pellers A, B, and C at 30% slip and at Reynolds 
numbers typical for testing behind 12 meter models. 


408 


PROPELLER A 


PROPELLER C 


These pictures were taken with UV-illumination. 

At the leading edge the paint is removed, due to 
high local velocities. The streaks are formed 
gradually, either in a nearly tangential direction 
(the turbulent region) or pointed outwards (the 
laminar region). The transition from laminar to 
turbulent boundary layer flow is shown by a change 
in direction of the streaks. 

Laminar boundary layer flow occurs in all cases 
near the leading edge. Transition in chordwise 
direction to turbulent boundary layer flow occurs 
gradually, but a transition region can be distin- 
guished and at the trailing edge the boundary layer 
is turbulent. When the paint streaks are nearly 
in the radial direction the flow is separated. At 
inner radii the boundary layer if often close to 
separation. Laminar separation was clearly present 


PROPELLER B 


FIGURE 10. Calculated pressure distribution on the 
suction side at 60% slip. 


on propeller D, as is shown in Figure 12. At 60% 
slip the radius where laminar separation is replaced 
by natural transition can be seen by the sharp 
corner in the paint streaks. 

At the suction side near the tip a turbulent 
region exists immediately from the leading edge 
(Figure 11). An increase in propeller loading 
showed a radial increase of the turbulent region at 
outer radii, as illustrated in Figure 12. The 
change in radial direction of the laminar region 
near the leading edge to the turbulent region at 
outer radii on the suction side is abrupt and 
nearly discontinuous, as sketched in detail in 
Figure 13. The laminar region is cut off and the 
region of natural transition at inner radii does 
not reach the leading edge. We will designate the 
radius where this discontinuity occurs, the critical 


SUCTION SIDE 


PROPELLER C , Rey= 0.66 x 10° 


radius of the propeller. Such a critical radius 
can also be observed from the paint pattern of 
Sasajima (1975) and of Meyne (1972). This critical 
radius turned out to be very important for cavita- 
tion inception and could be discerned in all cases. 
No photographs are shown because of the bad contrast 
of the monochrome prints. (Figure 16). 

On propeller B at 60% slip a separation bubble 
at the leading edge was observed, connected with a 


409 


PRESSURE SIDE 


FIGURE 1l. 
30% slip. 


Paint patterns at 


stagnation region near the tip on the suction side, 
which indicated the position of the tip vortex. In 
the direction of the hub the laminar separation 
bubble extended exactly until the critical radius. 
This lead us to the hypothesis that laminar sepa- 
ration near the leading edge was the cause of the 
discontinuous character of the paint streaks at the 
critical radius. To verify the hypothesis of laminar 
separation at the critical radius, boundary layer 


410 


30 °%e SLIP 


CRITICAL 
RADIUS 


LAM. SEPARATION 
NEAR MIDCHORD 


60 °%. SLIP 


Rey = 0.47 x10° 


FIGURE 12. Variation of the critical radius with 
propeller loading on propeller D (suction side). 


calculations were made, using the pressure distri- 
butions as calculated in Section 3. The laminar 
boundary layer was calculated with Thwaites' method 
[Thwaites (1949)]. Laminar separation was predicted 
using Curle and Skan's (1957) criteron. This cal- 
culation method does not take into account the 
delaying effect of rotation on laminar separation, 
but since laminar separation occurs very close to 
the leading edge the effect of rotation on the 
development of the boundary layer will still be 
small. The correlation between the calculated and 
the observed critical radius is given in Figure 14, 
and this correlation is quite good. The critical 
radius at all conditions and the variation of the 
critical radius between the propeller blades can 


FIGURE 13. Discontinuity of paint streaks at the 
critical radius. 


1.0 
" B 
Cc 
fa) 10) 
w 
a 
=>) 
3) 
< 
wi 
= 
O5 
fe) 


CALCULATED 


FIGURE 14. Correlation of calculated radius of 
laminar separation and measured critical radius. 


also be found from Figure 14. As can be seen, the 
variation of the critical radius per blade in one 
condition can be considerable, showing the sensi- 
tivity of laminar separation to the manufacturing 
accuracy. The critical radius per blade, however, 
reproduced remarkably. 

The position of laminar separation is independent 
of the Reynolds number. So another check on the 
hypothesis of laminar separation at the critical 
radius is the independence of the critical radius 
from the Reynolds number. Propellers A and C were 
therefore tested with about twice the original 
number of revolutions. Propeller A was also inves- 
tigated in a cavitation tunnel: the highest 
Reynolds number in the towing tank was repeated 
and another condition with about three times the 
original Reynolds number was tested. The paint 
tests in the cavitation tunnel were less accurate 
since turbulent spots occurred, which caused a 
wedge shaped tangential streak through the laminar 
pattern. This was strongest at the higher Reynolds 
numbers. 

Figure 15 gives the critical radius as a function 
of Reynolds number for the blades available for 
comparison. There is a slight trend for the critical 
radius to decrease with increasing Reynolds number, 
but this is only very slight. The critical radius 
is strongly dependent on the propeller loading and 
a slight increase of the propeller loading with 
increasing Reynolds number might cause the decrease 
of the critical radius. For comparison the obser- 
vations of Sasajima are also drawn in Figure 15. 

He observes a larger shift of the critical radius 
with Reynolds number, but his results from the 

tank show no variation with Reynolds number. The 
variations found in the cavitation tunnel might 
well be caused by variations in propeller loading 
or by wall effects. The conclusion seems justified 
that the critical radius is independent of the 
Reynolds number, at least until natural transition 
occurs close to the minimum pressure point. In 
that case a critical radius no longer exists. 


@ PROPELLER A’ SLIP =0.3 TUNNEL 
© " ” ” ” TANK 
x ” " n =0.6 ' 
q " B n =03 " 
1.0 © SASAJIMA (1975) TANK 
% ° 
0.9 


{e) os 1.0 15 2.0 4G 25 FIGURE 15. Effect of Reynolds number on the 
Rey x 10 critical radius. 


It is important to note that in Figure 12 at 60% 
slip the radius where laminar separation occurs 
near midchord is not the critical radius, although 
in this case the difference between both is small. 
With increasing Reynolds number, however, the region 
of laminar separation near midchord will decrease, 
while the critical radius will remain unchanged. 
The distance between the sharp corner in the paint 
streaks of Figure 12b and the critical radius will 
therefore increase with increasing Reynolds number. 

An increase of Reynolds number causes a shift 
in the chordwise position of the transition region 
at radii inside the critical radius, as is illus- 
trated in Figure 16. This was also observed on 
the pressure side. In Figure 17 the chordwise 
position of the transition region is given at 
r/R=0.7 as a function of the sectional Reynolds 
number, which is related to the entrance velocity 
and the chordlength of the propeller section at 
that radius. The transition region is averaged in 
Figure 17. This makes clear that a complete turbu- 
lent boundary layer at a radius of 0.7R requires 
sectional Reynolds numbers of about 5x106. At the 
suction side, turbulent flow at this radius also 
occurs when the loading is increased, i.e. the 
critical radius is smaller than 0.7. 

Empirical criteria for transition of the boundary 
layer to turbulence have been given as a relation 
between the Reynolds numbers based on the length 
from the stagnation point, Re,, and based on the 
momentum thickness, Reg. [Michel (1951), Smith 
(1956) ]. Van Oossanen used the Smith line 


Rey = 0.73 x10" 


Rey = 1.56 x 10° 


= 1 1740Re, aac 


SO ears (10) 


as a criterion. When the relation between Reg and 
Re, Over the chord was calculated, both on the 
suction side and on the pressure side, this relation 
was so closely parallel to the criterion of Eq. 10 
that no reliable intersection was possible. When 
there is a strong negative pressure peak at the 
leading edge the relation between Reg and Re, is 


such that Eq. 10 always predicts transition very $$ GRIMIENG RASIUS 

close to the leading edge. When the pressure £22 777 7~7~— TRANSITION 

distribution was nearly shockfree, the prediction 

was erroneous. FIGURE 16. Effect of Reynolds number on the transi- 


To calculate the transition region, calculation tion region. Propeller A at 30% slip. 


On propeller A and on propeller B at 30% slip 
the radial extent of the cavitation is clearly 
restricted by the observed critical radius. Some— 
PRESSURE SIDE times there is a small difference between the 

FF SUP =0.6 critical radius and the inception radius, which is 
probably caused by a change in the pressure distri- 
bution by the cavitation. 

The calculated ideal inception radii at 60% slip 
should be considered with caustion. They are close 
to the hub and the influence of the hub is not 
taken into account in the calculations. For example 
on propeller B at 60% slip the inception radius is 
larger than calculated. In that case the critical 
radius is smaller than the inception radius and 
does not cause any viscous effects on cavitation. 
The distance between the ideal inception radius 
and the critical radius on propeller C is small, 
so the scale effects due to the critical radius 
will be small too. 

We can conclude that no cavitation occurred in 
regions of laminar flow near the leading edge. The 
radial extent of cavitation can be seriously 
5x10° 10° 5 x10° restricted by the critical radius. Since the crit- 

Re (0.7) ical radius is connected with laminar separation 
this means that variation of the Reynolds number 
does not remove this restriction until very high 
Reynolds numbers. From Figure 17 the sectional 
Reynolds number at r/R=0.7 has to exceed 5x106, 
whereas a value of 3x105 is mostly considered 
enough to avoid Reynolds effects on thrust and torque. 


O PROPELLER C- SUCTION SIDE SLIP=0.3 
a cr PRESSURE SIDE co BOOKS 
O PROPELLER A SUCTION SLIP =0.3 
e@ 
a 


FIGURE 17. Chordwise position of natural transition 
inside the critical radius. 


of the stability of the laminar boundary layer 

might give better results [Smith and Camberoni (1956) ]. 
Since transition occurs far from the leading edge, 

the effect of rotation can be important. When the 
calculation scheme of Arakeri (1973) is used it is 
possible to take the effect of rotation into account 
using Meyne's (1972) results. This was beyond the 
scope of this paper. j 


Variation of Reynolds Number 


Propellers A and C were tested at a higher Reynolds 
number in the towing tank, while propeller A was 


Rey = 0.73x10" Rey =0.51x10° Rey =0.66x10° 


5. CAVITATION OBSERVATIONS 


The cavitation on propellers A, B, and C is sketched 
in Figure 18 for both slip ratio's. The cavitation 
index at the blade tip in top position, Oymp (Eq. 2) 
was always 1.5. The Reynolds numbers Rey, were 
about 5x105. At 30% slip the condition is not far 
from inception and a cavitating tip vortex is 
present in nearly all cases. However, in some cases 
at low Reynolds numbers, propellers A and C were ob- 
served without any cavitation. This was not due to 
intermittent cavitation during one test, but oc- 
curred when tests were repeated with time-intervals 
of some weeks. During one test the observations 
were quite consistent, indicating that the varia- 
tions are caused by factors which are still not 
under enough control, e.g., air content, nuclei 
content, turbulence. 


30% SLIP 


Correlation with Paint Test 


Of interest is the correlation of the radial extent 
of the cavity with the observed critical radius, 

found from the paint test. In Figure 18 the 

observed position of the critical radius is indicated, 
as well as the calculated ideal inception radius, PROP. A PROP. B PROP. C 
which is the radius where the minimum pressure on 

the blades equals the vapor pressure. Also indicated 
is the cavitation, observed when the leading edge 

was roughened, as will be discussed in the next 


section. FIGURE 18. Cavitation observations at Chom = 1.5. 


——— WITH ROUGHNESS 
= OBSERVED CRITICAL RADIUS 
— CALCULATED IDEAL INCEPTION RADIUS 


also tested in the cavitation tunnel at two Reynolds 
numbers. No differences in cavitation pattern due 
to variation of the Reynolds number were observed 
in the towing tank. Notably the radial extent of 
the cavity was unchanged, which confirmed that the 
critical radius restricted cavitation inception 
independent of the Reynolds number. The results 

of propeller A at 30% slip are shown in Figure 19. 
In this figure the observations of the tests in 

the cavitation tunnel are also shown. These show 
some differences requiring further attention. The 
cavity in the cavitation tunnel at Rey=1.56x106 is 
somewhat larger than in the towing tank, but the 
difference is not significant and is probably 
caused by a slight difference in propeller loading. 
(The tunnel condition was taken at a’K,-value 
derived from the open water measurements. The flow 
velocity was not measured). Remarkable are the 
spots of cavitation at Rey=l.56x106 which increased 
in number when time increased! 

At Rey=2-72*106 there is a sheet outside r/R=0.9, 
the same as at Rey=1.56x106. The spots however, 
have increased in number and they coalesce at some 
distance from the leading edge, forming a cavity 
until about r/R=0.8 with isolated spots until r/R=0.7, 
which is the ideal inception radius. The increase 


of the number of spots with time was not observed 
in this situation, but the time to reach a stable 
condition was much longer than at lower Reynolds 
numbers. 


TANK 
Rey, =1.36x10° 


a. TANK b. 
Rey =0.73x10° 


TUNNEL 
Rey =2.72x10° 


c. TUNNEL d. 
Rey = 1.56x10° 


FIGURE 19. Effect of Reynolds number on propeller A 


at 30% slip with On = 1.5. 


413 


The occurrence of cavitating spots in the laminar 
region agrees with the observation of turbulent 
streaks in the paint tests in the cavitation tunnel 
at higher Reynolds numbers. Therefore, it is 
conjectured that, in the tunnel, tiny particles 
were deposited on the leading edge of the propeller, 
thus creating turbulent streaks. The number of 
these streaks may increase with time, and these 
turbulent streaks cause spots of cavitation. 

Another possible effect is that the propeller 
is not hydrodynamically smooth. With increasing 
Reynolds number the boundary layer becomes thinner 
and more sensitive to local roughness. In this 
case the streaks would always be in the same position. 
Not enough observations were made to verify this, 
but the strongly reduced occurrence of turbulent 
spots in the towing tank points to the flow as the 
origin of the disturbances. The occurrence of 
these streaks was also apparent in the tank when 
the pressure was drastically lowered, as is shown 
in Figure 20. It is of course very important to 
recognize these cavitating spots since they indicate 
a region of laminar boundary layer flow anda 
possible restriction of the radial extent and the 
volume of the cavity. 

The effect of Reynolds number on cavitation in 
the region from the critical radius to the tip is 
small. In nearly all cases cavitation took place 
in this region at low Reynolds numbers. In some 
cases no cavitation was present in this region at 
a low Reynolds number, as shown in Figure 21. A 
paint test is included to show the critical radius. 
At a higher Reynolds number, cavitation was present 
until the critical radius. The ideal inception 
radius in this case is at r/R=0.7. A similar effect 
was sometimes seen at propeller C and can be 
explained by the fact that the reattachment region, 
where inception is assumed to occur, shifts to 
lower pressure regions with increasing Reynolds 
number. Calculations of such an effect are given 
by Huang and Peterson (1977). It is not certain, 
however, that the Reynolds number is the only 
variable since application of electrolysis also 
caused inception at low Reynolds numbers. Apparently 
the nuclei distribution becomes more critical with 
lower Reynolds numbers. 


Observations with Oyp = 0.5 


Laminar boundary layer flow was seen to prevent 
sheet cavitation at the leading edge. To see if 
there is some threshold for inception the cavitation 
index was drastically lowered to opy=0-5. This 
was only possible at high Reynolds numbers. In 
Figure 22 propeller A is shown at 30% slip, a 
condition comparable with Figure 19b, but at a low 
cavitation index. It is clear that even in this 
extreme condition no cavitation occurred in the 
laminar flow region. 
A comparison of the local cavitation index with 
the pressure coefficients as given in Figure 9 
shows that, e.g., at r/R=0.8, the minimum pressure 
coefficient is 0.54 while the cavitation index at 
that radius is 0.08 to 0.012, depending on the 
position of the blade. The cavitation index at this 
radius is lower than the pressure coefficient over 
most of the propeller section. When turbulent spots 
appeared inside the critical radius these spots 
were supercavitating, as is also shown in Figure 20. 
Bubble cavitation can be expected near midchord 


Ont =1.5 


Oyt =0-5 


Rey = 1.29 x 10° 


FIGURE 20. Turbulent streaks inside the critical 
radius at higher Reynolds numbers. Propeller C at 
30% slip. 


at inner radii, where the minimum pressure exists 
near midchord. At propeller A at r/R=0.6 the cavita- 
tion index is between 0.13 and 0.20, at Oyp-O-5, 
while the minimum pressure coefficient is 0.26. 

As can be seen in Figure 22 no bubble cavitation 
occurred. The cavitating spot at midchord is a 

dent in the propeller surface and illustrates the 

low local pressure. Similar observations were made 
with propeller C at Oyp=0.5. No threshold for sheet 


cavitation could be established and no bubble 
cavitation occurred near midchord at inner radii. 
Both phenomena are suspected to be caused by a lack 
of nuclei. So electrolysis was applied, as will be 
discussed in the next section. 


6. VARIATION OF NUCLEI CONTENT BY ELECTROLYSIS 


Some measurements in the NSMB Depressurized Towing 
Tank with the scattered light method indicated 
that the nuclei content of this tank was nearly 
independent of the pressure. The density of small 
nuclei (17 um) was 1.2x107 m7~3 and that of the 
largest available nuclei (45 um) was 1.2x10° Wo 
[This corresponds with nuclei number densities, as 
defined by Gates (1977) of 9x10!! ana 2.4x101° 
respectively]. A description of the measuring 
technique which was used is given by Keller (1974). 
A comparison with similar measurements in the NSMB 
large cavitation tunnel [Arndt and Keller (1976) J 
shows that the nuclei content is lower than that 
in the cavitation tunnel at the lowest air content 
by a factor of about 5. The nuclei content in the 
cavitation tunnel was very much dependent on the 
total air content of the water, showing variations 
of a factor of 10 between high (12.5 ppm) and low 
(6.3 ppm) air content. This dependency was absent 


PAINT OBSERVATION 
Rey= 0.73x10° 


CAVITATION OBSERVATION 
Rey=0.73x 10° 


Ont= 1.71 


CAVITATION OBSERVATION 
Rey = 1.56 x10° 


Gyt=2-09 


Effect of Reynolds number on_ cavitation 


FIGURE 21. 
inception outside the critical radius. Propeller A 
at 40% slip. 


Rey = 1.56 x10° 


Guz = 0.5 


FIGURE 22. Cavitation at very low cavitation index. 
Propeller A at 30% slip. 


in the Depressurized Towing Tank. So when cavita- 
tion observations in the tank are compared with 
observations in the tunnel, we can assume that the 
nuclei content in the tunnel is always larger than 
that in the tank by at least a factor of 10. Perhaps 
most important, however, is that in the tank nuclei 
greater than 60 wm are absent. 

The nuclei content in the tank has been varied 
using electrolysis, as described in Section 2. The 
nuclei size distribution from the wires of 0.2 mm 
diameter has not been measured. Exploratory 
photographic observations showed that the bubbles 
coming from the wires are in the range of 50 to 


100 um under comparable conditions. 
The influence of the wires on the propeller 


boundary layer was checked by a paint test on 
propeller A at 30% slip. The paint patterns with 
and without wires were identical. So we assume that 
the turbulence, coming from the wires, did not 
affect the propeller boundary layer. This assumption 
should be treated with some caution, because Gates 
(1977) showed widely different effects of flow 
turbulence on two headforms, both with laminar 
separation. 

Gates also showed that large amounts of nuclei 
can influence the boundary layer. Notably the 
laminar separation bubble on his hemispherical 
headform was removed. To see if this was also the 
case in our tests a paint test was carried out with 
propeller A at 30% slip. The cavitation index was 
just above inception, so cavitation was avoided. 

To correct for the higher pressure in this condition 
the current through the electrolysis wires was 
increased to produce the same volume of gas per 


second as in the cavitating condition. No effect 
on the paint pattern could be observed. Especially 
the critical radius remained unchanged. So we 


assume that the nuclei had no disturbing effect on 
the boundary layer. As to the effect of electrolysis 


415 


on the cavitation pattern, three regions on the 
suction side of the propeller blades can be 
distinguished: 

a. At radii larger than the critical radius, 
where, at least near the critical radius, 
laminar separation takes place. 

b. At radii smaller than the critical radius 
having a negative pressure peak at the leading 
edge. 

c. At radii smaller than the critical radius 
having a pressure distribution which is 
nearly shockfree. 

At radii larger than the critical radius no effect 
of electrolysis on sheet cavitation could be seen 
in those cases where it was present. In the few 
cases where no cavitation was present in this 
region application of electrolysis restored inception. 
An example of absence of cavitation, apparently due 
to a lack of nuclei, is shown in Figure 23, where 
blade 3 of propeller C at 60% slip showed consider- 
able cavitation , while blade 4 was free of sheet 
cavitation during the whole run (9 photographs in 

3 different blade positions). 

Absence of cavitation in regions of laminar 
separation, however, is an exception in the steady 
case. A possible explanation is that the water is 
never completely without nuclei and sooner or later 
a nucleus will expand in the separated region and 
cause inception. After inception cavitation seems 
to be more or less self-sustaining. This agrees 
with the observation of Gates (1977) that inception 
on a hemispherical body appeared to be insensitive 
to freestream nuclei content as long as laminar 
separation took place. The situation is different, 
however, in the unsteady case, when a blade passes 
a wake peak. Only a very restricted time is avail- 
able for inception at every propeller revolution 
and a high frequency of encounters with nuclei is 
necessary to obtain inception at every revolution. 
This can explain why the "stabilizing" effect of 
electrolysis is more pronounced behind a ship model 
than in the open-water tests of the current test 
program. 

At higher Reynolds numbers absence of cavitation 
in regions of laminar separation was not observed. 
Apart from viscous effects this can also be caused 
by an increase in encounter frequency of nuclei, 
since an increase in Reynolds number of the same 
propeller models always implied an increase in 
propeller revolutions. 

At radii smaller than the critical radius elec- 
trolysis surprisingly had no effect at all. No 
cavitation was initiated in the minimum pressure 
peak, although the pressure was far below the vapor 
pressure. Even the cavitation pattern at very low 
cavitation index, as shown in Figure 22, was 
unchanged. It is not clear yet why the nuclei do 
not expand. Possibly nuclei do not reach the 
minimum pressure region due to a screening effect 
as described by Johnson and Hsieh (1966). Ina 
situation as shown in Figure 23, however, nuclei 
promoted cavitation inception and were not pushed 
away. This is only possible when the critical 
size of nuclei in a laminar flow region is different 
from the critical size in the reattachment region 
of a laminar separation bubble. 

The third region which has to be considered is 
the region where the pressure distribution is 
nearly shockfree and has its minimum pressure near 
midchord. When the pressure is low in these regions 
bubble cavitation can be expected. A situation 


416 


Rey = 0.66 x 10° 


Ont Sikhs 


FIGURE 23. Inconsistency of cavitation inception 
outside the critical radius at low nuclei content. 
Propeller C at 60% slip. 


like this is shown in Figure 22, but none or only 
a few transient bubbles were seen. 

Electrolysis sometimes restores bubble cavitation 
in this region, but in many cases it does not. 
This inconsistency could even be found on the same 
propeller in virtually the same condition when 
tested repeatedly with long time intervals. In 
one case an abundant amount of large bubbles was 
visible without causing bubble cavitation, while 
an amount of invisibly small nuclei did cause 
bubble cavitation in the same condition. In Figure 
19d it was seen that in the cavitation tunnel 
cavitating spots at the leading edge were formed at 
high Reynolds numbers. When the cavitation index 
was lowered, bubble cavitation occurred in the 
wake of these spots, while at radii in between of 
the spots no bubble cavitation was observed. When 
the cavitation index was lowered to about Oyp=0.5 
the spots were connected with intense bubble cavita- 
tion, as shown in Figure 24. It can be seen that 
the bubble cavitation is related to the spots at 
the leading edge. Apparently the stream nuclei, 
which were abundant in the tunnel at this low 
cavitation index, did not create bubble cavitation, 
while nuclei, generated by a cavitating spot 
created intense bubble cavitation. The possible 
relation between pressure distribution, boundary 
layer, and nuclei distribution must be studied to 
analyse these phenomena. 


7. VARIATION OF THE BOUNDARY LAYER BY ROUGHNESS 
AT THE LEADING EDGE 


In all tests, at least one of the propeller blades 
was roughened at the leading edge, as described in 
Section 2. With paint tests, it was verified that 
the laminar regions were changed into turbulent 

ones. Although the grain size of 30 wm and 60 um 


is larger in comparison with the boundary layer 
thickness, there was a lower limit in the region 
which had to be covered with carborundum to cause 
turbulent flow. For thin sections an evenly dis- 
tributed layer of carborundum of say 0.5% of the 
chord was necessary to trip the boundary layer. 
There was little difference between the effect of 
30 um and 60 um carborundum. At thick sections 
to be effective roughness was necessary until about 
the minimum pressure point. At the pressure side 
the boundary layer remained increasingly laminar 
when the loading increased. At 70% slip the 
the pressure side of the roughened blades was 
completely laminar near the leading edge. 
Attention, given until now to the propeller 
boundary layer, was focussed on the effect on 
torque and thrust. Calculation methods to account 
for Reynolds effects on open-water characteristics 
are based on the assumption of turbulent boundary 
layer flow on the propeller model [Lerbs (1951) ] 
or on an empirical value in between fully turbulent 
and fully laminar, as compiled by Lindgren (1972). 
From the paint tests however, we saw that the 
turbulent region at the suction side strongly 
depends on the propeller loading. The difference 
between the dimensionless thrust and torque coeffi- 
cients, therefore, will not only depend on the 
Reynolds number, but also on the propeller loading. 
In order to eliminate the dependency of thrust 
and torque coefficients on the Reynolds number, 
turbulence stimulators have been used. Sasajima 
(1975) used studs, Yasaki and Tsuda (1972) and 
Tsuda et al. (1977) used trip wires at some distance 
from the leading edge. Apart from changing the 
boundary layer, these devices also have considerable 
resistance of their own. Effects both on thrust 
and torque are difficult to separate. The influence 
of roughness at the leading edge on thrust and 


6 
Rey = 2.72 x10 


Oyy = 0.5 


FIGURE 24. Bubble cavitation in the wake of spots at 
the leading edge. Propeller A at 30% slip in the cavi- 
tation tunnel. 


{e) 0.2 0.4 06 08 1.0 1.2 1.4 
ADVANCE COEFFICIENT J 


FIGURE 25. Effect of leading-edge roughness on torque 
and thrust coefficients. 


torque coefficients is given in Figure 25. These 
measurements were carried out with a special dyna- 
mometer inside the propeller hub to assure that the 
differences were not insignificant due to inaccuracy 
of the measurements. The accuracy in Figure 25 is 
still only about + 0.005. 

Using Lindgren (1972), the value of AK, between 
fully turbulent and fully laminar boundary layer 
flow on the propeller is 0.0035. The actual 
influence of the roughness at the leading edge is 
smaller, so that we can conclude that the resistance 
due to the carborundum was very small. An analysis 
of the effect of roughness at the leading edge on 
the performance of the propeller is beyond the 
scope of this paper. 

The effect of leading edge roughness on cavitation 
is sketched in Figure 18. The radial extent of the 
cavitation is increased in those cases where the 
critical radius was a limit for cavitation. The 
risk of scale effects on cavitation inception due 
to laminar boundary layer flow is largest at low 
propeller loadings, when the risk of laminar 
separation is smallest. But it still can be 
considerable at high loadings, as is shown in Figure 
26, where propeller A is shown with and without 
roughness at 60% slip. 

Application of roughness at the leading edge is 
expected to cause two problems. First the geometry 
of the leading edge may be altered, having a pro- 
found influence on the minimum pressure peak. 
Secondly, the local inception index may be changed 
due to roughness. The effect on the shape of the 
leading edge can only be minimized by using small 
grain sizes. However, to obtain a turbulent boundary 
layer the current 30 wm grainsize was about the 
minimum and no differences in cavitation behavior 
were observed between blades roughened with 30 um 
and 60 wm roughness. The effect of surface irreg- 
ularities on cavitation inception can be large, as 
was shown by Holl (1965). Moreover, Holl points 
out that "the most disastrous place to locate 
surface roughness is at the point of minimum 
pressure of a parent body". This is exactly what 
cannot be avoided at the rather sharp leading edge 
of thin propeller sections. The situation very 
close to the leading edge, however, is different 
from the situation of an isolated roughness at a 


417 


surface, as studied, e.g., by Holl (1965) and 
Benson (1966). Application of their results is 
also difficult, because the ratio between grainsize 
and boundary layer thickness without roughness, 
which is required for the calculations, varies 
rapidly in this region. The boundary layer thickness 
on the smooth blades near the end of the roughness 
was about 30 um in all conditions, when no separation 
took place. At the position of laminar separation 
the boundary layer thickness was only a few um. 
Thus, the ratio of grainsize to boundary layer 
thickness easily varies by a factor of ten. Appli- 
cation of inception calculations on distributed 
roughness [e.g., Arndt and Ippen (1968)] seems 
more appropriate, but this is difficult, because a 
friction coefficient is required for the calculations, 
as well as an "equivalent sandroughness". Both 
are strongly interrelated [Bohn (1972) ] and espe- 
cially near the leading edge these quantities are 
difficult to estimate. 

The roughness elements do form a massive distur- 
bance of the boundary layer and an increase in the 


ES 


SMOOTH 


60 [Lm CARBORUNDUM 
6 


Rey = 0.73 x10 
Ont = 1-5 


FIGURE 26. Effect of leading edge roughness on cavi- 
tation. Propeller A at 60% slip. 


Rey = 0.73 x 10° 


6 


Rey =1.36 x 10 


60 [Lm CARBORUNDUM 
Ont = 15 


FIGURE 27. Effect of Reynolds number on spot cavita- 
tion at roughness elements. Propeller A at 30% slip. 


cavitation inception pressure to a value greater 
than the vapor pressure is possible, which would 
create additional scale effects on cavitation 
inception. To estimate the importance of a possible 
increase in cavitation inception index, the ideal 
inception radius is also given in Figure 18. This 
is the radius where cavitation should start when 

the calculations of the pressure distribution were 
correct and when no scale effects would occur. As 


a 
SMOOTH 


60 [Lm CARBORUNDUM 


0.73 x10® 
Gur = 2-5 


a 

o 
z 

u 


FIGURE 28. Effect of roughness near inception. Pro- 
peller A at 30% slip. 


can be seen, no cavitation occurs inside the ideal 
inception radius, indicating that the pressure at 
inception with roughness is not far from the vapor 
pressure. Assuming that full scale inception takes 
place near the minimum pressure point at the vapor 
pressure [oj=-Cp (minimum)], application of sand- 
roughness can effectively simulate this situation 

at much lower Reynolds numbers. Further experiments 
are necessary to find out the precise effect of 


leading edge roughness on the flow and on the 
boundary layer. Holographic methods, as applied by 
van der Meulen (1976) in studying the effects of 
polymers can be attractive for these experiments. 

When the effect of roughness at the leading edge 
is studied three regions on the model propeller can 
again be distinguished. At radii larger than the 
critical radius, where inception on the smooth 
blades takes place due to laminar separation, the 
cavitation behavior is unaffected by roughness. 
Cavitation was always present on the roughened 
blades. It is unknown if the sensitivity to nuclei 
in the unsteady case increases, as is suspected on 
the smooth blade. Experiences with several other 
propellers behind a model indicate that this is 
not the case and that nuclei have very little effect 
when roughness is applied. 

In the laminar region, at radii smaller than the 
critical radius, roughness at the leading edge has 
its major effect, as described above. In some 
cases, however, problems appeared in the form of 
streaky cavitation as shown in Figure 27a. When 
the pressure on the blade sections was constant, 
as was the case for propeller A at r/R=0.7 and for 
propeller C at r/R=0.8, both at 30% slip (Figure 9), 
and when the Reynolds number was low, cavitating 
streaks were formed behind the roughness elements. 
In Figure 27b the same blade in the same condition 
at a higher Reynolds number is shown. Here a smooth 
cavity is seen. The roughness elements apparently 
suffer from laminar separation at low Reynolds 
number and cavitation occurs in the separated regions 
behind the roughness. The length of the spots is 
strongly dependent on the cavitation index, as is 
shown in Figure 28b, where the same situation as 
in Figure 27a is shown at a somewhat higher cavita- 
tion index. The spots disappeared and the propeller 
is near inception. Figure 28 also shows that in- 
ception of the sheet at the leading edge is not far 
from the vapor pressure, because the ideal inception 
radius in this case was 0.78. When roughness was 
applied, electrolysis had no further effect at 
radii smaller than the critical radius. 

In the region with shockfree nressure distribution, 
bubble cavitation was seen to be promoted in some 
cases by roughness at the leading edge. The influ- 
ence of roughness, however, was inconsistent again 


SMOOTH 60 [Lm CARBORUNDUM 
Rey = 2.72x10° 


Gut = 1.0 


“FIGURE 29. Effect of leading edge roughness on bubble 
cavitation. Propeller A at 30% slip in the cavitation 
tunnel. 


419 


in this region, as it was with electrolysis. When 
there was cavitation at the leading edge due to 

the roughness, again bubble cavitation appeared at 
midchord, as is illustrated in Figure 29, where 
nuclei generated by cavitation at the leading edge 
created bubble cavitation at midchord. The cavita- 
tion index at 0.7R in Figure 29 is 0.18 and the 
minimum pressure coefficient from Figure 9 is 0.20, 
so the situation with roughness seems to be the 
situation without scale effects on cavitation 
inception. Nuclei in the flow, however, did not 
create bubble cavitation. 


8. 


CONCLUSIONS 


The results of the present test program can be 
summarized as follows: 


ib5 


On the suction side of a model propeller a 
critical radius can exist outside of which 

the boundary layer is turbulent from the 
leading edge. This critical radius is due 

to laminar separation, as was seen from some 
observations, from calculations (Figure 14), 
and from the Reynolds independency of the 
critical radius. (Figure 15). 

To obtain natural transition near the leading 
edge on a propeller model, high Reynolds 
numbers (Reyy>2 510°) are required. 

The critical radius is a limit for the radial 
extent of sheet cavitation from the leading 
edge. An increase of nuclei by electrolysis 

is ineffective in the laminar region (Figure 22). 
Outside the critical radius, cavitation is not 
inhibited (the inception pressure was not 
accurately determined), but a lack of nuclei 

at low Reynolds number seems to decrease the 
frequency of inception (Figure 23). In the 
unsteady case the nuclei content of the water 
is probably important in this region. 

Roughness at the leading edge can effectively 
remove the critical radius, thus simulating a 
higher Reynolds number. Inception of cavitation 
at the roughness elements occurs close to the 
vapor pressure, which is assumed to be also 

the case on the prototype. 

When the pressure distribution is very flat 

and the Reynolds number is low, the roughness 
elements can induce spots of cavitation. The 
length of these spots is strongly dependent on 
the cavitation index and is different from the 
cavity length at high Reynolds numbers. This 
is probably due to laminar separation at the 
roughness elements (Figure 27). 

The inception of bubble cavitation near mid- 
chord at inner radii is not consistent. There 
seems to be an interaction between the pressure 
distribution, the nuclei distribution, and 

even the boundary layer. When cavitation at 
the leading edge is present, bubble cavitation 
occurs near midchord when the pressure is below 
or near the vapor pressure in that region. 
Lifting line and lifting surface calculations 
can adequately predict the open-water character— 
istics of a propeller. For the calculation of 
the pressure distribution, however, lifting 
surface calculations are necessary. The corre- 
lation between calculations and the results of 
paint tests and cavitation observations is good. 


420 


From the previous investigations in uniform flow 
some tentative explanations can be given of the 
scale effects on cavitation as shown in Figures 
1-3. The explanations can only be tentative since 
the unsteady pressure distribution on the propellers 
in the wake is not known. Propeller A in Figure 1 
apparently had a critical radius at r/R=0.9 in 
this blade position, which was removed by roughness 
at the leading edge. Also, behind the model in 
some situations no cavitation at all occurred in 
the wake peak, which is expected to be due to a 
lack of nuclei (as seen in Figure 21). 

The lack of nuclei is more apparent at propeller 
B. The critical radius is expected to be near the 
hub, but the low encounter frequency with nuclei 
of sufficient size makes cavitation inception more 
or less random. The irregular collapse of the 
cavity on propeller C is apparently due to a strong 
change in the pressure distribution, due to a sharp 
wake peak. The critical radius at the position of 
Figure 3 is near r/R=0.9 but the cavity at inner 
radii is still collapsing. This phenomenon could 
also be seen on high speed films, where the sheet 
cavity was seen to detach from the leading edge 
and collapse while moving with the flow. Some 
cavitating spots can be seen at r/R=0.8 on propeller 
(Cr 


ACKNOWLEDGMENT 


Part of this program was supported by the Dutch 
Ministry of Economic Affairs under the ICOSTE- 
program. 


REFERENCES 

Acosta, A. J., and B. R. Parkin (1975). Cavitation 
inception - a selective review. J. of Ship 
Research, 19, 193. 

Albrecht, K., and O. Bjorheden (1975). Cavitation 


testing of propellers in a free surface tunnel 
utilizing micro air bubble control. J. of Fluids 
IDK BASIC 

Arakeri, V. H. (1973). Viscous effects in inception 
and development of cavitation on axi-symmetric 
bodies. Thesis, Calif. Inst. of Techn. Report 
No. Eng. 183-1. 

Arakeri, V. H., and A. J. Acosta (1973). Viscous 
effects in the inception of cavitation on axi- 
symmetric bodies. J. of Fluids Eng., 519. 

Arakeri, V. H. (1975). A note on the transition 
observations on an axi-symmetric body and some 
related fluctuating wall pressure measurements. 
Trans A~SMsk., 82 

Arndt, R. E. A., and A. T. Ippen (1968). Rough 
surface effects on cavitation inception. 
AJS.M.E., J. Of Basic’ Eng., 249) 

Arndt, R. E. A., and A. P. Keller (1976). Free 
gas content effects on cavitation inception and 
noise in a free shear flow. Proc. of IAHR- 
Symposium, Paris, France, 3. 

Benson, B. W. (1966). Cavitation inception on 
three dimensional roughness elements. David 
Taylor Model Basin, Rep. 2104. 

Bindel, S. G. (1969). Comparison between model and 
ship cavitation. 12th ITIC, Cavitation Committee, 
App. III, 365. 


Bohn, J. C. (1972). The influence of surface 
irregularities on cavitation: a collation and 
analysis of new and existing data with application 
to design problems. Thesis. Penn State Univ., 
Ordnance Res. Lab., Tech. Mem., File No. 72-223. 

Casey, M. V. (1974). The inception of attached 
cavitation from laminar separation bubbles on 
hydrofoils. Inst. of Mech. Eng., Conference on 
cavitation. Edinburgh, Scotland. 

Curle, S. C., and S. W. Skan (1957). Approximate 
methods for predicting separation properties of 
laminar boundary layers. Aeron, Quart, 8, 257. 

Daily, J. W., and V. E. Johnson Jr. (1956). Tur- 
bulence and boundary layer effect on cavitation 
inception from gas nuclei. Trans. A.S.M.E., 

78, 1965. 

Gates, E. M. (1977). The influence of freestream 
turbulence, freestream nuclei populations and 
a drag-reducing polymer on cavitation inception 
on two axi-symmetric bodies. Calif. Inst. of 
Techn. Report No. Eng. 183-2. 

Gent, W. van (1977). On the use of lifting surface 
theory for moderately and heavily loaded ship 
propellers. Thesis, Netherlands Ship Model 
Basin, Publ. No. 536. 

Goldstein, S. (1948). Low-drag and suction airfoils. 
J. of the Aeron Sciences, 15. 

Hoekstra, M. (1975). Prediction of full scale 
wake characteristics based on model wake survey. 
Inst. Shipbuilding Progress, 22. 

Holl, J. W., and G. F. Wislicenus (1961). Scale 
effects on cavitation. A.S.M.E. J. of Basic 
imneteis pn Chip Sec 

Holl, J. W. (1965). The estimation of the effect 
of surface irregularities on the inception of 
cavitation. A.S.M.E. Symp. on Cavitation in 
Fluid Machinery, G. M. Wood et al. (eds), 3. 


Holl, J. W. (1970). Nuclei and cavitation. Trans. 
WNASctleltan Co. Cis EACHIC Iino, SeiL. 
Huang, T. T., and D. E. Hannan (1975). Pressure 


fluctuations in the regions of flow transition. 
David Taylor Model Basin, Report No. 4728. 

Huang, T. T., and F. B. Peterson (1976). Viscous 
effects on model/full scale cavitation scaling. 
Wig: (ere Riehl INAS6 5 BOR 2do 

Johnson, V. E., and T. Hsieh (1966). The influence 
of trajectories of gas nuclei on cavitation 
inception. Proc. 6th Symp. on Naval Hydrody- 
eee, MEG to Ie Cap To 

Keller, A. P. (1974). Investigations concerning 
scale effects of the inception of cavitation. 
Proc. Conference on Cavitation, Inst. of Mech. 
Eng., Edinburgh, Scotland. 

Kuiper, G. (1974). Cavitation testing of marine 
propellers in the NSMB Depressurized Towing 
Tank. Proc. Conference on Cavitation, Inst. of 
Mech. Eng., Edinburgh, Scotland. 

Kuiper, G. (1978). Cavitation scale effects, A 
case study. Int. Shipbuilding Progress, 25. 
Lerbs, H. W. (1951). On the effects of scale and 
roughness on free running propellers. J. of 

the Am. Soc. of Naval Arch., 63, 58. 

Lerbs, H. W. (1952). Moderately loaded propellers 
with a finite number of blades and an arbitrary 
distribution of circulation. Trans. SNAME, 60. 

Lighthill, M. J. (1951). A new approach to thin 
aerofoil theory. Aeron. Quart, 3. 

Lindgren, H. (1972). Ship model correlation method 
based on theoretical considerations. Proc. 13th 
ITTC, Report of Performance Committee, App. 2, 
181. 


Maltby, R. L., ed. (1962). Flow visualization in 
wind tunnels using indicators. AGARDograph 70. 

Meulen, J. H. J. van der, (1976). A holographic 
study of cavitation on axi-symmetric bodies 
and the influence of polymer additives. Thesis, 
Netherlands Ship Model Basin, Publ. No. 509. 

Meyne, K. (1972). Untersuchung der Propellergrenz- 
schichtstrémung und der Einfluss der Reibung 
auf die Propellerkenngréssen. Jahrbuch der 
Schiffbautechnischen Gesellschaft 66, 317. 

Michel, R. (1951). Etude de la transition sur les 
profils d'ailestablissement d'um point de tran- 
sition et calcul de la trainee de profil incom- 
pressible. ONERA report 1/1578A. 

Morgan, Wm. B., V. Silovic, and S. B. Denny (1968). 
Propeller lifting surface corrections. Trans. 
SNAME, 76. 

Noordzij, L. (1976). Some experiments on cavita- 
tion inception with propellers in the NSMB 
Depressurized Towing Tank. Intern. Shipbuilding 
Progress, 23. 

Okamoto, H., K. Okada, Y. Saito, and T. Takahei 
(1975). Cavitation study of ducted propellers 
on large ships. Trans. SNAME, 83. 

Oossanen, P. van (1974). 
and cavitation characteristics of propellers 
including effects of non-uniform flow and 
viscosity. Thesis, Neth. Ship Model Basin, 
Publ. No. 457. 

Pinkerton, R. M. (1934). Calculated and measured 
pressure distributions over the midspan section 
of the NACA 4412 airfoil. NACA Rep. No. 569. 

Sasajima, H., and I. Tanaka (1966). On the 
estimation of wake of ships. Proc. llth ITTC, 
Tokyo. 

Sasajima, T. (1975). A study on the propeller 
surface flow in open and behind conditions. 
ieareier, Iiekolag Tne, Sipe Tale 

Schiebe, F. R. (1969). The influence of gas nuclei 
size distribution on transient cavitation near 
inception. Univ. of Minnesota, St. Anth. Falls 
Hydr. Lab., Proj. Report No. 107. 

Smith, A. M. O., and Nathalie Gamberoni (1956). 
Transition, pressure gradient and stability 
theory. Douglas Aircraft Co. Rep. 26388. 

Smith, A. M. O. (1957). Transition, pressure 
gradient and stability theory. Ix Congres 
International de Mechanique Appliquée, IV. 
Bruxelles, Belgium. 

Sparenberg, J. A. (1962). 
surface theory to ship screws. 
Acad. of Sciences. Series B, 5. 


Application of lifting 
Royal Netherlands 


Calculation of performance 


421 


Theodorsen, Th. (1932). Theory of wing sections of 
arbitrary shape. NACA Report No. 411. 

Thwaites, B. (1949). Approximate calculation of 
the laminar boundary layer. Aeron Quart, 1, 
245. 

Tsakonas, S., W. R. Jacobs, and M. R. Ali (1976). 
Propeller blade pressure distribution due to 
loading and thickness effects. Stevens Inst. of 
Techn., Report S.T.T.-D.C.-76-1869. 

Tsuda, T., S. Konishi, and S. Watanabe (1977). On 
the application of the low pitch and high 
revolution propeller to the self propulsion test. 
ITTC Performance Committee. 

Wrench, J. W. (1957). The calculation of propeller 
induction factors. David Taylor Model Basin, 
Rep. No. 1116. 


APPENDIX 


The geometry of the four propellers, used in this 
study and shown in Figure 4, is given in this 
appendix. The output is from a propeller data 
base and is not dimensionless but in mm on model 
scale. Propellers A and C were stored in the data 
base on a different model scale than actually used 
in the tests, but this has no further impact. Cal- 
culations were made directly from this data-base. 

At each radius, R, the pitch, P, is given, 
together with the distance to the generator line of 
the trailing edge, TE, the leading edge, LE, and 
the position of maximum thickness, TM. The positive 
direction is from the generator line to the leading 
edge. 

The geometry of the propeller section is given 
by the thickness and the distance of the face of 
the propeller section above the pitch line. The 
ordinates of the section geometry are given as 
percentages of the distance between point of max- 
imum thickness and leading edge (positive) or 
trailing edge (negative). The origin therefore 
always is at the point of maximum thickness of the 
profile. 

The profile thickness at leading and trailing 
edge is finite in this appendix. The radii at the 
leading edge were determined by generating a spline 
through the profile contour or by interpolating in 
the transformed plane after conformal mapping. 

Both interpolating techniques gave nearly the same 
results and were very close to the actual propeller 
geometries. 


tv 3a16vL 


‘WW NI SSSNMOIHL 310d 


(WW 8'>Oocge SVM WVe9Oed LS3L 
SHL NI GASN TAGOW 3HL AO YSALAWVIO YsATTsdOud SHL AN) 


‘WWW OF BZL = YSALAWVIG 
(WG8ZlO0 340 YSaLaWVIG V IV) VY Y¥3711ed0ud 


422 


‘eV Savi 


OL- 


os- 


WwW NI 


O6- S6- OOL- Ade 


SSSNWDIHL 3Ws0¥8d 


WW €’SlLd = YALAWVIC 


‘G wamadOud 


423 


EV alagvl 


Ede be Shu eves GZo° 

Sp ve 6c° ve 2° ve ae jon O 

BOR GPCR Y= YR SEO et BOO = po ORR As? “oa 2 a50 * jx59 ec fee sue Bape nae. 
2u* 20%= ga* a DI oemocn 6e* Zp vs° BO) eee 00: OS a OG Uy Vee ence reyes 8Gg° 
20 pa°= pd I 63 2¢ cp 2s a9 177 94 tio eGGee SOG. Che hee CU hoe oAg? 
BUF RUE OO NU Ge REGS ODE GRR PR YP RP YG GaP RD ao a0 eoz° 
COrmaeec 6208 Gere z= atte aa%e ge° gt* “2° 2c? tee ze° gue ir BBR PO Garo ang? 
68 poe t9 8 49 = 99° £9 © bo'e F9°® gG°= BG°= Op°e Gr°e 6p°= pGr= form Byte Zore Ho! aes* 


ty’ p6°s ac*t= za°t= curze cerze go®ze od°2= getz= Z6°z™ Cotze deize CL°z= C9%e= Co®ze CEoze 92°20 tye qe= 


OOlL- G/k2 


‘WW NI SSSANHOIHL 31sOu’d 


“(WU O'OOE SVM WVe9OUd LS3L 
SHL NI GaSN TAaGOW AHL 4O YSLAWVIC YSaTTAdOud SHL GN) 


WW OO'Ore = YALAWVIG 
(W720 JO YSLEWVIG V IV) “D3 YAaTIadOdd 


424 


eeceeeeee 
FOMNN-— — 


14 
Gy 
Te 
ra 
26 
iw4 
96 
iw 


‘yV Alavl 


os- 


WW AI 


O6- S6- OOL- GA/e2 


SSSNYDIHL 31sdOu’d 


WwW O'?lLe = YSALAWVIG 


‘dG = Y¥371ad0ud 


425 


426 
Discussion 


SHIN TAMIYA, HIROHARU KATO, and YOSHIAKI KODAMA 


SETTLING SECTION —-NUCLE! GENERATION SECTION 


TEST SECTION (80x80x10G0) 


4 
=F 


EED TANK 
— FILTER TANK “ORIFICE 


FIGURE 1. General arrangement. 


The discussers appreciate the excellent re- 
search work on cavitation inception done at NSMB*. 
At the University of Tokyo the discussers also per- 
formed similar experiments using both hemispherical 
and ITTC headforms tested in our newly built cav- 
itation tunnel. This tunnel has a filtering tank 
containing 60 cartridge type filters, which contin- 
uously remove air nuclei and solid particles larger 
than ca. 1 um from water (Figure 1). |. 

Figures 2 through 7 show the effect of elec- 
trolysis on cavitation inception. The photograph 
in Figure 3 was taken a few seconds after that in 
Figure 2. The flow conditions are exactly the same 
for Figures 2 and 3; the only difference is the 
presence of hydrogen bubble nuclei. The photographs 
in Figures 4 and 5, as well as 6 and 7, were also 
taken under the same conditions. 

In the discussers' experiments the cavitation 
caused by electrolysis nuclei generates only bubble PICUREV Qe muitehoutl hydrogenabubblles\vasichsunver 
type cavitation. Even when sheet cavitation exists, 3 ='Osehh. 
the cavitation bubbles caused by the electrolysis 
nuclei seem to break up the sheet cavity. 


*Netherlands Ship Model Basin FIGURE 3. With hydrogen bubbles V = 6.8 m/s, 
#Statens Skeppsprovningsanstalt Ci = Machi, 


FIGURE 4. Without hydrogen bubbles V = 6.8 m/s, 
o = 0.71. 


FIGURE 5. With hydrogen bubbles V = 6.8 m/s, 
o=0.71. 


FIGURE 6. Without hydrogen bubbles V = 6.8 m/s, 
o = 0.60. 


O. RUTGERSSON 


I would like to congratulate the author of 
this interesting paper. As a complement to the data 
presented I think that some results obtained at 
SSPA# when testing high-speed propellers could be 
of some interest. A propeller of the supercavi- 
tating type was tested with three different gases 
in the water. Also, two different conditions of 
the blade surface were used, smooth polished and 
painted with a thin spray paint giving the surface 
some roughness. 

In Figure 1 the propeller characteristics from 
these tests for homogenous flow at the cavitation 
number, 0 = 0.6, are shown. In the partially cav- 
itating region (J > 1.0) there is a very pronounced 
influence due to gas content for the polished pro- 
peller. For the painted propeller no such influ- 


FIGURE 7. With hydrogen bubbles V = 6.8 m/s, 
o = 0.60. 


ence was found. Cavitation pictures at the advance 
ratio, J = 1.1, give the explanation for these 
differences. Figure 2 shows the cavitation at the 
lowest gas content (a/a, = 0.2) for the polished 
propeller. The cavitation pattern is divided into 
two parts. The first part is a sheet starting at 
the leading edge. The second part is an unstable 
sheet of bubble cavitation at the aft part of the 
blade. Tests at higher gas contents (Figure 3, 
a/a, = 0.4) show that the aft part cavitation now 
has a larger extension. The painted propeller 
(Figure 4) shows a rather different pattern for the 
aft part cavitation (the leading edge sheet is al- 
most uninfluenced by gas content and roughness) . 
The aft part cavitation now consists of a thin sheet 
of very small bubbles. The sheet also has a rela- 


428 


tively larger extension on the painted propeller 

than on the polished propeller. Obviously it is 

the changes in this aftpart cavitation that cause 
the changes in propeller characteristics. 

Full scale tests have also been conducted with 
this propeller design. In Figure 5 the full scale 
cavitation pattern corresponding to the model tests 
is shown. This cavitation pattern is very similar 
to that of the painted model propeller. 

In the author's Figure 29 bubble cavitation 
is shown very similar to that in tests with the 
painted propeller at SSPA. The author concludes 


08 a/Os 
0.4 Painted 
Ky —-—-— 0.2 Not painted No 
0.4 Not painted — 
10Kq ———— 0.6 Not painted 
n 
° 06 
0.4 
0.2 


all 
0.6 0.8 - 1.0 1.2 AIS 
ADVANCE RATIO 


FIGURE 1. Propeller characteristics at 6 - uU.6. 


FIGURE 2. J = 1.1 a/a = 0.2 polished blade.* 


that this cavitation is inconsistent. Based on our 
experience with full scale cavitation, however, I 
think that the pattern shown could very well repre- 
sent a full-scale case. 

The influence of nuclei content and blade 
roughness on the cavitation pattern is found to be 
rather similar in the tests at NSMB and SSPA. The 
main difference is the necessary amount of rough- 
ness. This difference is possibly due to the dif- 
ference in Reynolds number, about 10 times as high 
in the tests at SSPA as in those carried out at 
NSMB. 


FIGURE 3. J = 1.1 a/a = 0.4 polished blade.® 


FIGURE 4. J =1.1a/a = 0.4 painted blade. 


FIGURE 5. J = 1.1 6 = 0.65 full scale. 


429 


Author’s Reply 


G. KUIPER 


Both the hemisphere and the ITTC body are 
known to exhibit laminar separation in the zenge 
of Reynolds numbers (estimated at about 2 x 10°) 
used in the experiments of Tamiya et al. as was 
already shown by Arakeri and Acosta (1973). They 
now point to an apparent discrepancy between the 
results as described in my paper and their obser- 
vations: on the propellers nuclei were found to 
generate sheet cavitation in the very few cases 
where it was not yet present, and the nuclei never 
changed the appearance of the cavity. 

First of all, the cavitation patterns, both 
with and without electrolysis, on the headforms of 
the discussers show a remarkable resemblance to 
various patterns shown on the ITTC bodies at other 
facilities [Lindgren and Johnsson (1966) and also 
reproduced by Gates and Acosta in their paper on 
this program] illustrating that the nuclei content 
was at least one of the factors causing the varia- 
tion in type of cavitation observed at different 
facilities. 

From the observations of the discussers it can 
be concluded that the nuclei, generated by elec- 
trolysis, removed the laminar separation bubble in 
the same manner as shown very clearly by Gates and 
Acosta in their symposium paper. This phenomenon 
was found when there were many large free stream 
bubbles in the flow, as can also be observed in the 
pictures of the discussers. In our case, however, 
we verified with a paint test that electrolysis did 
not remove the laminar separation bubble by veri- 
fying that the critical radius was unchanged. 

The observations of the discussers show that 
an overdose of nuclei can change the situation 
considerably. Gates and Acosta assume that the 
free stream bubbles do trip the boundary layer. 
Another possibility, however, is that the dynamic 
behavior of the bubbles near the minimum pressure 
region changes the pressure distribution on the 
body, specifically by decreasing the low pressure 
peak. This would also remove laminar separation, 
leaving the boundary layer laminar over a longer 
distance. In fact the nuclei do not only affect 
cavitation inception but they change the free 
stream conditions, making a correct comparison of 
the inception phenomena impossible. 

Rutgersson, in his discussion, gives an illus- 
tration of a possible effect of nuclei and roughness 
on bubble cavitation. With the pictures alone, 


only some assumptions can be made as to what hap- 
pened on this propeller, but I will make an attempt 
to give an explanation. 

Although the Reynolds number was rather high 
it looks like the boundary layer within r/R = 0.8 
is laminar over a large portion of the chord, while 
the minimum pressure region is near midchord 
(Figure 2). An increase of the nuclei content 
leads to occasional cavitating spots, starting at 
the low pressure region (Figure 3). On the painted 
blade, however, the boundary layer seems to be 
turbulent and bubble cavitation starts there, near 
the minimum pressure region (Figure 4). : 

If my tentative description is correct there 
is a difference between the discussers' case and 
Figure 29 (and also Figure 24) from my paper, since 
there the boundary layer in the region of low pres- 
sure was turbulent, and still no bubble cavitation 
occurred. Only when cavitation, generated by rough- 
ness, at the leading edge took place, a separate 
region of bubble cavitation also appeared. 

Whatever may be the case, it must be kept in 
mind that these descriptions of phenomena do not 
explain them, because it is not clear to me why 
there should be any interaction between the bound-~ 
ary layer and the free stream nuclei and which 
parameters would control this. I think more sys- 
tematic research is necessary to be able to 
simulate bubble cavitation on model propellers 
in a reproducible way. 

I agree with the suggestion of the author that 
the increased amount of bubble cavitation, as shown 
many times by roughened propeller models, may well 
be representative for full-scale cases. Bubble 
cavitation seems to be inhibited on scale models 
very easily. When bubble cavitation does occur on 
scale models the situation is so bad that invari- 
ably erosion problems do occur on full-scale. 
Ironically a better simulation of bubble cavitation 
may not make the interpretation easier. 

In general both discussions have made it clear 
again that it is impossible to make general state- 
ments about the effect of nuclei or roughness. To 
make any interpretation and to avoid confusion the 
test conditions must be given as complete as pos- 
sible. Finally, I thank the discussers for their 
discussions and for their kind attention to my 
paper. 


yi 


uy 


“ 


Session VI 


CAVITATION 


BLAINE R. PARKIN 

Session Chairman 
Pennsylvania State University 
State College, Pennsylvania 


1 

F c* et 

meats . a 

< 
s < 
» z 

i 
= 


A Holographic Study of the Influence of 
Boundary Layer and Surface Characteristics 
on Incipient and Developed Cavitation on 
Axisymmetric Bodies 


J. H. J. van der Meulen 
Netherlands Ship Model Basin 
Wageningen, The Netherlands 


ABSTRACT 


This paper describes an experimental investigation 
of boundary layer flow and cavitation phenomena on 
three axisymmetric bodies. The bodies possess 
different boundary layer or surface characteristics. 
The importance of these features for incipient and 
developed cavitation are studied by using in-line 
holography. A good correlation is found between 
observations and calculations of laminar flow 
separation and subsequent transition to turbulence 
of the separated shear layer. The influence of 
polymer additives on laminar flow separation is 
studied in detail. The results of this study explain 
the effect of cavitation suppression by polymer 
additives on certain bodies. 


1. INTRODUCTION 


Axisymmetric bodies have often been used to study 
the inception of cavitation. These studies were 
usually made by systematically varying the parameters 
related to the liquid flow (velocity, turbulence, 
air content, pressure history) or to the body (size, 
surface roughness, wettability). Although a con- 
siderable knowledge of cavitation was obtained in 
this way, a complete understanding of many cavitation 
phenomena was still lacking. A breakthrough was 
achieved by Acosta (1974) who emphasized the need 
for a thorough understanding of the basic fluid 
mechanics of the liquid flow surrounding the bodies 
in which cavitation takes place. This statement 
was based on an earlier study by Arakeri and Acosta 
(1973) in which the boundary layer flow was visual- 
ized by the employment of the schlieren method. 
Cavitation inception could be correlated with the 
occurrence of laminar flow separation. Unawareness 
of this important flow phenomenon had obscured the 
results of comparative cavitation studies with 
axisymmetric bodies, made in the past. 

In general, it can be stated that cavitation 


433 


inception on a body is affected by nuclei, viscous, 
and surface effects. The present study deals with 
the two latter effects. The use of holography, a 
three-dimensional imaging technique, enabled a new 
approach. The employment of this method for the 
observation of cavitation inception phenomena has 
been reported before by Van der Meulen and Ooster- 
veld (1974). In the present study an extended 
version of the method has been used by which boundary 
layer flow phenomena also could be observed. Viscous 
effects were studied by comparing two axisymmetric 
bodies, a hemispherical nose having laminar flow 
separation and a blunt nose not having it. Surface 
effects were studied by comparing two hemispherical 
noses, one made of stainless steel, the other made 

of Teflon. 

The phenomenon of turbulent-flow friction reduc-— 
tion by polymer additives of high molecular weight 
has been known for about thirty years. In recent 
years an increased interest has been shown on the 
effect of polymer additives on cavitation. In the 
present work the influence of polymer additives on 
the flow about the test bodies is studied and 
related with the influence on cavitation. 


2. EXPERIMENTAL METHODS AND PROCEDURE 
Description of Test Facility 


The facility used is the high speed recirculating 
water tunnel of the Netherlands Ship Model Basin. 
Originally, the maximum speed in the 40 mm circular 
test section was 65 m/s and the maximum allowable 
tunnel pressure 35 kg/cm?. A detailed description 
of this tunnel and its air content regulation system 
is given by Van der Meulen (1971, 1972). For the 
present study a new test section was made. It has 
a 50 mm square cross section with rounded corners 
(radius 10 mm), to limit the influence of the walls. 
The models, having a diameter of 10 mm, occupy 3.25 
percent of the cross-sectional area of the test 


434 


MZ 


[_] purse -pamper 


ae 


POLYMER INJECTION LINE 


DEAERATION LINE 


CENTRIFUGAL Cc | 


[| [ {eeu 
— 


| {COOLING WATER 


c 


| 


eo; 
FIGURE 1. Schematic diagram of high fa Stass aeanee 
speed cavitation tunnel with polymer 


injection system. 


section. Injection of polymer solutions from the 
nose of the models was made by a Hughes Centurion- 
100 pump unit. The unit consists of a drive mech- 
anism fitted with two pump heads. A pulse-damper 
was used to minimize flow variations. Further 
details are given by Van der Meulen (1974b). A 
schematic diagram of the tunnel with the polymer 
injection system is shown in Figure 1. 

To measure the influence of polymer additives 
on the friction factor and the surface tension of 
the solutions, a turbulent-flow rheometer and a 
surface-tensionmeter have been used. Details on 
these measuring devices are given by Van der Meulen 
(1974a, 1976b) . 


Test Models 


According to Arakeri and Acosta (1973), most 
axisymmetric models used in cavitation inception 
studies, such as the hemispherical nose and the 
ITTC standard headform, exhibit laminar boundary 
layer separation. It means that the laminar boundary 
layer is unable to overcome the adverse pressure 
gradient and the flow separates from the wall. 
Schiebe (1972) introduced a standard series of 
axisymmetric models which, theoretically, should 
not exhibit boundary layer separation. To distin- 
guish between these two classes of axisymmetric 
models, a hemispherical nose and a blunt nose, 
selected from Schiebe's standard series, were used 
in the present investigations. Both models were 
made of stainless steel (SST). In addition, a 
third model (hemispherical nose) was used, made of 
Teflon. The contour of the blunt nose is derived 
from the combination of a normal source disk and 
a uniform flow. Schiebe (1972) calculated the 
dimensionless coordinates and pressure coefficients 
for a series of models in the range, Co din = Oods 
(point source) - 1.0. From this series a blunt 
nose with a minimum pressure coefficient of 0.75 
was selected. 

The diameter, D, of the cylindrical part of the 
hemispherical nose is 10.00 mm. Theoretically, 
the diameter of the blunt nose increases smoothly 


TUIUILILZ/7 


Oc MOTOR 


mtd 


100i 


y 


i pk { 


WATER SUPPLY 


ROTARY PUMP. 


VACUUM PUMP 


to an asymptotic value, D, with increasing axial 


distance, x. 
However, 


This value was set at 10.00 mm. 
for the manufacture of the blunt nose a 


minor deviation from the theoretical contour had 


to be permitted. 


HEMISPHERICAL NOSE 


FIGURE 2. 
(dimensions in mm). 


Thus, 


the actual contour coincides 


BLUNT NOSE 


Cross sections of stainless steel models 


435 


1.0 T T T T T 1 
0.9 irrotational 
Se irrotational flow with walls 
o/X 5 
yo” = 
08 / \ e Re=2.1 x10 


2) 
N 


fo} 
a 


Pressure Coefficient , Cp 
fe} 
h 


(eo) 
WwW 


0.2 


0.1 


0.4 0.6 0.8 10 1.2 1.4 


Surface Coordinate over Diameter, s/D 


with the theoretical contour over a distance, x/D 

= 0-1.6, and next changes smoothly into a circular 
cylinder with a diameter of 9.88 mm. The cross 
sections of the SST models are shown in Figure 2. 
For the Teflon hemispherical nose the dimensions 
are the same as for the SST hemispherical nose. 
However, the Teflon model was not made of solid 
Teflon but consisted of a Teflon nose slipped on 

a SST core. Extreme care has to be exercised in 
manufacturing models for cavitation studies. An 
accurate similarity of the model contour is essential, 
but a smooth surface is even more critical. The 
drastic effects of surface roughness, in particular 
isolated irregularities, on cavitation inception 
have been demonstrated by Holl (1960) and Arndt 

and Ippen (1968). The present models were made by 
Instrumentum TNO in Delft. The models were inspected 
by an optical comparator (magnification 50x). For 
the SST hemispherical nose the maximum deviation 
from the true contour was within 5 um, for the 
Teflon hemispherical nose within 10 pm. For the 
blunt nose, the maximum deviation for x/D < 0.3 

was within a few microns and for x/D > 0.3 within 
10 um. The mean surface roughness height for the 
SST models was 0.05 um; for the Teflon model this 
value was considerably higher. 

Computations of the pressure coefficient for the 
hemispherical nose and the blunt nose were made at 
the National Aerospace Laboratory NLR, The Nether- 
lands. The velocity potential for irrotational 
flow along the model contour was computed with the 
variational finite element method according to 


flow without walls _| 


FIGURE 3. Computed pressure co- 
efficient as a function of surface 
coordinate over diameter for hemi- 
spherical nose. Data points ob- 
tained from measurements by Rouse 
and McNown (1948) at Re = 2.1 x 
10° are included. 


20 


1.6 


18 


Labrujére and Van der Vooren (1974). This method 
“is suitable for axisymmetric flows. The relation 
between the pressure coefficient, Cp, and the 
velocity potential, $, is given by 


2 
CK) |W se) (1) 
where s is the streamwise distance along the model 
contour and V, the free stream velocity. The pres- 
sure coefficient was computed in the absence of 
tunnel walls and with tunnel walls. In the latter 
case, it was necessary to substitute the square 
cross section with rounded corners by a circular 
one (diameter 55.44 mm), having the same cross- 
sectional area. For the hemispherical nose, the 
results are plotted in Figure 3. Also given are 
data points obtained from measurements by Rouse 
and McNown (1948) at a Reynolds number of 2.1 x 10°. 
The computed Cp-values are claimed to be accurate 
within 0.1 percent. The Cp-value for irrotational 
flow in the absence of tunnel walls is 0.7746 at 
s/D = 0.6825 (y = 78.2°). With tunnel walls the 
Cpmin Value at the same location is 0.8367. For 
the blunt nose, the results are plotted in Figure 
4. The computed Cp-values are accurate within 1 
percent. The Cp-value for irrotational flow in 
the absence of tunnel walls is 0.750, which is 
consistent with the accurate computations by Schiebe 
(1972). With tunnel walls the Cpy;j,-value is 0.802. 
Tabulated values of Cp are presented by Van der 
Meulen (1976b) and Labrujére (1976). 


436 


Pressure Coefficient, Cp 


FIGURE 4. Computed pressure co- 
efficient as a function of surface 
coordinate over diameter for blunt ie) 02 
nose. 


Holographic Method 


In the present work, in-line holography has been 
used to study cavitation and flow phenomena about 
the test models. The method consists of making 
photographic records containing detailed information 
on the cavitation and flow patterns. Holography 
has become one of the most important areas of modern 
optics since the invention of the laser as a new 
light source. Holography is usually described as 
a method for storing wavefronts on a record from 
which the wavefronts may later be reconstructed. 
The record, formed in photosensitive material, is 
called a hologram. In forming holograms two sets 
of light waves are involved: the reference waves 
and the subject waves. In the present case of in- 
line holography only one set of waves is used 
basically. The undeflected light waves from this 
set of waves act as reference waves, the light 
waves deflected by the subject act as subject waves. 
A schematic diagram of the applied optical system 
is shown in Figure 5. The light source is the 
Korad K-1QH pulsed ruby laser of the Institute of 
Applied Physics TNO-TH. To improve the resolution 
of the system, the red light from the ruby laser 
is converted to ultraviolet light, with a wavelength 
of 0.347 um, in a KDP-crystal. The pulse duration 
is 25 nanoseconds and the maximum energy 4 mJ in the 
TEMg99 mode. A telescopic system (Ly and L3) is used 
to obtain a laser beam with a diameter of 30 mm. 
A mirror reflects the beam into the test section of 
the tunnel. In the walls of the plexiglass test 
section, two optical glass windows are inserted. 


irrotational flow without walls 


irrotational flow with walls 


06 08 1.0 1.2 14 1.6 18 20 


Surface Coordinate over Diameter, s/D 


The location of the body in the test section is 

such that the nose is illuminated by the laser beam 
over a length of about 20 mm, and the body contour 

is imaged on the hologram. A shutter is placed on 
the first window. The camera containing the holo- 
graphic plate is located close to the second window. 
Agfa-Gevaert Scientia Plates 8E56 and 8E75 with a 
resolution up to 3000 lines/mm were used as recording 
material. The ruby laser could also be used as a 
multiple switched laser. Two or three pulses with 


PULSED RUBY ral 
LASER v 


| 
KDP-CRYSTAL| UV-FILTER } 


Us 


GLASS WINDOW 


TUNNEL WALLS, 
\ 


GLASS WINDOW 


cog 
{ SSEREENS;| 
ISS =< 


CAMERA /HOLOGRAPHIC PLATE 


FIGURE 5. Schematic diagram of optical system for 
making holograms of cavitation or flow phenomena in 
test section of tunnel. 


Lo POLAROID FILTER 


Sea f alee a 


[MICROSCOPE 


STAGE WITH 
HOLOGRAM/ 


FIGURE 6. Schematic diagram of reconstruction 
set-up. 


pulse separations of 50 or 100 usec could be 
generated. This enabled multiple imaging of moving 
cavities on one hologram. 

Reconstruction of the holograms was made with a 
continuous-wave He-Ne gas laser (\ = 0.633 ym). A 
schematic diagram of the reconstruction set-up is 
shown in Figure 6. The diameter of the laser beam 
is enlarged by the lenses, L} and Lz. The intensity 
of the light can be adjusted by a polaroid filter. 
The hologram is placed on a stage, fitted with 
guides so that the hologram can be moved in two 
orthogonal directions. The movement of the stage 
is measured on vernier scales. The reconstructed 
image is studied with a microscope with a magnifi- 
Cation between 40x and 200x. 


Flow Visualization Technique 


A new technique had to be developed to visualize 
the boundary layer flow about the axisymmetric 
models. A description of the several methods in- 
vestigated is given by Van der Meulen (1976b). The 
ultimate method consisted of injecting a sodium 
chloride solution into the boundary layer from a 
hole located at the stagnation point of the model. 
The diameter of the hole is 0.08 mm. The sodium 
chloride solution has a slightly different index 

of refraction from the surrounding fluid. The 
light emitted from the pulse laser will be deflected 
and the deflections are recorded in the hologram. 
Optimum conditions for flow visualization are given 
by Van der Meulen and Raterink (1977). In the 
present study, the ratio of the injection velocity, 
Vi, to the velocity in the test section, Vo, was 
usually between 0.1 and 0.2. The sodium chloride 
concentration was 2 percent. At first, the fluid 
was injected with a hypodermic syringe, but later 
on, a plunger with a constant motion was used. 


Procedure 


The tests performed in the high speed tunnel com- 
prised flow visualization tests, cavitation tests 


437 


and cavitation inception measurements. Essentially, 
the flow visualization and cavitation tests consisted 
of making holograms at prescribed conditions. Prior 
to each series of tests the model was cleaned and 
the tunnel refilled. To adjust the air content, 

the water was passed through the deaeration circuit 
for a period of 1% h at a constant pressure in the 
deaeration tank. All tests were made at a constant 
air content, a, of about 5 cm?/2 (1 cm? of air per 
liter of water at STP corresponds to 1.325 ppm by 
weight). For each test the temperature of the 
tunnel water was measured to obtain the dynamic 
viscosity and the vapor pressure. The average 
value of the water temperature was 20°C. The flow 
visualization tests covered a velocity range of 2 
to 30 m/s. For the cavitation tests, the velocity 
ranged from 10 to 20 m/s. The effect of polymer 
additives on cavitation and cavitation inception 
was investigated by injecting a 500 ppm Polyox WSR- 
301 solution from the nose of the models. Polymer 
injection was provided by the Hughes Centurion-100 
pump unit. The holograms were made at the instant 
of maximum injection rate. The injection rate was 
such that the average value of Vi/Vo was 0.17. For 
the cavitation inception measurements, the velocity 
ranged from 10 to 24 m/s. Inception (or desinence) 
was observed visually. 


3. BOUNDARY LAYER STUDIES 
Newtonian Flow 


The holograms exhibited a distinct occurrence of 
laminar boundary layer separation on the hemispher- 
ical nose. The location of separation could be 
obtained quite accurately from the holograms. At 
this location the interference pattern usually 
showed a V-shape. This is shown in the photograph 
presented in Figure 7. This photograph also shows 
the laminar separation bubble itself and the 
subsequent transition to turbulence and reattachment 
of the separated shear layer. In the transition 
region, the flow is still visualized by the sodium 
chloride, but further downstream, where the turbu- 
lence becomes more developed, mixing of the sodium 
chloride prevents any further observations. The 
determination of the length and the maximum height 
of the laminar separation bubble from the holograms 
was somewhat complicated by the fact that the height 
of the bubble may show a maximum, as illustrated by 
case A in Figure 8, or that the outer flow line 
shows an inflexion point, as illustrated by case B 
in Figure 8. The location of separation for the 


FIGURE 7. Photograph showing laminar separation bubble and subsequent transition to turbulence on SST 
hemispherical nose. The flow is from left to right. At the position of separation the interference pattern 


shows a "V". YW, = 4 m/s. 


438 


A 


maximum 


Inflexion point 


FIGURE 8. 
on hemispherical nose (schematically) and definitions 
of length and maximum height of bubble. 


Observed shapes of laminar separation bubble 


hemispherical nose is given in Figure 9. In this 
figure the angular location of separation, yg, is 
plotted against the Reynolds number. Results on 
the length, L, the height, H, and the length to 
height ratio, L/H, of the separation bubble are 
presented in Figures 10, 11 and 12. Each data 
point refers to one hologram (values for the upper 
and lower side of the model are averaged). Most 
data points refer to the SST hemispherical nose, 

a few refer to the Teflon hemispherical nose. 

The present observations are in agreement with 
those obtained earlier by Arakeri (1973) and Arakeri 
and Acosta (1973). From Figure 9 it follows that 
the boundary layer separation angle is independent 
on the Reynolds number, which is consistent with 
theory (Schlichting, 1965). For the SST hemispher- 
ical nose the average value of yg is 85.43°. This 
value is claimed to be quite accurate. To compare 
this experimental value with the theoretically 
predicted one, laminar boundary layer calculations 
were made using the method derived by Thwaites 


(1949). With this method the parameter m is cal- 
culated, where m is defined as 
a2 dau 
ae oi wv) ds (2) 


and where 9 is the momentum thickness, U the velocity 
at the edge of the boundary layer, v the kinematic 
viscosity, and s the distance along the surface. 


iv) 
» ° 
o 90 T T 7, T T T T T T 
Te) 
[= st) 
4 Yo 
5) 86° NE 
ow 
c 4 
© 

fo} 
o ° ° 
w 86 a 
“ ¢ 20 @ ON 6) P.O, © fe) 
@ Ors @ oo, 8 oe | 
2 8 
= 84h 
a o SST 
3 e@ Teflon + 
3 82° rn ! et —————E eee 
o 02 04 06 o8 1 2 3eee4 


Reynolds Number x107> 


FIGURE 9. Boundary layer separation angle, Yc, 
as a function of Reynolds number for hemispherical 
nose. 


o SST 
04 @ Teflon J 


Length of Separation Bubble to Diameter, i/o 


02 04 06 O8 10 20 30 40 


Reynolds Number x1072 


FIGURE 10. Length of separation bubble to diameter, 
L/D, as a function of Reynolds number for hemispheri- 
cal nose. The solid lines refer to theoretical pre- 
dictions. 


Laminar boundary layer separation is said to occur 

for m = 0.09. The computations of yg were made 

with the accurate pressure distributions obtained 

earlier (Figure 3). For the actual case (with 

tunnel walls) yg was found to be 85.57°, and thus, 

in excellent agreement with the experimental value. 

The theoretical value of yg is hardly affected by 

the presence of the tunnel walls, since in the 

absence of tunnel walls we found yg = 85.53°. 

Arakeri (1973) found experimental and theoretical 

values of 87°. However, his computations were 

based on the experimental pressure distribution 

data by Rouse and McNown (1948), as shown in Figure 3. 
The length and the height of the separation 

bubbles decrease gradually with increasing Reynolds 

number. The variations in length and height for a 

given Reynolds number are partly due to the different 


0.05 STs T T Tiana Vemerl T T 

‘ ° o SST 

0.04 @ Teflon 4 
fo) 

003 =| 


Height of Separation Bubble to Diameter, H/o 


002 fo) 4 
8 9° 

O_o 
001 e@ @o = 

0 & 

YU © 
° 
° 

fo) [nee ee 


02 04 OG O8 2 3 4 


-5 
Reynolds Number x 10° ~ 
FIGURE 11. Height of separation bubble to diameter, 


H/D, as a function of Reynolds number for hemispheri- 
cal nose. 


20 - <j T sr * T T T 
oS o- 
o> ° © 
aes e 
& © o oO ‘ e ° 
» 8 @ 00 ° o@ 
tm = 10 0° 7 
=a fo} fe) ° ° 
O) 0 QO 
ag) {s 

2 
os 
~ 0 
(& 
= x 4 
Do 
th o SST 
So @ Teflon 


02 04 06 o8 1 2 3 4 
Reynolds Number x107° 


FIGURE 12. Length to height ratio of separation bubble, 
L/H, as a function of Reynolds number for hemispherical 
nose. 


appearances of the separation bubbles near transition, 
as illustrated in Figure 8. The length to height 
ratio of the separation bubbles (Figure 12) is not 
very dependent on the Reynolds number. For the SST 
hemispherical nose an average value of 10.8 is found. 
To compare experimental values of L with theoret- 
ical ones, it is necessary to calculate the location 
of transition on the separated shear layer. Recently, 
Van Ingen (1975) presented a calculation method for 
the laminar part of separation bubbles in which 
also the location of transition is predicted. The 
method is based on a solution of the Navier-Stokes 
equations, valid near the separation point. A 
relation is found for the separation streamline 
leaving the wall at an angle, 6. By using constant 
values of B, Oar and Msepr to be obtained experi- 
mentally, a formula is derived to calculate the 
length of the separation bubble. It is assumed 
that the separation streamline is straight and 
that the angle 6 is given by 


tan 6 = aL (3) 
Re, 


sep 


where REQsep is the Reynolds number based on the 
momentum thickness at separation, given by 


Re = (Gao) : (4) 


TABLE 1. Separation Streamline Angle 6 For SST 
Hemispherical Nose, Derived from Holograms. 


Re x 107° Regsep 8 B 

0.21 56 14.6° 14.6 
0.36 74 9.0° 11.8 
0.62 97 710° 11.9 
0.97 121 52° it 
1.35 144 4.4° Tied 
1.41 146 4.2° 10.8 
1.87 169 A390 WD7 
2.40 191 3.9° 13.0 
3.40 228 2.8° Died 


439 


The amplification factor, o is defined as 


a’ 
a 


a 
neutral 


where a/apeytra, 18 the ratio of the amplitude of 

a disturbance to its amplitude at neutral stability. 
Meo is the value of m at separation (Msep = 0.09). 
According to Dobbinga et al. (1972), B is usually 
between 15 and 20, but lower values are also found. 
To obtain B for the present case, the separation 
streamline angle, &, was derived from a series of 
holograms. The results are presented in Table 1. 
The average value of B is 12.0. 

With Van Ingen's method, the location of transi- 
tion has been calculated for oa = 7 and go, = 8, 
using Msep = 0.09 and B= 12. The results are 
plotted in Figure 10. It is found that most experi- 
mental data points lie between the two theoretical 
curves. The best fit would be obtained for og = 7.5. 
It should be noted that the present experimental 
data refer to the beginning of transition. Ina 
recent paper, Van Ingen (1976) attempted to corre- 
late the amplification factor with the turbulence 
level, Tu. For og = 7.5, predicting the beginning 
of transition, we find Tu = 0.15%. Although the 
turbulence level in the high speed tunnel has not 
been measured, it is possible to obtain an approx- 
imate value (without considering noise aspects) . 
Arakeri (1975a) measured the location of transition 
on a 1.5 caliber ogive in the axisymmetric test 
section of the CIT high speed water tunnel. The 
turbulence level in this tunnel was 0.2%. Recently, 
Arakeri (1977) performed similar measurements in 
the NSMB high speed water tunnel. The agreement 
between the transition data indicates that the 
turbulence level in both tunnels was approximately 
the same. Hence, the turbulence level in the NSMB 
tunnel may have been close to 0.2%, which is con- 
sistent with the value derived earlier. The above 
considerations on the turbulence level are, however, 
not confirmed by the measurements of Gates (1977), 
who found that the turbulence level had no effect 
whatsoever on the location of transition on a 
hemispherical nose. 

As shown in Figures 9 through 12, the appearance 
of the laminar separation bubble on the Teflon 
hemispherical nose is the same as for the SST 
hemispherical nose. From Figure 10 it is found 
that the higher surface roughness of the Teflon 
body has no effect on transition. Apparently, the 
amplification of disturbances mainly occurs down- 
stream of separation. 

The blunt nose exhibited a laminar boundary layer 
with normal transition to turbulence. Laminar flow 
separation did not occur. A photograph showing 
transition is presented in Figure 13. A plot of 
the transition data is given in Figure 14. Since 
the outflow of the sodium chloride solution from 
the nose of the model was in some cases quite 
unstable, the determination of the precise location 
of transition provided some difficulties, but an 
upper or lower bound could still be indicated. [In 
Figure 14 these data points are marked with an 
arrow. When the arrow is pointing upward the data 
point is considered to be the lower bound; when the 
arrow is pointing downward the data point is con- 
sidered to be the upper bound. Silberman et al. 
(1973) made laminar boundary layer calculations for 
a series of blunt noses having CPpin Values ranging 


FIGURE 13. Photograph showing transition (T) 


from 0.333 to 1.0. The calculations showed that 
none of the blunt noses were subjected to laminar 
separation. The present observations are in 
agreement with these theoretical predictions. 


Non-Newtonian Flow 


The influence of polymer additives on the boundary 
layer flow about the models was investigated by 
injecting a 500 ppm (parts per million by weight) 
Polyox WSR-301 solution from the nose of the models. 
To visualize the flow, the injection fluid contained 
2 percent sodium chloride. For the SST hemispher- 
ical nose, the holograms showed that laminar flow 
separation was no longer present. An example is 
given by the photograph presented in Figure 15. At 
or shortly downstream from the location where 
Newtonian flow separation occurred, transition from 
laminar to turbulent boundary layer flow is observed. 
From the holograms made in the velocity range 4 to 
20 m/s, it could be derived that transition to 
turbulence occurred close to the location of 
Newtonian flow separation. It was difficult, however, 
to indicate the precise location of transition. 


24 


20 


over Diameter, St/D 
52 
@ 


Streamwise Distance to Boundary Layer Transition 
fo} 
rs 


FIGURE 14. Streamwise distance to boundary 


layer transition over diameter, S,/D, asa 03 


function of Reynolds number for blunt nose. 


from laminar to turbulent boundary layer on blunt nose 
(Sp/D = 1.68). The flow is from left to right, W. = 8 m/s. 


Another important observation was that the sodium 
chloride was not completely mixed in the turbulent 
region, but was still able to show the existence of 
waves and streaks further downstream, till the end 
of the hologram. An example of this phenomenon 

has been given by Van der Meulen (1976b). For the 
Teflon hemispherical nose it was found that the 
influence of polymer additives on laminar flow 
separation was the same as for the SST hemispher- 
ical nose. Although the observations made with the 
blunt nose were somewhat obscured by the irregular 
outflow from the nose of the model, the main con- 
clusion to be derived from the holograms is that 
the polymer causes early transition to turbulence. 
The approximate locations of transition are plotted 
in Figure 14. 

The polymer concentration used during the above 
observations is rather high when compared to the 
most effective concentration for turbulent-flow 
friction reduction. From Figure 16, where the 
friction factor, f, for flow through a circular 
tube is given as a function of the Reynolds number, 
it can be derived that a Polyox WSR-301 concentration 
of about 20 ppm gives a maximum friction reduction. 
Additional holograms for the SST hemispherical nose 


O 2percent NaCl injection 
@ 2percent NaCl +500ppm Polyox WSR-301 injection 


06 O7 O8 1.0 15 20 30 


04 foe) 


Reynolds Number x 1075 


ee Imm =| 


FIGURE 15. Photograph showing boundary layer flow about SST hemispherical nose when a solution of 500 ppm 


Polyox WSR-301 is injected. The flow is from left to right. We 


were made at polymer concentrations of 100, 50, 

and 20 ppm. The injection rate was such that Vj/Vo 
= 0.2. The phenomena observed at these lower 
concentrations were the same as those found at 500 
ppm. Recently, Gates (1977) studied the influence 
of polymer additives on laminar flow separation at 
low injection rates, and was able to find inter- 
mediate stages of separation suppression. 

The study on the influence of polymer additives 
on laminar flow separation has been limited so far 
to the case where the polymer is present only in 
a thin layer adjacent to the body (the "inner part" 
of the boundary layer). To study the influence of 
polymer additives present in the "outer part" of 
the boundary layer, additional tests with the SST 
hemispherical nose were made in which the tunnel 
was filled with a 50 ppm Polyox WSR-301 solution. 
To prevent polymer degradation, the water speed in 
the test section was set at a low value of 4 m/s. 
Three different solutions were injected: a solution 


0.04 ey al LU T 
ae Water "| 
0.03 ae °885es + 
Peg, d 
fo 0° 2°00 00 ©0000 Soo go0a, | 
© 
® eo © © 00 0 00 | QrdQp000DmpDLD 
oO iy 
02 =| 
Ga So “ov, Op 20 0B °° | 
% 7 VOOo 00 0 Go 00 4 
@ ee a vv 4 
os 4 gv ea Ea Vo =| 
ve a v ] 
8 P chaos dt 8 Cowie | 
0 Foy o*," V ,dd4an | 
0 4) Bal 
o we? 6 Pag, 
av) esta! a We | 
5 FOS? ol — vy 
6) A. 
c 0.01 
F hee 
eO01ppm 010 ppm Cash 4 
202 v 20 
@e@05 @50 = 
f= 84 o1 4100 
e °02 B 200 a] 
a5 v 500 


6 10 15 20 25 30 
Reynolds Number x1072 


FIGURE 16. Friction factor of Polyox WSR-301 solu- 
tions in water as a function of Reynolds number, 
according to Van der Meulen (1974a). 


= 4nm/s. 


of 2 percent NaCl, a solution of 2 percent Nacl + 
50 ppm Polyox and a solution of 2 percent NaCl + 
500 ppm Polyox. The injection velocity Vj was 

0.8 m/s (Vi/Vo = 0-2). Photographs showing the 
boundary layer flow are presented in Figure 17. 

For comparison a photograph is included showing 

the influence of polymer injection when the tunnel 
is filled with water (Figure 17a). When a 2 per- 
cent NaCl solution is injected (Figure 17b), the 
boundary layer first shows a tendency to become 
unstable but further downstream the instabilities 
are suppressed and the boundary layer is laminar 
again. When a 2 percent NaCl + 50 ppm Polyox 
solution is injected (Figure 17c), the boundary 
layer first shows a slight tendency to become 
unstable, but further downstream the boundary layer 
is laminar. When a 2 percent NaCl + 500 ppm Polyox 
solution is injected (Figure 17d), the boundary 
layer remains completely laminar, till the end of 
the hologram. The conclusions to be derived from 
these observations are that the presence of the 
polymer in the "inner part" of the boundary layer 
leads to destabilization, whereas the presence of 
the polymer in the "outer part" of the boundary 
layer leads to stabilization, and the latter effect 
is predominant. In all cases considered (Figure 
17), laminar flow separation is suppressed. 

An explanation of the various phenomena observed 
can, as yet, not be given. Apparently, some of the 
phenomena are in agreement with those reported 
elsewhere, others may not have been observed before. 
This is mainly due to the fact that numerous studies 
have been made on drag reduction in turbulent flow, 
but only a few were made on the influence of polymer 
additives on laminar flow. In studying laminar 
flow around circular cylinders, James and Acosta 
(1970) found that the streamline patterns with 
dilute polymer solutions were significantly different 
from those with Newtonian fluids because of visco- 
elastic effects. These effects may also play a 
dominant role in eliminating flow separation in 
those cases that the boundary layer remains laminar. 
In those cases where the boundary layer becomes 
turbulent due to the presence of the polymer in the 
"inner part" of the boundary layer, it is still 
questionable whether flow separation is eliminated 
by early turbulence by viscoelastic effects, or by 
a combination of these. The occurrence of early 
turbulence as found in the present study and reported 
before [Van der Meulen (1976a, 1976b), Gates (1977) ] 
is consistent with the findings of others. According 
to Lumley (1973), polymer solutions producing drag 


(a) 


> ; hee (d) 


1mm Ress 
FIGURE 17. Photographs showing boundary layer flow about SST hemispherical nose. The flow is from right to 
left. V_ = 4 m/s. (a) Injection of 50 ppm Polyox in water. 


(b) Injection of water in 50 ppm Polyox. (c) Injec- 


tion of°50 ppm Polyox in 50 ppm Polyox. (d) Injection of 500 ppm Polyox in 50 ppm Polyox. 


reduction display a positive Weissenberg effect for 
which destabilization is predicted analytically. 
Destabilization is also predicted by the numerical 
analysis of Kiimmerer (1976) on the stability of 
boundary layers in an idealized viscoelastic fluid. 
Experiments by Forame et al. (1972) and Paterson 
and Abernathy (1972) also suggest destabilization. 
On the other hand, Castro and Squire (1967) and 
White and McEligot (1970) found that polymer solu- 
tions in water cause a delay in transition to 
turbulence. According to Lumley (1973), drag- 
reducing polymers tend to increase the thickness 
of the viscous sublayer. Experimental evidence 


08 T T T T T T T T 
) ° O. 2 
° @ e 
e eo f® ee 
© O6r oe @ ee - 
« 5 °o (Oo 
: 0 @ 00 
3 
e } 
c & 
oc 4 A & 
= 4 4 4 4 
i] a 4 a & a a a 
5 04- 5 : =| 
” inception 3 
=43cm-/1 

o e pieces S S, / 

| & inception, 500ppm Polyox WSR —301 injection 3 al] 

a=47¢em>/1 
& desinence,500ppm Polyox WSR -301 injection 
02 . 1 4 L Jt 1 1 4 


08 12 16 20 24 


Reynolds Number x 10-2 


IGU Cavitation inception and desinence number 
as a function of Reynolds number for SST hemispherical 
ose with and without polymer injection. 


for this phenomenon has been provided by Rudd (1972), 
who measured velocity profiles in a polymer solution 
by using a laser dopplermeter. By examining the 
expansion behavior of isolated polymer molecules 

in a flow field, Lumley (1973) postulated a mech- 
anism which predicted a decreased intensity of 
small-scale turbulence in the buffer layer and which 
also predicted that, in the maximum drag reduction 
regime, the turbulence should consist primarily of 
larger eddies. The present observations of waves 
and streaks along the surfaces of the models seem 

in agreement with the above predictions. They also 
agree with the observations made by Hoyt et al. 
(1974) on the structure of jets of polymer solution 
discharged in air. 


4. CAVITATION STUDIES 
Inception 


Cavitation inception data for the SST hemispherical 
nose are plotted in Figure 18. Inception was 
measured by gradually lowering the pressure until 
the first appearance of cavitation was observed. 
Desinence was measured by starting from developed 
cavitation and gradually raising the pressure until 
cavitation just disappeared. The type of cavitation 
mostly observed at inception was sheet cavitation. 
Also plotted in Figure 18 are cavitation inception 
data when a 500 ppm Polyox WSR-301 solution was 
injected from the nose of the model. The type of 
cavitation observed in this case was travelling 
bubble cavitation. Cavitation inception data for 
the Teflon hemispherical nose are plotted in Figure 


Llama hae aaa Ta T T T oo ] 
O Inception 
Aa=44c mA | 
e @ desinence | 
: | 
© ee 
b é ° é | 
> 2 } 
6 ee e 
rs) e e el 
E | 
3 | 
2 °o 4 
5 ie) 
S © ro) OO 
= ° 
5 O68 ©) 4 
C ° 9 6 | 
(S) ° | 
Os ° 
1 
O04 1 1 1 i 1 1 1 i 
os 12 16 20 24 


Reynolds Number x 10-2 


FIGURE 19. Cavitation inception and desinence number 
as a function of Reynolds number for Teflon hem- 
ispherical nose. 


19. The type of cavitation observed at inception 
was spot cavitation. The spots were usually located 
between the pressure minimum (y = 78°) and the 
transition of hemisphere and cylinder (y = 90°). 

The most striking differences between the inception 
data for both models are: (a) the inception data 
for the Teflon model are much higher than for the 
SST model and (b) the Teflon model exhibits a strong 
cavitation hysteresis [Holl and Treaster (1966) ] 
whereas the SST model exhibits no hysteresis. Such 
observations have been reported before by Reed 
(1969), Gupta (1969), and Van der Meulen (1971). 
Since the viscous flow behavior of the Teflon model 
is the same as for the SST model (see Section 3), 
the above differences can only be explained by 
surface effects. Teflon is a porous material and 
has a high contact angle. Both properties are 
essential features of the Harvey nucleus [Harvey 

et al. (1944) ]. Hence, the Teflon surface acts as 

a host for surface nuclei, from which (gaseous) 
cavitation is initiated. The mechanism most probably 


08 
b 
co O6 
© a 
€ inception (or desinence) 
=) 
z 
6 
2 
oO 
2 
204 
oO 
oO Cp (Irrotational flow) 

T 
4 
02 er mes | ES 1 “=| 
10 14 18 22 
Reynolds Number x1072 

FIGURE 20. Comparison of cavitation inception (or 


desinence) number with pressure coefficient at sepa- 
ration, Cp_, and at transition, Cpr for SST hem- 
ispherical’ nose. 


443 


involved with inception on the SST hemispherical 

nose has been described by Arakeri (1973). Inception 
takes place in the transition and reattachment 

region of the separation bubbles, where high pressure 
fluctuations occur [Arakeri (1975a)]. The nuclei 
may either originate from the surface (Arakeri) 

or from the stream where they become trapped in. 

the strong vortices occurring in the reattachment 
region. 

When o; (or og) for the SST hemispherical nose 
is to be compared with the pressure coefficient, 
several problems arise. The most obvious pressure 
to compare 0; with would be the pressure coefficient 
at transition, Cp,, since the onset of cavitation 
takes place at the location of transition. Accord- 
ing to Arakeri (1973), however, the important 
pressure coefficient to compare 0; with would be 
the pressure coefficient at separation, Cpg- This 
opinion is probably based on the assumption that 
the pressure within the separation bubble is con- 
stant (and thus Cpg = Cp) but, according to Van 
Ingen (1975), this is a good approximation only 
at low values of Re. A mean curve of the present 
inception (and desinence) data is plotted in Figure 
20. Also plotted are Cp, = 0.76 and Cp, for 
irrotational flow, derived from Figures 3 and 10 
(with og, = 7.5). The real (or viscous) values of 
Cin are unknown and should be obtained from pressure 
measurements. It can be estimated that the real 
values of Cp,, are considerably larger than those 
for irrotational flow, but still smaller than Cpe: 
Thus it would seem that oj; (or oq) can be correlated 
with the real value of Coe eEnnthatecase waktacan 
be argued that the peak pressure fluctuations, 
measured by Arakeri (1975a), are creating the 
negative pressures necessary to overcome the sta- 
bilizing pressure in stream nuclei, caused by the 
surface tension. 

Cavitation inception data for the blunt nose are 
plotted in Figure 21. Also plotted are inception 
data with polymer injection. At inception, a 
region of travelling bubbles was observed. The 
approximate location of this region was x/D = 0.2 
- 1.0. In Section 4, a further analysis will be 
given of the type of cavitation occurring. The 
inception data show that the o,;-and og-values are 
almost identical and nearly constant (Sia = 0.46, 
in the absence of polymers). When oj is to be 
compared with a suitable pressure coefficient, the 


ean T T T T T T T 
b 
¢ 0.6F a) 
F 
€ 4 
2 e A a A | 
ee = a 
2 oo8 e ag A aA & 
& ° @o ® On Sie A @ ) 
3 ° ° o Me 
5 Os), © inception 7 
ry) F X= 4.8 cm3/1 
@ desinence 
4 inception, 500ppm Polyox WSR-—301 injection J] 
- aly ob a=51 em3/L 
& desinence, 500ppm Polyox WSR -—301 injection 
St Sa ete FE ak wi shee St Nl 
08 U3 1.6 20 2.4 


Reynolds Number x 10-2 


FIGURE 21. Cavitation inception and desinence number 
as a function of Reynolds number for blunt nose with 
and without polymer injection. 


444 


06 = aaa =e T T 
| 

04 “/_inception (or desinence) =i} 
b 
C 
© 
Qa 
E 
=) 
z 
c 
° 
~ O02 
i) 
= 
> 
% 
16) 

Oo 

10 
Reynolds Number x 107° 

FIGURE 22. Comparison of cavitation inception (or 


desinence) number with pressure coefficient at tran- 
sition, Cpe for blunt nose. 


best choice would seem the pressure coefficient at 
the location of cavitation inception. However, 
this location can not precisely be indicated. For 
bodies with attached boundary layers, Arakeri (1973) 
suggested correlating 6; with the pressure coeffi- 
cient at transition, Cp,. For a 1.5 caliber ogive 
a close correlation was found between measured 
values of og and computed values of Cp,. The same 
comparison can be made for the blunt nose. In 
Figure 22, o5 q and Cp,, derived from Figures 4 and 
14, are plottéd agesinee the Reynolds number. In 
this case it may be assumed that the real (or viscous) 
values of Cp,, are the same as those for irrotational 
flow. It is evident from Figure 22 that oj (or 
Og) cannot be correlated with Cp,,. The location 
where Cp, = 0.46 (= 53a) is well in the laminar 
region of the boundary layer for the Reynolds numbers 
considered. 

The influence of polymer additives on cavitation 
inception is a rather new phenomenon. Darner (1970) 


investigated the addition of polymers to water on 
Ellis 
reported on the effect of polymer 


acoustically induced cavitation inception. 
et al. (1970) 


Surface Tension S, dyne/cm 


lo) 100 200 300 400 500 600 


Polyox WSR - 301 Concentration , ppm 

FIGURE 23. Surface tension as a function of Polyox 
WSRT301 concentration in water, as measured in surface 
tensionmeter. 


solutions on flow-generated cavitation inception. 
The effect of the polymer was to suppress cavitation 
inception. An explanation for the effect could, 

as yet, not be given. Ting and Ellis (1974) studied 
the growth of individual gas bubbles in dilute 
polymer solutions but concluded that the polymers 
hardly affected bubble growth. From Figure 23 it 

is found that the surface tension is slightly 
reduced by small additions of Polyox WSR-301, but 
according to Hoyt (1973) this effect should cause 
earlier cavitation instead of cavitation suppression. 
From Figure 18, a considerable effect on oj and 

Og is found when a 500 ppm Polyox solutuion is 
injected from the nose of the SST hemispherical 
model. For Re above 1.2 x 10°, the reduction 
amounts to 30 percent. For the mean value of oj 

and 0g we have Siva = 0.445. The o;- and og-values 
are independent of Re. From Figure 21 it is found 
that oj; and og are hardly affected by the injection 
of a 500 ppm Polyox solution from the nose of the 
blunt model. For Re above 1.2 * 10°, the mean 

value of oj and og in the absence of polymers is 
i,q = 0-45. Hence, inception on the SST hemispher- 
ical nose with polymer injection takes place at the 
same cavitation number as inception on the blunt 
nose in the absence of polymers. 

As found in Section 3, the influence of the 
polymer is to suppress the laminar boundary layer 
separation on the hemispherical nose. Hence, the 
strong pressure fluctuations, occurring at the 
position of transition and reattachment of the 
separated shear layer [Arakeri (1975a) ] and being 
the principal mechanism for cavitation inception, 
are eliminated and cavitation will start at a much 
lower cavitation number. The flow visualization 
studies described in Section 3 do not only explain 
the suppression of cavitation inception by polymer 
injection, but also by having a polymer ocean 
[Ellis et al. (1970)]. Earlier studies by Van der 
Meulen (1973, 1974b) showed that polymer injection 
had hardly any effect on cavitation inception on a 
Teflon hemispherical nose. The reason for this 
finding is clear now, since cavitation inception on 
a Teflon hemispherical nose is related to surface 
effects and not to viscous effects. 


Appearance on Hemispherical Models 


The appearance of cavitation on the SST hemispher- 
ical nose is closely related to the occurrence of 
laminar boundary layer separation. Arakeri (1973) 
showed that cavitation bubbles are first observed 
at the location of transition and reattachment of 
the separated shear layer. This type of cavitation 
is usually called bubble cavitation. An example 

is shown in Figure 24a. The larger bubbles at the 
location of transition are preceded by smaller ones 
which, according to Arakeri (1973), are travelling 
upstream with the reverse flow in the separated 
region. With a reduction in o, the larger bubbles 
create a single cavity as shown in Figure 24b. 

With a further reduction in o, the cavity is filling 
the separated region, and a smooth attached cavity 
is observed (Figure 24c). This type of cavitation 
is usually called sheet cavitation. When o is 
further reduced, the length and the height of the 
cavity extend, but the first part of the cavity 
remains smooth (Figure 24d, e). By analyzing 
double exposure holograms made of developed cavita- 
tion, it could be established that the first smooth 
part of the cavity is stable. 


Smm 


The appearance of cavitation on the Teflon 
hemispherical nose is closely related to the presence 


of weak spots on the surface. Discrete cavities 
originate from points located on the hemisphere. 
The cavities develop cone-shaped in the downstream 
direction. The first part of the cavity surface 
is smooth; the cavity leaves the wall at a very 
small angle. Some of these features can be observed 
on the photographs presented in Figure 25. The 
cavitation separation angle Ycg for both hemispher- 
ical models is plotted in Figure 26. For the Teflon 
model it is found that the cavities start upstream 
of the minimum pressure point (Yes < YPmin)! when 
0 is sufficiently low. For the SST model it is 
found that the cavities always start downstream 
of the minimum pressure point (Yoo < Wien Jo. Yes 
is both a function of o and Re. For a given Re, 
Yes decreases with decreasing o and for a given 
5, Yes decreases with increasing Re. These tenden- 
cies for the SST model are in agreement with the 
observations by Arakeri (1975b). 

The shape of the cavity nose on the SST model 
has been analyzed further. A schematic drawing of 
the geometry of the cavity nose is presented in 
Figure 27. From a detailed study of the holograms 
it could be established that the cavity nose was 
circularly shaped. It was found that the nose 
angle 8 varied between 70° and 120°, but was 
independent of o or Re. An average value of 90° 
was obtained from 28 cavity noses. Since the cavity 
nose is immersed in the separation bubble and the 
flow comes to a standstill near the cavity nose, 


445 


FIGURE 24. Photographs showing progressive 
development of cavitation on SST hem- 
ispherical nose. The flow is from left to 
right. V, = 13.2 m/s. (a) o = 0.60; 

(b) o = 0.59; (c) o = 0.56; (d) o = 0.47; 
(e) o = 0.39. 


it is to be expected that the nose angle equals 

the contact angle for the present liguid-gas-solid 
system. This is confirmed by the fact that, accord- 
ing to Adamson (1966), the contact angle for a 
water-air-steel system is 70°-90°. The nose radius 
ry was independent of o but, as shown in Figure 28, 
the radius decreases with increasing Re. The length 
of the sheet cavity (the smooth part preceding the 
developed cavity) is more or less independent of 

oO but decreases with increasing Re. In Table 2, 
mean values of Lo¢/D are compared with corresponding 
values of L/D, obtained from Figure 10 (with 

5, = 7-5). From this table it can be concluded 

that transition to turbulence on the cavity surface 
is closely related to transition to turbulence on 
the fully wetted separated shear layer. The shape 
of the developed cavity is determined by the total 
length to maximum height ratio of the cavity, 

L¢/Her (in most cases the cavity reached its maxi- 
mum height close to the trailing edge of the cavity). 
Values of this ratio are given in Figure 29. The 
mean value of Lc/He is 10.2. Since the mean value 
of the length to height ratio of the separation 
bubble is 10.8, it may be concluded that the shape 
of the developed cavity appearing on the SST hemis-— 
pherical nose is strongly governed by the shape of 
the separation bubble. 

With polymer injection, the cavities on the SST 
hemispherical nose are either attached or may show 
the appearance of travelling bubbles, resembling 
the type of cavitation observed on the blunt nose. 
Details are given by Van der Meulen (1976b). 


446 


FIGURE 25. Photographs showing progres- 

sive development of cavitation on Teflon 

hemispherical nose. The flow is from left 
to right. Ve = 13.2 m/s. (a) o = 0.96; 

(b) o = 0.63; (c) o = 0.40. 


10 T T T T 
Teflon Pp Pp 
— Re=134x 10° 
O9fF : p P 7 
SST S 
Oo Re =094 x102 Sp. =78.2 
os @ Re =127 x102 aes ea 4 
5 
b d Re=154 x102 
is a Re =2.03 x 105 ve) 
rs 
E O7P 4 
=) 
2 - 
c p p A 
) L 4 
. 0.6 aN 
= @® wo 
S pP pP ° ° 
oO 
6) m 44 00 J 
0.57 ey 6 
foe} 
ay A 
0.4 p p 00 J 
p Pp 
03 ——— He 1 
65° 70° 75— 80° 85° 90° 95 


Cavitation Separation Angle Ocs 


FIGURE 26. Cavitation separation angle, Ycs, as a 
function of cavitation number and Reynolds number for 
hemispherical nose. 


water 
3 VN cavity me 
ip \ 
Sas ELLIE 7 
TL, Wate TA VSAM TSE Ut lof HA 7 
model 
JRE 7 schematic diagram of cavity nose on SST 


hemispherical model. 


fa) 
” 
= 
° 
a 
xo 
& 
vo o 
ome 
2.9 
a 
> 
SS (5 
> © 
o > 
oO 0 
FIGURE 28. 


a function 
model. 


0.606 
re) fo) 
O0.004F- fo) | 
0.002} © 4 
oO LL} 


08 1.2 1.6 20 24 
Reynolds Number x 107° 


Cavity nose radius over diameter, r/D, as 
of Reynolds number for SST hemispherical 


065 T T = T T T = 
a 5 
o Re=094 x10 a a 
= 5 
e@ Re=1.27 x10 aura | 
O60 
[ a Re=154x 10° iy & 
A Re = 2.03 x 10° 
eo e fe} 
sal 4 
055 ° ° 
6 
& 
ve) ray 
E osob ° a A 4 
=} 
Zz 
{= 
3 e e 
S 
o 
= O45 fo) ° 5] 
> 
oO 
oO 
Aa 
is 8 
O40-r (o) ©) | 
e e 
035 : ss = = 
(0) 2 4 6 8 10 12 14 
Length to Height Ratio of Cavity, Lc /He 
FIGURE 29. Total length to maximum height ratio of 


cavity, Lo/HAe as a function of cavitation number and 
Reynolds n 


Ser for SST hemispherical nose. 


TABLE 2. Length of Sheet Cavity Over Diameter, 
Lsc/D, and length of Separation Bubble Over 
Diameter, L/D, for SST Hemispherical Nose. 


Re x 107° Sc L/D 

0.94 0.156 0.124 
1627 0.124 0.096 
1.54 0.070 0.084 
DNOS 0.074 0.068 


Appearance on Blunt Model 


The type of cavitation occurring on the blunt nose 
is typically travelling bubble cavitation. An 
example is shown in Figure 30a (o = 0.33). When 
o is reduced, a single transient cavity may develop, 
as shown in Figure 30b (o = 0.28). The transient 
character of the cavities occurring on the blunt 
nose is clearly observed in the photographs taken 
from multiple exposure holograms. Figure 31 shows 
a photograph taken from a hologram, where three 
pulses were generated by the ruby laser with pulse 
separations of 50 usec and 100 usec respectively. 
The flow is from right to left. The picture shows 
the growth of a cavity near the nose of the model. 
The cavity is attached to the model and its shape 
is a spherical segment. The cavity grows (its 
radius increases) and, at the same time, travels 
along the surface with a velocity slightly below 
that of the surrounding fluid. When the cavity 
reaches a certain height, its shape becomes more 
like an attached bubble, as shown in Figure 32. 
In this figure, the flow is from left to right. 
The attached bubble hardly grows, travels along 
the surface, and finally collapses. 

The streamwise distance to cavitation separation 


on the blunt nose obtained from a series of holograms 


taken at various values of o and Re, is plotted in 
Figure 33. Also plotted are data points where no 
cavitation was observed in the hologram on either 
one or both sides of the model. It is found that 


| x/p =0247 


| x/o=0267 


FIGURE 30. Photographs showing cavitation on blunt nose. 
(2) @ = Oo33p (Gey) @ = Wolo 


447 


the streamwise distance to cavitation separation 
decreases with increasing Re (apart from the scatter, 
typical for travelling bubble cavitation). For 

Re = 2.08 x 10°, cavitation separation is located 

at a short distance from the pressure minimum 
[(s/D) pis, = 0-371. 

The Observations of the cavity growth as 
represented in Figure 31, enables a comparison with 
theory. Plesset (1949) analyzed experimental 
observations by Knapp and Hollander (1948) and 
compared the growth and collapse of bubbles on a 
1.5 caliber ogive with the equation of motion for 
a bubble. The agreement was quite satisfactory. 
Recently, Persson (1975) introduced some refinements 
in the comparison. The present analysis is based 
on the so-called Rayleigh-Plesset equation according 
to Hsieh (1965). For a vapor bubble, the motion 
of the bubble wall is given by the equation 


0 ae 25 _ 4uk 
OPIS teh) (6) 


where p is the liquid density, R the instantaneous 
bubble radius, Py, the vapor pressure, P the instan- 
taneous ambient pressure, S the surface tension, 
and uw the dynamic viscosity. The dots indicate 
differentiation with respect to time t. The 
multiple exposure hologram (Figure 31) provided 
data on Ro(to), Ri (tot50us), and R2 (to+150us), 
whereas P(t) could be derived from Figures 31 and 
4. Equation (6) was solved numerically to obtain 

a theoretical value of Ro. The results of the 
computations are given in Table 3. To compare the 
significance of the right-hand side terms of Eq. 
(6), numerical values of these terms are presented 
in Table 4. The main conclusion to be derived from 
Table 3 is that the experimentally observed growth 
of the cavity on the blunt nose is fairly well 
represented by the Rayleigh-Plesset equation of 
motion. This is mainly due to the fact that the 
blunt nose does not exhibit laminar flow separation 
and viscous effects seem to be small. 

The appearance of developed cavitation on the 
blunt nose with polymer injection was essentially 
the same as that without polymers. Details are 
given by Van der Meulen (1976b). 


f= 5mm 


(b) 


The flow is from left to right. We = 218) m/si. 


448 


FIGURE 31. Photograph of multiple exposure hologram showing three stages of cavity growth near nose of blunt 
model. The pulse separations are At) = 50 usec and Atp= 100 usec. The flow is from right to left. Ye = 10 m/s; 
1 = 0.31. The radii of the growing cavity are indicated on the lower figure. 


x/D=0543 = Imm =| 


FIGURE 32. Photograph of multiple exposure hologram showing three stages of travelling bubble along blunt nose. 
The time separations are: At, = 50 c and At2 = 100 usec. The flow is from left to right. VA = 10 m/s; 


0.50 + T a T =T T T T 
o Re =094x10° 
(s/D)p =037 @ Re =1.22 x10° 
045+ min is 4| 
a Re=155 x10 
A a 
A Re = 2.08x10° 
fe} 
oaol J 
b 
e i Ir 
3 6 a 
L a 4 
5 O35+ Q 
e A Ae 
5 A A 
2} fe) fe) 
= o30} e a a 44 e | 
> 
3 fats 
c 
© 
o25s- 2 4 
5S 
oO 
8 
fe} 
c 
Oro —— —— fin 1 r i 1 —_ 
fo) 02 04 06 08 10 12 14 16 


Streamwise Distance to Cavitation Separation 
over Diameter, S¢/D 


FIGURE 33. Streamwise distance to cavitation separa- 
tion over diameter, s_/D, as a function of cavitation 
number and Reynolds number for blunt nose. Also plotted 
are some data points where no cavitation was observed 
on one or both sides of the model. 


5. CONCLUSIONS 


The application of in-line holography and injection 
of a 2 percent sodium chloride solution from the 
nose of the axisymmetric bodies are useful methods 
to visualize the boundary layer and to obtain 
detailed information on boundary layer phenomena 
and cavitation patterns. 

Laminar boundary layer separation and transition 
to turbulence of the separated shear layer on the 
hemispherical nose can be predicted quite accurately 
by existing approximate calculation methods. 

Cavitation on axisymmetric bodies may be strongly 
influenced by boundary layer effects. For the SST 
hemispherical nose, inception and appearance of 
cavitation are both related to the location and 
appearance of the separation bubble. For the 
blunt nose, however, cavitation is apparently more 
related to nuclei effects than to viscous effects. 
The type of cavitation occuring in this case is 
travelling bubble cavitation. The growth of a 
cavity on the blunt nose is adequately described 
by the Rayleigh-Plesset equation of motion for a 
cavitation bubble. 


TABLE 3. Theoretical (R) and Experimental (R 
Values of Bubble Radius for Cavity Growth on 
Blunt Nose (Figure 31). 


exp) 


R R R R 


D exp 
t m/sec m/sec mm mm 
ee -5710 259)3} 0.84 0.84 
te so 50 we -5290 2.66 0.98 0.98 
-4540 Qos Le 23} 1.28 


449 


TABLE 4. Influence of Vapor Pressure, Py, Liquid 
Pressure, P, Surface Tension Pressure, 2 S/R, and 
Viscosity Pressure, 4y R/R, on Cavity Growth on 
Blunt Nose (Figure 31). 


Py P DESVAR 4u R/R 
t N/m? N/m? N/m? N/m? 
om 1940 -6320 170 15 
t + 50 us 1940 -3590 150 12 
t + 150 us 1940 + 330 120 8 


Surface effects on the Teflon hemispherical nose 
play a dominant role in both inception and appearance 
of cavitation. 

The presence of polymers in the "inner part" of 
the boundary layer on the SST hemispherical nose 
leads to destabilization, whereas the presence of 
the polymer in the "outer part" of the boundary 
layer leads to stabilization, and the latter effect 
is predominant. For all cases considered, laminar 
boundary layer separation is suppressed. 

Since the influence of polymer additives is to 
suppress laminar boundary layer separation on the 
hemispherical nose, the strong pressure fluctuations, 
occurring at the position of transition and reattach- 
ment of the separated shear layer and being the 
principal mechanism for cavitation inception, are 
eliminated and cavitation will start at much lower 
pressures. As a consequence, the cavitation charac- 
teristics of the SST hemispherical nose with polymer 
injection are approximately the same as those of 
the blunt nose without polymer injection. 


NOTATION 
B Constant in Equation (3) 
Cp Pressure coefficient 
CPmin Minimum pressure coefficient 
CPs Pressure coefficient at separation 
Cpm Pressure coefficient at transition 
D Model diameter 
H Height of separation bubble 
He Height of cavity 
L Length of separation bubble 
Le Length of cavity 
Lsc Length of sheet cavity 
P Static pressure 
Po Free stream static pressure 
Pmin Minimum static pressure 
2 Vapor pressure 
R Bubble radius 
Re Reynolds number, VoD/v 
REQ sap Equation (4) 
iS Surface tension 
Tu Turbulence level 
U Velocity at edge of boundary layer 
Vo Free stream velocity 
Vi Injection velocity 
a Amplitude of disturbance 
Aneutral Amplitude of disturbance at neutral 
stability 
if Friction factor 
m Equation (2) 


ie Nose radius of cavity 


450 


s Surface coordinate 

Sc Streamwise distance to cavitation 
separation 

Sp Streamwise distance to boundary layer 
transition 

x Axial coordinate 

a Air content 

8 Nose angle of cavity 

Y Angular coordinate 

Ys Boundary layer separation angle 

Yes Cavitation separation angle 

6 Angle at which separation streamline 


leaves wall 

Momentum thickness 

Dynamic viscosity 

Kinematic viscosity 

Liquid density 

Cavitation number, (Po-Py) /40Vo2 
Incipient cavitation number 
Desinent cavitation number 
Amplification factor 

Velocity potential 


ab 


ar pre ne we ee 


REFERENCES 


Acosta, A. J. (1974). Cavitation and fluid machin- 
ery. Cavitation Conf., Edinburgh, Scotland. 

Adamson, A. W. (1966). Physical chemistry of 
surfaces. Interscience Publishers, New York. 

Arakeri, V. H. (1973). Viscous effects in inception 
and development of cavitation on axisymmetric 
bodies. Cal. Inst. Tech. Rep. E-183-1. 

Arakeri, V. H. (1975a). A note on the transition 
observations on an axisymmetric body and some 
related fluctuating wall pressure measurements. 
J. Fluids Engng., Trans. A.S.M.E. 97, 82. 

Arakeri, V. H. (1975b). Viscous effects on the 
position of cavitation separation from smooth 
bodies. J. Fluid Mech. 68, 779. 

Arakeri, V. H. (1977). Experimental investigation 
of laminar boundary layer stability with surface 
heating and cooling. N.S.M.B. Rep. 01054-2-CT. 

Arakeri, V. H., and A. J. Acosta (1973). Viscous 
effects in the inception of cavitation on 
axisymmetric bodies. J. Fluids Engng., Trans. 
Me Soicitie G5, Sale) 

Arndt, R. E. A., and A. T. Ippen (1968). Rough 
surface effects on cavitation inception. J. 
Basic Engng., Trans. A.S.M.E. 90, 249. 

Castro, W., and W. Squire (1967). The effect of 
polymer additives on transition in pipe flow. 
Appl. Scientific Res. 18, 81. 

Darner, C. L. (1970). Sonic cavitation in water. 
N. L. R. Rep. 7131. 

Dobbinga, E., J. L. van Ingen, and J. W. Kooi (1972). 
Some research on two-dimensional laminar separa- 
tion bubbles. AGARD CP-102, Lisbon. 

Ellis, A. T., J. G. Waugh, and R. Y. Ting (1970). 
Cavitation suppression and stress effects in 
high-speed flows of water with dilute macro- 
molecule additives. J. Basic Engng., Trans. 
AOS-MoE. 927 405). 

Forame, P. C., R. J. Hansen, and R. C. Little (1972). 
Observations of early turbulence in the pipe 
flow of drag reducing polymer solutions. 

A LeCiiek . JOULNAL LG m2 Sis 

Gates, E. M. (1977). The influence of freestream 
turbulence freestream nuclei populations and a 
drag-reducing polymer on cavitations inception 
on two axisymmetric bodies. Cal. Inst. Tech. 
Rep. E-183-2. 


Gupta, S. K. (1969). The influence of porosity 
and contact angle on incipient and desinent 
cavitation. M.S. thesis, Pennsylvania State 
University. 

Harvey, E. N., D. K. Barnes, W. D. McElroy, A. H. 
Whiteley, D. C. Pease, and K. W. Cooper (1944). 
Bubble formation in animals. J. Cellular and 
Comparative Physiology 24, 1. 

Holl, J. W. (1960). The inception of cavitation 


on isolated surface irregularities. J. Basic 
Engng., Trans. A.S.M.E. 82, 169°. 

Holl, J. W., and A. L. Treaster (1966). Cavitation 
hysteresis. J. Basic Engng., Trans. A.S.M.E. 
88, 199. 


Hoyt, J. W. (1973). Jet cavitation in polymer 
solutions. A.S.M.E. Cavitation and Polyphase 
Flow Forum, 44. 

Hoyt, J. W., J. J. Taylor, and C. D. Runge (1974). 
The structure of jets of water and polymer 
solution in air. J. Fluid Mech. 63, 635. 

Hsieh, D. Y. (1965). Some analytical aspects of 
bubble dynamics. J. Basic Engng., Trans. 
A-S.M.E. 87, 991: 

James, D. F., and A. J. Acosta (1970). The laminar 
flow of dilute polymer solutions around circular 
cylinders. J. Fluid Mech. 42, 269. 

Knapp, R. T., and A. Hollander (1948). Laboratory 
investigations of the mechanism of cavitation. 
TRANS) VAvsS). Mare Oya Os 

Kiimmerer, H. (1976). Numerische Untersuchungen 
zur Stabilitat laminarer Grenzschichten in 
viskoelastischen Fliissigkeiten. Rheol. Acta 15, 
579 

Labrujére, Th.E. (1976). Berekening van de 
wandinvloed voor twee lichamen in een water- 
tunnel. National Aerospace Laboratory Rep. 

NLR TR 76017 L. 

Labrujére, Th.E., and J. van der Vooren (1974). 
Finite element calculation of axisymmetric sub- 
critical compressible flow. National Aerospace 
Laboratory Rep. TR 74162 U. 

Lumley, J. L. (1973). Drag reduction in turbulent 
flow by polymer additives. Macromolecular 
Reviews 7, 263. 

Paterson, R. W., and F. H. Abernathy (1972). Trans- 
ition to turbulence in pipe flow for water and 
dilute. solutions of pclyethylene oxide. J. Fluid 


Mech. 51, 177. 
Persson, B. O. (1975). Reconsideration of a problem 


in cavitation bubble dynamics. J. Appl. Mech., 
GRINS NoSoNGite CA Ts 

Plesset, M. S. (1949). The dynamics of cavitation 
bubbles. J. Appl. Mech., Trans. A.S-M.E. 16, 
277. 

Reed, R. L. (1969). The influence of surface 
characteristics and pressure history on the 
inception of cavitation. M. S. thesis, Pennsyl- 
vania State University. 

Rouse, H., and J. S. McNown (1948). Cavitation 
and pressure distribution; head forms at zero 
angle of yaw. State University of Iowa, Eng. 
BUaOlyaees aie 

Rudd, M. J. (1972). Velocity measurements made 
with a laser dopplermeter on the turbulent pipe 
flow of a dilute polymer solution. J. Fluid 
Mech. 51, 673. 

Schiebe, F. R. (1972). Measurément of the cavitation 
susceptibility of water using standard bodies. 
St. Anthony Falls Hydraulic Lab. Rep. 118. 

Schlichting, H. (1965). Grenzschicht-Theorie, 
Verlag G. Braun, Karlsruhe. 

Silberman, E., F. R. Schiebe, and E. Mrosla (1973). 


The use of standard bodies to measure the 
cavitation strength of water. St. Anthony Falls 
Hydraulic Lab. Rep. 141. 

Thwaites, B. (1949). Approximate calculation of 
the laminar boundary layer. Aero. Quart. l, 
245. 

Ting, R. Y., and A. T. Ellis (1974). Bubble growth 
in dilute polymer solutions. Phys. of Fluids 
17, 1461. 

Van der Meulen, J. H. J. (1971). 
hemispherical nosed teflon bodies. 
Symposium, Leningrad. 

Van der Meulen, J. H. J. (1972). Incipient and 
desinent cavitation on hemispherical nosed bodies. 
Int. Shipb. Progress 19, 21. 

Van der Meulen, J. H. J. (1973). Cavitation supp- 
ression by polymer injection. A.S.M.E. Cavita- 
tion and Polyphase Flow Forum, 48. 

Van der Meulen, J. H. J. (1974a). Friction reduction 
and degradation in turbulent flow of dilute 
polymer solutions. Appl. Scientific Res. 29, 

161. 

Van der Meulen, J. H. J. (1974b). 
polymer injection on cavitation. 
Conf., Edinburgh, Scotland. 

Van der Meulen, J. H. J. (1976a) 


Cavitation on 
IUTAM 


The influence of 
Cavitation 


Holographic study 


451 


of polymer effect on cavitation. A.S.M.E. 
Cavitation and Polyphase Flow Forum, 4. 

Van der Meulen, J. H. J. (1976b). A holographic 
study of cavitation on axisymmetric bodies and 
the influence of polymer additives. Ph.D. thesis, 
Enschede. 

Van der Meulen, J. H. J., and N. B. Oosterveld 
(1974). A holographic study of cavitation 
inception on a hemispherical nosed body. Symp. 
High Powered Propulsion of Large Ships, Wagen- 
ingen. 

Van der Meulen, J. H. J., and H. J. Raterink (1977). 
Flow visualization of boundary layers in water 
by in-line holography. Int. Symp. on Flow 
Visualization, Tokyo. 

Van Ingen, J. L. (1975). On the calculation of 
laminar separation bubbles in two-dimensional 
incompressible flow. AGARD, CP-168, Gdttingen. 

Van Ingen, J. L. (1976). ‘Transition, pressure 
gradient, suction, separation and stability 
theory. Workshop Rand Corporation, Santa Monica. 

White, W. D., and D. M. McEligot (1970). Transition 
of mixtures of polymers in a dilute aqueous 
solution. J. Basic Engng., Trans. A.S.M.E. 92, 
411. 


Mechanism and Scaling of 
Cavitation Erosion 


Hiroharu Kato 
University of Tokyo 


Toshio Maeda 


Mitsubishi Heavy Industries Ltd. 


Atsushi Magaino 
University of Tokyo 


Tokyo, Japan 


ABSTRACT 


Recently cavitation erosion has been primarily 
treated experimentally. However a need exists for 
both a theoretical cavitation erosion model and 
more quantitative erosion test methods. As a 
contribution to the state of the art, the authors 
have summarized their research at the University 

of Tokyo using the soft surface erosion test method 
(the aluminum erosion test). 

Two test series were completed, the first using 
the NACA 16021 foil section and the second using 
the NACA 0015 foil section. Two-dimensional erosion 
tests were systematically made at various velocities 
and cavitation numbers to obtain a correspondence 
between the erosion and the hydrodynamic character- 
istics of the cavitation pattern. It was found that 
the estimation of the cavity length and its fluctua- 
tion are important factors in the prediction of the 
cavitation erosion. 

The results of these tests are used to illustrate 
the effectiveness of Mean Depth of Deformation Rate, 
MDDR, aS a Cavitation Erosion Index. These test 
results also served as a background for extending 
the cavitation erosion scaling theory, previously 
proposed by Kato, to include differences in the 
cavitation number. 

After determining two empirical constants, the 
resulting predicted MDDR Cavitation Index was shown 
to be in good agreement with both Thiruvengadam's 
(1971) and the authors' test results. 

In addition to this basic research, two additional 
studies are summarized. The first is a comparative 
test of the aluminum erosion test and the paint 
test and the second is a study in the influence of 
air injection in reducing the cavitation erosion 
intensity. The test results obtained from the paint 
and aluminum tests were found to be in good agreement 
and for routine cavitation erosion checks, the paint 
test should be adequate. It was found that small, 
air injection rates reduced the cavitation erosion 
intensity dramatically and large injection rates 


452 


did not result in substantial reduction of the 
cavitation erosion intensity. 


1. INTRODUCTION 


Erosion is one of the largest problems caused by 
cavitation. Cavitation tests of model propellers 
have been made for the purpose of predicting cavita- 
tion erosion, especially for low-speed merchant 
ships. However, the prediction was mainly based 
on the observer's "feeling" of the cavitation 
pattern on the propeller blade. Recently a new 
testing method, i.e., paint test, was developed at 
several laboratories [Sasajima (1972) and Lindgren 
and Bjdrne (1974) ]. In this test the erosion inten- 
sity is judged by the area of paint peeled off. 

At the University of Tokyo in the authors' 
laboratory, erosion tests of soft aluminum test 
pieces have been made for several years [Sato et 
al. (1974) and Sato (1976)]. The main purpose for 
developing the soft aluminum method are: 

(1) Development of a quantitative prediction 

method for cavitation erosion. 

(2) Obtain a deeper insight into the mechanism 

of cavitation erosion by the observation 
of eroded metal surface. 

(3) Establishment of cavitation erosion scaling 

laws. 

The test piece is usually made of pure aluminum, 
which is easy to obtain, has stable quality, good 
machinability, and is relatively cheap. Its 
mechanical properties can be roughly established 
by hardness and tensile tests. The erosion resist- 
ance of pure aluminum is very low and its surface 
is roughed by cavitation attack within one half 
hour of test exposure which is similar to the testing 
time of the paint test. The increase in roughness 
is a first indication of erosion [e.g., Young and 
Johnston (1969) ]. It can be measured by a roughness 
tester and the quantitative erosion intensity can 
be obtained with sufficient accuracy. 


Micro-appearances of the eroded surface such as 
the pit shape, can also be qualified by examination 
of roughness records and microscopic pictures of 
the surface. 

The erosion intensity has been evaluated by 
mean depth of penetration (MDP) GoGo 5 Hammitt 
(1969) ] or energy absorbed by the material eroded 
l@aGio p Thiruvengadam (1966) ]. In addition one of 
the authors recently proposed a new concept of 
erosion intensity, mean depth of deformation (MDD) 
which functions as a bridge between surface rough- 
ness, SR, and MDP [Kato (1975) ]. Thus MDD corres- 
ponds to SR at the initial stage and MDP at the 
final stage of erosion. 

This paper discusses the experimental results 
of two-dimensional aluminum foil sections (pure 
and aluminum alloy), various considerations of the 
erosion mechanism in connection with the hydrodynamic 
characteristics of the foil section along with the 
modeling and scaling of erosion, and summarizes 
experiments using an air injection system which the 
authors found very effective in cavitation erosion 
preventation. Nomenclature is shown at the end of 
this paper. 


2. FOIL SECTION EROSION TEST 
High Speed Cavitation Tunnel at University of Tokyo 


Erosion tests of two-dimensional foil sections 

were made using a high speed cavitation tunnel at 
University of Tokyo. The test sections of this 
tunnel can be changed according to the experiment. 
For the present test two test sections were used. 
One was the rectangular high speed section with 
cross section dimensions of 100mm x 10mm. Test 
Series I was carried out using this section in 
1976. Since the side wall effect was so large that 
the two-dimensionality of the flow was almost lost 
near the trailing edge of the foil section, it was 
concluded that the 10mm width was too narrow. There- 
fore the test section was modified to a 80mm x 15mm 
cross section prior to starting Test Series II in 
1977. The maximum velocity of the section was 
about 50m/s. 

The second test section was the rectangular low- 
speed section used only in Test Series II (1977). 
It has cross section dimensions of 120mm x 25mm 
and a maximum velocity of 35m/s. 


Foil Section 


Two foil sections (NACA 16021 and NACA 0015) were 
tested. The NACA 16021 foil section used in Test 
Series I (1976), was the same foil section used in 
Kohl's experiment [Kohl (1968) ]. Kohl made his 
tests at an attack angle of a = 0°. Since this 

foil section has no camber, when it is set at a = 
O°, the inception point of cavity appears around 

60% chord. Thus, testing at a = 0° was not suitable 
for cavitation erosion tests, so the authors chose 

a test condition of a = 4°. Since its chord and 
Span are 40mm and 10mm respectively, the aspect 
ratio A = 0.25, was so small that the spanwise 
pattern of the cavity was not uniform. The cavity 
closed at midspan appearing as a kind of streak 
cavitation. Another disadvantage of using the NACA 
16021 section is its chordwise pressure distribution 
which is the "roof-top" type. The cavity length 


453 


drastically changes with only slight changes in the 
cavitation number. While this characteristic is 
desirable in practical applications, it was found 
to be undesirable in the present study since erosion 
would occur only in a narrow range of cavitation 
numbers which makes the experiment difficult. 

Therefore prior to starting Test Series II in 
1977, two major improvements were made. From wind 
tunnel tests the minimum aspect ratio necessary to 
maintain two-dimensional flow was found to be about 
4 = 0.4 and an aspect ratio, A = 0.5, was chosen 
for Test Series II. The smaller foil was designed 
with a 30mm chord and a 15mm span and the larger 
foil section was designed with a 50mm chord and 
a 25mm span. 

The second improvement was to change the foil 
section, from the NACA 16021 to the older NACA 0015, 
which has a chordwise pressure distribution of the 
"triangular" type. The experimental chordwise 
pressure distribution of this foil is compared in 
Figure 1 with the calculated pressure values. It 
can be seen that the agreement between the experi- 
ment and calculation is satisfactory. 


Test Condition 


In Test Series I (NACA 16021) the following items 
were tested: 

(1) Relationships between the mean depth of 
deformation (MDD), mean depth of penetration 
(MDP), and surface roughness (SR). 

(2) Effect of cavitation number, velocity, and 
the water's air content on the erosion 
intensity. 

(3) Comparison between the results obtained by 
the soft aluminum erosion test and paint 
EeSice 

(4) Influence of air injection on erosion pre- 
vention. 


-2.0 
—— THEoRY 
EXPERIMENT 
S55) 
- @= U4 pec, 
a= 
Cp 2 DEG. 
@= 0 pec. 
-1.0 $ = - 4 Dec. 
-0.5 
0 
0.5 
1.0 
FIGURE 1. Comparison of suction side Cp for NACA 


0015 foil section. 


454 


In Test Series I, the size and material of the foil 
section were not changed. The material was pure 
aluminum, JIS H2102-2 (AL > 99.5%). 

In Test Series II (NACA 0015) the following items 

were tested. 
(1) Effect of cavitation number, velocity, and 
chord length, and the hydrodynamic character- 
istics of the cavity flow, on the erosion 
intensity. 
(2) Effect of material properties on the erosion 
intensity. 
(3) Comparison of the soft aluminum and paint 
test results. 
The test conditions are summarized in Table 1. 
The attack angle was a = 4° throughout Test Series 
I and II. In Test Series I, the air content was 
a/adg = 0.5, while in Test Series II it was initially 
0.2 and increased gradually during the experiment 
to a value of 0.4 by the end of the experiment. 

In Test Series I-D, before the test began, air 
bubbles were injected into the cavitation tunnel 
to control air bubble content of the water. Then, 
the erosion test was completed to study the effect 
of air content on erosion. 

Experiments with air injection from the foil 

surface were also carried out to study the positive 


Table 1 


utilization of erosion prevention effect of air 
bubbles. 

At the start of the tests, the water temperature 
was about 25°C which increased during the high speed 
tests, reaching a maximum temperature of 50°C. 

In addition to the erosion tests, measurements 
of the hydrodynamic characteristics such as cavity 
length, pressure distribution etc., were completed 
using a similar foil section made of stainless steel. 


Material and Heat Treatment 


In Test Series I the foil section material was pure 
aluminum (JIS H2102, 99.5%), while in Test Series 
II pure aluminum and two kinds of aluminum alloy, 
JIS H4163-2 (AA 5056) and JIS H4163-5 (AA 6063) 
were used. These materials were selected for their 
low erosion resistance, good corrosion resistance, 
and good machinability. The foil sections tested 
were machined by a NC-milling machine and the surface 
was smoothed by a buffing machine. The foils' 
surface roughness was found to be less than 1 um 
in the virgin state. 

Since the foil surface was work-hardened, a thin 
layer of the foil surface had a large degree of 


Experimental Conditions 


Series I 


ie 


Cav. No. & Flow Vel.| Duration 


Material JIS* 
Attack Angle 


Flow Velocity 


Cavitation 
Number 


Exposure Time 


Air Content 


Material 
JIS (AA)* 


Chord 30 mm, Span 15 mm 


HA4LE3—5 Hh163—2 
(5056) 


(6063) 


Flow Velocity 


Cavitation 
Number 


Air Content 


4v8 ppm (a/as** =0.200.4) 


* JIS Japanese Industrial Standards 
AA : The Aluminum Association 
**¥ Qo : Saturated Air Content at 25°, 1 ata. 


455 


Table 2 Chemical Composition and Mechanical Properties 


Pure Aluminum 


JIS H2102-2 


Aluminum Alloy 


JIS H4163-5| JIS H4163-2 
(AA 6063) (AA 5056) 


v0.10 


Chemical 
Composition 


Stress 


Tensile 


2 
Mechanical Strength (kg/mm 


Vickers 


Properties 
Hardness (kg/mm? 


Young's 
Modulus 


(kg/mm? 


hardness, requiring heat treatment to remove this 
work hardened layer. Following the Japanese 
Industrial Standards, pure aluminum and aluminum 
alloy H4163-2 were annealed for 1 hour at 400°C 
and foils made of aluminum alloy H4163-5, were 
annealed for 1 hour at 205°C. 

The surface hardness before and after annealing 
are shown in Figure 2. This test was made using 
a micro Vickers hardness tester. The tensile test 
results are shown in Figure 3 and summarized with 
the composition of the materials in Table 2. 


Surface Roughness (SR), Mean Depth of Penetration 
(MDP), and Mean Depth of Deformation (MDD) 


For this study a NACA 16021 foil section was 
tested for 9 hours to find the relation among SR, 
MDP, and MDD. The result is shown in Figure 4. 
When a ductile material such as aluminum is exposed 
to cavitation, small pits detected by an increase 


60 


Pure Acuminum (H2102-2) 


Non-ANNEALED Pas 


7 
ANNEALED 


Hv (KG/MM) 
8 
9? 
° 
| 
° 


MEAN RoucHNess (gem) 


0 200 400 60( 
Static Loap (G) 


FIGURE 2. Result of Vickers hardness test [pure 
aluminum (H2102-2)]. 


in SR are formed at the first stage of erosion. 
At this stage there is no weight loss. This initial 
period is called the incubation period where after 
an initial increase, the SR value asymptotically 
approaches a larger value. 

It is well known that MDP remains zero during 
the incubation period. The time rate of MDP/(MDPR) 
increases to the maximum (acceleration period) then 
decreases gradually (deceleration period). Asa 
measure of erosion intensity the value of MDD, pro- 


30 


25 


ind 
oOo 


rH 
al 


Stress (KG/MM”) 


0 5 10 15 20 25 
STRAIN (%) 


FIGURE 3. Comparison of tensile test result. 


456 


i 
Oo 


WN 
Oo 


10 


ho 
oO 


x 
iS) 
Wwe 


o——o—°- 


Mean ROUGHNESS 5 


joo 
f=) 


We1GHT Loss (mG) 


Mean DEPTH oF DEFORMATION (gem) 


we 


0 — @—«@ 0 
0 2 4 € 8 19 
Time (HR) 
FIGURE 4. Extended duration cavitation test [NACA 


Pure Al (H2102-2), C = 40 mm, a = 4 deg.]. 


posed by one of the authors, seems to be more 
suitable than MDP. The advantages of using MDD 

are that it increases almost linearly over a wide 
range of exposure time as well as the fact that MDD 
corresponds to SR in the incubation period and to 
MDP after long exposure. 

In the present tests, SR was measured to shorten 
the testing time. Usually the test was completed 
within 1 hour so the SR value coincides with MDD. 
The degree of erosion after a long exposure can be 
estimated using the measured SR. 


3. HYDRODYNAMIC CHARACTERISTICS OF CAVITATION ON 
NACA 0015 FOIL SECTION 


Cavity Length 


Because erosion occurs at the collapsing point of 
the cavities namely the end of the cavity, it is 
important to know the cavity length for predicting 
cavitation erosion. 
tests, the cavity length and pressure distribution 
along the back surface of the NACA 0015 foil were 


PROBABILITY (2) 


Therefore, prior to the erosion 


measured. At the test condition 50 photographs 
were taken to measure the cavity length. 

The results are shown in Figure 5. As seen in 
the figure, above o > 0.8 the distribution of cavity 
length is characterized by a peak, but below o < 0.8 
the fluctuation becomes so large that there is no 
characteristic peak. For the supercavitation 
condition (o = 0.45) the fluctuation is reduced and 
a characteristic peak can again be observed. The 
mean value of cavity length and its standard devia- 
tion are shown Figures 6 and 7. The cavity length 
increases linearly with smaller cavitation number, 
and the standard deviation begins to increase 
rapidly about o = 0.85 as clearly seen in the figure. 

It is well known that the cavity length of a 
partially cavitated foil can not be determined 
theoretically by linear cavity models. The cavity 
length predicted by a closed type cavity model is 
usually longer than the observed length. If we 
adopt a open type cavity model., the situation 
becomes reversed and the predicted cavity length 
becomes shorter than the observed length. Conse- 
quently a half-closed type model is usually adopted, 
but this model requires the opening of the cavity 
end to be determined experimentally. 

In this study the cavity length was calculated 
using the half-closed type model by Nishiyama and 
Ito (1977). This method is based on linear theory 
using singularities (source and vortex) distributed 
on the cavitated foil. The calculated results are 
shown in Figure 7 where the opening de was system- 
atically changed. The contour of de = O coincides 
with the closed cavity model. The circles in this 
figure represent the "mean" value of the observed 
cavity length. Using this mean value, the opening 
6e can be calculated showing that de increases 
with smaller values of o (see Figure 8). 


Pressure Distribution and Cavity Shape 


The theoretical pressure distribution and cavity 
shape for the back side of NACA 0015 foil section 
are shown in Figure 9 along with the corresponding 
experimental result. Here the Nishiyama-Ito's half- 
closed model was used with the de values taken 


0 
0 


PROBABILITY (2) 


FIGURE 5. Fluctuation of cavity length (NACA 
0015, «a = 4 deg., V = 35.9 m/s). 


100 
REGS eso 
& 
Ena 
= 
rd 
S 
nig nO Le 

50 100 150 0 50 100 150 


Cavity LenctH (&CHoRD) 


PROBABILITY (4) 


50 100 150 


Cavity LENGTH (%CHoRD) 


= 
oO 


© V = 35 M/s 
@ V= 25 m/s 


Ww 
oO 


4 V=15 m/s 


ip] 
oO 


= 
oO 


STANDARD DEVIATION (ZCHoRD) 


oO 0.6 0.8 1.0 


CavITATION NUMBER 


FIGURE 6. Standard deviation of measured cavity length 
(NACA 0015, a = 4 deg.). 


from Figure 8. The pressure distribution diverges 
to a positive infinite value at the end of cavity 
because of singularity at this point. This singu- 
larity makes the agreement between theoretical and 
experimental results very poor. 

The cavity shape is also compared in Figure 9. 
The observed leading edge of the cavity is about 
10% chord position. Whereas, in the theory the 
leading edge of the cavity begins at the leading 
edge of the foil. This appears to be one of the 
reasons why the calculated cavity thickness is 
much thicker than the experimental thickness even 
though the cavities have similar profiles. 


4. EROSION TEST 

Cavity Length and Position of Erosion 

The roughness increment on the foil was measured 
for various exposure times. Spanwise roughness 


measurements were made over the entire chord at 
intervals corresponding to 5% the chord length. 


120 
THEORY 


100 —-— EXPERIMENT 


0..08C 
30 0.06C 


0. 04C 


60 


457 


Two examples of the roughness distribution are 

shown in Figure 10. Arrow marks in this figure 
indicate the position of cavity end and the standard 
deviation of its fluctuation. 

The figure clearly shows that the peak of erosion 
appears slightly downstream of the cavity end, and 
the erosion distribution agrees well with the cavity 
fluctuation. Namely, there is an obvious peak in 
the region of o > 0.8, but in the region of o < 0.8 
the surface roughness distribution spreads over a 
wider range. This result indicates that the esti- 
mation of cavity length and its degree of fluctuation 
are important factors in the prediction of erosion 
intensity. 


Effect of Hydrodynamic Factors on Erosion 
Cavitation Number 


The mean increment of surface roughness, SR, and 
its time rate of change can be determined from the 
roughness distribtuion shown in Figure 10. It 
corresponds to the mean depth of deformation rate 
(MDDR) because the test was finished within the 
incubation period. While Thiruvengadam has proposed 
adopting the rate of energy absorbed by the eroded 
material, which can be calculated by multiplying 
MDP by the energy absorbing capacity of the material 
per unit volume, the present research uses MDDR as 
a measure of erosion intensity in order to find 
which property is responsible for cavitation erosion. 
It is known that the erosion intensity, MDDR, 
has a peak at the certain cavitation number. The 
change of measured MDDR to cavitation number is 
shown in Figure 11, where plots (a) and (b) refer 
to the NACA 0015 foil tests while plot (c) refers 
to the NACA 16021 foil tests. The test result of 
Kohl and Thiruvengadam are also presented in plot 
(a) [Kohl (1968) and Thiruvengadam (1971) ]. As 
mentioned earlier, while the same foil section 
(NACA 16021) was tested in Test Series I, a different 
attack angle was used. 
There are several differences in the results 
obtained in the NACA 0015 foil tests and the NACA 


= 35 m/s 
= 25 m/s 
= 15 m/s 


o.czc _8,-0 


40 


Cavity LENGTH (ZCHORD) 


20 


0.4 0.6 0.8 1,0 
CaviITATION NuMBER 


FIGURE 7. Comparison of calcu- 
lated and observed mean cavity 
length (NACA 0015, a = 4 deg.). 


1.2 1.4 


a 
i: 
So 
3S -1.0 
ss 
o 
Ce 
4.0 gl 
“0.60 0.80 1.0 1,20 HEORY 
CavITATION NUMBER 
EXPERIMENT 
: , : ; LE. 
FIGURE 8. Derived relationship between Se and cavita- Ghern 


tion number for NACA 0015 foil, a = 4 deg. 


FIGURE 9. Comparison of NACA 0015 foil calculated 
cavity shape and Cp distribution with experiments at 
a= 4 deg. 


STANDARD 
DEVIATION 


J PEND one 
30 MIN Cavity °° 
° 
STANDARD 
<—— DEVIATION 


it oF Cavity 


—?P 


100 


MDD (pom) 


FIGURE 10. Illustration of MDD 
(Mean Depth of Deformation) data 

: 0 20 40 60 80 100 0 20 49 60 80 100 
koi NITES, (IGM OR EOS Ipsithas L.E,  CHorD Position (CHORD) T,£, LE, CHORD Position (%CHORD) T,E, 


aluminum (H2102-2), a = 4 deg., 
V = 35 m/s]. (a) C = 50 mm (6) C = 30 mm 


459 


SON Z102—2 

— e— H4163-2 (5056) 

—4— H4163-5 (6063) 
° 


S 
= 
= 
a =—_ 
= cs 
= < 
2 st 
a 
a1 
(= 
— 
i=) 
= 
0.6 0.8 1.0 ney? 
Cavitation NuMBER 
(a) NACA 0015, H2102-2, 0.6 


a= 4deg., V = 35 m/s 


1.0 oe 


CAVITATION NUMBER 


(b) NACA 0015, C=30 mm 
a=4deg., V = 45 m/s 


41,7 m/s 
32.6 m/s 


are 


MDDR (pe m/min) 


2 
Peak Erosion INTENSITY (WATT/M’ ) 


0.6 0.25 0.30 


0.2 0.4 
CAVITATION NuMBER 


(c) NACA 16021, H2102-2, 
C = 40 mm, a= 4 deg. 


16021 foil tests. First, the width of the peak of 
NACA 16021 is narrower than the NACA 0015 peak. 

This is caused by the difference of pressure distri- 
bution between the two foil sections. The NACA 
16021's distribution is flat, resulting in a larger 
change of cavity length with small changes in 
cavitation number. In contrast, the NACA 0015 
section has a triangular pressure distribution so 
the difference between the inception cavitation 
number and supercavitation number is large. Since 
erosion occurs only when the cavity bubbles collapse 
on the foil surface, it seems quite reasonable that 
NACA 0015 has a much wider peak than the NACA 16021. 
Here the authors would like to point out that due 

to side wall effects the measured pressure distri- 
bution of the NACA 16021 foil and the peak value 

of o = 0.4 can not be obtained directly by a two- 
dimensional calculation. 


0.35 


CavITATION NUMBER 


(d) NACA 16021, 1100F-A1 
(Thiruvengadam, 1971) 


0.40 0.45 


FIGURE 11. Summary of MDDR erosion 
index and 'test results. 


Another difference between these two results is 
the value of the maximum MDDR. It is much larger 
for the NACA 16021 foil when compared with the 
NACA 0015 foil results, even though the chord length 
and test velocity are not that much different. The 
main reason lies in the difference of cavity pattern. 
With the NACA 16021 section, the cavity inception 
is concentrated at the mid-span position and the 
cavity was a streak type. Correspondingly, the 
erosion pattern was a streak type, where a narrow 
and deep eroded groove was formed along the middle 
of foil. A picture of this groove taken by a 
scanning electron microscope is reproduced in Figure 
12. Streak cavitation can induce severe erosion in 
comparison to sheet cavitation erosion which occurred 
in the NACA 0015 foil tests. The difference in the 
cavity patterns seems to cause this large difference 
in MDDR. 


460 


FLOW 
DIRECTION —> 


FIGURE 12. Scanning electron 
microscope photographs of eroded 
surface (NACA 16021, H2102-2, 

Cc = 40 mm, a = 4 deg., V = 41.7 
m/s, o = 0.450). 


Referring to Figure 11 (b) in the 3 test series 
where only the material of the foil was changed, 
the position of maximum MDDR changes. This seems 
irrational because the flow condition is not changed 
by the material. The reason of this shift is the 
occurrence of the foil's bent trailing edge. Ona 
full scale propeller, cavitation erosion is some- 
times accompanied by a bent trailing edge. The 
same thing happened in the present test. The: foil 
section made of pure aluminum is much weaker than 
those made from an aluminum alloy, and it was bent 
more at the trailing edge causing the shift of 
peak MDDR to the larger cavitation number. 

An example of a bent trailing edge is shown by 
the profile view in Figure 13. The amount of bend 
is large at the corner of the trailing edge, which 
exaggerates considerably the shape shown in this 
figure. The bent trailing edge was observed on 
every NACA 0015 foil sections when the erosion 
occurred. On the contrary, it hardly appeared on 


“ Poe 
Eropep REGION 


NACA 16021 foil section because of its thicker 
trailing edge. 


Velocity 


It is well known that the erosion intensity is af- 
fected very much by the mean velocity since Knapp's 
suggestion of 6th power law [Knapp et al. (1970) ]. 
The effect of velocity on the peak value of MDDR 

is shown in Figure 14. Usually the exponent obtained 


‘experimentally, has a large spread falling somewhat 


between 3 and 9. In the present tests with the 
NACA 16021 foil the exponent, n, was 9 and for the 
NACA 0015 foil tests the exponent, n, was 6. 


Chord Length 


The chord length of a foil also has a large ef- 
fect on the erosion intensity. This is very 


(a) BEFORE EXPERIMENT 


(b) AFTER EXPERIMENT. 


FIGURE 13. Impression of bent trailing edge. 


important for marine propellers where the scale 
ratio between a full scale propeller and its model 
is large. Sometimes this ratio exceeds 30. As 
mentioned above, while the effect of the velocity 
difference is very large, we can still make a model 
test with the same tip speed as full scale by 
increasing the revolution of the model propeller. 
However it is very difficult to reduce the scale 
ratio of chord length. 

Experimental verifications on this problem are 
also very poor. Thiruvengadam (1971) made his 
erosion tests using two chord lengths, 1.5 and 3 in. 
His result shows that the erosion intensity increases 
proportional to the chord length. The result 
obtained in the present test is shown in Figure 15. 
In the present tests the erosion intensity increases 
proportional to the square of chord length. The 
effects of hydrodynamic factors such as cavitation 


x10"? 
15 


MDDR ( pem/min) 
ine) 


0,2 
20 40 70 100 


VELocITY (m/s) 


FIGURE 14. MDDR vs. velocity 
(NACA 0015 : H2102-2, C = 30 mm, 

a = 4 deg.) (NACA 16021 : H2102-2, 
C = 40 mm, a = 4 deg.). 


461 


MDDR ( pem/min) 


0,3 
20 40 
CHorD LENGTH (mM) 


70 100 


FIGURE 15. MDDR vs. chord length 
(NACA 0015, H2102-2, a = 4 deg., 
V = 35 m/s). 


number, velocity, and chord length can be explained 
universally by a model of erosion mechanism. The 
details of this model will be given in Section 5. 


Air Content 


The effect of air content was examined using the 
NACA 16021 foil section results. The air content 
was controlled as follows. As a pretreatment, the 
water was degassed to about 8ppm in a vacuum chamber 
and introduced into the cavitation tunnel. Then a 
certain amount of air was injected into the tunnel 
through an injection port before the test. In this 
case the ratio of gaseous air to total air content 
is much greater than found in ordinary water where 
the amount of air is an order of parts per million 
of total air content [Ahmed and Hammitt (1969) ]. 
With increase of air content the value of MDDR 
decreases as seen in Figure 16. This tendency 
agrees with the test results of SSPA [Lindgren and 
Bjarne (1974) ] and those of Stinebring et al. 
[Stinebring et al. (1977)]. The reason is attributed 
to the damping effect of air in a collapsing cavity 
bubble, attenuation effect of tiny air bubbles to 
shock wave, or a combination of both. 


Material Properties 


The effects of material properties on erosion are 
usually tested by accelerating devices such as 
vibrators, rotating discs, water jets etc. Summa- 
rizing these results, Heymann has made the chart 
shown in Figure 17 where the hardness of the 

material was taken as a factor governing the erosion 
[Heymann (1969)]. As seen in the figure the slope 
differs according to the material group, namely 

the slope of the steel group is steeper than that 

of aluminum and copper and brass group. This implies 
that the erosion resistance cannot be fully repre- 
sented by hardness alone. Thus other material 
properties such as strain energy absorbed to material 
(engineering strain energy) [Thiruvengadam (1966) J, 
ultimate resilience [Hobbs (1966) ], or their com- 


MDDR ( pen/min) 


0 20 40 60 
Air Content (PPM) 


FIGURE 16. Effect of air content on MDDR 
(NACA 16021, H2102-2, C = 40 mm, a = 4 deg., 
V = 41.7 m/s, o = 0.443). 


bination [e.g. Hammitt et al. (1969) ], have also 
been proposed by several researchers. 

The present test results are also compared with 
those material properties, i.e., hardness, engi- 
neering strain energy, and ultimate resilience. 
Hardness seems to give the best representation as 
seen in Figure 18. This will be discussed in Section 
5 dealing with modeling the erosion mechanism. 


5. THEORETICAL CONSIDERATIONS 
Review of Erosion Scaling Theory 


Thiruvengadam has made several theoretical. consid- 
erations on scaling of erosion. In 1971, he 
introduced a scaling formula [Thiruvengadam (1971) ]. 
He assumed a statistical distribution of air nuclei 
and derived the efficiency of erosion, 6, as, 


_ o& 1 -2.67 
6 = 5 (As) €XD F(A), 


where 6, 0, Ao, and W are nondimensional nuclei 
size, cavitation number, degree of cavitation, and 
Weber number respectively. Equation (1) is very 
attractive because it has no empirical constants. 
However the calculated values are quite different 
from the experimental values. While 9 should be 
the order of 10° by the calculation, the 8 obtained 
from model tests typically has an order of 10710, 
This discrepancy comes from the assumption that the 
total energy of the cavity bubbles generates the 
erosion. The theory shows that when the cavitation 
number is reduced, the efficiency, 8, increases 
from the point of cavitation inception to a maximum 
and then decreases to zero when cavitation number 
reaches zero. This tendency agrees qualitatively 
with experiments. It is expected, since the actual 
cavity becomes a supercavity at a certain cavitation 
number causing the erosion intensity to decrease 
greatly and in a practical sense reach zero. 

One of the authors has proposed a model of erosion 
mechanism in which the discharged energy of the 
collapsing bubble is assumed to be distributed 
statistically as: 


a6), === - — 
) =a € exp € (2) 


where f is the distribution function of energy 
density, €, reached on the material surface. Then 
a scaling law for cavitation erosion was derived 
using an empirical formula for the erosion resis- 
tance of materials. A comparison with only the 
peak erosion intensity taken from Thiruvengadam's 
tests showed good agreement [Kato (1975)]. 


Consideration on Effect of Cavitation Number 


As mentioned before, MDDR has a peak value of a 
certain cavitation number. This is due to a 
combination of the following two reasons. There is 
an increase in the collapsing cavity volume as the 
cavitation number decreases which causes increased 
erosion. On the other hand, the decrease of cavita- 
tion number causes an increase in the cavity length 
so the eroded area shifts towards the trailing edge 
of a foil. Also when the cavity length exceeds the 
chord length, the cavity does not collapse on the 
foil surface, causing no cavitation erosion. 
Usually the cavity length fluctuates and the erosion 
intensity will change continuously with the cavita- 
tion number. Although there seems to be a consider- 
able decrease in the collapsing pressure of cavity 
decreasing cavitation number, the control factor 
of erosion intensity is the change of cavity length 
as mentioned above. 

The decrease of erosion intensity at the right 
hand side of the MDDR peak in Figure 19 is caused 
by the lack of cavity and by too long a cavity on 
left hand side. By increasing the cavitation number, 
the cavity becomes intermittent, and if the cavity 
is stabilized by roughing the leading edge, the 
MDDR peak shifts to a higher cavitation number 
where the peak value is increased. This was verified 
in the authors' experiments as shown in Figure 20. 


100 


0.1 
0,06 


0), 5010) 


10 Hv (k6/MM?) 


(b) CopPER AND Brass 


NorRMALIZED EROSION RESISTANCE (NE ) 


0.01 
100 1000 15 100 300 
Hv (kG/MM*) Hv (kG/MM-) 


(a) STEEL (c) ALUMINUM ALLOY 


FIGURE 17. Vickers hardness vs. erosion 
resistance [Heymann (1969)]. 


V/MDDR (r1N/ pe) 


1/MDDR (MIN/ gm) 


20 40 
Hv (Ke/MM") 


70 100 ] 2 ih 10 


S.(kG/mm" ) 
(a) Vickers Hardness (b) Engineering Strain Energy 


300 
200 


/MDDR (mIN/ pe m) 


20 


10 
0.007 0.01 0.02 0.04 0,070.1 0.2 


UR (k6/Mm*) 


(c) Ultimate Resilience 


FIGURE 18. MDDR vs. various mechanical 
properties of material (NACA 0015, C = 30 mm, 
a = 4 deg., V = 45 m/s). 


Modelling of Cavitation Erosion and Scaling Factors 


As mentioned above, one of the authors developed 
a model of the cavitation erosion mechanism. How- 
ever it is limited to only constant cavitation 
numbers and the effects of material properties were 
derived empirically from accelerated tests. In the 
present paper, this model is developed further to 
treat differences in the cavitation number. The 
effect of the material's mechanical properties is 
also studied and a simple model is introduced. 

The total energy of collapsing bubbles per unit 
is given as: 


E. = n(p-pv)O , (3) 


probability of bubble collapses on a 
foil’ surface, 


where nN 


P-Py : pressure difference at the collapse 
point, 
Q : volumetric flow rate of cavitation 
bubbles. 


Equation (3) can be modified: 


463 


GovERNING Factor 


<— 
Cavity LENGTH 


—_ 
Cavity VoLUME 


EROSION INTENSITY ——» 
(MDDR) 


CavITATION NuMBER —»> 


FIGURE 19. Illustration of MDDR peak characteristic 
(test data given in Figure ll). 


E, * 1(Po-Py) 5 BV 


« nopv? eS) es (4) 
where 6 displacement thickness of cavitation 
bubbles, 
B : foil] span, 
de : cavity thickness at the cavity end, 
V : velocity, 
L : reference length. 


Assuming that a cavity bubble grows according to 
Knapp's similarity law, the volume,V, is: 


Vera (wg = ye (5) 


where T = ay where A is the cavity length. 
The pressure difference, Ap, is assumed as 


x10°? 
3.0 


© SmooTH SURFACE 


@ RouGHEeD aT LeaDING EDGE 


2.0 


1.0 


MDDR ( pem/miNn) 


0.3 0.4 0.5 
CavITATION NUMBER 


FIGURE 20. Impression of effect of roughened leading 
edge [NACA 16021, tested by Ozaki and Kiuchi (1975)]. 


464 


Ap © Py-Pmin 


Ga (ot Gam) OF - (6) 


Combining Eqs. (5) and (6), the following equation 
is derived. 


3 
v « 3 [-(o + Cryin) | (7) 


The number of cavity bubbles per unit time is then 
given as, 


de V 
va @e 


7 (8) 


= 3 
r3n [-( © + CPmin) ]2 


where \ as the nondimensional cavity length, X= 
A/D Se is the nondimensional cavity thickness at 
the end, and de = Se/L. 

Here, we make the same assumption as in the 
previous paper [Kato (1975) ] on the statistical 


energy distribution of cavity bubbles. The distri- 
bution is given as, 
n = cE exp (-aE) 9 (9) 


where n is the number of bubbles per unit time 
whose energy is between E and E + dE. Total number, 
N, and total energy of bubbles, E;, are given as 
follows: 


foe) 
c 
N= a ndE = a2 (10) 
0 


Constants a and c can be decided by combining Eqs. 
(4), (8), and (10). 


1 
ES 


pp 
Il 
No 


nov2L2r3[-(o + CPmin) | 


oom (alah) 
ki de 


Se Ras Se 9 
n2p2v3L7A902[-(0 + CPrmin) ]” 


where Kj} and K} are constants independent of the 
chord length, velocity and cavitation number. From 
Eq. (9), the distribution function of energy density, 
f, is derived as a function of energy density, €. 
The detailed discussion of this point is given in 
the previous paper [Kato (1975) ]. 

Substituting Eqs. (9) and (11) into the relation 


£(E) coon (ele) i, (12) 
the final expression for f is 


f = C € exp (-Ae) f (13) 


where A= 


= x) 
n2p2v3L32\ 202 [-(o0 + Cp.) ]2 
min 


In the present case the chord length is taken as a 
suitable reference length, L. 

Equation (13) is similar to Eq. (2), but it is 
extended to include differences in the cavitation 
numbers. 

The next problem is the modelling of deformation 
of a material surface caused by the attack of 
collapsing bubbles. For the present tests, hardness 
seems the best property to express the erosion 
resistance of a material. However it was found to 
be insufficient as seen in Figures 17 and 18. 

The methods of hardness testing can be divided 
into two types. One is the measurement of a dent 
size caused by the static load of a sphere or a 
pyramid on the material surface. The other method 
is the measurement of absorbed energy from dropping 
a certain test body on the surface. The Vickers 
hardness test made in the present study belongs to 
the first type. 

When a pyramidal dent whose depth is d, is 
formed by a static load F (Figure 21), the energy 
used to the deformation is 


iy Gaiotel 9B (14) 


The hardness has the following relation by its 
definition. 


(15) 


The increase of surface roughness (SR) by the single 
dent is given as 

wv . ae 

Se SS 16 

SRES aia) rayne (16) 


where Y and S are the volume of the dent and refer- 
ence area, respectively. 
Combining equations (14) ~ (16), 


i 
a p 
Gos aera vie STE 17 
oe EL S H ee 


Diamond PYRAMID 


FIGURE 21. Model of Vickers hardness test method. 


where e 
atecraieulo 


is the energy density absorbed by the 


s plastic deformation. 


If e is small 


enough, the deformation is within the elastic limit 

and no permanent dent will be formed. When e 

exceeds a certain limit, e,, the plastic deformation 

of surface occurs and a pernament dent is formed. 
Then the following relation is derived: 


a) = 0 fore < ee 


Qa. = © 2° 


Pp @ for e > Se 9 


(18) 
The above mentioned argument is valid for the actual 
case of erosion where many cavity bubbles collapse 
in a certain period if e is substituted tO}, Er, en 
these equations. 


Then, 
—e =0 for € <e 
P 
€ =e-e LOnReescme * (19) 
p c 
and 
oO 
MDDR « — € £(e)de 
Vv 
Eo 
ao 
ay (e-e)c (-ae)d 20 
a EQ e exp €)de (20) 
Vv 
we 
Integrating Eq. (20), 
K ov3 € 
MDDR = —~ —— g (a)( 2+ —1"_) 
Loy dele oV LF (o) 
ic 
ex | = FI (21) 
OV2LF(o) 


where 


= 3 
F(o) = naso [-(o + Cp in) J? 


G(o) = node o 


Here F and G are functions of cavitation number, 
where G is proportional to the total energy of the 
cavity reaching to the surface, and F is related 
to the individual energy of each cavity bubble. 

The probability of the bubble collapse on the 
foil surface, n, is calculated using the estimated 
mean position of collapse and its fluctuation. In 
the case of the NACA 0015 foil section, the position 
was estimated as 1.3 A from Figure 10 and the 
fluctuation is assumed to be the same as the cavity's 
fluctuations. The thickness at the end of cavity 
is taken from Figure 8. The value of F andG for 
NACA 0015 section were calculated at a = 4°. The 
results are shown in Figure 22. 

While the critical value of energy density, Er 
should be expressed by the mechanical properties of 


465 


material such as yield strength, Young's modulus 
etc., at the present stage, for lack of data we 
assume the following relation, 


Yield strength (22) 


aC y o oy 
and determine the power, n, from the erosion experi- 
ments. 


Comparison with Test Result 


The results of this theoretical model are compared 
with the erosion test of NACA 0015 section in 
Figure 23 where the two constants, K, and Ky, in 
Eq. (21) were determined using two different test 
points. In this figure those points are shown by 
dashed marks. The value of the power, n, was taken 
as n = 1/4 from the experimental results. The 
agreement between this theory and the test results 
is satisfactory. 

The theory was also compared with Thiruvengadam's 
test result [Thiruvengadam (1971)]. In this case, 
no data about the cavity was measured, so only the 
peak value of erosion intensity was used in this 
comparison with the present theory. The agreement 
is almost perfect as seen in Figure 24 where one 
set of data was used to determine two constants. 
Photos in Figure 23 (b) also show the paint test 
results discussed in the next section. 


Paint Test and Soft Aluminum Erosion Test 


Recently the paint test has been routinely used at 
several research laboratories to predict erosion 
intensity, in contrast to the present research using 
the soft aluminum erosion test to predict erosion. 
Both of these two test methods have merits and 
demerits. The soft aluminum erosion test is some- 
what troublesome and the surface of the material 


5 O26 0.7 0.8 0.9 1.0 Itai 
CAVITATION NUMBER 


FIGURE 22. Derived F and G values for NACA 0015 foil 
section at a = 4 deg. 


$/W Gp = A ‘WwW Og = 0 ‘(Z-ZOLZH) WnuIWN|y and (4) 


YdadWNN NOTLVLIAW) 


al OT 60 3'0 Z'0 


s}[nsey qsey juted 


(NIN/Wm™) YCaW 


“(“bep 7 = 2 


“STOO WOWN) S3INsexr [TeqUueUTrsedxs 
UQTM XO8puT UOTSOZe YddW PezOTp 
-o01d Jo uostzeduod *ez Tena 


YBEWNW NOLLVLIAV) 
weil Olea nO 8'0 Z'0 Oi) 0) 


wy 
S 
(NIW/WTT) YW 


YIGWNN NOILVLIAV) 
els OT 6'0 8'0- £0 oH Sh) 


0 
S 
3 
> 
= 
9 S'0 = 
J 
oT 
7-01 
UBAWNN NOLLVLIAV) 
iCall O'T 6'0 8°0 E0950 S'0 
0 
0'T 
Ss 
o 
= 
zr 
= 
= 
072 


S/W Gp = A ‘‘bapp=0 
‘wl O€ = 9 ‘(9S0S 
‘C-EQLPH) AON WnuiWNiy (Pp) 


s/W Gp = A ‘bapp=0 
‘WwW Of = 9 ‘(£909 
"G-EQLb~H) AOly WnuiwWNn|y (9) 


s/W GE =A 
“(Z-ZOLZH) Wnuiuinjy aing (2) 


466 


nN 
f=) 


a 
(=) 


oO 
MNS 


S) 
JS 


9,2 THIRUVENGADAM 


2 
Peak Erosion INTENSITY , IE (waTT/M’ ) 


o 1.5 IN.FoIL 


© 3,0 In.FoIL 


0.1 
100 (F/s) 200 300 


a a 
40. 60 80 
V (m/s) 


FIGURE 24. Comparison with 
Thiruvengadam's data (NACA 
16021). 


is destroyed, as a matter of course, after a long 
exposure to cavitation. But as mentioned in Section 
1, it has the merits of yielding quantitative and 
reliable erosion data, a similar appearance of the 
full scale eroded surface, etc. 

The paint test has just the opposite merits. It 
is a cheap and handy method. And although the 
conditions under which the paint is removed changes 
with very small changes in the paint composition, 
test procedure, etc., it appears that by developing 
standards, the paint test can be used to represent 
relative differences between similar models. 

From this discussion of the paint test merits 
and demerits, the paint test appears suitable for 
daily routine tests of usual propellers. The soft 
aluminum test is suitable for making standard com- 
parative tests at different research laboratories 
as well as for different types of propellers and 
for situations where critical erosion predicitions 
are required. 

It is valuable to make a comparison of these 
test methods using the same foil section. After 
testing several kinds of paint, a marking paint 
"AOTAC" was found to be the best. Figure 23 (b) 
shows appearances of the painted surface after 5 
min. test. They can be compared with the theory 
and the soft aluminum erosion test results shown 
in the same figure. The cavitation number of 
maximum erosion intensity is slightly different 
between the paint test and theory. But the general 
tendency agrees well and the paint test seems very 
useful especially for a comparative testing. 

The position of maximum erosion intensity esti- 
mated from the paint test also agrees well with 
the chordwise distribution of MDD shown in Figure 
10. 


6. AIR INJECTION SYSTEM 


Tiny air bubbles in the free stream reduce the 

erosion intensity by the action of their damping 
effect as mentioned in Section 4. To achieve a 
positive damping effect an air injection system 


467 


with air bubbles injected from holes on the foil 
surface is sometimes adopted. This system has 

been used very effectively to prevent erosion on 

the inner surface of a full-scale ducted propeller 
(Ooo p Okamoto et al. (1975) and Narita et al. 
(1977) ]. However the mechanism of prevention is 

not yet fully explained, and the best injection 
position and/or the necessary amount of air injection 
have not been clarified. 

The authors made the air inject test using NACA 
16021 foil sections with three air injection holes 
of 0.5mm dia. drilled at 10% or 37.5% chord position 
(Figure 25). The tests were made at a = 4°, 

V = 41.9m/s, and o = 0.438. The previous test 
showed that the peak MDDR value falls somewhere 
between 40 ~ 45% chord. The injection position 

of 10% chord represents the injection near the 
leading edge of the section, and that of 37.5% chord 
represents the injection which insures effective 
coverage of the eroded area. Air was then injected 
at 2, 5, and 10 cc(normal)/min. The quantity of 
air was so small that separate air bubbles were 
found even at the 10 cc/min, and consequently the 
air jet column typical at high flow rate was not 
observed. As seen in Figure 26, the injection 

from 10% chord gives better performance and even as 
small a rate of the injection as 2 cc/min results 
in drastic decrease in the erosion intensity. With 
injection the MDDR value reduced to 1/5 of non- 
injection level. Increasing air volume, the value 
of MDDR decreases but the effect seems to become 
saturated with a larger rate of air injection. 


7. CONCLUSIONS 


(1) The purpose of the present research was to 
find the mechanism of cavitation erosion and its 
scaling laws with special reference to the relation- 
ship between the appearance of cavitation and the 
erosion intensity. 

(2) Detailed observations of the cavity pattern 
were made on a two-dimensional foil section (NACA 
0015). Then erosion tests, using the same foil 
section of pure aluminum and aluminum alloy, were 
made to measure the increase of surface roughness. 
The erosion intensity was also compared with the 
observed cavity pattern and other hydrodynamic 


0.1C or 0.375C 


ERoDED REGION 
(0.45 - 0.5C) 


FIGURE 25. 


Location of air injection. 


468 


©0 cc/MIN 


Inuection Position 0,375C 


MDDR (ypm/mIN) 


oD Ce/MIN 


wn 
o 


Invection Position 9.1C 


ye 


2 cc/MIN 
‘4. 


ae cc/MIN 


4 
10 cc/MIN 


0 5.0 10.0 | 
INJECTION RaTE (A/V - S) x10 


FIGURE 26. Effect of air injection on MDDR erosion 
index [NACA 16021, pure Al (H2102-2), C = 40 mn, 
a = 4 deg., V = 41.9 m/s, o = 0.438). 


factors such as cavitation number, water velocity, 
etc. 

(3) Modelling of cavitation erosion has been made 
assuming a statistical distribution of cavitation 


bubble. Using the model, a theory of erosion scaling 


was established which contains two constants given 
by the experiment. The erosion scaling of cavita- 
tion number, velocity, chord length, and material 
can be made by the theory. The theory has been 
shown to give good agreement with the authors' and 
Thiruvengadam's tests. 

(4) Another two-dimensional foil section (NACA 
16021) was also tested, but in this case the side 
wall effect was so large that the results were not 
compared with the theoretical calculations. 

(5) The paint test also was made with the same 
foil section (NACAOQ0O15). The results of paint 
test agreed with that of the aluminum erosion test 
although it gives qualitative data. 

(6) The effect of air content and air injection 
method was also investigated experimentally. The 
air injection was found to be very effective in 
preventing erosion. 


ACKNOWLEDGMENTS 


The authors would like to express their acknowledg- 
ments to Prof. S. Tamiya and members of the High 
Speed Dynamics Laboratory, University of Tokyo, for 
their many valuable discussions and help during 

the research work. They also wish to thank Mr. T. 
Komura, Mr. R. Latorre, and Miss N. Kaneda for 
their sincere help during the preparation of the 
Paper. 


This research work was financially supported by 
the Grant in Aid for Developmental Scientific 
Research (2), Ministry of Education, Japan (Research 
No. 185087) and the authors are grateful for the 
support. 


NOMENCLATURE 
A, a : constants 
B : span 
ic : constant, chord length 
c : constant 
Cp : pressure coefficient 
d : depth 
E : energy 
Et : total energy of bubbles 
e : energy density 
1 : force 
£ : energy density distribution function 
Ish re: Vickers hardness 
Ky, Kj constants 
Ko, Ki : constants 
L : reference length (chord length) 
al : length 
MDD : mean depth of deformation 
MDDR : mean depth of deformation rate 
MDP : mean depth of penetration 
MDDR : mean depth of penetration rate 
N total number of cavity bubbles 
n distribution function of bubble number 
p pressure 
Q volumetric flow rate 
R bubble radius 
Ss area 
SR surface roughness 
T time 
Vv velocity 
a attack angle, air content 
6 thickness 
Se cavity thickness at the end 
€ energy density rate 
n probability 
A aspect ratio 
nN cavity length 
fo) density 
oO cavitation number 
Sy : yield stress 
Vv : volume 
SUBSCRIPTS 
critical 
min > minimum 
Pp plastic deformation 
Ss : sSaturate 
Vv vapor 
©0 infinity 
— : nondimensional value 
REFERENCES 


Ahmed, O., and F. G. Hammitt (1969). Determination 
of particle population spectra from water tunnel 
using Coulter counter. Univ. of Michigan, Cav. 
and Multiphase Flow Lab. Rep. 01357-4-I, 08466-7-I. 


Reina I Con Mo Ca Ilbetoe, (5 ah allies beS WIE 
Mitchell, Jr., and L. P. Solomon (1969). A 
statistically verified model for correlating 
volume loss due to cavitation or liquid impinge- 

ASTM, STP 474, 288. 

Toward quantitative pre- 

ASTM, STP 


ment. 
Heymann, F. J. (1969). 
diction of liquid impact erosion. 
474, 212. 
Hobbs, J. M. (1966). 
tion erosion test. 


Experience with a 20-kc cavita- 
ASTM, STP 408, 159. 


Kato, H. (1975). A consideration on scaling laws 
of cavitation erosion. Int. Ship. Prog., 22, 
25)3)305i 


Knapp, R. T., J. W. Daily, and F. G. Hammitt (1970). 
Cavitation. McGraw-hill Book Co., New York. 

Kohl, R. E. (1968). Experimental studies to 
establish scaling laws for modelling cavitation 
damage. Hydronautics Tech. Rep. 233-12. 

Lindgren, H., and E. Bjarne (1974). Studies of 
propeller cavitation erosion. Conf. on Cavita- 
iesicra, Beaks 

Narita, H., Y. Kunitake, K. Holden, and B. Mugaas 
(1977). Investigation on the ducted propeller 
cavitation and the duct erosion prevention by 
the air injection system. Symp. Hydrodynamics 
of Ship and Offshore Propulsion Systems, Paper 
W/3¥ 


Nishiyama, T., and J. Ito (1977). Calculation of 


469 


partially cavitating flow by singularity method 
(1st Report: Two-dimensional isolated hydro- 
foil). Trans. JSME, 43, 370; 2165, (in Japanese). 

Okamoto, H., K. Okada, Y. Saito, and T. Takahei 
(1975). Cavitation study of ducted propeller 
on large ships. Trans. SNAME, 83. 

Ozaki, H., and D. Kiuchi (1976). Cavitation erosion 
of two-dimensional foil sections. Graduation 
thesis, Dept. Naval Arch., Univ. of Tokyo, (in 
Japanese) . 

Sasajima, T. (1972). On cavitation erosion test 
method by using a painted model propeller. 
Mitsubishi Tech. Rep., 9, (in Japanese). 

Sato, R. (1976). Study on cavitation erosion. 
3rd Lips Prop. Symp., 19. 

Sato, R., S. Tamiya, and H. Kato (1974). Study on 
cavitation erosion. Selected Papers from J. 

Soc. Naval Arch. Japan, 12, 21. 

Stimebrng yD. oh Ra Ee AGe Arndt) and di Wel Hole 
(1977). Scaling of cavitation damage. J. Hydro- 
nauticsy, Lis) oike 


Thiruvengadam, A. (1966). The concept of erosion 
strength. ASTM, STP 408, 22. 
Thiruvengadam, A. (1971). Scaling laws for cavita- 


tion erosion. Proc. IUTAM Symp. Lenigrad, 405. 

Young, S. G., and J. R. Johnston (1969). Acceler- 
ated cavitation damage of steels and superalloys 
in sodium and mercury. ASTM, STP 408, 186. 


Experimental Investigations 
of Cavitation Noise 


G6ran Bark and Willem B. 


van Berlekom 


The Swedish State Shipbuilding Experimental Tank, 


Goteborg, Sweden 


ABSTRACT 


The requirement of low or acceptable noise levels 
onboard ships as well as low levels of radiated 

noise for special purpose ships can cause large 
problems for the naval architect. Low noise levels 
onboard ships are required in living quarters and 
also in some working spaces. The radiated noise 
field is of concern for instance for fishing vessels 
and ships with acoustical dynamic positioning systems. 

One important source of noise in ships is cavita- 
tion and especially cavitating propellers. The 
cavitation noise can have a quite varying character. 
It may for example sound like a hiss or like sharp 
hammer blows. For the naval architect it is impor- 
tant to be able to predict and, if possible, to 
reduce undesired cavitation noise. 

In this paper some of the research and develop- 
ment work on cavitation noise at the Swedish State 
Shipbuilding Experimental Tank (SSPA) will be 
described. This work at SSPA is mainly experimental 
and two projects will be described here in detail. 
One concerns the relation between cavity dynamics 
and cavitation noise. This work was carried out 
using an oscillating hydrofoil in the No. 1 SSPA 
cavitation tunnel. The other project concerns the 
relation between types of cavitation and cavitation 
noise. Different types of cavitation were generated 
in the tunnel using axisymmetric head forms and 
hydrofoils. 

A great deal of effort has been made at SSPA to 
develop adequate methods for measuring cavitation 
noise in cavitation tunnels. A short review of 
the measuring techniques now in use is given in an 
introductory chapter. Besides the two projects 
mentioned above several other projects are, or 
have been, carried out at SSPA. 


1. REVIEW OF MEASUREMENT TECHNIQUES AT SSPA 


Measurements of cavitation noise started at SSPA 
as early as 1958. The first tests concerned cavita- 


470 


ting axisymmetric head forms and were carried out 
in the SSPA cavitation tunnel No. 1. The measuring 
equipment was a waterfilled box attached to one of 
the plexiglass windows of the tunnel. A hydrophone 
was lowered into this box and could thus pick up 
the noise emanating from the source (propeller etc). 

The transmission path from the noise source is 
through water, plexiglass, and water to the hydro- 
phone. The transmission loss due to the presence 
of the plexiglass window is low. The drawbacks to 
this arrangement are reflected acoustic waves and 
vibrations in the box. The problem with the 
reflected waves may partly be overcome by carefully 
calibrating, or rather comparing, results from the 
hydrophone in a free field and in the box using the 
same known noise source. Vibration problems (from 
the vibrating tunnel plating) may be cured by using 
a pair of rubber bellows between the box and the 
window (see Figure 1). 

The signal from the noise source is, however, 
still distorted as can be seen in Figure 2. This 
figure shows the noise from a cavitating propeller, 
as measured by the hydrophone in the box and a 
hydrophone near the propeller. The differences in 
the curves are striking and show that the general 
shape is seriously altered by the box. It is in 
fact almost impossible to analyse the signal in 
time-domain using the hydrophone in the box. Com- 
paring results from 1/3 octave band analysis also 
shows differences, especially as regards the 
frequency dependence. These differences are, 
however, not as striking as those for signals in 
time-domain. 

The arrangements for noise measurements at SSPA 
are at present: 

1. Flush mounted pressure tranducers on the hull 
(Figure 3) 

2. Flush mounted pressure transducers on the tunnel 
wall 

3. Hydrophones in the flow field near the propeller 
(Figure 3) 

4. Hydrophone in the water-filled box outside the 
tunnel 


Tunnel No | Test section 700 x700 mm 


~— = blade frequency period 


pressure 


[| | 


i\, thy H 


sal 

/ Why, 

ain Yelle yh if 1 YAN 
yee AeA eye 
NanenleesNynt 
, 
a? 

un | 


Leen oe 


iia 
Hl | | 
: Maya ei bial ja il | iM ! 
| 


lyk 


time 


Hydrophone in external box 


Hydrophone box 1(2mm steel 

connected to plexiglass 

window by rubber bellows 
/ 


Free water surface 


Plexigldss 


| | 


FIGURE 1. 
noise measurement. 


= 


Ny ot WM vii) 
VW eaynt "Oy It Yat When 
ty 


vonege somone HE 


FIGURE 2. Pressure signals at different 
hydrophone positions. 


First arrangement for 


471 


472 


All, 2 pine, AnD, Sale, 
are pressure transducers 


Arrangement 1 is intended to be the standard 
measurement procedure at SSPA and results are easily 
compared with full scale measurements using the 
same equipment. This arrangement gives essentially 
the near field noise from the propeller. 


i 5 WL 
4 
339 
| 
6 
——— fr, = 
Le 190 
i oa 

1 | 

| 

| 

| 

7) 7 
| 
! 
BL | le 
0 1/2 1 
FIGURE 3. Arrangements for noise measurements on complete ship model. (Tunnel No. 2) 
@ Non-cavitating propeller 
OA Cavitating propeller 
Sound pressure level re 1[Pa in 1/3 octave band 
1 OF rTpst onlin lin Tne] in Li ict eT sion poo 


If it is of interest to know the radiated noise 
into the farfield, arrangement 2 can be used. 
Arrangement 4 also gives the farfield noise, but 
has its problems, as discussed above. Arrangement 
2 has less problems with reflected acoustic waves 
and vibrations than arrangement 4. The main reason 
why arrangement 4 is still used is to compare results 
directly with older measurements. Arrangement 3 
(Figure 3) has been especially developed for explor- 
ing the influence of variation in cavitation and 
the effect on the near field noise. Other arrange- 
ments of hydrophones have also been used for special 
purposes. 

Since the main concern in the noise measurements 
is cavitation noise, the effect of flow noise due 
to the turbulent boundary is of minor importance. 
Usually the increase in noise level due to cavita- 
tion is quite substantial, as can be seen in Figure 
4, which shows a typical example for a propeller in 
non-cavitating and cavitating condition. 


2. EXPERIMENTS WITH AN OSCILLATING HYDROFOIL 
Background to Experiments with Oscillating Hydrofoil 
A typical example of the pressure signal froma 


cavitating propeller model is shown in Figure 5. 
The pressure was measured by a hydrophone near the 


180 z “ | 


170 -—t-—- 
L ho“Ta, 
160 [ Fal Toe 
. rage eo 


| e | 
NOE + s ali = 
F | ian 
yay ett SU LIL 
30 50 100 200 500 tk 2k 5k 10k 20k 40k 
Hz 


FIGURE 4. Noise measurements on propeller-model. 
(Tunnel No. 2) 


pressure 


| 104 Pa | 


(a 
time (ms) 


0 5 


propeller. The signal corresponds to a spectrum of 

the the type shown in Figure 4 and typically is a 

rather slow variation of pressure interrupted by 

sharp and fairly infrequent pulses. The pulses 

are presumed to be generated during the final cavity 

collapse and they provide the main contribution to 

pressure levels at high frequencies. The pulses 
are often higher than the low frequency variations, 
but because of their low repetition frequency and 
wide frequency content the spectrum levels at high 
frequencies are lower than at low frequencies. 

To understand the scaling of cavitation noise 
and how different types of cavitation noise are 
generated, and perhaps can be reduced, it is 
important to study the mechanism generating different 
types of noise. A suitable way to obtain such 
knowledge is to carry out high speed filming and 
synchronous measurement of the cavitation noise. 
The first idea was to carry out such measurements 
with a propeller model. Because of high tip speed, 
small dimensions, and the complicated geometry of 
a propeller it was decided to take the first step 
by performing such experiments with oscillating 
hydrofoils. By suitable oscillation of a hydrofoil 
it is possible to generate cavitation with approxi- 
mately the same dynamic behavior as obtained from 
a propeller operating in a wake. The experiments 
with oscillating hydrofoils were supposed to shed 
some light on the following questions that originated 
from the search for methods of prediction and 
reduction of propeller cavitation noise: 

1. Which are the characteristic properties of the 
pressure pulses from some special types of 
cavitation? 

2. Are strong pulses generated by an orderly 
collapse of the whole cavity (e.g., a sheet 
cavity) or do they originate from large or 
small parts that separate from the main cavity? 
What is the geometry before and during collapse 
of cavities generating strong pulses? 

3. How is the pressure pulse related to the size 
of the cavity? Is there, for example, any 
relation between the maximum extension of a 
sheet cavity and the final pressure pulse? 

4. Is rebound of cavities important for generation 
of sharp pulses? 

5. What part of the cavitation period is of main 
importance for the generation of different 
types of noise (slow pressure variations, sharp 
pulses, etc.)? 

6. Which are the characteristic properties of the 


Blade frequency period 


473 


FIGURE 5. Pressure signal from a cavitating 


10 propeller model. 


flow field, oscillation frequency, etc., causing 
cavitation with violent collapse? 

7. To what extent is collapse time determined by 
the oscillation frequency of the hydrofoil? 

8. To what extent does the cavity behavior seem 
predictable by theoretical methods? How 
realistic is it to think that a sufficiently 
good scaling from model to full scale is 
obtained for the most important cavitation 
events? 


Experimental Set Up 
Cavitation Tunnel 


The tests were carried out in SSPA cavitation 
tunnel No. 1 (the samller one) equipped with test 
section No. 1 (500 x 500 mm). 


Oscillation Apparatus 


The hydrofoil was located horizontally in the test 
section and attached to an oscillation apparatus 
fixed to the test section wall (Figure 6). The 
hydrofoil was supported only at one end and forced 
to oscillate (rotate) around an axis fixed spanwise 
through the midchord point, i.e., the geometric 
angle of attack oscillated around an adjustable 
mean value, a9, (Figure 7). The axis was driven 
by a connecting rod and an adjustable crankpin. 
By setting the crankpin the oscillation angle, a, 
could be varied from 0 to 6°. With the hydrofoil 
used in these tests the oscillation frequency, foccr 
was varied from 0 to 15 Hz. The limits of water 
speed, ag, G, and f,,, were set by the strength of 
the hydrofoil and the background noise generated by 
the apparatus. One part of the background noise 
from such an apparatus is knocking in shaft bearings. 
To minimize this knocking, adjustable bearings were 
used. The motor, which was not dimensioned for this 
experiment, could deliver 16 kW at a maximum speed 
Cpe 5X0) 16/435 

The dynamic angle of attack, experienced by the 
leading edge of the hydrofoil, is composed of the 
geometric angle and of an angle caused by the motion 
of the leading edge. The angle is also affected by 
induced velocity. In the following only the geomet-— 
ric angle is considered (Figure 7). 

The system with connecting rod and crankpin 
results in an approximately sinusoidal oscillation 
of the geometric angle of attack. This manner of 


474 


Hydrofoil 


Hydrophone strut 


FIGURE 6. Experimental set up. 


oscillation does not cause a time variation of the 
angle of attack that is completely similar to that 
of a propeller blade in a wake. The reason for 
using this sytem was that, due to its strength, 
high oscillation frequencies with large hydrofoils 
could be obtained. If similarity with propellers 
is most important it is probably better to use 
oscillation systems of the types constructed by 
Ito (1962) and Tanibayashi and Chiba (1977). 


Hydrofoil 


In these introductory experiments an existing hydro- 
foil, earlier used for studies in two-dimensional 
flow, was used. The profile has NACA 16 thickness- 
distribution and is typical of a relatively thick 
propeller blade at about 0.7 of propeller radius. 
The hydrofoil data are 


Mean line a = 0.8 

Camber ratio = fy/c = 0.0144 
Thickness ratio = s/c = 0.0681 
Chord length = c = 120 mm 
Span = 200 mm 

Profile shown in Figure 7. 


Noise Measuring Equipment 


Two hydrophones (Briiel and Kjaer Type 8103 with 
frequency response 0.1 Hz - 140 kHz +2 dB) were 
placed in notches in a tube supported by two hydro- 
foils in such a way that photographing of cavitation 
was permitted (Figure 6). The frequency response 

of the hydrophones mounted in this manner was 
checked by white noise. No significant change in 
the frequency response was detected. 

The hydrophone signals were recorded on FM- 
channels on a Honeywell 5600-C tape-recorder (0-40 
kHz at 60 ips tape speed). Recordings were also 
made on direct channels (300 Hz - 300 kHz at 60 ips). 


It was then possible to write out the complete signal 


(0-40 kHz) by use of tape speed reduction and UV- 
recorder. 


Simultaneous with the hydrophone signals, a 


FIGURE 7. 


signal showing the events of maximum angle of attack 
was also recorded. 


High-Speed Film Equipment 


The requirements set up for the filming were that 
the film had to be synchronous with the noise 
recordings and permit measurements of cavity size 
as a function of time. The intention was not to 
measure the detailed behavior of small or very fast 
events. The minimum duration of the filming was 
set to about one second. 

These requirements were met by a Stalex VS 1C 
camera capable of 3,000 frames/s. This is a 16 mm 
rotating prism camera taking rolls of 30 m film. 
Lenses with focus lengths of 9.8 and 50 mm were 
used. For synchronization the camera could release 
a flash at a preset time. The flash trigging 
signal was recorded on tape together with hydrophone 
signals and the flash was placed within the frame. 
Only one flash was released during each filming. 


The camera was also equipped with a crystal-controlled 


timée-marker, making one light marking every milli- 
second on the edge of the film. This, together 
with the synchronization flash, made it possible 
to identify and follow cavitation behavior on the 
film together with the corresponding pressure 


Geometric angle of attack = QA~A +A sin 2T t fosc 


Oscillating hydrofoil. 


behavior recorded on tape. An example of the 
recorded signals is shown in Figure 8. 

As light sources, two 1,000 Watt spotlights were 
used. To get a proper background without reflections 
the hydrofoil was painted with a red matte paint. 

A test was performed with black and white film 
(Kodak 2479 RAR Film). The result was not very 
good, the contrast between hydrofoil and cavitation 
being too small. Color film (Kodak Vide News Film) 
was then used, with very good results. 


Evaluation of Films and Pressure Signals 


The pressure pulse generated by a cavity is related 
to the volume acceleration of the cavity and thus 
it is desirable to measure the cavity volume as a 
function of time. With complex cavities this is 
not very simple. An estimate of the cavity volume 
could be obtained if both cavity extent (area) and 
thickness were filmed synchronously. This is 
possible by the use of optical systems reflecting 
the two pictures into the same frame [Lehman (1966) ]. 
No such attempts were made. Most photographs were 
taken in order to measure the cavity area on the 
suction side of the hydrofoil. To obtain information 
about the cavity thickness some photographs were, 
however, taken from the free end of the hydrofoil. 
A method of estimating the relative thickness, 
synchronous with the cavity area, was to measure 
the length of a cavity shadow generated by the 
directed light. The method, which was calibrated 
by use of spherical bubbles, was rather rough, but 
some general information of thickness behavior was 
obtained. 

The photographs were studied by use of an analysis 
projector permitting single-frame projection on a 
focusing screen, where the area of the cavities 
could be measured by summing up elements in a 
pattern. For identification of cavitation events 
on the films and noise recordings the synchronization 
flash was the primary starting point. To increase 
the accuracy of identification of events far from 
the flash easily identifiable events, such as 
single bubble collapses, were used as reference 
points. 


Experiments 


The experiments with an oscillating hydrofoil 
presented in this paper are the first of this kind 
carried out at SSPA and they are to be regarded as 
introductory in several respects. 
Only one hydrofoil was used. The following 
flow parameters were held constant during the tests: 
Relative gas content (at atmospheric pressure) 
of the tunnel water was 25% 
Water velocity in test section = U = 5.0 m/s 
Cavitation number at the center of test section 


ey), = 12 
Bag oe TT. eee ys 
1 5 w 
where 2 


Po = surrounding pressure = 11.850 Pa 
Py vapor pressure of water (20°C) = 2.338 Pa 
po = density of water = 998 kg/m? 
The following oscillation parameters were varied 
in the experiments (see Figure 7): 
ag = mean angle of attack of the hydrofoil 


475 


& = oscillation angle 
1 eye = oscillation frequency 
In the figures the reduced frequency k, is used: 


TE c 
wo _ osc 


where 


WwW = 2M eos 

c = chord length of the hydrofoil 

U = water velocity 
After some introductory tests the following con- 
ditions of hydrofoil oscillation were selected from 
high-speed filming: 


ao a £ k 
osc (co 

(o) (o) (Hz) 

3 3 3 0.23 
Ad Y 7 O53) 
ui rT 10 0.75 
HW rf 5 at a abs} 
3 4 il 0.08 
i W 2 OFMES 
uw ui 3 0.23 
ui 1 4 0.30 
" ¥ 7 0.53 
wv uv 10 0.75 
Me " 14 1.06 
4 3 3 0.23 
" WY 7 0.53 
ui 10 Os 75 
" u 15 dL 5 ALg} 
4 5 3 OR23) 
u oe 7) @553} 
Hi i 10 0.75 


Results 


Primary results are presented as pressure signals 
from cavitating and non-cavitating hydrofoils, 
measurements of cavity area, and sketches of the 
cavitation pattern at various oscillation parameters. 


Presentation of Results 


In Figures 9 - 14 a survey of pressure signals 

and cavitation patterns at various oscillation 
conditions is shown. All pressure signals shown 

in these and other figures are from the hydrophone 
(Hl) near the leading edge of the hydrofoil. For 
each condition some oscillation periods are shown. 
The length, Toso = 1/fosc, Of an oscillation period 
is identified by the markings of maximum angle of 
attack, Omax- The figures show primarily cavitating 
conditions (cavitation number = 0.76) but in some 
cases signals from the corresponding non-cavitating 
condition is sketched (without the fine structure, 
which is apparatus noise). The pressure scale is 
given as a number of Pascal (Pa) per scale unit 
(su) defined at the top of the figures. The time 
scale is 6.15 ms/scale unit in all signal examples 
in Figures 9-13. For one of the oscillation periods 
the number of the oscillation period (relative to 
the synchronization flash) is shown in a circle, 
and for this period some additional data is given 
to the right. In the cavitation sketches are shown 
the maximum area extent, the maximum chordwise 
cavity length, &max, and the cavitation extent at 


FIGURE 8. 


hydrofoi 


ale 


NN ee EEEEyeEeEeEeEeEEEEeE~ILE _—————————EEEKEVaaa 


» 
Max. angle of attack Qmaqx | 
HI | | 
\ \ \ \ i \ 
iy ‘( iA \ | 1 
ay ) i \ hy 4 A \ fa iy 
ae | whet , 4 VW \r\ lg \ 
yo \ yf wit, ees | vi ee \ Ny r i 1 \ "| i 4 \ mil \ ; | \ 
Le eh WA \ ‘hy / Whth | Ven yy 7 
ne i Av Wis i Mis i (! WW \ Moat \ | 
vaniar ( i" t yl Mi Vi i" 5, 
Hydrophone signals t 
i} \ 
H2 A I, 
ag \\ NN 1A * i H 
uo vey WE ; WN Hy 4 V\ fh 4 IVI Ma 
‘ my eft lt A ‘i at h ey AL YP OM Yi Wy fy AM H\\ \W\ /\ 1 \ 
, ine, ste HedN My \ | wel) We \ 7S wh f | 
: Mt W , ef 
Sea re eae Os cr tating =—Synchronization flash End of film —= 


Recorded signals. ee EEE 


Freire ar : 
[1 1 1 1 4 50 scale units (su) | Max. extent. |Collapse(=max p) 


== pressure 25 Pa/su 


79 Pa/su 


79 Pa/su 


FIGURE 9. 


c | 


: == ——max extent.q 


A 


if h 


Ml 


—— 


max. p before final collapse 


cav. starts 


{i \ 


8 
c 
£ 
F 
a 
roy 
1S) 
c 
= 
o 
2 
3 
& 
o 
a 
a 
x 
cS] 
€ 
3° 
o 
c 
iS 
a 
3 
x 
5 
2 
a 
a 
o 
a 
bad 
cS] 
E 


Oscillating hydrofoil. Pressure signals and cavitation. O =m3era=u4on 


Pressure signals 


50 scale units (su) 


125 Pa/su 


pressure 


265 Pa/su 


265 Pa/su 


FIGURE 10. 


(approximately) that moment when maximum pressure 
is generated. For rapidly collapsing cavities the 
cavitation patterns shown existed 1/3-2/3 milli- 
seconds before the sharp pressure pulse. A note 

is also made as to whether or not the maximum 
pressure increase coincided with the final collapse 
(i.e., the complete disappearance of the cavity). 
The collapse velocity during the last stage is 
indicated by arrows: 


> = slow motion of the cavity boundary in 
the direction of the arrow 

>> = fast motion of the cavity boundary in the 
direction of the arrow 

>>> = very fast motion of the cavity boundary 


in the direction of the arrow 

At collapses with more or less spherical symmetry, 
arrows are placed opposite each other. 

To the right is shown the cavity growth time, Tg: 
and the collapse time, T,, for the complete cavity, 
measured by use of the time markings on the high- 
speed film. The’collapse time is measured from the 
time of maximum area extent to that time when the 
cavity generated the maximum positive pressure. For 
rapidly collapsing cavities this event coincides 
with complete disappearance of the cavity. This 
was not the case for slowly collapsing cavities; 


477 


max. p at final co 


a 
AC) 
i 

Cc 
= 
~ 

5 

ro 

x 

So 

E 


Oscillating hydrofoil. Pressure signals and cavitation. a = so 6 S 20. 


for these cavities the collapse times for complete 
disappearance are also given (in parenthesis). 


General 


The general character of noise and cavitation be- 
havior when the frequency of oscillation is varied 
is shown in Figures 9-14. The pressure signals 

from the cavitating hydrofoil are to be compared 
with signals from the non-cavitating hydrofoil 
(Figure 15) and with the curve in Figure 16, showing 
the schematic behavior of the pressure generated by 
a growing and collapsing cavity. 

In comparisons of generated pressure from non- 
cavitating and cavitating hydrofoils the most 
striking difference is often the high and sharp 
pulses generated at the cavity collapse. The 
generation of such pulses is obtained especially 
when fog, exceeds a certain value. Also the pressure 
increase corresponding to cavity growth and the 
pressure dip generated near maximum cavity extent 
are detectable. 

The generated pressure pulses were classified 
into three main types: 


478 


e Slow pressure increase at cavity collapse 
(normally obtained at fog, = 1-3 Hz) 

e Fast pressure increase (fosc = 4-7 Hz) 

e Very fast pressure increase, i-.e., sharp 
pulses (fos¢ = 7-15 Hz) 


Generation of High Frequency Noise 


Sharp pulses (i.e., high frequency noise) were 
generated in three main ways: 
A. By violent collapse of the main cavity (or 
a large part of it). 

B. By collapse of small spherical bubbles 
occurring independently of the main cavity. 
The bubbles generated rather strong pulses. 

Cc. By collapse of rather small irregular cavities 

separating continuously from the main cavity. 

Of greatest interest is the generation process 

A, which was obtained at high fo,,- The high and 
sharp pulses were generated in three somewhat 
different ways: 

Al. Separation of a rather large part of the 
main cavity at an early stage of the 
collapse. Thick cavity formations often 
separated in this way, especially if the 
cavity was long (large %may) and broken up 
by disturbances. At the end the collapse 
was often very violent and often followed 
by a violent rebound. Also the rebounded 
cavities (complex in form) cometimes 


Pressure signals 
50 scale units (su) 


+——Tose ———. @) 
Omax 


‘ee A, ye 
y W “hr a uA W 0 aly 
Va, ns Te sions area =O" un, 


cav. Pane 


ressure 156 Pa/s 


P - 
a 


time 


492 Pa/su 


492 Pa/su 


FIGURE 11. Oscillating hydrofoil. Pressure signals and cavitation. 


collapsed violently. An example of this 
behaviour is shown in Figure 13 for fog, = 
7 Hz (oscillation period 5). 

A2. Sharp pulses were also generated when a 
sheet collapsed towards the leading edge. 
The upstream cavity boundary was attached 
to the leading edge during the whole collapse. 
This process was normal at the conditions 
shown in Figures 11 and 12 and especially 
in cases where the main cavity was rather 
small. In these cases the whole collapse 
was orderly and without extensive separa- 
tions of cavity parts from the main sheet. 
After the collapse was completed a rebound 
of small cavities occurred about 10 mm 
downstream from the leading edge and not at 
the center of collapse as in the case of 
more symmetrical collapses. Also in cases 
where large cavities separated from the 
main cavity the remaining, rather smooth 
sheet often collapsed in this way (Figure 
10, 10 and 14 Hz, Figure 13, 7 and 10 Hz). 

A3. In cases where the smooth sheet attached 
to the leading edge was long and narrow it 
was also cut off from the leading edge. For 
the downstream part, the collapse then be- 
came more symmetric and violent and with 
a violent rebound (Figure 11, 10 Hz and 
Figure 13, 7 Hz). This process often 


occurred near the end of collapse. 
Spherical bubbles were very effective as genera- 


eo cs nn 


50 
(71) 
0.48 


Collapse oman nl 
3) 


max p at final coll max p at final coll, |max p before final coll! 9 


max p at final coll 


Pressure signals 
L111 1 4 50 scale units (su) 


156 Pa/su presse 
——cav. Starts 


131 Pa/su 


131 Pa/su 


FIGURE 12. 


tors of high frequency noise. This is discussed 
later in the text together with cavity area measure- 
ments. 

The generation of high frequency noise by small 
irregular cavities, continuously separating from 
the main cavity is the only generation process when 
fosc = 0. Also at low fos, (about 1-2 Hz) this 
process generated pulses. The separation of small 
cavities from the main cavity decreased with 
increasing fosc- 

When the high frequency noise was obtained it was 
always generated during the last part of collapse 
of the generating cavity (i.e., a bubble could col- 
lapse and generate high frequency noise during the 
growth of the main cavity). This is not surprising, 
but it should be mentioned that at studies of pro- 
peller cavitation it has been noticed that the growth 
of cavities in some cases also generates rather fast 
pressure variations which indicates that high volume 
acceleration can also occur during growth. 


Generation of Low Frequency Noise 


The generation of low frequency noise (vibration 
generating pressure disturbances at multiples of 
propeller blade frequency) can be identified by 
inspection of signals from non-cavitating conditions, 


479 


[tavitation nde 
Max extent. [Collapsetnax fe] [aii 


lmax=105 mm @) 


cpl| max p before final col 


c 
£ 
G 

3 

2 

a 
a 
5 
a 
x 
cS) 
E 
3G 
eS 
= 
~ 
ro] 
a 


Imax % 60 mm 


max p at final coll. 


Oscillating hydrofoil. Pressure signals and cavitation. ae = 2 eg, 


cavitating conditions, and the schematic pressure 
behavior shown in Figure 16. This is especially 
easy in cases where cavitation start is marked 
(Figure 9, 3 Hz, Figure 10, 7 Hz, Figure 11, 3 Hz, 
Figure 12, 7 Hz, Figure 13, 3 Hz) or where a non- 
cavitating period is followed by a cavitating one. 
In several cases it can be seen that a rather slow 
pressure increase is generated during the growth. 

When the volume acceleration is directed inwards, 
during a period around the maximum cavity volume, 
negative pressure is generated (for example Figure 9, 
3 Hz). This pressure variation is rather slow and is 
an essential part of the low frequency disturbance. 
Because of inertia effects in the motion of cavity 
walls this part of the motion will probably never 
contribute to really high frequencies. 

In most of the figures it can be seen that con- 
tribution to the low frequency pressure is also 
obtained from the collapse. Especially at low foc. 
the collapse seems important. The pressure increase 
during collapse is due to the outward-directed 
volume acceleration existing during the final part 
of collapse. This acceleration depends on the 
cavity geometry and the velocity of the cavity walls 
and it is in principle possible to obtain a collapse 
with constant volume velocity (no pressure generation), 
as well as a collapse with decreasing volume velocity, 
in which case a pressure increase is generated. It 
is supposed that both types of collapse can occur 


480 


Pressure signals 
L111 1 4 50 scale units (su) 


cay, starts 


=! 
4) 
o 
a 
w 
—t 
N 
o 
L 
=! 
w 
w 
oO 
—_ 
a 


1310 Pa/su 


Cavitation 


f 
a cls) 
| Max. extent. | Collapse («max plz 


) 


5 = 
a é 
3 : 
S fo} 
o a 
re] 
E 
FIGURE 13. Oscillating hydrofoil. Pressure signals and cavitation. Ce 4° @ = 5°. 
on propellers, depending on cavity geometry and medium-high frequencies from a propeller 
time variation of the surrounding pressure. (5-20 x blade frequency). The pressure 
The contribution from collapse obviously exists fluctuation seems related to the dynamics 
(see Figure 9, 1 and 2 Hz) but the quantitative of the main cavity, which at this stage was 
results especially at fo,, = 3-7 Hz must be used quite orderly. 
with much prudence, because of the resonant character 3. During the last part of collapse very sharp 
of the signal in these conditions. This is discussed pulses with durations less than 0.1 milli- 
in the Appendix. second were generated. At this scale of 
time, measurements and detailed observations 
of cavity behavior were not possible. Some 
Area Measurements of Some Cavities observations indicated, however, that the 
sharp pulses sometimes were generated by a 
For the condition a9 = a = 3° and rage = la ee rather well-ordered collapse. Figure 17 
some results from measurements of cavity area are shows an example of this behavior. The 
shown in Figures 17-23. The main cavity includes cavity was in this case attached to the 
the sheet and some small bubbles at the downstream leading edge during the whole collapse. 
edge, which follow the behavior of the sheet. 4. More complex cases are shown in Figures 18, 


Although the cavities in this condition were rather 
simple, with no large separations from the sheet, 
quite complex events often occurred during the 
last 1/2 millisecond of the collapse. 

Some comments on the figures will be made: 


1. From the shape of the area curves it can be 
seen that the growth of cavities was rather 
similar in all cases, while there are 
differences in the collapses. Compare, for 
example, Figures 17 and 20. 

2. It is seen that 1-2 milliseconds before 


final collapse a slow or moderately fast 
pressure increase was obtained. During this 
time collapse is fast, but measurable. This 
pressure fluctuation corresponds to low or 


21, 22, and 23. Several pulses were generated 
during a short time and it is impossible to 
separate the generating events (collapses 

and rebounds of several small cavities). 
Typical of these oscillation periods is 

that when the downstream cavity wall moves 
towards the leading edge, the cavity separates. 
into two parts, both attached to the leading 
edge. This separation was caused by a growing 
disturbance on the cavity surface. The 
disturbance grew from the downstream edge 
towards the leading edge. (See also Figure 
11). During the collapse some bubbles also 
separated from the downstream cavity edge 

and the disturbed area. These three cavity 


p 125 Pa/su A, 
Mi 


| W I Hu 

| | Wn rf 
At) 

yi” 

We) 


‘i 


i} 
{ 
Nghe PI) stl daimat q 


fosc= 


i wt 
fe A hy: 


a p 265 Pa/su | 
(=) 
w 
| B 
Withee ' 
A | ao 
ain 


p 265 Pa/su A 


Nh 


é an 1 Ma 


0 5 10 


FIGURE 14. Pressure signals during collapse. Expanded 
signals from Figure 10. 


15 t (ms) 


groups seldom collapsed exactly simultaneously 
or with the same violence. For example, in 
the cases shown in Figures 18 and 21 a part 

of the cavities was cut off from the leading 
edge during the last millisecond of the 
collapse. This resulted in violent collapse 
of the cut-off parts. 

From these examples it is understood that 
in a single oscillation period the character 
of the pressure signal is very sensitive to 
such things as simultaneousness and violence 
of separate cavitation events. Over many 
periods, normally used in measurements, the 
quantities are smoothed out to a mean value, 
which often is less sensitive to small dis- 
turbances. 

5. In some cases small bubbles and irregular 
parts separated from the main cavity and 
collapsed rather fast. In the case shown 
in Figure 22 a group of small bubbles behind 
the main cavity (cavity B) collapsed violently, 
simultaneously with the main cavity, and it 
is impossible to determine which of the 
cavities generated the main pulse. Examples 
of cavities that seemed rather fast, but 
only generated small pulses are shown in 
Figure 18 (B) and 19 (C). 

6. The most extensive rebounds resulted from 
cavities that were cut off from the leading 
edge and then collapsed fairly symmetrically. 
The cut-off normally occurred during the 
last one or two milliseconds and it often 
resulted in two cavities, one of which 
remained attached to the leading edge. The 


481 


p 26 Pa/su fosc = 1Hz 
aaa 50 su n 
| wih m4 
\ i i ea ft Ma Ah i 
II bint | \ "Ny th 
ail ne haiti 


i ii My Whi W, 


aa hh 


Time 


82 Pa/su n 


if 
H Le 


" Ae 
Maha th N 


82 Pa/su n n ices 2 Oh 
Ny svt 
Se, out o aye 
See “tin 
82 Pa/su fosc = 4Hz 
' i 
v Avs ve 
131 Pa/su ffosch=/AiZ, 
i] 1 
ca ‘ 
131 Pa/su fose = 10 Hz 
1 i 1 
fs *, oy d 4 ra 
rp ee \ye Sf 


265 Pa/su flosc = 116) Hz 
eae Miu ies clened 


FIGURE 15. Pressure signals from non-cavitating hydro- 


Poni, GO = 9 Ge a, 
(eo) 


Cavity volume 


4 


= Time 


Radiated pressure 


A 


= Time 


FIGURE 16. Schematic behavior of cavity volume and 
radiated pressure. 


482 


FIGURE 17. Cavity area and generated pres-— 
sure. Oscillation period -1l. 


Cavity area (cm?) 


70 


60 p 131 Pa/scale unit 


Cavity area (cm?) p ‘131 Pa/scale unit ag=3° 
50 a=? 
r fosc=15 Hz 
Ir 50 scale units 
mh hava, 
[ | 
[ on ty at 
ant a a a a 
L 0 5 10 15 20 25 30 35 
L t (milliseconds) 
20 
10 | 
| 
| 
i | 
0 L 
0 5 10 15 20 25 30 


50 scale units 


t (milliseconds) 


FIGURE 19. 
sure. Oscillation period 6. 


Cavity area and generated pres- 


30 


35 40 45 
t (milliseconds) 


FIGURE 18. Cavity area and gen- 
35 40 erated pressure. Oscillation 
t (milliseconds) period 4. 
Cavity area (cm?) p 131 Pa/scale unit Gp = 3° 
507 a=3° 
L fosc= 15 Hz 
50 scale units 

I leg 
30+ be +1 J SSS SS SS eee) 

0 5 10 15 20 25 30 


t (milliseconds) 


25 30 
t (milliseconds) 


Cavity area (cm?) dp =3° 
° 
50, p 131 Pa/scale unit @=3 
f fosc=15 Hz 
40+ 50 scale units 


30 


a frie] 
10 15 20 25 30 


20 — ees = {VTMIUUER GHEE 


Bmax =/22 mm | | 


a 


| | 
A | 
25 30 
t (milliseconds) 


Cavity area (cm?) 
707 


t P 


483 


FIGURE 20. Cavity area and generated pres- 
sure. Oscillation period 7. 


= 48mm 


Imax 


131 Pa/scale unit 


50 scale units 


t (milliseconds) 


FIGURE 21. Cavity area and gen- 
erated pressure. Oscillation OF 5 0 
period 12. 


- 2 4 
Canty aca (Gar) p 131 Pa/scale unit CoP o 
60 a =3 
fosc= 15 Hz 
j | 
50 50 scale units | 
40 


3 
t (milliseconds) 


a 
15 20 25 30 35 40 
t (milliseconds) 
FIGURE 22. Cavity area and generated pressure. 


Oscillation period 13. 


484 


FIGURE 23. Cavity area and generated pres- 
sure. Oscillation period 14. 


rebounded cavity (often a group of small 
cavities) collapsed after three to four 
milliseconds. Compared with the main cavity 
the area of the rebounded cavity was small 
(Figure 18 cavity C, Figure 22 cavity C and 
Figure 23 cavity B). The rebounded cavity 
often generated pulses of nearly the same 
height as the main cavity. 

7. The equipment was not designed to measure 
small and fast collapsing cavities such as 
small bubbles, but an example.of a diameter 
measurement of a bubble is shown in Figure 
24. The area (1a2/4) of the same cavity is 
plotted in Figure 19 (cavity A), where the 
sharp collapse pulses are also visible. 
Other examples of bubble collapses are shown 
in Figure 17 (time = t = 5 ms), 18 (t = 10), 
20 (t = 0, cavity A), and 23 (t = 0). Bubble 
collapses are also shown in Figures 9-13. 


Diameter (mm) 
8 


-2 -1 0 1 2 


-3 
Time (milliseconds ) 


FIGURE 24. Diameter of a spherical cavity. 
in Figure 19.) 


(Cavity A 


Cavity area (cm?) p 


70 


60 


50+ 


131 Pa/scale unit (cm2) 


50 scale units 


Nea 
See 


vo 


0 5 10 15 20 25 
t (milliseconds) 


-5 0 5 10 15 20 25 
t (milliseconds) 


The bubbles studied appeared just before or 
during the growth of the main cavity and the 
pressure pulses were then easy to identify. 
The bubbles normally rebounded once or twice. 
From the size of the bubbles and the generated 
pressure it is obvious that the bubbles are 
very effective as sources of high frequency 
noise. During the first life cycle, the 
bubble surface was smooth, but in the rebound 
cycles it became rough as reported by other 
authors. 


Dimensionless Presentation of Some Results 


The pressure generation at collapse is related 
to the violence of the collapse and it is then 
natural to study the collapse time, Te, for cavities 
generating different types of pressure pulses. Tc, 
given in Figures 9-13, is measured for the complete 
cavity, but in several cases it is only a separted 
part of the cavity that generates the main pressure 
pulse. Because of this simplification T. is probably 
not significant for the generated pressure in all 
cases. The intention was, however, to study the 
relevance of parameters for the complete cavity. 

In Figure 25 dif (its + Tg), (Ty = growth time), 
is plotted for the cavities shown in Figures 9-13. 
As seen the steepness of the curves tends to 
stabilize at a lower value for foo, resulting in 
sharp pulses. The growth and collapse are, however, 
not generally related to each other and Figure 25 
may thus give a distorted picture of T_-behaviour. 
In an effort to remove this drawback Wea also 
was plotted, where Ti is a hypothetical collapse 
time given by the formula for spherical cavities 
(Rayleigh 1917): 


Te 
Te +1g 
(A)a) ny a 
06 45) 8 
a 3 4 
(a) E : Z 
(_) complete collapse 
(0) no( ) collapse to max pressure 
0.5 


7 slow pressure increase 

” fast pressure increase 

A very fast pressure increase 
@ mean value of 8 samples 


o4p 


- bubbles 


03> 

O2-F 
By 
0.0 05 10 Reduced freq 1.5 
L n eS Cee es ees) 
0) 5 10 15 fose (Hz) 20 


FIGURE 25. Normalized collapse time. 


where 
Pg = surrounding pressure 
Py = vapor pressure 
U = undisturbed velocity 
pe = density of water 
Oo = cavitation number 


Of course this formula at best gives a time 
proportional to the collapse time of the sheet with 
maximum length, 2£m3y- As is shown in Figure 26 
the tendency is simlar to that in Figure 25. The 
conclusion is that at high f the collapse is 
mainly regulated by a surrounding pressure consider- 
ably higher than the pressure inside the cavity, 
which results in T./Tc' = constant and a violent 
collapse of the type predicted by classical theory 
[Rayleigh (1917)]. At low f5,, it can be supposed 
that during collapse the pressures outside and 
inside the cavity are approximately equal. Then a 
violent collapse will not occur and T./T,' becomes 
considerably larger than for a "free" collapse. 

If the cavity is considered as a monopole source 
the generated pressure, p, in the far field is 


d2v(t - =) 
P= G 2 (1) 
4tr at2 
where 

V = cavity volume 
r = distance between cavity and hydrophone 
c¢ = velocity of sound 
t = time 


Applying this and classical theory of cavity 
collapse it can be shown [Ross (1976) ] that the 
generated maximum pressure, Pmax, at certain con- 
ditions is given by 


max 
p = const 
max 


485 


where 
Rnax = the maximum radius of a spherical cavity 
AP = Po - Py 
Py = surrounding pressure 
Py = vapour pressure 


According to this 
" fs, AP 
P Y/ "max (2) 


would be an appropriate coefficient to study for 
different cavities in our case. 


+ : : 
Pp = maximum pressure increase at collapse 


oe = maximum chord-wise extension of the sheet 
cavity (for bubbles 25, = diameter) 


The parameters are: 


The distance r is measured individually for 

every collapse. 
1 DB 

MPSS OU G 2 Op) ma 

Inherent in the coefficient above is an assumption 
about the collapse dynamics and, as the dynamics 
are dependent on cavity type, there is no universal 
value for the coefficient (2). For our purpose 
the coefficient may be seen as a measure of the 
pressure generation efficiency of different types 
of cavities. For spherical cavities this coefficient 
was used by Harrison (1952) and Blake et al. (1977). 

Another treatment which leads to a dimensionless 
pressure coefficient is to suppose that a constant 
part of the potential energy available for collapse 
is radiated as noise [Levkovskii (1968)]. The 


dimensionless parameter derived from this assumption 
is 


6.0r 
Te max p 
1." 
: bn 
5.0 jo 9 
(ss 4 
Oo 4 3 
@ 4 5 
7 slow pressure increase 
7 fast pressure increase 
4.07 “A very fast pressure increase 
“1m mean value of 8 samples 
3.0 
2.0 
1.0 
0.0 
0.0 05 1.0 Reduced freq. 1.5 
L 4 4 aa Lt 
0 5) 10 15 foge (Hz)20 


FIGURE 26. Normalized collapse time. 


486 


ptr [Pam] ky & a co bubble 
4000 SNS) 5 a3 Sie ES: 
[ ATSS) 4 ptr [a] ¢) 3 
Oo 4 3 (yi 4 
e4 5 imax®F fA 8 
7 slow pressure increase [ @ 4 5 
7. fast pressure increase 7 slow pressure increase 
A very fast pressure increase 7. fast pressure increase c® bubble 
A very fast pressure increase 


3000 [ 


3.0 [ 


2000 
2.0 
1000 
10F 
=A bubbles 
0 
0.0 05 1.0 Reduced freq. 1.5 
o 5 10 1 fosc (Hz) 20 0.0 05 10 Reduced freq. 1.5 
+ 
FIGURE 27. Pressure p at collapse. Different 1 
conditions. ; ) 5 10 15 fog¢ (Hz) 20 
+ 
FIGURE 28. Pressure p at collapse. Different 
n va be 
3 7 IRE conditions. 
29/2) ae manera 
R ¥v 9 cAP 
max 
fy pe 93 
y por ‘ce max eh max? (5) 
p = density of water 
c = velocity of sound From the films it was observed that the cavity 
Other symbols as above thickness seemed proportional to the length rather 
Here it is necessary to know a time At propor- than to the square root of the cavity area and the 
tional to the duration of the pressure pulse. following coefficient was obtained in cases where 
With the use of At some information about the the area was measured. 
real collapse dynamics is introduced and therefore 
coefficient (3) may be somewhat more universal than ‘i 
(2). Note, however, that for the original use of Pit Te max p 
(3) similarity in cavitation was assumed. teal 
Of interest for future work is to what extent 3.0 a a a 
the final pressure behavior can be described by a Fi a t 
measured cavity data. In this case it is more NSB bubbles 
natural to think of methods to estimate a2v/at? in @ 6 3 
(1). It is then necessary to know V(t) or to assume Oo 5 bs 
a relation between a*v/at? and measured parameters, 7 slow pressure increase 
nh llapse time and vity size In thi 2.0 7 fast pressuresrincrease 
SSS ao. P A ea yo i a s : A very fast pressure increase 
paper only the cavity area A(t) is presented. As 
a first approximation it will be assumed that V(t) 
is proportional to a3/2 ox (see. From the measure- 
ments of A(t) attempts were made to estimate a*v/at? 
by difference ratios in the conventional manners. (&) 
This failed, due to uncertainty in A(t) during the 10- extreme 
final collapse. Then as a very rough assumption 
Vv 
GAY ~ 
=~ = const —* (4) 
at? an 
Cc 0 SSE 
0.0 05 1.0 Reduced freq. 1.5 


was tested. 

This is true only at very special circumstances. 
The assumption was, however, used and from (1) and 
(4) the following dimensionless pressure coefficient FIGURE 29. Pressure p- at collapse. Different 
is obtained conditions. 


(ft 
0 5 10 15 foge (Hz) 20 


487 


+r [Pa:m] reached the low value region (Figure 26). There 
is considerable scatter in generation efficiency. 
It must, however, be remembered that the plot is 
based on single cavitation events probably not 
always typical, the results must only be seen as a 
first hint of tendencies. The coefficient (3) gave 
tooo} + = a t aoratry results rather similar to those from (2) but with 
somewhat smaller dispersion. In Figure 29 it can 
be seen that with coefficient (5) the dispersion 
of the points was considerably decreased. 

In Figures 30-32 results from Figures 17-23 are 
plotted. Only the dimensionless coefficients (2) 
and (6) are shown and it is seen that both attain 
approximately the same values for similar pulses, 
but neither of them brings the values of oscillation 
periods 6 and 7 into agreement with the others. 

100 | (tees a Le The other coefficients give similar results. Also 
if the coefficients are based on values of area, 
| time, etc. closer to the final collapse, the scatter 
is not decreased drastically. The conclusion of 
| ical Iai this is that, in the prediction of noise by theory 
0 10 20 30 40 50 60 7, or model tests, good similarity in certain cavitation 
Maximum cavity area(em’) events is important, and that these important events 
are not generally described by such simple parameters 
as To and Vpax- 
Because it was not possible to estimate d2v/at2 


500 


_——— a on 1 
osc.per -1 7 6 13 4 2 «4 


aP 4 5 
FIGURE 30. Pressure p from different oscillation 


pamiests € 248 Hu, te ago 20: directly from measured values of V(t) functions of 
osc the type: 
es ge A Q V(t) = const[1 - cos o(t)] [o(t) is a polynomial 
Ee c max BA max max” (6) with six variable para- 
In Figures 27, 28, and 29 results are shown for meters ] 
the different conditions shown in Figures 9-13. were closely matched to nearly the whole collapse. 
ptr is shown in Figure 27 only to provide a reference The pressures then calculated by use of these 
for the other parameters. functions agreed fairly well with measured values 
Figure 28 shows that the generation efficiency in many cases. These simple computations also 
increased strongly at a certain fogc~ (or reduced demonstrated how sensitive the generated pressure 
frequency). The increase normally coincided with often was to the final behavior of V(t) and it was 
generation of very sharp pressure pulses and at easy to realize that parameters of the types dis- 
these f5,, the relative collapse time had also cussed above can only be "universal" if they are 
applied to fairly similar cavitation events. 
ptr 
[PAR 
wee Bitdlemaxp 
30 Amax'max 
20) 2.0 . ey 
1.0 10 ree a 
05 
| 
| 
YS Saal aaa: 
i 2 le 
0 10 20 30 40 50 60 70 


Maximum cavity area (cm?) 


tt" 


oscper -1 7 6 130 14 124 osc.per -17 6 3 (4 12 
bite + : P . + 
FIGURE 31. Pressure p from different oscillation FIGURE 32. Pressure p from different oscillation 
periods f S15 fe, Oo ses 3%, periods f = 9 fe, @. = fs =o. 
osc © osc ° 


488 


3. SUMMARY AND CONCLUSIONS FROM EXPERIMENTS WITH 
AN OSCILLATING HYDROFOIL 


1. The generation of sharp pulses was dependent 
of the oscillation frequency. At low 
frequencies no high and sharp pulses were 
generated and above a certain frequency very 
high pulses were generated. 

2. The sharpest and highest pulses were generated 
by cavities which separated from the main 
cavity and underwent a rather symmetrical 
and orderly collapse. Detailed studies 
showed, however, that a series of pulses was 
often generated, indicating that the collapse 
was not always simple at the very end. 

3. Very high pulses could also be generated by 
cavities that were attached to the leading 
edge during the whole collapse. 

4. The highest pressure generation efficiency 
was observed for spherical bubbles, which 
despite their smallness generated rather 
strong pulses. 

5. The sharp pulses were generated during the 
very last part of the collapse. 

6. Rebound of cavities was an important process 
for generation of sharp pulses. The most 
violent rebounds were obtained for separated 
cavities. 

7. Low frequency noise was generated during the 
growth, near the time of maximum cavity 
extent and during the rather late stage of 
collapse. Because of a disturbing resonance 
the importance of collapse was, however, 
difficult to determine. 

The basis of existing scaling laws for cavitation 
noise is mainly [see for example Levkovskii (1968) 
and Baiter (1974)]: 

1. Ideas from theory and experiment concerning 
the dynamics and radiation properties of a 
single cavity. 

2. Ideas concerning statistical properties of 
the pulse-generating events. 

The dynamics and radiation depend on cavity 
geometry, cavity size, and the surrounding pressure. 
Scaling laws based on simple theory deal with model 
scale and magnitude of surrounding pressure, while 
similarity has to be assumed in cavitation behavior. 

It has to be accepted that complete similarity 
in cavitation behavior will not occur, but if it is 
known which events in the cavitation process are 
crucial for generation of important pulses this 
will provide an indication of to what extent 
similarity is necessary for proper application of 
scaling laws. 

Of course these introductory experiments cannot 
supply the final and complete answer, but the results 
indicate that one of the most important factors is 
that the separation of a cavity into parts is 
correctly scaled, the reason being that these 
separations are often the starting points for violent 
collapses. 
this often begins at an early stage of the collapse, 
or is even initiated by disturbances during the 
growth of the main cavity. 

Parameters that determine tendencies to separation 
of cavities have only been studied to a limited 
extent, but it is clear that the combination of a 
long (chord-wise) cavity and high reduced frequency 
causes extensive separation of large parts from the 
main sheet. From the plots of collapse times and 


pressure generation efficiency, DeC/ AP ene: as 


Especially when large parts are separated, 


functions of reduced frequency it can be concluded 
that within special regions it is important that 
the time variations of the surrounding pressure be 
properly scaled. Such a scaling may be critical 
for the onset of separation of large cavity parts 
from the main cavity. 


4. NOISE FROM DIFFERENT CAVITATION SOURCES 
Introduction 


In order to gain more information concerning the 
noise emitted from a cavitating source, tests with 
four axisymmetric head forms and two hydrofoils 

have been carried out in SSPA cavitation tunnel No. 
1. The aim of these tests was to obtain well-defined 
and unambiguous types of cavitation, as bubble, 
sheet, and vortex cavitation. Comparisons of the 
noise levels from these different types of cavitation 
were made, as well as some investigations of the 
effect of free-stream velocity and gas content. 

The results reported here will only concern effects 
of the type of cavitation. 


Test Set-Up 


The tests were carried out in SSPA cavitation tunnel 
No. 1 test section, 0.5 m x 0.5 m. The noise was 
measured using arrangement 4 (hydrophone in water- 
filled box), see also Figure 1. In some of the 
later tests a flush-mounted hydrophone in the 
tunnel wall (arrangement 2) was used as well as a 
hydrophone in the flow field. Signals from the 
hydrophone(s) were registered by a tape recorder, 
but also directly analysed by a 1/3 octave band 
analyser and a narrow-band analyser. Main results 
given here are from the 1/3 octave band analysis. 

Tests were carried out for a water speed 9 m/s, 
but with some additional tests at 7.5 m/s and 11 
m/s. The gas content of the water at the tests 
was 10% and 40%, with some additional tests at 
higher gas content. 


Test Set-Up 


The first series of tests was carried out with 
axisymmetric head forms. The reason for this 

choice was that cavitation patterns for these bodies 
were well-known and well-defined from rather exten- 
sive tests [Johnsson (1972)]. The head forms used 
are given below, see also Figure 33. 


Head form Shape Cavitation Type of 
SSPA iden- of nose number for cavita- 
tification contour cav inception tion 
U1A hemispherical 0.67 sheet 
N39 flatt+elliptic 3:1 0.4 bubble 
N3 flat+elliptic 6:1 0.42 sheet 
N10 flattelliptic 4:1 0.43 sheet 


The head forms were attached to a cylinder and 
a faired afterbody, which were suspended from the 
tunnel roof via a thin wing. The main difficulty 
at the tests was the low cavitation numbers needed. 
At cavitation numbers below 0.4 fairly extensive 
cavitation occurred at the wing-tunnel roof junction 
and at other imperfections along the tunnel walls. 
This cavitation caused rather excessive background 


U1A 


Hemispherical 


N 39 (flat nose ) 


Elliptic 3:1 
N 
Q 
s 
oo 
c=) 
N3 (flat nose) 
Elliptic 6:1 
i 
\ 
(a) 
w 
Oo 
oe! 


N10 (flat nose ) 
Elliptic 4:1 


\ 


— 


@ 05 


FIGURE 33. Axisymmetric head forms. 


noise and made noise measurements almost impossible 
at low cavitation numbers. There is also some 
question whether such background noise from undesired 
cavitation was obtained at higher cavitation numbers 
than o = 0.4, when cavitation numbers are increased. 
With regard to these findings the results given 

here are limited to cavitation numbers o > 0.6 and 
only for decreasing pressure. 

In Figure No. 34 1/3 octave band noise spectra 
for cavitation numbers o = 1 and o = 0.6 are given. 
At o = 1.0 no visual cavitation was obtained and 
the noise levels are almost the same as for the 
empty tunnel (at the same velocity and cavitation 
number). At o = 0.6 the cavitation is well developed 
for the hemispherical nose, for the other head forms 
no cavitation can be visually observed. There are, 
however, rather large differences in noise spectra 
for the three "non-cavitating" head forms. Thus 
head forms N3 and N1O have noise levels 10 to 20 
dB above N39, for which the noise level is equal 
to non-cavitating or empty tunnel conditions. These 
differences cannot be attributed to unwanted cavita- 
tion on the wing or tunnel walls. In that case the 
noise levels for head form N39 should also have 
increased. The conclusion is thus that head forms 
N3 and N1O have audible but not visible cavitation. 

From the tests with axisymmetric head forms it 
can be concluded that the cavitation numbers will 
be low, which implies that effects of unwanted 
cavitation will increase background noise levels 
and violate results for the cavitating head forms. 


Tests with Hydrofoils 


In order to obtain cavitation at higher cavitation 
numbers tests with two wings have been carried out. 
Using wings, vortex cavitation can also be obtained. 
The problem is here rather to obtain other types of 
cavitation without getting vortex cavitation. 


489 


One of the wings tested has cambered sections 
and elliptical planform, and the other has symmetric 
sections and trapezoidal planform, see Figure 35. 


Wing Angle Cavitation Type of 

(SSPA ident- of number for cavitation 

ification) attack, a cav inception 

Elliptic, =e =2 sheet 

cambered 

(16-12.12) the De) vortex 
LW22 3 vortex 

Trapezoidal, 

symm rounded (0) 0.5 bubble 

tip (K7 Vbl1*) Bo SL 5) vortex 

Trapezoidal 

symm with 

end plate Sse =1.2 sheet 

(K7 Vp3*) 


(*The wing K7 was tested with rounded tip, Vbl, 
and a small end plate, Vp3, see also Figure 35). 


For the comparison of noise emitted from different 
types of cavitation it is important that these 
comparisons be made at the same cavitation number. 
One inherent difficulty is that pure bubble cavita- 
tion seems to be possible to obtain only at rather 
low cavitation numbers compared with the other 
cavitation types. 


dB re10°° Pa 
150 r 
Cav. number O=1 


140 - 


WD No cavitation 
19 — T 
05 2 5 10 40 f (kHz) 
dB re 10° Pa 
150 
Cav. number O=0.6 
140+ Sheet cav. (U1A) 


130 IP supra lice = 


(N10) 
sedli (N3) 
Nob 
No cav. (N 39) 
100 is 1 
05s 2 5 10 40 f (kHz) 
FIGURE 34. Axisymmetric head forms, cavitation noise 


(1/3 octave band). (Free stream velocity 9 m/s, 
gas content 10%.) 


490 


Wing 16-12.12 
Elliptic , cambered 


Wing K7 


Trapezoidal , symmetric 


Tip shape: 
Rounded K7 Vb1 
S—— 


End plate K7 Vp 3 


FIGURE 35. Wings. 


Results from the tests are given here for five 
cavitation numbers, o = 3, 2.5, 2, 1.5, andl. 

The free stream velocity was 9 m/s and the gas 
content ratio was 10%. Results are given as faired 
curves for the noise levels from 1/3 octave band 
analysis. 

For cavitation number o = 3 (Figure 36) only 
the cambered wing 16-12.12 at a = 172° cavitates 
with vortex cavitation. Noise levels for the wings 
with no cavitation are of the same order as for the 
empty tunnel. The vortex cavitation at a = 172° 
gives an increase in noise levels of 15 to 20 dB 
compared with non-cavitating conditions. 

At o = 2.5 the wing 16-12.12 has vortex cavita- 
tion at a = 2° and a = 172°, Figure 36. It is of 
interest to note that the vortex cavitation at 
a = 2° is not attached to the wing tip but starts 
behind the wing. This vortex can only be obtained 
when the pressure in the tunnel is increased 
(increasing cavitation number). The increase in 
noise level due to vortex cavitation here is also 
15 to 20 GB. 

For the cavitation number o = 2 the wing 16-12. 
12 has vortex cavitation at a = 172°, intermittent 
vortex cavitation at a = 2° and sheet cavitation 
at a = -2°. The vortex cavitation gives an increase 
in noise level of the order of 15 dB. The sheet 
cavitation at a = -2° increases the noise levels 
at higher frequencies (f > 5 kHz), 10 to 15 dB 
above the level for vortex cavitation, see Figure 
Bike 

At o = 1.5 it can be noted that in some cases 
no pure types of cavitation can be obtained. Thus, 
wing 16-12.12 gives sheet cavitation at a = -2°, 
vortex cavitation at a = 2° and vortex and bubble 
Cavitation at a = 172°. Results in Figure 37 show 
the largest increase of noise levels for sheet 
cavitation. Note also the differences between 


dB 
150 


140 


130 


120 


110 


100 


dB 
150 


140 


130 


120 


110 


100 


FIGURE 
(Free 


re 10° Pa 
Cav. number O=3 
Vortex cav. (16-12.12 , @=172°) 
No cavitation 


f(kHz) 
re 10° Pa 
Ir 

Cav. number O=2.5 
Vortex cav. (16-12.12, 
Fe os 5 
_- <— Vortex not Q=172°, a= 2°) 
<a attached (A= 2°) 


No cavitation 


05 10 f (kHz) 


36. Wings, cavitation noise (1/3 octave band). 
stream velocity 9 m/s, gas content 10%.) 


dB re 10° Pa 
150 
Cav. number O= 2 
140 
Sheet cav. (16-12.12,@=-2°) 
130- 


Vortex cav.(16-12.12, 


120 
@=172°, a=2°) 
eile No cavitation 
1 
100 05 2 5) 10 40 f(kHz) 
dB re 10° Pa 
150 Cav. number O=1.5 
Sheet cav. (16-12.12 ,@=-2°) 
10F 
_- Vortex and bubble - 
= cav. O decreasing. 
IO (16 -12.12 , @=172") 
120 Vortex cav. (16-12.12,@=172", 
O increasing ,@= 2°) 
110 
No cavitation 
108 0.5 2 5 10 40 f (kHz) 
FIGURE 37. Wings, cavitation noise (1/3 octave band). 
(Free stream velocity 9 m/s, gas content 10%.) 


decreasing and increasing cavitation number for 

a = 172°. For decreasing 0 small cavitation bubbles 
are obtained, which increase the noise level about 
15 dB compared with increasing o. 

From the results at cavitation number o = 1.0 
(see Figure 38) it is obvious that bubble cavitation 
gives the largest increase in noise levels from 
25 dB at low frequency (500 Hz) to 55 dB at high 
frequency (40 kHz). Sheet cavitation gives less 
increase but depends on the intensity of the cavita- 
tion. Thus for wing 16-12.12, a = -2°, the sheet 
cavitation is extensive and gives an increase from 
20 dB at low frequencies to 50 dB at high frequencies 
compared with non-cavitating condition. For wing 
K7 Vp3 the sheet cavitation is concentrated at the 
leading edge and an increase in noise level is only 
obtained for higher frequencies (> 2 kHz) and the 
increase at 40 kHz is of the order of 25 dB. The 
differences in noise level for wing K7 Vbl for 
increasing and decreasing cavitation numbers can be 
attributed to differences in cavitation patterns. 

No pure vortex cavitation could be obtained at 
cavitation number o = 1.0. 


Conclusions from Tests with Head Forms and Hydrofoils 


Tests with head forms are less suited as rather low 
cavitation numbers are needed. This may cause 
problems with high background levels due to undesired 
cavitation on tunnel walls etc. Tests with hydrofoils 
can be used to obtain effects on noise levels from 
different types of cavitation. There may, however, 
be some problems in obtaining pure cavitation types. 

Vortex cavitation gives an increase in noise 
level of about 20 dB. It should be noted that 
differences in vortex cavitation can be obtained 
for increasing and decreasing pressure, which also 
show as differences in noise level. Also a vortex 
not attached to the wing causes increases in noise 
level. The increase in noise level due to vortex 
cavitation seems to be less for lower cavitation 
numbers. 

Sheet cavitation gives substantially higher 
levels than vortex cavitation. The extent of the 
sheet has some influence on the noise level. For 
a fairly large sheet increases in noise level of 
20 dB at 500 Hz to 50 dB at 40 kHz are obtained. 

For a small, leading edge sheet the increases in 


dB re 10° Pa Cav. number O=1 
160 Bubble and vortex 
cav. (16 -12.12 ,a=172°) 
150 
Sheet cav 
140 - (16-1212, @=-2°) 
130 
al Sheet cav.(K7 Vp 3 a=5) 
Vortex and sheet cav. 
increasing O. 
NOP (K7 Vb1 a=5) 


No cavitation 


1 
ge f (kHz) 


0.5 2 5 10 40 


FIGURE 38. Wings, cavitation noise (1/3 octave band). 
(Free stream velocity 9 m/s, gas content 10%.) 


491 


noise level are obtained for higher frequencies 
(£ > 2 kHz) and for 40 kHz the increase is 25 dB. 

Bubble cavitation gives the largest increases 
in noise level. Levels are for this case 5 to 10 
dB above the levels for sheet cavitation. 


ACKNOWLEDGMENT 


This work is part of the research program at the 
Swedish State Shipbuilding Experimental Tank and 

the authors are indebted to Dr. Hans Edstrand and 
Mr. H. Lindgren for making this study possible. 

Part of the work reported here has been carried 

out with financial support from The Defence Material 
Administration of Sweden. 

The authors would also like to express their 
sincere thanks to those members of the staff at SSPA 
who have taken part in the investigations and the 
analysis of the material. 


REFERENCES 


Baiter, J.-H. (1974). Aspects of Cavitation Noise. 
Symposium on High Powered Propulsion of Large 
Ships, Part 2, December 1974, Wageningen, The 
Netherlands. Publication No. 490, Netherlands 
Ship Model Basin, Wageningen, The Netherlands, 
pp. XXV 1-39. 

Blake, W. K., M. J. Wolpert, and F. E. Geib (1977). 
Cavitation Noise and Inception as Influenced by 
Boundary-Layer Development on a Hydrofoil. J. 
Fluid Mech. 80, 4, pp. 617-640. 

Harrison, M. (1952). An Experimental Study of 
Single Bubble Cavitation Noise. J. Acoust. Soc. 
Am. 24, 5; 776-782. 

Itd6, T. (1962). An Experimental Investigation into 
the Unsteady Cavitation of Marine Propellers. 
Proceedings of IAHR-Symposium, Sendai, Japan, 
1962, Cavitation and Hydraulic Machinery edited 
by Numachi, F., Institute of High Speed Mechanics, 
Tohoku University, Sendai, Japan, 439-459. 

Johnsson, C.-A. (1972). Cavitation Inception Tests 
on Head Forms and Hydrofoils. Thirteenth Inter- 
national Towing Tank Conference. Proceedings 
Volume 1 edited by Schuster. S., and M. Schmiechen. 
Versuchsanstalt fur Wasserbau und Schiffbau, 
Berlin, Germany, 723-744. 

Lehman, A. F. (1966). Determination of Cavity 
Volumes Forming on a Rotating Blade. Eleventh 
International Towing Tank Conference, Tokyo 1966, 
Proceedings edited by Kinoshita, M., Yokoo, K. 
The Society of Naval Architects of Japan, Tokyo, 
Japan, 250-253. 

Levkovskii, Y. L. (1968). Modelling of Cavitation 
Noise. Sov. Phys.-Acoust. 13, 3; 337-339. 

Rayleigh, Lord, (1917). On the Pressure Developed 
in a Liquid During the Collapse of a Spherical 
Cavity. Phil. Mag. 34, 94-98. 

Ross, D. (1976). Mechanics of Underwater Noise. 
Pergamon Press Inc., New York, USA, p. 218. 

Tanibayashi, H., and N. Chiba (1977). Unsteady 
Cavitation of Oscillating Hydrofoil. Mitsubishi 
Technical Bulletin 117. Mitsubishi Heavy 
Industries, Ltd. Tokyo, Japan. 


492 


APPENDIX 


LOW OSCILLATION FREQUENCIES MAINLY GENERATING 
RATHER SLOW PRESSURE PULSES 


The following observations were typical for fy, = 
1-3 Hz and ag = 3°, & = 4° (Figure 9) but most of 
the results are also valid for other angle conditions: 

1. The maximum pressure increase is generated 
before the sheet cavity has disappeared 
completely. At the moment of maximum pressure 
increase the collapse Slowed significantly 
and the rest of the collapse was very slow. 
Due to hysteresis the total collapse time 
was sometimes longer than the growth time, 
T.. Typical for the collapse from maximum 
extent to maximum pressure was Way (ls + Tg) 
> 0.4. The sheet cavities were attached to 
the leading edge during the whole collapse 
and only small parts were separated from the 
downstream cavity edge. 

Already during growth a large part of the 

cavity is disturbed and consists of one part 

with a smooth surface and one with thick 
irregular cavity formations. From this total 
connected cavity, small parts were separated 
both during growth and collapse. Only a few 
of these parts collapsed violently, which is 
also confirmed by the pressure signals which 
do not contain many sharp pulses during 
growth and first part of collapse. 

At very low f,., (1-2 Hz) these contin- 
uously occurring collapses of small cavities 
were, however, the only source of high- 
frequency noise. At these conditions also 
most sharp pulses were obtained in the 
hydrophone (H2) near the trailing edge. 

3 AG Eose = 2 and! 4) Hz the pressure ancrease 
often ends with a sharp pulse. The pulse 
was, however, not caused by an orderly and 
violent collapse of the main cavity, but 
instead by small cavities that separated 
from the main cavity and then collapsed 
separately. It was also observed that these 
rather violent collapses of small cavities 
mainly occurred during the time when the 
pressure was high owing to main cavity col- 
lapse. 

On a more expanded time scale it can also 
be seen that the sharp pulse is superimposed 
on a slower pressure increase. If not very 
clear, this tendency is still detectable in 
the 7 Hz-condition in Figure 14. This figure 
shows the pulse (oscillation period 6) in 
the 7 Hz-condition shown in Figure 10, but 
with the time axis expanded 40 times. 

4. The cavitation sketches in Figures 9-13 show 
that for f,., $ 4 Hz the cavitation extent 
was approximately independent of fosc, but 
that at higher fog, the cavity did not develop 
to the full size. One reason for this may 
be that the time variation of the dynamic 
angle of attack is altered with OSG 

5. Characteristic of low fosc is also the fact 
that collapsing cavities show little tendency 
to rebound. Rebound is only obtained in 
small bubbles. 


to 


HIGH OSCILLATION FREQUENCIES MAINLY GENERATING 
SHARP PRESSURE PULSES 


Below some observations are reported regarding the 
conditions ag = 3°, @ = 4° and £,,, = 10 and 14 

Hz (Figure 10). Many of the results are also valid 
for other similar conditions. Typical observations 
are: 

1. The sharp pulses are often much higher than 
slow pressure variations. 

2. The duration of the final part of the sharp 
pulses seems (as long as can be determined 
in the recording) independent of fosc (Figure 
14). For the earlier parts of the cavitation 
period the dependency on ERS is more complex 
due to different cavity sizes etc. 

3. For this condition (ag = 3°, &@ = 4°) the 
complete change of cavity dynamics and 
pressure generation occurred between fog, = 
7 and 10 Hz (Figure 14). At 7 Hz the cavity 
mainly collapsed towards the leading edge. 
At 10 Hz a large part consisting of thick 
formations separated and performed a violent 
collapse at the middle of the hydrofoil (B 
in Figure 14). This collapse occurred about 
1.4 milliseconds later than the collapse of 
those two parts (A) of the sheet that were 
attached to the leading edge during the 
whole collapse. Also these two parts col- 
lapsed rather violently, but a small pulse 
was generated. The thick separated cavity 
(B) consisted of several parts that did not 
collapse exactly simultaneously and, thus, 

a series of collapse and rebound pulses was 
generated. A significant rebound was only 
obtained from the separated cavity. The 
group of rebounded cavities collapsed rather 
slowly, resulting in a small pulse about 

3.5 milliseconds after the collapse of the 
separated cavity (B'). In some oscillation 
periods the separated cavities and those 
attached to the leading edge collapsed almost 
simultaneously and it also happened that 
high pulses were generated at the collapse 
of rebounded cavities. 

4. The cavitation behaviour at fy., = 14 Hz is 
approximately similar to that at 10 Hz 
(Figures 10 and 14). The thick formation 
(C) separated and collapsed at a later stage. 
The first pulse (Figure 14) was generated 
by the outer cavity (A) attached to the 
leading edge. About 1.4 milliseconds later 
the other cavity (B) attached to the leading 
edge collapsed. This cavity was complex 
and generated a series of pulses. First 
about 3.5 milliseconds after the first pulse 
the thick formation (C) collapsed, generating 
a sharp pulse. No violent collapses were 
experienced by rebounded cavities in this 
case. The overall impression from these 
two conditions with fog, = 10 and 14 Hz is 
that normally the separated thick cavities 
generated the highest pulses, but that in 
some cases pulses of almost equal height 
were generated by cavities attached to the 
leading edge. 

Another behavior of the signal from the cavitating 
hydrofoil is a low frequency variation (about 23 Hz) 


that seems rather independent of f,,,- Sequences 


containing cavitating as well as non-cavitating 
periods indicate that the fluctuations were generated 
by cavitation (Figure 10, 7 Hz, Figure 11 and 12). 
Inspection of the films shows, however, that no 
cavitation is visible on the hydrofoil or about 

0.5 chord-lengths downstream it (Figure 9, 11, 3 Hz). 


493 


Here the most probable cause is that the cavitation 
started resonance vibrations in some structure. 
These vibrations probably cause disturbing errors 
in the pressure signal at some f,,,, mainly in the 
region 3-7 Hz, and quantitative results from such 
conditions must be used with care. 


Cavitation Noise Modelling at 
Ship Hydrodynamic Laboratories 


Gavriel A. Matveyev 
and 


Alexei S. Gorshkoff 


Krylov Ship Research Institute 


Leningrad, USSR 


ABSTRACT 


Theoretical and experimental correlation of visual 


and accoustical effects of cavitation are considered. 


The Froude similarity is treated critically because 
of the pressure effects on the coefficient of cavity 
energy transformation into cavitation noise as well 
as because of the increase of noise absorption or 
cavitation resistance of water. Though in large 
cavitation tunnels which have no free surface the 
nonstationary boundary conditions can be reproduced 
less perfectly, their capability of simulation 

at full-scale pressure is regarded as the leading 
factor. It is suggested that extrapolation formulae 
should take into account the effect of the rate of 
pressure increase (or pressure gradient) in the 
cavity collapse area. This corresponds to an 
increase in the square of acoustic pressure on the 
model, compared to the prototype, inversely pro- 
portional to the linear scale of modelling. 


1. COMPARISON OF VISUAL AND ACOUSTIC EFFECTS OF 


CAVITATION 


The occurrence of strong visual and noise effects 
of cavitation are usually considered to be coinci- 
dent. When this coincidence is actually the case, 
it provides certain conveniences. The measurement 
of noisiness makes it possible to detect cavitation 
on structural elements not easily accessible for 
inspection. Visual observation of cavitation on 
models is used for the prediction of noisiness of 
various prototypes. However, the experiments 
involving visual and acoustical recording of 
cavitation indicate that there may be a considerable 
discrepancy between these two manifestations of 
cavitation. It is interesting to discover the 
nature and the cause of the discrepancy by means 

of a mathematical model of an elementary cavitation 
process which is described by the well-known 
differential equation of a single spherical cavity 
growth 


494 


jw 


(1) 


i] 

i 
Ulr 
LAS 

ue} 
fo) 

i} 

ue) 
Q 
es 

i} 

ne} 


Here R and R, are the cavity radius and its initial 
value, respectively; p and py are the variable 
component and the initial value of ambient pressure; 
Par ©, Y, and v are vapor pressure, density, and 
the surface tension and kinematic tension coefficient, 
respectively. 

Computations were made by equation (1) for the 
negative pressure pulse 


1-1 t 
INE) SES! 7 ee 


(2) 
which is characterized by the time scale, T, and 
the amplitude, Pm: Such a pulse represents the 
region of negative pressure having the length, L, 
on the profile with the maximum negative pressure 
coefficient, Coms in the flow with velocity, U: 


2 
it 
pus is) T 


Linearization of Eq. (1) with respect to 6 = z2-z 
for the small-amplitude oscillation frequency gives 


(3) 


According to (3), oscillatory properties of the 
cavity disappear at pp = 1 ata when Ry < 10-© m. 
Bearing in mind that the natural period is limited 
by the pulse duration, T, computations were made 
for 


Po = 10" kg/m?; R, = 107°+1073 m; ~ = 1073410-1m; 


V = 10+20 m/s; Cc =o 


pm 


The following value is taken as a measure of the 
accoustical effect of an elementary cavitation 
process to an accuracy of the potential energy 
transformation coefficient for the maximally expanded 
cavity: 23 5/43 

G_ = 10 Log a (4) 
R? 
n 


Here, Rn is the threshold value of the cavity radius 
which, for the sake of convenience and without 
limiting the generality of conclusions, is taken as 
10-6 m. For large Rp/Ro values this measure differs 
only slightly from a simpler measure used in Figure 1 


Vy as 5 
Gh 302g (5) 


The threshold of the visual observation of cavita- 
tion is taken as Re = 1073 m, which coincides with 
the upper limit to the size of the cavitation nuclei 
under study. For the chosen measure of acoustic 
effect this threshold corresponds to 90 dB. Since 
the resolution of vision is limited by angular 
dimension, the measure of the visual effect where 
the distance to the object of observation remains 
constant is the first order, linear dimension of 
the cavity. Hence, when the origins coincide 

al 


Giese 
B 3.7 @) 


and the processes below the level of G, = 90 dB are 
out of visual observation. 

Thus, leaving out of account the actual signal- 
noise ratios, the acoustical recording makes it 
possible to penetrate much deeper (by 2-3 orders) 
into the "microcavitation" region. 

Worthy of notice is the qualitative similarity 
of the curves shown in Figure 1 to the experimental 
curves of cavitation noise increase against velocity 
which are given below, as well as by Sturman's data 
(1974). It is evident that at an early cavitation 


p= 4°) 


FIGURE 1. Calculated comparison of visual and acous- 
tical effects of elementary cavitation process in a 
limited region of negative pressure. 


495 


stage the predicted levels drop by 20 dB with the 
velocity decreasing 10-fold. This stage is usually 
regarded as free of cavitation. 

With the increase of velocity there comes a stage 
which is sometimes referred to as "true" cavitation 
and in which the most intensive growth of cavities 
and cavitation noise is observed. This stage corre- 
sponds to a decrease and loss of static equilibrium 
of the cavity. 

At the third stage the intensive cavity growth 
ceases and asymptotic saturation of the acoustic 
effect occurs due to the fact that the size of the 
cavity is nearing that of the zone of negative 
pressure. The asymptotic values of saturation shown 
in Figure 1 correspond to the rough estimation 


L 
Gre = 15 +15 Hee + 302g a (6) 
n 

As to the relationship between visual and noise 
manifestations of cavitation, Figure 1 allows one 
to assert that: 

- at sufficiently high levels of ambient noise 
the acoustic detection of cavitation may coincide 
with the visual detection or takes place even later; 

- potentially, at a fairly low level of the 
ambient noise, the acoustic manifestation of cavita- 
tion must be detected much earlier than the visual 
one. 

In particular, the acoustic effect of cavitation 
can be rather strong (e.g., an increase of noisiness 
by several dozens of decibels) in the case of "micro- 
scopic" cavitation invisible to the eye. 

The indicated values are largely conditional as 
the threshold of visual detection may differ under 
different conditions. Nevertheless they are close 
to those obtained under laboratory conditions. 

It is of interest that Figure 1 reveals such a 
contradictory phenomenon as vagueness in respect 
to cavitation inception. At high levels the curves 
for various nuclei coincide, so for a more correct 
determination of cavitation inception one should try 
to reduce rather than to increase the accuracy of 
recording methods. The increase of accuracy, as is 
shown in Figure 1, brings about increasing ambiguity 
of cavitation inception and expansion of the vague- 
ness region to cover an increasing range of veloci- 
ties. However, as the accuracy decreases, more and 
more small zones of cavitation inception are left 
out of control. 

The above analysis simplifies the actual processes 
and can be at variance with them mainly due to the 
fact that the coefficient of cavity potential energy 
transformation into acoustic energy is not constant 
being a complex function of many parameters [Benia- 
minovich et al. (1975)]. Specifically it may have 
a much greater value for small cavities as compared 
to larger cavities. 


2. EXPERIMENTAL STUDY ON MODELS 


There is an urgent need for an effective and well- 
founded classification of a great variety of forms 
and types of cavitation which substantially differ 
in the mechanism of nonstationarity giving rise 
to noise and having other practical consequences 
of cavitation. 

The following brief list of the forms and types 
of cavitation represents a more or less established 
practice with respect to marine propellers [Goncharov 
et all. (1977) ])- 


496 


——-— - bubble cavity 
- sheet cavity 


Uniform flow 


FIGURE 2. Development of noise and visual manifesta- 
tion of cavitation at constant pressure vs. velocity, 
in- reference to the conditions of cavitation noise 
detection. 


According to the location of cavitation zones: 

- vortex cavitation (in the cores of tip and 
axial vortices), 

- leading edge cavitation (on the suction side 
and pressure side at the leading edge), 

- blade-profile cavitation (in the region of 
large blade thicknesses) . 


—-—-— - bubble cavity 
- sheet cavity 


Nonuniform flow 


FIGURE 3. Development of noise and visual manifesta- 
tion of cavitation at constant pressure vs. velocity, 
in reference to the conditions of cavitation noise 
detection. 


- root cavitation (at the blade roots). 

According to cavity pattern: 

- bubble cavitation (with cavities moving with 
the flow through negative or increased pressure 
zones), 

- sheet cavitation (with cavities which on the 
average are motionless in relation to the propeller). 

By steadiness (uniformity) of the incoming flow: 

- steady cavitation (noise and other effects 
result from the inner unsteadiness of the cavity 
which is steady on the average), 

- unsteady cavitation (noise and other effects 
result from the cavity pulsations at an almost 
regular frequency, the phenomenon of cavitation 
buffeting), 

- cavitation in an unsteady flow (noise and other 
effects here again result from the pulsations as 
well as from the probable disappearance of cavities 
with the frequency of flow condition change). 

It seems extremely difficult to provide a com- 
parative description of noisiness for about three 
dozen cavitation types characterized only by the 
above-mentioned features. Some guidance is given 
by the experimental data presented in Figures 2 to 
Bs 

In steady-state conditions the bubble cavitation 
types are the most noisy (Figure 2). Among cavita- 
tion zones of different locations, vortex cavitation 
types are the least noisy (Figure 3), whereas 
pressure-side, leading-edge cavitation types are 
the most noisy (Figure 4). 

In an unsteady (non-uniform) flow the relation 
between the noisiness of sheet cavitation and that 
of the bubble type is different (Figure 5). 

The higher noisiness of the pressure-side leading- 
edge cavitation is accounted for by the rapid 
increase of pressure behind the suction zone (high 
gradient), which is typical of these conditions. In 
case of the bubble structure of a cavity this rapid 
pressure increase is accompanied by the increase 
of acceleration during the collapse. In case of 
the sheet structure it is accompanied by the 


~o- - vortex cavitation V 


—e - leading-edge 
cavitation 


FIGURE 4. Development of noise and visual manifesta- 
tion of vortex and leading-edge cavitation appearing 
in succession in a uniform flow at constant pressure. 


- cavity on the v 
pressure-side of 
the leading edge 


—o 


—eo - cavity en the 
suction-side of 
the leading edge 


FIGURE 5. Development of noise and visual manifesta- 
tion of cavitation on both pressure- and suction-side 
in a uniform flow at constant pressure. 


unsteadiness of even small size cavities due to 
closure behind the maximum suction zone. 

The change in the relative noise intensity of 
sheet and bubble cavities depends upon the fact 
that in the case of the bubble cavity structure the 
unsteadiness varies but slightly, whereas the volume 
of sheet cavities begins to severly pulsate. Passing 
over to the unsteady flow, we may even observe the 
reduction of bubble cavitation noise. This occurs 
when one portion of the propeller gets free from 
the cavity whereas, on the other portion thereof, 
the intensive development of cavitation is not 
accompanied by an increase of noise due to a satura- 
tion effect. 

Individual points on the graphs shown in Figures 
2 to 5 indicate moments of the first visual detection 
of cavitation. As is seen, in a large cavitation 
tunnel where the measurements were made, the above 
conclusion that the noise comes ahead of the visual 
detection of cavitation is to a variable degree 
valid for any type of cavitation. 


3. MODEL-PROTOTYPE CORRELATION AND COMPARISON OF 
MODEL-TEST RESULTS WITH FULL-SCALE DATA 


It is usually assumed [Levkovsky (1968) and 

Sturman (1974)] that the fraction of the cavity 
potential energy converted into cavitation noise 
(coefficient of transformation) is the same for 

the model and the full scale ship. Experience con- 
firms the validity of the conflicting conclusions 
[Beniaminovich et al. (1975)] that are indirectly 
confirmed in some works. The coefficient of cavity 
energy transformation into noise proves to be 
strongly dependent on the absolute pressure, Po- 

It is this fact, that was used by Beniaminovich 

et al. (1975) for explaining the reduction of the 
transformation coefficient by several orders with 


497 


a decrease of pressure, Por from 1 to 0.4 ata. ie 
is also emphasized that at sufficiently low Pp, the 
collapse of cavities is not necessarily accompanied 
by shock wave generation. 

Vacuum noise measurements, when performed in 
ship hydrodynamics laboratories engaged in cavitation 
research, show inadmissible noise absorption in 
the facility water unless measures are taken to 
insure additional removal of gaseous nuclei of cavi- 
tation from the water. By intensified water degassing 
the absorption may be reduced to an acceptable level, 
but the resulting growth of cavitation resistance 
of water leads to a drastic change of conditions 
for inception and development of cavitation [Gorsh- 
kof£ and Lodkin (1966)]. In view of the complicated 
character of absolute pressure effects on the 
coefficient of cavity energy transformation into 
noice it appears to be good practice to perform 
cavitation noise measurements at a full-scale value 
of pressure. 

That the Froude similarity will not be fulfilled 
under these conditions, can be accepted provided 
that adequate means are available for the description 
and reproduction of the conditions of flow non- 
uniformity behind the hull. This approach, used 
in a large cavitation tunnel in combination with 
correlation methods recommended by Levkovsky (1968), 
Sturman (1974), has shown that overestimated cavita- 
tion noise levels are predicted in this case. This 
was found to arise from the fact that the coefficient 
of cavity energy transformation into noise is 
approximately proportional to the rate of pressure 
growth leading to the cavity collapse. In modelling 
by the Froude method this pressure growth rate 
decreases as VL. 

In case of large-scale modelling the comparison 
of model-test and full-scale data may not have 
revealed this discrepancy among other more pro- 
nounced ones. One can use pressure gradient instead 
of the rate of pressure variation with time. Then, 
for Froude similarity, the noise level model-to-full- 
scale extrapolator coincides with that used by Sturman 
(1974). Not so with modelling at full-scale absolute 
pressure. Here the proportionality of the transform- 
ation coefficient both to the velocity of pressure 
variation with time and to the pressure gradient in- 
volve the same extrapolator. Giving up the construc- 
tion of dimensionless parameters of which, with a 
great number of constants involved, there is ample 
freedom of choice, the extrapolator suggested by 
Sturman (1974) 

3 
A. = 2808 (7) 
<p*> —SS SS 
2 


can be substituted by the following: 


is the square of the acoustic pressure, 
T is the distance to the point of noise 
measurement, 
No is the number of cavities collapsing in 
unit time. 
If we assume in the regular way that the similar- 
ity of cavity patterns is observed and the noise is 
measured at similar points of the flow, then 


V w/e 


- model-test data 
coo - full-seale test data 


FIGURE 6. Comparison of noise levels extrapolated 
from model with measured full-scale data in a wide 
band of frequencies. 


and t = L. The table below shows extrapolators for 
scaling the square of the acoustic pressure during 
cavitation from model to full-size with reference 
to the assumptions of constant and variable coeffi- 
cients of cavity energy transformation into noise, 
nN, and to fit the cases of constant Froude number 
and constant absolute pressure. 

That the frequencies vary inversely in proportion 
to linear dimensions in modelling at a constant 
pressure may turn out to be a significant advantage, 
so the acoustic wave lengths change in proportion 
to linear dimensions of the model and wave inter- 
ference patterns remain unchanged. In modelling by 


the Froude method wave lengths on the model are 


© model-test and full-scale results for the wide 
frequency band 
@ model-test and full-scale results for the 1/3- 


octave noise frequency band of the model - 80 kHz 


* instant of visual detection of vortex cavitation 
on the model 


FIGURE 7. Comparison of noise levels extrapolated 
from model with measured full-scale data. 


TABLE 1. Cavitation Noise Levels Scaling 
Extrapolator 
P P 
=e ike 
nN = const L T 
Fo = const i3/2 Be72 Le 
Bs = const iL 1/L 1/L 


known to be VI, times larger than the model linear 
dimensions. 

Figure 6 shows the comparison between the model- 
test data (solid line) scaled the comparison between 
the model-test scale data (dotted line) for the 
noise level in a wide band of frequencies. Figure 
7 gives a similar comparison with another prototype. 
Curves of cavitation noise increase are also compared 
in a 1/3-octact band for the model at the frequency 
of 80 kHz. In Figure 7 the moment of visual detec-— 
tion of cavitation is marked on the general level 
curve with an asterisk. Full-scale data are given 
here for individual rates of speed. 

The scaling extrapolator (8) needs to be verified 
under full-scale conditions and is likely to be 
refined. However, the need for stability of the 
coefficient of cavity energy transformation into 
cavitation noise appears to be an indisputable 
argument for cavitation noisiness scaling with the 
full-scale pressure retained. 


CONCLUSION 


The two major conclusions can be formulated as 
follows: 

- Scaling for cavitation noise measurements with 
the full-scale pressure retained gives a high value 
coefficient of cavity energy transformation into 
noise and substantial advantages in respect to: 

a) obtaining high levels of cavitation noise; 

b) similarity of sound waves to the model. 

- Large-scale modelling with the full-scale 
pressure retained confirmed the possibility brought 
out by the analysis of an elementary cavitation 
process of acoustic detection of cavitation long 
before the cavity reaches the size that can be 
detected visually. 


REFERENCES 


Beniaminovich, M. B., K. A. Kondratovich, and I. V. 
Krutetsky (1975). On experimental determination 
of acoustic radiation energy during cavitation 
bubble collapse. (in Russian). Symposium on 
Physics of Acoustics-Hydrodynamics Phenomena, 
Reports, Izd. "Nauka", Moscow. 

Goncharov, O. N., A. S. Gorshkoff, and V. J. Vaniu- 
khin (1977). Cavitation and its noise radiation 
in steady and unsteady flows. 9th All-Union 


Symposium on Physics of Acoustics-Hydrodynamics 
Phenomena, (in Russian), Reports, Vol. 2, Izd. 
Academy of Sciences of the USSR, Moscow. 
Gorshkoff, A. S., and A. S. Lodkin (1966). The 
inception of cavitation under symmetrical stream- 


lining round a body of revolution with blunt nose. 


llth ITTC, Tokyo. 


499 


Levkovsky, Yu. L. (1968). Scaling of cavitation 
noise, (in Russian). Akustichesky zhurnal, 13, 
3. Moscow. 

Sturman, A. M. (1974). Fundamental aspects of the 
effect of propeller cavitation on the radiated 
noise. Proceedings of Symposium on High-Powered 
Propulsion of Large Ships, Wageningen. 


Fluid Jets and Fluid Sheets: 
A Direct Formulation 


P. M. Naghdi 


University of California 


Berkeley, California 


ABSTRACT 


The object of this paper is to present an account 

of recent developments in the direct formulation of 
the theories of fluid jets and fluid sheets based 

on one and two-dimensional continuum models origi- 
nating in the works of Duhem and E. and’ F. Cosserat. 
Following some preliminaries and descriptions of 
(three-dimensional) jet-like and sheet-likeé bodies, 
the rest of the paper is arranged in two parts, 
namely Part A (for fluid jets) and Part B (for fluid 


sheets), and can be read independently of each other. 


In each part, after providing the main ingredients 
of the direct model and a statement of the conserva- 
tion laws, appropriate nonlinear differential equa- 
tions are derived which include the effects of 
gravity and surface tension. Application of these 
theories to various one and two-dimensional fluid 
flow problems, including water waves, are discussed. 


1. INTRODUCTION 


Jets and sheets are a class of three-dimensional 
bodies whose boundary surfaces have special charac- 
teristic features. To this extent they are, respec- 
tively, similar to another class, namely that of 
rods and shells (or plates), although the nature of 
the specified surface (or boundary) conditions in 
the two classes may be different. Moreover, the 
kinematics of jets and rods are identical, as are 
the kinematics of sheets and shells. Indeed, it is 
only through their constitutive equations that a 
distinction appears between rods and jets on the 

one hand, and shells and sheets on the other. It 

is natural to inquire as to the possible utility of 
methods of approach in the construction of theories 
in the class of rods and shells for that of jets 

and sheets and vice versa. The main purpose of this 
paper is to call attention to the possible utility 


of a direct approach for jets and sheets, an approach 


500 


which has met with considerable success in the case 
of rods and shells. The direct approach for fluid 
jets is based on a one-dimensional model, called a 
Cosserat (or a directed) curve which is defined in 
Section 3; and the direct approach for fluid sheets 
is based on a two-dimensional model, called a Cos- 
serat (or a directed) surface which is defined in 
Section 5. It should be emphasized that a Cosserat 
curve and a Cosserat surface are not, respectively, 
just a one-dimensional curve and a two-dimensional 
surface; but are, in fact, endowed with some struc-— 
ture in the form of additional primitive kinematical 
vector fields. 

The concept of 'directed' or 'oriented' media 
originated in the work of Duhem (1893) and a first 
systematic development of theories of oriented media 
in one, two, and three dimensions was carried out by 
E. and F. Cosserat (1909). In their work, the Cos- 
serats represented the orientation of each point of 
their continuum by a set of mutually perpendicular 
rigid vectors. The purely kinematical aspects of 
oriented bodies characterized by ordinary displace- 
ment and the independent deformation of N deformable 
vectors in N-dimensional space has been discusssed 
by Ericksen and Truesdell (1958), who also intro- 
duced the terminology of directors. 

A complete general theory of a Cosserat surface 
with a single deformable director given by Green 
et al. (1965) was developed within the framework of 
thermomechanics; and their derivation (Green et al. 
1965) is carried out mainly from an appropriate 
energy equation, together with invariance require- 
ments under superposed rigid body motions. A re- 
lated development utilizing three directors at each 
point of the surface, in the context of a purely 
mechanical theory and with the use of a virtual work 
principle, is given by Cohen and DeSilva (1966). A 
further development of the basic theory of a Cosserat 
surface along with certain general considerations re- 
garding the construction of nonlinear constitutive 
equations for elastic shells is given by Naghdi 


(1972), which also contains additional historical 
remarks relevant to oriented continua and to the 
theory of thin elastic shells. A parallel develop- 
ment in the theory of a Cosserat curve with two 
deformable directors begins with a paper of Green 
and Laws (1966) whose derivation is carried out 
mainly from an appropriate energy equation, together 
with invariance requirements under superposed rigid 
body motions. A related theory of a directed curve 
with three deformable directors at each point of the 
curve, in the context of a purely mechanical theory 
and with the use of a virtual work principle, is 
given by Cohen (1966). A further development of 

the basic theory of a Cosserat curve along with 
certain general developments regarding the construc- 
tion of nonlinear constitutive equations for elastic 
rods is given by Green et al. (1974a,b). 

In general, two entirely different approaches may 
be adopted for the construction of one-dimensional 
and two-dimensional theories of mechanics pertain- 
ing to certain motions and (three-dimensional) media 
responses which are effectively confined, respec-— 
tively, to one-dimensional and two-dimensional re- 
gions. For example, the theory of slender rods and 
that of fluid jets are both one-dimensional theories; 
and, similarly, the theory of thin shells and that 
of fluid sheets are both two-dimensional theories 
in the context of the particular classes of three- 
dimensional bodies mentioned earlier. 

Of the two approaches just mentioned, one starts 
with the three-dimensional equations of the classi- 
cal continuum mechanics and by applying approxima- 
tion procedures strives to obtain one-dimensional 
(in the case of jets and rods) and two-dimensional 
(in the case of sheets and shells) field equations 
and constitutive equations for the medium under 
consideration. In the other approach, the particu- 
lar medium response mentioned above is modelled as 
a one-dimensional and a two-dimensional directed 
continuum, namely a Cosserat curve and a Cosserat 
surface introduced earlier; and one then proceeds 
to the development of the field equations and the 
appropriate constitutive equations. If full inform- 
ation is desired regarding the motion and deforma- 
tion of the continuum under study in the context of 
the classical three-dimensional theory, then there 
would be no need to develop a particular one- 
dimensional and a two-dimensional theory. In fact, 
the aim of one-dimensional and two-dimensional theo- 
ries of the type mentioned above is to provide only 
practical information in some sense: for example, 
in the case of fluid sheets information concerning 
quantities which can be regarded as representing 
the medium response confined to a surface or its 
neighborhood as a consequence of the (three- 
dimensional) motion of the body, or the determina- 
tion of certain weighted averages of quantities 
resulting from the (three-dimensional) motion of 
the body. A parallel remark may be made, of course, 
in the case of fluid jets. The desire for obtain- 
ing limited or partial information if the basic 
motivation for the construction of such one- 
dimensional and two-dimensional theories as those 
for slender rods and thin shells and for fluid flow 
problems of jets and sheets. 

The nature of difficulties associated with the 
development of both the shell theory and the theory 
of water waves on the one hand, and that of rods and 
jets on the other, from the full three-dimensional 
equations is well known and has been elaborated upon 


501 


on various occasions.* In view of these, it is rea- 
sonable to attempt to formulate one-dimensional and 
two-dimensional theories of the types described above 
by replacing the continuum characterizing the (three- 
dimensional) medium in question with an alternative 
model which would reflect the main features of the 
response of the three-dimensional medium and which 
would then permit the formulation of appropriate 
one-dimensional and two-dimensional theories by a 
direct approach and without the appeal to special 
assumptions or approximations generally employed in 
the derivation from the three-dimensional equations. 

Of course, the introduction of an alternative 
model and formulation of one-dimensional and two- 
dimensional theories by the direct approach do not 
mean that one ignores the nature of the field equa- 
tions in the three-dimensional theory. In fact, 
some of the developments of the field equations by 
direct procedures are materially aided or influenced 
by available information which can be obtained from 
the three-dimensional theory. For example, the inte- 
grated equations of motion from the three-dimensional 
equations provide guidelines for a statement of one 
and two-dimensional conservation laws in conjunction 
with the one and two-dimensional models, and also 
provide some insight into the nature of inertia terms 
and the kinetic energy in the direct formulation of 
the one-dimensional and two-dimensional theories. 

Inasmuch as most of the difficulties associated 
with the derivation of the one-dimensional and two- 
dimensional theories from the three-dimensional equa- 
tions occur in the construction of the constitutive 
equations, it is in fact here that the direct ap- 
proach offers a great deal of appeal. This construc- 
tion, as well as the entire development by the 
direct approach, is exact in the sense that they 
rest on (one-dimensional and two-dimensional) pos- 
tulates valid for nonlinear behavior of materials 
but clearly they cannot be expected to represent all 
the features that could only be predicted by the 
relevant full three-dimensional equations. Theories 
constructed via a direct approach necessarily sat- 
isfy the requirements of invariance under superposed 
rigid body motions that arise from physical consider- 
tions and, of course, they are also consistent and 
fully invariant in the mathematical sense. More- 
over, the development by the direct approach is con- 
ceptually simple and does not have the difficulties 
involving approximations usually made in the devel- 
opment of the theory of thin shells and the theory 
of water waves (or the theories of slender rods and 
jets) from their corresponding three-dimensional 
equations. 

Following some general background information 
and definitions of jet-like and sheet-like bodies 
in Section 2, the remainder of the paper is arranged 
in two parts which can be read independently of each 
other: one part (Part A) is concerned with the 
theory of fluid jets and the other (Part B) is de- 
voted to the theory of fluid sheets and its applica- 
tion to water waves. In our discussion of the 
direct formulation of these two topics, considerably 


“the nature of these difficulties with particular reference 
to shells is discussed by Naghdi (1972, Secs. 1,4,19,20,21). 
Some of the difficulties associated with both nonlinear and 
linear theories of water waves are noted by Naghdi (1974) and 
are also discussed in the first and final sections of the 
paper of Green et al. (1974c). 


tsee the remarks following Eqs. (26) and (50). 


502 


more space is devoted to fluid sheets and water 
waves. This is partly due to the fact that, in the 
context of the direct formulation, the theory of 
fluid sheets has to date received more attention 
than that of fluid jets. Thus, in Part A (Sections 
3-4) , we summarize the basic theory of a Cosserat 
curve and briefly discuss a restricted form of the 
theory for straight jets which are not necessarily 
circular. The resulting system of nonlinear ordi- 
nary differential equations includes the effects of 
surface tension and gravity and has been derived for 
both inviscid and viscous jets. We do not record 
these here; but we call attention in Section 4 toa 
number of existing solutions, which serve as evidence 
of the relevance and applicability of the direct 
formulation of the theory of fluid jets. 

In Part B (Sections 5-8), after briefly describ- 
ing the basic theory of a Cosserat surface in Sec- 
tion 5, we present in outline a derivation of a 
restricted theory in Section 6, and then obtain a 
system of nonlinear partial differential equations 
for the propagation of fairly long waves in a homo- 
geneous stream of variable depth (Section 7). This 
system of differential equations, which includes 
the effects of surface tension and gravity, is de- 
rived for incompressible inviscid fluids. Some ex- 
tensions of these results to nonhomogeneous and 
viscous fluids are available but these are not dis- 
cussed here. In the final section of the paper we 
make a comparison between the differential equations 
derived in Section 7 and the systems of equations 
for water waves that are often used in the litera- 
ture; and, on the basis of compelling physical con- 
siderations, argue as to why the system of equations 
of the direct formulation should in general be pre- 
ferred to others. In Section 8, we also call at- 
tention to a number of existing solutions, which 
serve as further evidence of the relevance and ap- 
plicability of the direct formulation of the theory 
of fluid sheets. 

In the course of our development, sometimes the 
same symbol is utilized in Parts A and B to denote 
different quantities; but this should not give rise 
to confusion, as the two parts can be read indepen- 
dently of each other. Throughout the paper, Latin 
indices (subscripts or superscripts) take the values 
1, 2, 3, Greek indices take the values l, 2 only, 
and the usual convention for summation over a re- 
peated index is employed. 


2. GENERAL BACKGROUND 


In this section, we provide appropriate definitions 
for jet-like and sheet-like bodies. To this end, 
consider a finite three-dimensional body, 8, ina 
Euclidean 3-space, and let convected coordinates, 

61 (i = 1, 2, 3), be assigned to each particle (or 
material point) of 8. Further, let tr* be the posi- 
tion vector, from a fixed origin, of a typical parti- 
cle of § in the present configuration at time, t. 
Then, a motion of the (three-dimensional) body is 
defined by a vector-valued function, £*, which as- 


tone use of an asterisk attached to various symbols is for 
later convenience. The corresponding symbols without the 
asterisks are reserved for different definitions or designa- 
tions to be introduced later. 


signs position, r*, to_each particle of 8 at each 
instant of time, i.e., 


fa (DE O07) (1) 


We assume that the vector function, £*,--a 1- 
parameter family of configurations with t as the 
real parameter--is sufficiently smooth in the sense 
that it is differentiable with respect to 6+ and t 
as many times as required. In some developments, 
it may be more convenient to set 93 = — and adopt 
the notation 


Fiz Wem ie gel ta oc (2) 


Ae 
CES ay 7 Gay Chi Gin gee Celera) 
56 = z 
gt 0 J; = 8; " gt = g”9, E gt 6 gJ = gtJ , (3) 
L, 
dv = g*aelae7ae? (4) 
and further assume seal 
Xs 
Ga S ig _g Gill SO co (5) 
~1=2=3 


In (4), g and g are the covariant and the contra- 
variant base vectors at time, t, respectively, Si5 
is the metric tensor, gtJ is its conjugate, 6+ is 
the Kronecker symbol in 3-space and dv the volume 
element in the present configuration. 

The velocity vector, v*, of a particle of the 
three-dimensional body in the present configuration 
is defined by 


Wi Sat (6) 


where a superposed dot denotes material time dif- 
ferentiation with respect to t holding 6+ fixed. 
The stress ‘vector t across a surface in the present 
configuration with outward unit normal y* is given 
by 


(7) 


Vos are 4 
au ey a 


oct 
i] 


ico 


where 


§ F . 
Recall that when the particles of a continuum are referred 


to a convected coordinate system, the numerical values of 
the coordinates associated with each particle remain the 
same for all time. Although the use of a convected coordi- 
nate system is by no means essential, it is particularly 
suited to studies of special bodies (such as sheets, jets, 
shells, and rods) and often results in simplification of 
intermediate steps in the development of the subject. 


Wine choice of positive sign in (5) is for definiteness. 
Alternatively, for physically possible motions we only need 
to assume that g@ # 0 with the understanding that in any 
given motion [g}g293] is either > 0 or < 0. The condition 
(5) also requires that 01 be a right-handed coordinate 
system. 


* ky kG 
Vi vg =v “Gy (8) 


Wo ENGiter cia = Gea |p 
~ 3] , 

and where tik are the contravariant components of 
the symmetric stress tensor. In terms of quantities 
defined in (5)-(8), the local field equations which 
follow from the integral forms of the three- 
dimensional conservation laws for mass, linear 
momentum and moment of momentum, respectively, are 


* 1 
eo a 


+ = 
in Os OS DSI 


x lox i 
Z o Gy T =0 (9) 


where p* is the three-dimensional mass density, £* 
is the body force field per unit mass, and a comma 
denotes partial differentiation with respect to 6?. 
A material line (not necessarily a straight line) 
in 8 can be defined by the equations, 6% = 6%(&); 
the equation resulting from (1) with 6% = e"(E) rep- 
resents the parametric form of this material line in 
the current configuration and defines a l1-parameter 
family of curves in space, each of which we assume 
to be smooth and nonintersecting. We refer to the 
space curve, 6% = 0, in the current configuration 
by c. Any point of this curve is specified by the 
position vector, xr, relative to the same fixed ori- 
gin to which r* is referred, where 


v= Z(E,t) = £ (0,0,8,t) . (10) 


Let a3 denote the tangent vector along the E-curve. 
By (10) and (3), 


@ 
> 


ag = a3(E,t) = DE = g3(0,0,€,t) (11) 


and the unit principal normal, a,, and the unit bi- 
normal vector, aj, to c may be introduced as 


da,/dE 

a; = a, (&,t) = |[BaaDET P 

on = (Ee) ek a 12 
a2 a2 Tas] l 2 (12) 

lL, 

la3| = (a33)7 , 

Glee) Seley 2 Ele) 0 

[ajaga3] >O , (12) 


where the notation Ja3| stands for the magnitude of 
a3. The system of base vectors, aj, are oriented 
along the Serret-Frenet triad and satisfy the dif- 
ferential equations 


day ky 

Be ~ 'l@ag) Gao ke3 

dag i, 
Te | Sega) ano 

Be3 1 2233 

— = ————— : 13 
ae 7 259 Dean OCS ane 


503 


where K and Tt denote, respectively, the curvature 
and the torsion of c. In the special case that c 
is a plane curve, we may choose aj as the unit 
normal to the curve and then ag will be perpendicu- 
lar to the plane of a, and a3. TEC isvarstravight 
curve, then there is no unique Serret-Frenet triad 
and a, may be chosen as any orthogonal triad with 
a1,a2 as unit vectors. Equations (13) are not 
identical to the formulas of Frenet because the pa- 
rameter, €, is not necessarily the arc length of c. 
It may be noted here that the convected coordinate, 
&, may be chosen to coincide with the arc length in 
any one configuration of the material curve, e.g., 
in the present configuration. However, in a general 
motion (involving different configurations) the arc 
length between any pair of particles changes while 
the convected coordinates of each particle must re- 
main the same. Therefore, arc length would not 
qualify as a convected coordinate. 

A material surface in 8 can be defined by the 
equation, —€ = &(6%); the equation resulting from 
(1) with € = €(6%) represents the parametric form 
of this material surface in the current configura- 
tion and defines a l-parameter family of surfaces 
in space, each of which we assume to be smooth and 
nonintersecting. We refer to the surface, € = 0. 
in the current configuration by s. Any point of 
the surface, s, is specified by the position vector, 
r, relative to the same fixed origin to which r* is 
referred, where 


BS (Ope) = BY Ore) (14) 


Let a, denote the base vectors along the 6°-curves 
on the surface, s. By (14) and (3)j, 


‘asa (0. ,t) = 
~O. 


=g (6',0,t) , (15) 
~O =a 


and the unit normal, a3 = a3(0’,t), to s may be 
defined by** 


Ey eG ee lO) hb. Crs cia le a 
Bn E  lepeeeal > Os (XE) 


In the next four paragraphs we provide appropri- 
ate definitions for jet-like and sheet-like bodies 
in fairly precise terms. 


Definition of a Jet-like Body. A Representation 


for the Motion of a Slender Jet. 


Consider a space curve c defined by the parametric 
equations, e* = 0, over a finite interval, € SESE>. 
Let r be the position vector of any point of c and 
let aj,a2 and a3 denote its unit principal normal, 
unit binormal, and the tangent vector, respectively. 
At each point of c, imagine material filaments ly- 
ing in the normal plane, i.e., the plane perpendicu- 


kk 
The use of the same symbols for base vectors of a surface 


in (15)-(16) and for the triad of a space curve in (11)-(12) 
should not give rise to confusion. The main developments 
for jets and sheets are dealt with separately in the rest 
of the paper; this permits the use of the same symbol for 
different quantities in the case of jets and sheets without 
confusion. 


504 


lar to a3, and forming the normal cross-section'tT, 
Qn - The surface swept out by the closed boundary 
curve, 0Q,, Of Gp is called the lateral surface. 
Such a three-dimensional body is called jet-like if 
the dimensions in the plane of the normal cross- 
section are small compared to some characteristic 
dimension, L(c), of c (see Figure 1), e.g., its 
local radius of curvature 1/K, or the length of c 
in the case of a straight curve. A jet-like body 
is said to be slender if the largest dimension of 
Q@y is much smaller than L(c). If a, is independent 
of €, the body is said to be of uniform cross- 
section, otherwise of variable cross-section. Since 
a material curve in the three-dimensional body, 8, 
can be defined by the equations, e% = e%(E), it 
follows that the equation resulting from (1) with 
e% = 9%(&) represents the parametric form of the 
material curve in the present configuration and de- 
fines a curve, c, in space at time, t, which we as- 
sume to be sufficiently smooth and nonintersecting. 
Every point of this curve has a position vector 
specified by (10). Let the (three-dimensional) jet- 
like body in some neighborhood of c be bounded by 
material surfaces, § = &), € = &9, (indicated in 
Figure 1) and a material surface of the form 


me ,O=) =O 5 (17) 


which is chosen such that & = constant are curved 
sections of the body bounded by closed curves on 
this surface with c lying on or within (17). In 
the development of a general theory, it is preferable 
to leave unspecified the choice of the relation of 
the curve, c, to one on the boundary surface (17). 
In special cases or in specific applications, how- 
ever, it is necessary to fix the relation of c to 
the surface (17). @ 

Suppose now that £* in (1) is a continuous func- 
tion of eit and has continuous space derivatives 
of order 1 and continuous time derivatives of order 
2 in the bounded region lying inside the surface (17) 
and between € = €], € = 9. Hence, to any required 
degree of approximation f£* may be represented as a 
polynomial in el, 62 with coefficients which are con- 
tinuously differentiable functions of &, t. Instead 
of considering a general representation of this kind, 
we restrict attention here to the approximation. 

ak (o} 


BE + 6 qa p (18) 


where r is defined by (10) and a = qd (6 rt) - 


Definition of a Sheet-like Body. A Representation 


for the Motion of a Thin Sheet. 


Consider a two-dimensional surface, s, defined by 
the parametric equation, & = 0, over a finite co- 
ordinate patch, a' = 6! Sa", Bg" S 62 = p".. Let ig 
and a3 denote, respectively, the position vector and 
the unit normal to s. At each point of s, imagine 
material filaments projecting normally above and 
below the surface, s. The surface formed by the 
material filaments constructed at the points of the 
closed boundary curve of s is called the lateral 
surface. Such a three-dimensional body is called 


tt , A é : 

The normal cross-section of a jet is a portion of the 
normal plane to the curve, c, i.e., the intersection of the 
body and the normal plane. 


FIGURE 1. A jet-like body in the present configuration 
showing the line of centroids with position vector r 
and the end normal cross-sections & = &;, § = &. Also 
shown are the unit principal normal aj), the unit binor- 
mal az and the tangent vector a3 to the curve with po- 
sition vector r. 


a sheet if the dimension of the body along the nor- 
mals, called the height and denoted by h, is small. 
A sheet is said to be thin if its thickness is much 
smaller than a certain characteristic length, L(s), 
of the surface, s, for example, the local minimum 
radius of curvature of the surface, or the smallest 
dimension of s in the case of a plane sheet. If h 
is constant, the sheet is said to be of uniform — 
thickness, otherwise of variable thickness. Since 
a material surface in the three-dimensional body can 
be defined by the equation, — = &(6%), it follows 
that the equations resulting from (1) and (2) with 

— = £(6%) represent the parametric forms of the 
material surface in the present and the reference 
configurations, respectively. In particular, the 
equation, € = 0, defines a surface in space at time, 


e 


~ 


FIGURE 2. Sketch of the cross-section (y = const.) of 

a sheet of vertical thickness $ showing a wave motion 
propagating over a bottom of variable depth. Also shown 
is the surface 6? = 0 (with position vector r and height 
Y) chosen such that the center mass of the (three- 
dimensional) fluid region lies on this surface. The top 
and bottom surfaces of height 8 and a are specified by 
93 = 1/2 ana 63 = -1/2, respectively. 


t, which we assume to be smooth and nonintersecting. 
Every point of this surface has a position vector, 

x, specified by (14). Let the boundary of the three- 
dimensional continuum be specified by the material 
surfaces 

B= ei@=ne) 9 § = Bald 0-) Ei < Em p (is) 
with the surface, & = 0, lying either on one of the 
two surfaces (19)1,2 or between them (see, for ex- 
ample, Figure 2), and a material surface 


f(0!,02) =o , (20) 


which is chosen such that € = const. forms closed 
smooth curves on the surface (20). As pointed out 
previously [Naghdi (1975)], in the development of 

a general theory, it is preferable to leave unspeci- 
fied the choice of the relation of the surface, s, 
(—§ = 0) to the major surfaces, st ands. In spe- 
cial cases of the general theory or in specific ap- 
plications, however, it is necessary to fix the 
relation of s to the surfaces (19)1,2- 

Suppose now that r in (1) is a continuous func— 
tion of 61,t, and has continuous space derivatives 
of order 1 and continuous time derivatives of order 
2 in the bounded region, &)S&S&>. Hence, to any 
required degree of approximation, oe may be repre- 
sented as a polynomial in € with coefficients which 
are continuously differentiable functions of 6%,t. 
However, instead of considering a general represent- 
ation of this kind, we restrict attention here to 
the approximation 


ce En sei (Bel (21) 


where r is defined by (14) and d = d(9%,t). 


PART A 


In Part A (Sections 3-4), we summarize the basic 
theory of a Cosserat (or a directed) curve and then 
briefly discuss a restricted form of the theory ap- 
propriate for straight fluid jets. Although we are 
concerned here mainly with the purely mechanical 
theory involving appropriate forms of the conserva- 
tion laws for mass, linear momentum, and moment of 
momentum, we also include the conservation of energy. 
The latter is useful in some applications and sup- 
plies motivation for some requirements in the de- 
velopment of certain solutions. 


3. THE BASIC THEORY OF A COSSERAT CURVE 


Having defined a (three-dimensional) jet-like body 
in Section 2, we now formally introduce a direct 
model for such a body. Thus, a Cosserat (or a 
directed) curve, R, comprises a material curve, L, 
(embedded in a Euclidean 3-space) and two deformable 
directors attached to every point of the curve, 
The directors which are not necessarily along the 
unit principal normals and the unit binormals of 
the curve have, in particular, the property that 
they remain unaltered under superposed rigid body 
motions. Let the particles of | be identified by 
means of the convected coordinate, &, and let the 


505 


curve occupied by | in the present configuration of 
R at time, t, be referred to as 2. Let r and dy 

(a = 1,2) denote the position vector of a typical 
point of & and the directors at the same point, 
respectively, and also designate the tangent vector 
to the curve, &, by a3. Then, a motion of the Cos- 
serat curve is defined by vector-valued functions 
which assign a position, r, and a pair of directors, 
dy to each particle of R at each instant of time, 
i.e. 
xr = valle) , 


cL SCL) 9 MGC ea) = © (22) 


and the condition (22)3 ensures that the directors, 
dy, are nowhere tangent to % and that d),d2 never 
Change their relative orientation with respect to 
each other and a3. The velocity and the director 
velocities are defined by 


He 7m ee (23) 


and from (23), and (11) we have 


ay =a 4 (24) 


where a superposed dot denotes material time dif- 
ferentiation with respect to t holding € fixed. 

Consider an arbitrary part of the material curve, 
L, in the present configuration, bounded by & = &j 
and 1G )= 75 (EG seo) jandy let 


1. 
ds = (a33)*d& , a33 = a3 ° a3 (25) 


be the element of the arc length along the curve, 

&. It is convenient at this point to define the 
following additional quantities: The mass density, 
p = p(&,t), of the space curve, %; the contact 
force, n = n(&,t), and the contact director couples 
is = p"(—,t), each a three-dimensional vector field 
in the present configuration; the assigned force, 

£ = £(€,t), and the assigned director couples, 

ga = g%(£,t), each a three-dimensional vector field 
and each per unit mass of the curve, 2; the intrin- 
sic (curve) director couples, 1% = mt (Epe) o per unit 
length of & which make no contribution to the supply 
of momentum; the inertia coefficients, y® = y%(&) 
and yB = y%8(z), with y°8 being components of a 
symmetric tensor, which are indenpendent of time; 
the specific internal energy, €« = €(§,t); the spe- 
cific heat supply, r = r(&,t), per unit time; and 
the heat flux, h = h(&,t), along 2, in the direction 
of increasing §, per unit time. The assigned field, 
£, represents the combined effect of (i) the stress 
vector on the lateral surface (17) of the jet-like 
body denoted by f,, and (ii) an integrated contri- 
bution arising from the three-dimensional body force 
denoted by fp, e.g., that due to gravity. A parallel 
statement holds for the assigned fields, ge Sim- 
ilarly, the assigned heat supply, r, represents the 
combined effect of (i) heat supply entering the 


5 Spor convenience, we adopt the notation for r in (10) and 
(18) also for the surface (22);. This permits an easy iden- 
tification of the two curves, if desired. The choice of 
positive sign in (22)3 is for definiteness. Alternatively, 
it will suffice to assume that [d)d a3] # O with the under- 
standing that in any given motion the scalar triple product 
[djdja3] is either > 0 or < 0. 


506 


lateral surface (17) of the jet-like body from the 
surrounding environment, denoted by r,, and (ii) an 
integrated contribution arising from the three- 
dimensional heat supply denoted by rp. Thus, we may 
write 


1 OLs es) sO ett Olam ny i ry= tat COU (26) 


The various quantities in (26) are free to be spec- 
ified in a manner which depends on the particular 
application in mind and, in the context of the the- 
ory of a Cosserat curve, the intertia coefficients, 
yo, yo8 and the mass density, 9, require constitu- 
tive equations. Indeed, fo, 22 and r,, as well as 
fb, &£% and Xp, Can be identified with the corre- 
sponding expressions in a derivation from the three- 
dimensional equations [see, for example, Green et 
al. (1974a)]. Likewise, the inertia coefficients, 
yo, yos , and the mass density, p, may be identified 
with easily accessible results from the three- 
dimensional theory. 

With the above definitions of the various field 
quantities and with reference to the present con- 
figuration, the conservation laws for a Cosserat 
curve are: 


a iy) * - 
dt p as = ’ 
fe A 
Bi Eo a Eo Eo 
ae o(v +y w)ds = i pfds + [n] ' 
G1 ei ey 
Bo 
a ap 
ae oly We ae SY Welds 
1 
£2 ie 2y) 
= i) (OH = gs) es * tee, 
el 1 
£2 
ae | Ol wy G <y, Fa xy) 
Ey 
+ qe x yew las 
b2 


£2 
zs f p(x +£-v+ fa w)ds 
€1 = y = ~OL 


ls Onypsde By ie) py (27) 


where we have used the notation 


=) 


£ 
[ (E,t) 1, 


Srp) = wep) (28) 


The first of (27) is a statement of the conservation 
of mass, the second is the conservation of linear 
momentum, the third that of the director momentum, 
the fourth is the conservation of moment of momentum, 
and the fifth represents the conservation of energy. 

Under suitable continuity assumptions, the first 
four equations in (27) are equivalent to 


1 6 r) 
h = A(E) = p(a33)? or 6a33 + pag ° = =0, (29) 
an 
pe + AE =A(v + yw.) , (30) 
ap. 
WS ae SNES oes) (31) 
0€ ~ ~ ~ ~B 
om 
a3 x n+ Be @ Key SO (32) 
dh : 
Ue pe 7 er 1 =O , (33) 
where 
ia a , a 
LSS eS = it er liey Gr ’ 
(o} a aie ap 
QRS SV ES (34) 
and 
ov ow, 
MI fees a. @). Seley 
ALE = n DE +7 Wo +p DE (35) 


is the mechanical power. With the help of (34), the 
local form of the moment of momentum equation (31) 
can be reduced to 


ad 
ag Xn 4d" x qe 4 ie = Oy (36) 


u dE P 


It may be noted that the local field equations 
in the mechanical theory of a Cosserat curve have 
the same forms as those that can be derived from the 
three-dimensional equations; the latter can be de- 
rived by suitable integration of (9)],9,3 with re- 
spect to 61 and 62 and in terms of certain definitions 
for integrated mass density and resultants of stress 
[for details, see Green et al. (1974a)]. Moreover, 
given the approximation (18), there is a 1-l corre- 
spondence between the one-dimensional field equations 
that follow from the conservation laws of a Cosserat 
curve and those that can be derived from the three- 
dimensional equations provided we identify the 
director dy in (18) with (22) 9 and adopt the defini- 
tions of the resultants mentioned above. A similar 
1-1 correspondence can be shown to hold between (33) 
and an integrated energy equation derived from the 
three-dimensional energy equation. 

The above results include the local form of con- 
servation of energy derived from (27)5. For the 
purely mechanical theory in which the law of con- 
servation of energy is excluded, the appropriate 
conservation laws are the first four of (27). In 
the context of the purely mechanical theory, it is 


worth recalling that the rate of work by all contact 
and assigned forces acting on the curve, %, and its 
end points minus the rate of increase of the kinetic 
energy can be reduced to: 

E2 j 4 &2 


-w,)ds + In - wi 


a 
3) 
A 
iq 
+ 
IwD 
ig 
+ 
OD 


where P is defined by (35). 

Before closing this section, we note that the 
restriction imposed on the motion of the medium by 
the condition of incompressibility reduces to |||| 
[see Green (1976) ] 


5, [didza3] = 0 (38) 


and can alternatively be expressed in the form 


eax Yau yaniecy k= (39) 
° x ° — i= 

~B a3 We Sil a2) 0& t 

where 8 is the permutation symbol in 2-spaces. To 


complete the theory of a Cosserat curve under the 
constraint condition (39), eee assume that each of 
the functions, n, a, and p is determined to within 
an additive constraint essen ee so that 


f =f 9 7 » Sa ep pCO 


n=n+n , 
where fi, #%, and p% are determined by constitutive 
equations and the functions n(&,t), Tt (Ep) p and 
pa(é,t) are the response due to the constraint; the 
latter quantities are arbitrary functions of &,t and 
do no work. For an incompressible inviscid fluid 
jet, which models the properties of the three- 
dimensional inviscid fluid at constant temperature, 
we introduce the constitutive assumption that n,74, 
ig do not depend explicitly on the kinematic quan- 
tities, dv/dé, Wor W/E, and are furthermore work- 
less, i.e., 


ov ow 


a, 
110% erent WON A Se eaife) 


DO ae a) ge ee is (41) 


provided w., dv/d— satisfy the constraint condition 


(39). It can then be shown that [Green and Laws 
(1968) and Green (1976) ] 


p=0, (42) 


lll In. general, there are three conditions of incompressibility 
in the theory of incompressible directed fluid jets; for a 
discussion of these, see Caulk and Naghdi (1978a, Appendix) . 
In restricted forms of the theory discussed in the next sec- 
tion, two of the three conditions are satisfied identically. 
The specification (38) is motivated from an examination of 
the incompressibility condition in the three-dimensional 
theory when the position vector is approximated by (18). 


507 


where isya is an arbitrary scalar function of &,t. 
For an incompressible viscous jet, the constraint 
response,n, mo, po, are determined similarly with the 
use of the Constraint condition (39), but constitu- 
tive equations are required for n, ister 1p in (40). We 
do not record here the results for a viscous jet and 
refer the reader to Green (1976) and Caulk and 
Naghdi (1978b) . 


4. STRAIGHT FLUID JETS. ADDITIONAL REMARKS 

We now specialize the results of the previous sec- 
tion to straight jets of elliptical cross-section. 
In order to display some details of the kinematics 
of a straight jet, including the rotation of the 
directors in a plane normal to the jet axis, it is 
convenient to introduce a fixed system of rectangular 
Cartesian coordinates (x,y,z) with the z-axis paral- 
lel to the jet. Further, let the unit base vectors 
of Ene rectangular Cartesian axes be denoted by 

(i,j rk) and introduce, for later convenience, the 
eee lowell base vectors 


Qy = ab cos © w 3 Sain F 


ep ie San Ol ta ICOSMON my §Cae Kner, (43) 


where 8 is a smooth function of z and t. We assume 
that the directors are so restricted that they de- 
scribe an elliptical cross-section of smoothly vary- 
ing orientation along the length of the jet and that 
at each z = const., the base vectors, e] and eg, 

lie along the major and minor axes of the ellipse, 
respectively. Then, the angle, 8, called the 
sectional orientation, specifies the orientation of 
the cross-section as a function of position. With 
this background, henceforth we restrict motions of 
the directed curve, R, such that in the present con- 
figuration at time, t, 


SS] B(Eraes oo Ghee 7 ca = wen) (24) 


where $)] and ¢9 measure the semiaxes of the ellipti- 
cal cross-section. In the case of a circular jet, 
¢1 = 62. 

The complete theory also requires the specifica- 
tion of explicit values for Ary, yOB, £ and £%. In 
particular, the values for d,y%,yoB may be obtained 
by an appeal to certain results from the three- 
dimensional description of the jet. Thus, recall- 
ing (18) and the remark made following (17), here 
we choose the curve, 6% = 0, as the line of centroids 
of the jet-like body and identify this curve with 
the curve, 2, in the theory of a Cosserat curve. 
This leads to the identification 


L eh 
\ = p@aq)e = i g°ae!ae? 


a 
* 4s a 
Ay = (Ni tefl) cls) Cl) 9 yp 
a 
x 1 i 2 
pF iE g*e"e"aa ao, (45) 
where p* is the three-dimensional mass density in 


(9) and the determinant g defined by (3)3 is cal- 
culated from the approximation (18). Again, with 


508 


the use of (18) and the equations of motion (9)2, 3, 
the expressions for f and 2% can be identified in 
terms of the integrated body force, £*, over the 
cross-sectional area, a, and specified pressure and 
surface tension over the boundary, da of a [for 
details, see, for example, Caulk and Naghdi (1978a)]. 
We observe that since y® = 0 hy (45)9, the equations 
of motion (30) and (31) assume a slightly simpler 
form. We do not record here the details of the 
system of ordinary differential equations which can 
be obtained from (29)-(33) for both inviscid and 
linear viscous fluids. They are readily available 
in the papers cited: see Green and Laws (1968), 
Green (1975, 1976, 1977), and Caulk and Naghdi 
(1978a, b). 

In the rest of this section, we briefly call 
attention to some available evidence of the relevance 
and applicability of the direct formulation of the 
fluid jets. Available solutions obtained to date 
are limited to those for straight jets and among 
these most of them deal with jets of circular cross- 
section. Some general aspects of compressible 
inviscid jets, including a discussion of ideal gas 
jets in the context of a thermodynamical theory, 
have been studied by Green (1975). Applications to 
incompressible circular jets for both inviscid and 
viscous fluids are contained in the papers of Green 
and Laws (1968) and of Green (1976). Green (1977) 
has also studied a steady motion of an incompressible 
inviscid fluid jet which does not twist along its 
axis. A more detailed analysis of the motion of a 
straight elliptical jet of an incompressible inviscid 
fluid in which the jet is allowed to twist along its 
axis is contained in a recent paper by Caulk and 
Naghdi (1978a). This study, which includes the ef- 
fects of gravity and surface tension, utilizes the 
nonlinear differential equations of Section 3 with 
r and dg at time, t, specified in the form (44). A 
number of theorems are proved in the paper of Caulk 
and Naghdi (1978a) which pertain to the motion of a 
twisted elliptical jet and some special solutions 
are obtained which illustrate the influence of twist. 
Further, a system of linear equations, derived for 
small motions superposed on uniform flow of an in- 
compressible circular jet, is employed by Caulk and 
Naghdi (1978b) to study the instability of some 
simple jet motions in the presence of surface ten- 
sion, i.e., the so-called capillary instability that 
leads to disintegration of the jet. In particular, 
they [Caulk and Naghdi (1978b)] consider the breakup 
of both inviscid and viscous jets: in the case of 
an inviscid jet excellent agreement is obtained with 
the three-dimensional results of Rayleigh (1879a,b); 
and for a viscous jet, through a comparison with 
available three-dimensional numerical results 
[Chandrasekhar (1961)], the solution obtained is 
shown to be an improvement over an existing approxi- 
mate solution of the problem by Weber (1931). A 
related study by Bogy (1978), concerning the insta- 
bility of an incompressible viscous liquid jet of 
circular section, partly overlaps with the work of 
Caulk and Naghdi (1978b) on the temporal instability 
of a viscous jet, and considers the spatial insta- 
bility of a semi-infinite jet formulated as a 
boundary-value problem. 


PART B 


In Part B (Sections 5-8), after briefly describing 
the basic theory of a Cosserat (or a directed) sur- 


face, we summarize a special case of the theory 

which is particularly suited for applications to 
problems of fluid sheets and to the propagation of 
fairly long water waves. For the sake of simplicity, 
we confine attention here to homogeneous fluids; but 
note that, as in Green and Naghdi (1977), the deriva- 
tion can be modified to allow for variation of mass 
density with depth. Although we are concerned mainly 
with the purely mechanical theory involving appropri- 
ate forms of the conservation laws for mass, linear 
momentum, and moment of momentum, we also include 

the conservation of energy. The latter easily sup- 
plies motivation for some requirements in the devel- 
opment of certain solutions. 


5. THE BASIC THEORY OF A COSSERAT SURFACE 


Having introduced the notion of a (three-dimensional) 
sheet-like body in Section 2, we now formally define 
a direct model for such a body. Thus, a Cosserat 
(or directed) surface, C, comprises a material sur- 
face, S, (embedded in a Euclidean 3-space) and a 
single deformable vector, called a director, attached 
to every point of the surface, S. The directors 
which are not necessarily along the unit normals to 
the surface have, in particular, the property that 
they remain unaltered under superposed rigid body 
motions. Let the particles of the material surface 
of C be identified by means of a system of convected 
coordinates, 0% (a = 1,2), and let the surface oc- 
cupied by S in the present configuration of C€ at 
time, t, be referred to as Jd. Let ry and d denote 
the position vector of a typical point of J and the 
director at the same point, respectively, and also 
designate the base vectors along the 6%-curves on 
Jd by ay: Then, a motion of the Cosserat surface is 
defined by vector-valued functions which assign posi- 
tion, x, and director, d, to each particle of C at 
each instant of time, imens 

r= r(0,t) , d= d(0",t) , [ajard] > 0 (46) 
and the condition (46), ensures that the director, 
d, is nowhere tangent to Jd. The base vectors, ay, 
and their reciprocals, ar the unit normal, a3, and 
the components of the metric tensors, aap and ars, 
at each point of & are defined by 


ax a a 
furs a | = | 8g SOB See Zoe 
30 
OS 2. 8 gah 
adie aa Mee at ene le feel 
5 
a = det at, ane [ajaja3] > O , (47) 


where 67 is the Kronecker delta in 2-space. The 
velocity and the director velocity vectors are de- 
fined by 


43) 


ig 
ll 

ine 

2 Que 


9 YS 


*For convenience, we adopt the notation for ry in (14) and 
(21) also for the surface (46),. This permits an easy 
identification of the two surfaces, if desired. The choice 
of positive sign in (46)3 is for definiteness. Alterna- 
tively, it will suffice to assume that [ajajd] # 0 with the 
understanding that in any given motion the scalar triple 
product [a,ajd] is either > 0 or < 0. 


where a superposed dot denotes differentiation with 
respect to t holding 0° fixed. 

Let P, bounded by a closed curve, 0P, be a part 
of J occupied by an arbitrary material region of 
S in the present configuration at time, t, and let 


Vow =| vie (49) 


be the outward unit normal to 3P. It is convenient 
at this point to define certain additional quantities 
as follows: The mass density, p = p(0’,t), of the 
surface, J, in the present configuration; the con- 
tact force, NES N(8Y,t;v), and the contact director 
forcel, M = M(6Y,t;v), each per unit length of a 
curve in the present configuration; the assigned 
force, £ = £(0Y,t), and the assigned director force, 
L= R(0Y,t), each per unit mass of the surface,J ; 
the intrinsic director force, m, per unit area of 

di the inertia coefficients, k = k(@Y) and k = k(6Y), 
which are independent of time; the specific internal 
energy, € = e(O8Y,t); the heat flux, h = h(6’,t;v) 

per unit time and per unit length of a curve, OP; 

the specific heat supply, r = r(6Y,t), per unit time; 
and the element of area, do, and the line element, 
ds, of the surface, J. The assigned field, £, may 
be regarded as representing the combined effect of 
(i) the stress vector on the major surfaces of the 
sheet-like body denoted by f., e.g., that due to the 
ambient pressure of the surrounding medium, and (ii) 
an integrated contribution arising from the three- 
dimensional body force denoted by fy, e.g., that due 
to gravity. A parallel statement holds for the as- 
signed field, 2. Similarly, the assigned heat sup- 
ply, r, may be regarded as representing the combined 
effect of (i) heat supply entering the major surfaces 
of the sheet-like body from the surrounding environ- 
ment, denoted by Yo, and (ii) a contribution arising 
from the three-dimensional heat supply, denoted by 
Yp- Thus, we may write 


7 Sse, ab ae A & = fl) a A ae ES Ge tb ry o (SO) 


The various quantities in (50) are free to be speci- 
fied in a manner which depends on the particular ap- 
plication in mind and, in the context of the theory 
of a Cosserat surface, the inertia coefficients, k, 
k and the mass density, p, require constitutive equa- 
tions. Indeed, forko and Yor as well as fpr ep and 
Yp, can be identified with corresponding expressions 
in a derivation from the three-dimensional equations 
[for details, see Naghdi (1972,1974)]. Likewise, p 
and the coefficients,k,k, may be identified with 
easily accessible results from the three-dimensional 
theory. 

In terms of the above definitions, the conserva- 
tion laws for a Cosserat surface can, be stated in 
fairly general forms. We do not record these here 
since they are available elsewhere [Naghdi (1972), 
p. 482) or Naghdi (1974)]. Instead, we turn our 
attention to the relatively simple theory of the 
next section. 

It may be noted that the local field equations 
in the mechanical theory of a Cosserat surface have 


+ 
The terminology of director couple is also used for M depend- 
ing on the physical dimension assumed for the director, d. 


Here we choose d to have the physical dimension of length so 
that M has the same physical dimension as N. For further 


discussion see Naghdi (1972, Ch. C) and Green and Naghdi 
(1976). 


509 


the same forms as those that can be derived from the 
three-dimensional field equations () 1 4) Jey Suites 
able integration between the limits, &, and €5, and 
in terms of certain definitions for integrated mass 
density and resultants of stress [for details, see 
Naghdi (1972, Sections 11-12) or Naghdi (1974)]. 
Moreover, given the approximation (21), there is a 
1-1 correspondence between the two-dimensional field 
equations that follow from the conservation laws of 

a Cosserat surface and those that can be derived from 
(9)1,2,3 provided we identify the director, d, in 
(21) with (46)9 and adopt the definitions of the re- 
sultants mentioned above. As similar 1-1 correspon- 
dence can be shown to hold between the two-dimensional 
energy equation in the theory of a Cosserat surface 
and an integrated energy equation derived from the 
three-dimensional energy equation. 


6. A RESTRICTED THEORY OF A COSSERAT SURFACE 


Special cases of the general theory can be obtained 
by the introduction of suitable constraints, thereby 
resulting in constrained theories. Alternatively, 
corresponding special cases can be developed in which 
the kinematic and the kinetic variables are suitably 
restricted a priori and then restricted theories are 
constructed by direct approach. Such special cases 
of the general theory have been discussed previously 
by Naghdi (1972, Sections 10 and 15) and by Green 
and Naghdi (1974) and are of particular interest in 
the context of elastic shell theory. We provide here 
an outline of a restricted theory developed by Green 
and Naghdi (1977) mainly for application to problems 
of fluid sheets. The resulting equations can also 
be obtained as a constrained case of those given for 
directed fluid sheets [Green and Naghdi (1976)], but 
it is more convenient tq restrict the kinematic and 
the kinetic variables at the outset and construct a 
corresponding restricted theory from an appropriate 
set of conservation laws in integral form. 

Let the director, d, while deforming along its 
length, always remain parallel to a fixed direction 
specified by a constant unit vector, b. It should 
be kept in mind that b is fixed relative to the body 
and not relative to the space. Thus, recalling (46) 9 
and (48)5, we write 


d= $(0,t)b , w=w(0,t)b , w= . (51) 


Further, in view of the assumed form of (51), for 
the director, it is convenient to decompose M,m and 
2 into their components along and perpendicular to 
the) unit vector, b, iJe., 


MS MOV ews Bis Oey) 7 SSDS O 2 


=n(Ob be s@e) > s ° 


n=} 

I 
ow 

Il 
) 


= DOM oa Hs oO ae) eS) 


i) 
| 
2Q0 
° 
ey 
Il 
je) 


where M,m and % are scalar functions and S,s,c are 
vector functions of their arguments. According to 
the decomposition (52) ; the vector, M, is resolved 
into two parts. One pat is along b and the other 
part is the perpendicular projection of M onto the 
plane defined by S * b = O which is perpendicular to 


510 


b. Parallel statements hold for vectors, m and L, 
in (52)9,3- 

Also, it is convenient to decompose the assigned 
fields, f and 2, into two parts, one of which repre- 
sents the three-dimensional body force acting on the 
continuum which is assumed to be derivable from a 
potential function, 2(r,$), and the other which 
represents the effect of applied surface loads on 


the major surfaces of the fluid sheet. Thus, we 
write 
dQ 8Q 
i i fo ’ oe (— nen Lo) O (53) 


With the foregoing definitions of the various 
field quantities and with reference to the present 
configuration, the conservation laws for a restricted 
theory of a Cosserat surface [different from the re- 
stricted and constrained theories discussed previ- 
ously by Naghdi (1972) and by Green and Naghdi (1974) ] 


are: 
d 
ae do = 
s |e (o} OQ 4» 
d 


—= ne = + 
ae exe + kwb) do J ,eta0 lhe N dst; 


ae : p (kv + kwb) wdo = bt J (or-mas ae Mds] 


+b x if (pc-s) do 
Diese 


+ J gp80s] : 


A 


ae px x v + k(x x wh + d x v)]do 


12 


=[ ple x £ + ax (b x c)]do 
5 C 


+ [ope x N+ax x s)las , 


aE ple +2 + 4(v + v + 2kv + wh + kw?)]do 
x, avant 2 


= [ pirte “y+ 2 w)ao 
P CLO! ~ (© 


+ foe ONY ar Iie) > TNCIS= 5G (54) 


In the above equations (54) ; is a statement of con- 
servation of mass, (54) 9 the conservation of linear 
momentum, (54)3 that of the conservation of the 
director momentum, (54), the conservation of moment 
of momentum and (54)5 represents the conservation 
of energy. It should be noted that the quantities, 
M and 2% .¢ no contributions to the moment of momen- 
tum equation, and the quantities, ¢ and S, make no 
contribution to the equation for conservation of 
energy in the present restricted theory. 

Under suitable continuity assumptions, the curve 
force, N, the director force, M, and the heat flux, 
h, can be expressed as S 


(er Ch. a 
Erie Mabeaicbia aU Doki 0 


(55) 


where q is the heat flux vector and the fields, N°, 
seme, q%, are functions of eY,t. The five conserva- 
tion equations in (54) then yield the local equa- 
tionst 


pa- = y(@)) (56) 
a) eyes 
(aN) PAE = Wy > Ky) (57) 
; z v £ 
Bw : 
Dr). ey = ue + y(kv * b+kw) , 
‘ s 
74.0 5 : 
(a°s’) Bi Netra ek MR or bx ykv , (58) 
Qa Qa 
ay x N + dq x (b 1) <5 Gl " x) (6b) Si) Oe 9) 
r - div =pe + N° OW + + Mw = 0 (60) 
f s@ p = ~, a. ie pe) u 


where "div is the surface divergence operator de- 
fined by divs q = q,q ° ao and a comma denotes par- 
tial differentiation with respect to the surface 
coordinates,0". It should be noted that the vector 
fields, Sc and s, are workless and do not contribute 
to the reduced energy equation (60). 

The above results include (60), which is derived 
from (54)5. For the purely mechanical theory in 
which the law of conservation of energy is excluded, 
the appropriate conservation laws are the first four 
of (54). In the context of the purely mechanical 
theory, it is worth recalling that the rate of work 
by all contact and assigned forces acting on P and 
on its boundary, dP, minus the rate of increase of 
the kinetic energy in P can be reduced to [see 
Naghdi (1972,1974)]: 


fpolt “vt 22 * w)do + f (N > v +M ° w)ds 
pois Yves we ys” s 


bgt (vy + v + 2kv + w + kw2)ao = [Pao 5 (Gib) 
dt Yp~ ~ ~ ~ = P 
where 
P= N° OF AY + mw + Mw 


is the mechanical power. 

Before closing this section, we also note that 
the restriction imposed on the motion of the medium 
by the condition of incompressibility, in the context 
ache restricted theory under discussion, reduces | 
to 


i line with a remark made at the end of the previous 
section, we note that equations (56)-(60) can also be 
derived by suitable integration across the thickness of 
the sheet, respectively, from the three-dimensional equa- 
tions (9)1 2,3 and the three-dimensional energy equation. 


am general, there are two conditions of incompressibility 
in the theory of incompressible directed fluid sheets; for 
a discussion of these, see Naghdi (1974, Section 3). In 
our present discussion, since d is assumed to have the form 
(51) ,, the second condition is satisfied identically and 
the corresponding pressure (arising from the constraint 
response) is a part of the response functions for Se and s. 
The specification (62) is motivated from an examination of 
the incompressibility condition in the three-dimensional 
theory when the position vector is approximated by (21). 


d 
ae [ajagd] = 0 (62) 
and can alternatively be expressed in the form 
a a 
(Glo aa” = (el oa aglow way °o we @ ofGs) 


= A! 2 


For an incompressible inviscid fluid sheet, which 
models the properties of the three-dimensional in- 
viscid fluid at constant temperature, we introduce 
the constitutive assumption that N*,m,M% do not de- 
pend explicitly on the kinematical quantities, Yiar 
WW ye and are furthermore workless, i.e., 


lo} 
IN 


+mw+Mw =O , (64) 


, ' 


provided v g and w satisfy the constraint condition 
(63). With the use of (51), it can then be shown 
that [see Green and Naghdi (1976, 1977)] 


NO = - p!{(d + a3)a" = (d + a°)ag} 
So g ap 
ne no 
Mes wa, om kr SOG (65) 


where PS is an arbitrary scalar function of ey,t 

and e¢8 is the alternating tensor in 2-space. With 
the help of the energy equation (60) and the fact 
that the mechanical power vanishes identically for 

an incompressible inviscid fluid at constant tempera- 
ture, it can be shown that [see the appendix of Green 
and Naghdi (1976) ] 


7. WATER WAVES OF VARIABLE DEPTH 


Within the scope of the restricted theory of the 
previous section, we include here an outline of a 
derivation of a system of nonlinear differential 
equations governing the two-dimensional motion of 
incompressible fluids for propagation of fairly long 
waves in a stream of water of variable initial depth. 
Our developments include the effects of gravity and 
surface tension but we assume that the mass density 
of the fluid does not vary with depth. However, a 
more general derivation for a nonhomogeneous inviscid 
fluid in which the mass density is allowed to vary 
with depth is given by Green and Naghdi (1977). Let 
e€1,e2,e3 be a set of right-handed constant orthonormal 
base vectors associated with rectangular Cartesian 
axes and choose the unit vector, b, to coincide with 
e3.- Then, the position vector, xr, in (46), and the 
director, d, in (51), can be represented as 

(66) 


BS son > Yep Sg GOSS sg 


where x,y,,> are functions of 9! ,62,t. The velocity, 
v, and the director velocity now take the forms 


YS Us 1 Vena Ney op We ue; a (67) 


where 


511 


u=x ,v=y , _=0 ,w= (68) 
and we note that the velocity components, u,v,A,w, 


may be regarded as functions of either 6) ,02,t or 


of x,y,t. From (67) follow the expressions 
Vue qt ven + he; " w = wes (69) 
and 
u POUL act UL Ulta A qv =v, + uv VV ' 
t y t x y 
Noh. “oh. 2 a i wo we & uw + vw , (70) 
t x y x 


where the subscripts, x,y,t, designate partial dif- 
ferentiation with respect to x,y,t, when u,v,A,w are 
regarded as functions of x,y,t. With the use of (67) 
and (70), the incompressibility condition (64) as- 
sumes the simpler form 
OGL AP A)! Ey SO) 5 (71) 
x y 
In order to complete our development, we need to 
specify values for the assigned force, £, and the 
assigned director force, 2, and to identify the co- 
efficients,y,k and k, which, in general, require 
constitutive equations. For this purpose we consider 
the corresponding fluid sheet in the three-dimensional 
theory in which an incompressible homogeneous fluid 
under gravity|| ,-g*e3, flows over a bed specified by 
the position vector 


r* = xe; + yep + a(x,y)e3 (72) 
and we specify the surface of the fluid by 
r* = xe, + yeo + B(x,y,t)e3 (73) 


In (72), a is a given function of x,y but 8 in (73) 
depends on x,y,t. At the surface (73) of the stream 
there is constant pressure, Por a constant normal 
surface tension, T. At the bed the (unknown) pres- 
sure, Pr depends on x,y and t. Thus, the normal 
pressure, p*, at the top surface (73) is 


w= 
19) 15) 


Sone neg 7 
T{(1 + 82)8 - 2888 + (1 + 62)6 3 
& WL 28 eV MKS EXammVs ; 
(Le BE & Bayeve 
x y 


(74) 


At the bed (72) the normal velocity of the fluid is 
zero and the pressure, p* takes the value 

p* = p(x,y,t) , (75) 
where p is to be determined. 

To proceed further, we recall the notation ais ((3})) 5 
let the surface, & = 0, defined by (15) coincide with 
the surface, J, and consider the three-dimensional 
region of space between the surfaces (72) and (73) 
occupied by the fluid. Any point in this three- 
dimensional region is then specified by 


I We use g* (instead of g) for gravity, since the letter, g, 
is used for a different quantity in (3), (5) and elsewhere in 
the paper. 


sae Fert 3he3 = xje]+ yep + (p + 0%p)e3 Pen GG) 7% 


where the surfaces, a and $8, in (26) or (72) and (73) 
correspond to 93 = Bale 93 = Eo, respectively. Also, 
x,y,W and $¢ in (76) are functions of 61,62 and t and 
GS Waren, g SW sp Saw o (77) 
Next, in order to obtain explicit values of y,k,k,f 
and 2 in relation to the top and bottom surfaces of 
the fluid, we choose the surface, 93 = 0, so that the 
center of mass of the three-dimensional fluid region 
under consideration always lies on this surface and 
we then identify this surface with the surface,d, 
in the theory of Cosserat surface. Without loss in 


generality, we may choose &) = -, E> = +5 (see 


Figure 2). This leads to the identification: 
1 *; ol * 3 ( ) 
= cakee= Zola) \3 eee x1Y 
Y = pa = ij Gel" =.) Sa 
5 aanes) 


1 
eo 


Ss 4 x L 
k = [ o g*e3ae3 =0 , 


es 7. Be) 
% > Sp 
p g7(6%)2ae2 = FS - (78) 
= 3 (61,67) 
where p* is the three-dimensional mass density in 
(9) and the determinant g defined by (3)3 is cal- 
culated from the approximation (21) so that 


1 
3 


a (x,y) 
= ae SS (79 
y 3(6!,62) : 


Substitution of (78) and the appropriate expressions 
for f and QR into (57) to (59) results in the dif- 
erential equations of motion 


ne 0 
p gu = = (Py = DB = pa, 


*ov=-p + (p -@)® - pa , 
p> I 1D. = Gi Be Pa, 
eta = Rok 
PRON ESP eet etd ¢ <7 
iL oO = 
ao OY SAG = 2.) = ape : é (80) 


where 


P=plo . (81) 


Moreover, Since the bed of the stream is stationary, 
from (77) and (70) 3,4 we have 


a = ua, + va = V -}i $ =o PAW 6 (82) 


The above system of equations is independent of the 
remaining equations (58) which involve S%,s. The 
fields, $%,s, correspond to appropriate constraint 
responses for the restricted motion (51). 


ak 
In (76) to (78), we have returned to the notation 


6° instead of & introduced in (2) 


The questions of continuous dependence upon the 
initial data and uniqueness for solutions of initial 
boundary-value problems for a class of symmetric 
flows characterized by a special case of the system 
of nonlinear partial differential equations given 
by Green et al. (1974c) has been discussed by Green 
and Naghdi (1975). A similar procedure may be used 
to establish uniqueness for the more general system 
of equations (80). 

For later reference, we consider here the reduc-— 
tion of the system of nonlinear differential equa- 
tions (71) and (80) for unidirectional flow in the 
absence of surface tension, T. Without loss in 
generality, we set the ambient pressure, py = 0, 


and consider flows in the x-direction only. Then, 
with q = 0, from (71) and (80) we obtain 
Qn 3? COBY) =O" 
OMS. st 
ep dh =p-o*g% , 
pe ow=-spee (83) 


We may solve (83)3,, for p and p and obtain the ex- 
pressions 


= * * ° 
DSO Oe SP WY) 
oF D 5 io 
bp=%p ¢ (g+dAt aw 6 (84) 


Introduction of (84); 9 into (83);,9 yields a system 
of two partial differential equations in u and w but 
we do not record these here. A further simplifica- 
tion of these equations results for a horizontal bed. 
For a horizontal bottom a may be taken to be zero and 
(77)1,2 and (68) 3,4 reduce to 


a=O , B=o , W=%o ,rA=4W . (85) 


8. FURTHER REMARKS 


The system of nonlinear differential equations (71) 
and (80)),2,3,4, which include the effects of gravity 
and surface tension, govern the two-dimensional mo- 
tion of incompressible inviscid fluids for the propa- 
gation of fairly long waves in a stream of variable 
initial depth. They are derived here by a direct 
approach as consequences of the conservation laws 
(54) subject to the incompressibility condition (64) . 
Upon specialization to unidirectional flow, the non- 
linear differential equations (71) and (80) reduce 
to those for inviscid fluids over a bottom of vari- 
able initial depth given by Green and Naghdi (1976a, 
Sections 5-6), while the equations for two- 
dimensional flow over a horizontal bottom were de- 
rived earlier [Green et al. (1974c)]. 

The differential equations governing the motion 
of a viscous fluid sheet are discussed briefly by 
Green and Naghdi (1976a, Section 11) and a similar 
development can be given within the framework of the 
restricted theory of Section 6, but we do not con- 
sider this aspect of the subject here. The system 
of differential equations obtained in Section 6 is 


valid for incompressible, inviscid, and homogeneous 
fluids. A more general derivation for propagation 
of fairly long waves in a nonhomogeneous stream of 
variable initial depth in which the mass density is 
allowed to vary with depth is contained in a recent 
paper of Green and Naghdi (1977). 

In the case of incompressible inviscid fluid 
sheets, the nonlinear equations for wave propagation 
in water of variable depth can also be derived from 
the three-dimensional theory: the procedure involves 
the use of the (three-dimensional) equation for con- 
servation of energy, the incompressibility condition, 
invariance requirements under superposed rigid body 
motions, along with a single approximation (21) for 
the position vector. Then, by (6) and (21), the ap- 
proximation for the (three-dimensional) velocity 
field is given by 


Vay Oy (86) 


where v and w in (86) have the same forms as those 
in (67). A derivation of this kind has been carried 
out by Green and Naghdi (1976b). It is important, 
however, to note that this derivation is limited to 
incompressible inviscid fluids which do not require 
constitutive equations. tt 

It is natural to ask what are the relationship and 
advantages (if any) between the above system of equa- 
tions and those which are currently employed by other 
investigators. To provide a ready comparison, we 
list below from Whitham (1974) alternative forms of 
equations for water waves moving in the direction of 
a fixed x-axis for a stream of initial constant depth, 
h. Let the elevation of the stream be h + yn. Then, 
for unidirectional flow and in terms of n and the 
horizontal velocity, u, we recall from Whitham 
(1974, pp. 460-463) the system of equations 


nt + f{u(h + nyt =OMine 


* L 2 2 
us + uu, +g es + c“hn =O© > (87) 


and the pair of equations attributed to Boussinesq, 
namely 


nte(h+tnu =0 , 
x 


° x it 
ar += = 
wi i. 3 Wuleare ORS, (88) 


where the notations in (87) and (88) are the same as 
those in (70), g* is the acceleration due to gravity 
introduced in Section 7 and c@ = g*h. Both systems 
of equations (87) and (88) allow for wave propagation 
in either direction along the x-axis. For waves moyv- 
ing along the positive x-direction only there is the 
Korteweg-deVries (1895) equation--hereafter referred 
to as the K.dV. equation--i.e., 


De & On =0 (89) 
h x 


3 
+ — 
Ne OMe se 6 XXX 


2 


tt A , E 2 
Recall that in the three-dimensional theory of incompress- 


ible inviscid fluids the stress vector is specified in terms 
of a pressure which is determined by the equations of motion 
and the boundary conditions. 


513 


or an equation due to Benjamin et al. 
by 


(1972) given 


Ss 
+ 
Q 
Ee 
+ 
LOH es) 


Ab uo) = 
yn, 6 ch eee ='0) 3 (90) 


‘gh 
h 
As already remarked by Green and Naghdi (1977), 
it may immediately be verified that the set of equa- 
tions (88) and (90) only have steady state solutions 

if n and u are both constants. Also, although the 
K.dV. Eq. (89) admits a solitary wave in which the 
velocity at infinity is zero and the stream there is 
at its undisturbed height, h, it does not admit a 
steady state solution with u constant and n = 0 at 
infinity. This fact is related to another property 
of (89) which is also shared by (88) and (90): the 
three sets of equations (88) to (90) are not invari- 
ant in form under a constant superposed rigid body 
motion of the whole fluid. To see this, suppose 
that a constant superposed rigid body translational 
velocity is imposed on the whole fluid so that the 
particles at the place, x, are displaced to x? at 
time, t , specified by 


+ 
See tb ee a te ue aE (91) 


where a and a are constants. The variables that oc- 
cur in the differential equations (87)-(90) are n = 
n(x, )) and w= wtx,t)). Let ni = nica t) and ut = 
ut(xt,tt) be the corresponding scalar quantities de- 
fined over the region of space occupied by the fluid 
after the imposition of the superposed rigid body 
motion (91),. Then, from (68), and (8.6) we obtain 


+ + + 
Dex) = el (Ge pie) Se 


+ —_ 
U 63 + ae, oe > am a «2 (92) 


We expect the elevation, h + n, of the fluid to re- 
main unaltered by superposed rigid body motions; and, 
since h remains unaltered also, this leads us to re- 
quire that 


n(x,t) = a’ Gee) = a” (GR eye, oh Gy) 5 (ES) 


From (92) and (93), we calculate expressions of the 
type 


= + = 
ue U) re en 2 ike Win, 9 
t x x 
ae ar 2, + 
= + 
xtt. Vee ae” etl). fa CW ne te 
ae me oS 18 oS 2S 
. + + o+ e+ 
Hh, FM Sy, Fe Hh = i , (94) 
te x + 
ste x 


with similar results for uz,uy and u in terms of ut 
and their derivatives. It was noted by Green and 
Naghdi (1977) that if the independent variables, x, 
t, in (88) to (90) are changed to (91), the equations 
for u,n in terms of xt+,tt+ are different from those in 
terms of x,t and this was illustrated explicitly with 
reference to the K.dV. equation (89). Here, we con- 
sider the pair of equations (88) 1 2- After substi- 
tuting (92)-(94), they become 


514 


x 
Pe Gap ees hg 

Ue Ul at, oe 

x Gta 

+ + 
=— 2a) - at : 

i tre Ie ca) oP ar op 2) 

BS) Bre) 1S Re “Bie ge 


The first of (95) is of the same form as (88), and 
hence invariant but clearly the second of (95) dif- 
fers from (88). This means that the character of 
the solutions of (88), (89) and (90) is substantially 
altered by superposing a constant rigid body trans- 
lational velocity on the fluid, which is contrary 

to what happens if we use the full three-dimensional 
equations of motion for an inviscid fluid. On the 
other hand, the set of equations (87) is not subject 
to this drawback, and the equations do have useful 
steady state solutions. It may be argued that be- 
cause of the nature of the approximation in obtain- 
ing (88) to (90) from the three-dimensional theory 
we should not expect these equations to be invariant 
under a superposed constant translational velocity, 
but this then leaves in doubt which version of any 
of the sets (88) to (90) are to be chosen as basic. 
The difficulty disappears if we linearize any of 

the above sets since the resulting equations are 
then invariant under a small superposed constant 
translational velocity, as we would expect. 

From the above discussion, it might appear that 
the equations (87) may be preferable to any of (88) 
to (90), but arguments are put forward by Whitham 
(1974, p.462) to suggest that the system (88) is to 
be preferred to (87). Although considerable use has 
been made of some of the equations (87) to (90), it 
would seem that they all rest on a somewhat shaky 
physical foundation. By contrast, the system of 
equations (71) and (80) do not possess the undesir- 
able features of the type noted above: they are 
properly invariant under superposed rigid body mo- 
tions, admit general steady state solutions, and are 
free from anomalies mentioned earlier. 

For the purpose of providing a more explicit com- 
parison with the system of equations (87) to (90), 
we specialize the system of equations (83) to that 
for a horizontal bottom for which (85) j 2,3,4 hold. 
Then, denoting again the elevation of the stream by 
h +n, the differential equations (83) 2 can be re- 
corded in the form 


i) ar (a a) SO) A 
beg. = iH (96) 
Pep) ube = 3 tt | Be 0 
where 
2 
BF Uae > sy Vaden 
11 i42 2 
-== + oo 
3 > p (Aes se Bh BU Shae i 


b=h+tn . (97) 


Clearly if R on the right-hand side of (96) can be 
neglected, then (96), 2 reduce to those of Boussinesq 
given by (88)],2. It should be emphasized, however, 


that the nonlinear equations (96); 9 are invariant 
under a constant superposed rigid body translation 
while (88) 9 are not.#+ Within the scope of the 
nonlinear theory, it does not seem reasonable to 
neglect the quantity, R, in (96) on the basis of 
either physical considerations or mathematical argu- 
ments. It may be, however, that in some special 
circumstances the solution of (88) is a good approxi- 
mation to the solution of (96), but this is a dif- 
ferent question than that discussed above. In this 
connection, it is worth noting that a solution to a 
system of differential equations, which results from 
neglecting certain terms in a more general system of 
equations, in general, will not be the same as a 
solution obtained by approximation from a correspond- 
ing solution of the more general system of equations. 
We close this section by calling attention to some 
available evidence of the relevance and applicability 
of the direct formulation for fluid sheets. The sys- 
tem of equations (71) and (81), or a special case of 
it, has already been employed in some detailed stud- 
ies of a number of two-dimensional problems of in- 
viscid fluid sheets, as well as in some comparisons 
with known previous solutions on the subject. We 
mention here some of these studies and refer the 
reader to the papers cited for additional informa- 
tion: (a) the nonlinear differential equations admit 
a solitary wave solution [see Green et al. (1974c)] 
which is the same as that attributed by Lamb (1932, 
Section 252) to Boussinesq and Rayleigh; (b) this 
solitary wave solution, as well as appropriate jump 
conditions and certain results derived from the 
energy balance for an inviscid fluid sheet at con- 
stant temperature [Green and Naghdi (1976a, Appen- 
dix)], has been used by Caulk (1976) to discuss the 
flow of an inviscid incompressible fluid under a 
sluice gate; (c) the steady motion of a class of 
two-dimensional flows in a stream of finite depth 
in which the bed of the stream may change from one 
constant level to another, and the related problem 
of hydraulic jumps, both for homogeneous and non- 
homogeneous incompressible fluids [Green and Naghdi 
(1976a, Section 7) and Green and Naghdi (1977)]; 
and (d) a class of exact solutions [Green and Naghdi 
(1976a, Section 9)] which characterize the main fea- 
tures of the time-dependent free surface flows in 
the three-dimensional theory of incompressible in- 
viscid fluids [Longuet-Higgins (1972)]. 


ACKNOWLEDGMENT 


The results reported here were obtained in the course 
of research supported by the U.S. Office of Naval 
Research under Contract NO0014-76-C-0474, Project 

NR 062-534, with the University of California, 
Berkeley. 


REFERENCES 


Benjamin, T. B., J. L. Bona, and J. M. Mahony (1972). 
Model equations for nonlinear dispersive systems. — 
Phil. Trans. Royal Soc. Lond. A272, 47. 

Bogy, D. B. (1978). Use of one-dimensional Cosserat 
theory to study instability in a viscous liquid 
jet. Phys. Fluids 21, 190. 


Fone difference between the two disappears, of course, upon 
linearization. 


Caulk Dey Ate (LOG) 
under a sluice gate. 
FIBIBIES Ye 

Caulk, D. A., and P. M. Naghdi (1978a). The influ- 
ence of twist on the motion of straight ellipitcal 
jets. Arch. Rational Mech. Anal. (to appear) = 
Report No. UCB/AM-77-5, Office of Naval Research 
Contract NOO014-76-C-0474, Project NR 062-534 
(June 1977), Dept. of Mechanical Engineering, 
University of Calif., Berkeley. 

Caulk, D. A., and P. M. Naghdi (1978b). The onset 
of breakup in inviscid and viscous jets, Report 
No. UCB/AM-78-3, Office of Naval Research Contract 
NO0014-76-C-0474, Project NR 062-534 (July 1978), 
Dept. of Mechanical Engineering, University of 
Calif., Berkeley. 

Chandrasekhar, S. (1961). Hydrodynamic and Hydro- 
magnetic Stability. Oxford: Clarendon Press. 

Cohen, H. (1966). A nonlinear theory of elastic 
diimectedmcurves!) int.) Wi. Engng. SCi- 4,05). 

Cohen, H., and C. N. DeSilva (1966). Theory of 
directed surfaces. J. Math. Phys. 7, 960. 

Cosserat, E. and F. (1909). Théorie des corps 
déformables. A. Hermann et Fils, Paris; also 
Theory of deformable bodies [transl. from original 
1909 edition], NASA TTF-11, 561, Washington, D.C., 
1968. 

Duhem, P. (1893). 
pression hydrostatique. 
3p ausi7/ 5 

Ericksen, J. L., and C. Truesdell (1958). Exact 
theory of stress and strain in rods and shells. 
Arch. Rational Mech. Anal. 1, 295. 

Green, A. E. (1975). Compressible fluid jets. 
Rational Mech. Anal. 59, 189. 

Green, A. E. (1976). On the nonlinear behavior of 
fluid jets: Int. TF. Engng. Sei. 14, 49). 

Green, A. E. (1977). On the steady motion of jets 
with elliptical sections. Acta Mechanica 26, 
ALWAYS 

Green, A. E., and N. Laws (1966). A general theory 
of rods. Proc. Royal Soc. Lond. A293, 145. 

Green, A. E., and N. Laws (1968). Ideal fluid jets, 
Inte Wi. Engng. Sel. 6), 30 

Green, A. E., and P. M. Naghdi (1974). On the 
derivation of shell theories by direct approach. 
J. Appl. Mech. 41, 173. 

Green, A. E., and P. M. Naghdi (1975). 


On the problem of fluid flow 
WBNS Wis ligleperess enh, gel 


Le potentiel thermodynamic et al 
Ann. Ecole Norm. 10, 


Arch. 


Uniqueness 


and continuous dependence for water waves. Acta 
Mechanica 23, 297. 
Green, A. E., and P. M. Naghdi (1976a). Directed 


515 


fluid sheets, Proc. Royal Soc. Lond. A347, 447. 

Green, A. E., and P. M. Naghdi (1976b). A deriva- 
tion of equations for wave propagation in water 
of variable depth. J. Fluid Mech. 78, 237. 

Green, A. E., and P. M. Naghdi (1977). Water waves 
in a nonhomogeneous incompressible fluid. J. 
Appl. Mech. 44, 523. 

Green, A. E., P. M. Naghdi, and W. L. Wainwright 
(1965). A general theory of a Cosserat surface. 
Arch. Rational Mech. Anal. 20, 287. 

Green, A. E., P. M. Naghdi, and M. L. Wenner (1974a). 
On the theory of rods. I. Derivations from the 
three-dimensional equations. Proc. Royal Soc. 
Lond. A337, 451. 

Green, A. E., P. M. Naghdi, and M. L. Wenner (1974b). 
On the theory of rods. II. Developments by 
direct approach. Proc. Royal Soc. Lond. A337, 


485. 

Green, A. E., N. Laws, and P. M. Naghdi (1974c). On 
the theory of water waves. Proc. Royal Soc. Lond. 
A338, 43. 


Korteweg, D. J., and G. deVries (1895). On the 
change of form of long waves advancing in a 
rectangular channel, and on a new type of long 
stationary waves. Philosophical Magazine (Fifth 
Ser.) 39), (422). 

Lamb, H. (1932). Hydrodynamics, 6th edn. 
University Press. 

Longuet-Higgins, M. S. (1972). 
time-dependent, free surface flows. 
Mech. 55, 529. 

Naghdi, P. M. (1972). The theory of shells and 
plates. S. Fliigge's Handbuch der Physik, VIa/2, 
C. Truesdell, ed., Springer-Verlag, Berlin, 425- 
640. 

Naghdi, P. M. (1974). Direct formulation of some 
two-dimensional theories of mechanics. Proc. 
7th U.S. National Congr. Appl. Mech., Amer. Soc. 
Mechanical Engineers, New York, N.Y., 3-21. 

Naghdi, P. M. (1975). On the formulation of contact 
problems of shells and plates. J. Elasticity 5, 
S725 

Rayleigh, Lord (1879a). On the instability of jets. 
Proc. Lond. Math. Soc. 10, 4. 

Rayleigh, Lord (1979b). On the capillary phenomena 
of jets. Proc. Royal Soc. Lond. 29, 71. 

Weber, C. (1931). Zum Zerfall eines Fltissigkeits- 
strahles, ZAMM 11, 136. 

Whitham, G. B. (1974). Linear and Nonlinear Waves, 
John Wiley and Sons. 


Cambridge 


A class of exact, 
J. Fluid 


516 


Discussion 


G. L. CHAHINE 


I would like to congratulate the author on his 
very fine work and to comment on his conclusion 
that the Rayleigh-Plesset equation represents fairly 
well the growth of bubbles attached to a wall. As 
is well-known, the Rayleigh-Plesset equation relates 
the growth and collapse of a spherical bubble, with- 
out relative motion with respect to the unbounded 
surrounding fluid, for a given variation of pressure 
far from it. It then seems really surprising that 
such an equation could describe so well the growth 
of the bubble on a blunt nose as shown in Figure 31. 
None of the requirements for the validity of the 
Rayleigh-Plesset equation are fulfilled: 
a. the bubble is non-spherical, even if we 
agree that the shape in the figure plan 
is a portion of a circle, 
b. presence of a wall, 
c. shear flow around the bubble, 
d. yrelative motion between the bubble and the 
fluid (as pointed out by the author). 
Moreover, the presence of gas inside the 
bubble is not taken into account, while the gas 
behavior has been shown to be very important 
(Chahine (1974, 1976)]. We believe that the good 
agreement between experimental results and analyt- 
ical computations shown in this paper is mainly 
due to: 
a. the time of observation is too small com- 
pared to the hypothetical lifetime of the 
bubble. (For a bubble radius of 1.3 mm 


and an external pressure of 5,000 N/m2, 


the Rayleigh time is about 0.7 ms and the 
lifetime is greater than 1.5 ms; say 10 
times the observation time.) 

b. in order to integrate numerically the 
Rayleigh-Plesset equation one needs two 
initial conditions: an initial radius 
and an initial growth rate. If R, and R 
replace these initial conditions it is 
not surprising that the result deduced 
for Ry differs only 4% from the experi- 
mental result. 

Concerning Table 4, the calculated relatively 
small effect of surface tension and viscosity is 
in good agreement with previous asymptotic studies 
{Chahine (1976) and Poritsky (1952) ]. 


REFERENCES 


Chahine, G. L., (1974). Etude Asymptotique et 
Experimentale des Oscillations et du Collapse 

des Bulles de Cavitation. EWSTA Report 042, 
CEDOCAR, MF 50831. 

Chahine, G. L., (1976). Etude Asymptotique du 
Comportment d'une Bulle de Cavitation dans un 
Champ de Pression Variable. Jl. de Mecanique, 16 
(2), pp. 287-306. 

Poritsky, H., (1952). The Collapse or Growth of 
a Spherical Bubble or Cavity in a Viscous Fluid. 
Proceeding of the First U.S. Nattonal Congress in 
Applted Mechanics, ASME, pp. 813-821. 


517 


Author’s Reply 


J. H. J. van der MEULEN 


The author appreciates Dr. Chahine's comments 
and would like to point out that the principal aim 
of comparing the cavity growth on the blunt nose 
with theory was to show that the travelling bubble 
type of cavitation is more related to bubble dynam- 
ics than to boundary layer phenomena. 

The surprising observation (Figure 31) that 
the shape of the attached, growing cavity is a 
spherical segment is, to a certain extent, consis-— 
tent with the observation by Dr. Chahine (1977) 
that the growth of the lower part of a bubble below 
a free surface is not influenced by the presence 
of the free surface. 

It seems most unlikely that the presence of 
gas originating from a small stream nucleus or from 


diffusion may have affected the growth of the cav- 
ity during the observation period. Oldenziel (1976) 
has shown that such effects can be neglected for 
explosive bubble growth. 


REFERENCES 


Chahine, G. L., (1977). Interaction between an 
Oscillating Bubble and a Free Surface. J. Fluids 
Engng., Trans. ASME, 99, p. 709. 

Oldenziel, D. M., (1976). Gas Transport into a 
Cavitation Bubble during an Explosion. JLAHR Symp. 
on Two Phase Flow and Cavitation tn Power Genera- 
tton Systems. Grenoble, France. 


518 


Discussion 


R. LATORRE 


Our lack of understanding of cavitation noise 
and its measurement technique is an area of recent 
concern and the authors' experiments and discussion 


will hopefully aid other researchers with these 
problems. 


The correlation of cavitation noise and the 
observed cavitation is a complicated research topic. 
In my dissertation I am studying tip vortex cav- 
itation noise and as a contribution to the authors’ 
paper, I would like to present some illustrative 


A 


UNIV, TOKYO CAVITATION TUNNEL 


2 3 4 
1 
KEY: 
A B 
hie) / A FOIL / PROPELLER 
. B TIP VORTEX CAVITATION 
— om 1 B& K 8103 HYDROPHONE 
2 B&K 2626 COND. AMP. 
3 RION 1/3 BAND PASS FILTER 
ee 4 — RION HIGH SPEED LEVEL 


FIGURE 1. Tip vortex cavitation noise measure- 
ment. 


SHIP RESEARCH INSTITUTE 
CAVITATION TUNNEL 


RECORDER 


SOUND PRESSURE LEVEL, dB IN 1/3 OCTAVE 


0.5 


1/3 OCTAVE BAND CENTER FREQUENCY 


SRI MEASUREMENT 
PROPELLER No. 121 
J = 35, N= 20 RPS 
1 - 6, = 23.06 
2-6, = 20.0 3-6,= 19. 


1/4 SRI FOIL, 6.5 M/S, 10° 


= J 
1-6, 8 49 
2 = Oy 3 

3 hy 0 


a, 


SOUND PRESSURE LEVEL, dB IN 1/3 OCTAVE 


2 5 10 KHz OSL 2 b) 10 KHz 


1/3 OCTAVE BAND CENTER FREQUENCY 


D) 3.15 KHz BAND CENTER FREQ, 

Ww 

=> 

= 

3 10 oB 

= 40 

a 

a ' ' ' \ \ I \ 

= 30 ha SO 8 Bis RO PY BUS. aay 

7 N/ 

a AIR CONTENT: 

ef a 26 %, 2.4 PPM 

w UNIV. TOKYO TEST 

= 1/4 SRI FOIL 

wo 12 M/S, 10° 10 dB 

a 0 7 1-6, =3. 
FIGURE 2. Tip vortex cavitation noise measure- = NO =1 INTERMITTENT = 2 STEADY = 3 
ments of propeller and foil tests comparison of = TIP VORTEX CAVITATION TEST (REF. €) 
intermittent and steady tip vortex cavitation 0.5 2 5 10 KHz 


noise. 


1/3 OCTAVE BAND CENTER FREQUENCY 


DEVELOPMENT OF TIP VORTEX CAVITATION NOISE 
TRACE OF NOISE SIGNAL AND NOISE ENVELOPE 


noise measurements made at the University of Tokyo's 
and the Ship Research Institute's (SRI) cavitation 
tunnel - 

Figure 1 shows the measurement apparatus. The 
hydrophone was set in a 50 mm acrylic cup mounted 
on the tunnel's observation window and filled with 
water. The measurements were made in uniform flow 
at constant speed with the section pressure lowered, 
using propellers and foils. The propeller was SRI 
No. 121 (D = 250 mm, z = 6, area ratio = 0.8, 
constant P/D = 0.75). The foil (1/4 SRI Foil) was 
a scaled version of Dr. Ukon's (SRI) design using 
NACA 4412 wing section and a planform of c(n) = 
c,(1-n*)%. The 1/4 SRI Foil had an aspect ratio 
of 3, semi-span = 50 mm, and base chord, c, = 40 mn. 

The measurements are briefly illustrated in 
Figures 2, 3, and 4. In Figure 2 the noise spec- 
trum and envelope of tip vortex cavitation noise 
is shown for SRI and Tokyo University tests. The 
intermittant tip vortex noise appears as spikes in 
the spectrum between 2 and 6.3 kHz, as denoted by 
"2" in this figure. Using the complete test 
record it is possible to construct the envelope 
shown in Figure 2D. The shifts in the frequency 


TRIGGER SIGNAL : 6.3 KHz BAND CENTER FREQUENCY 


519 


appear to be a function of both the low pressure 
vortex core and the condition of the water. 
In an attempt to gain an understanding of the noise 
mechanism, additional experiments were performed. 
In Figure 3, the intermittant tip vortex noise 
signal at 6.3 kHz was used to trigger the camera 
shutter to photograph the intermittant tip vortex 
cavitation. It appeared that the noise mechanism 
is due to the pressure wave caused by the filling 
of the low pressure vortex core by dissolved gases. 

To test this hypothesis of the tip vortex cav- 
itation noise mechanism, air was injected from the 
1/4 SRI Foil tip and the noise spectrum measured. 
Figure 4 shows the results of the initial tests 
illustrating a qualitative agreement in the actual 
tip vortex cavitation noise spectrum and the sim- 
ulated tip vortex using air injection. At the time 
of writing, it has been possible to improve this 
technique and duplicate the intermittant "spikes" 
in the noise spectrum. 

Thus by the experimental results a basis for 
understanding the low frequency aspects of tip vor- 
tex cavitation noise has become possible. 


40 1 o= 3.8 


NO. CAVITATION 


DGy= 3.8 
AIR INJECTION 
FROM FOIL TIP 


3 Gye Sul 
STEADY TIP 
VORTEX 
CAVITATION 


1/4 SRI FOIL 
V = 12 M/S5) 105 


AIR CONTENT: 
24%, 2 PPM 


nN WN 
=) =) 


ex 
(ao) 


SOUND PRESSURE LEVEL, dB IN 1/3 OCTAVE 


%0,8 i 2 5 10 KHz 


1/3 OCTAVE BAND CENTER FREQUENCY 


1/4 SRI FOIL, 10 M/S, 10°, 6, 
AIR CONTENT: 23%, 1.9 PPM 


© PHOTO 


= 3,36 
FIGURE 3. Intermittent tip vortex cavitation 
noise signal and photo. 
FIGURE 4. Comparison of tip vortex cavitation noise 


spectrum trace and simulated tip vortex using air 
injection. 


Authors’ Reply 


GORAN BARK and WILLEM B. van BERLEKOM 


It is very interesting to hear of this hypoth- 
esis concerning generation of noise by tip vortex 
cavitation. We have performed experiments with tip 
vortex cavitation at propellers and hydrofoils and 
found that intermittant tip vortex cavities were 
noisiest. However, we have not performed high speed 
filming or other more advanced attempts to study 
the real mechanisms involved in the volume fluctu- 
ations of the tip vortex cavity. In the case of 


bubble cavitation and unsteady sheet cavitation, 
which we have studied in more detail, we are of 

the opinion that the highest pulses are generated 
during the final part of a collapse, which often 

is rather symmetrical, and that filling the cavities 
with gas is of minor importance as a primary gen- 
eration mechanism. However, some results indicate 
that this gas decreases the violence of the collapse. 


Session VIT 


GEOPHYSICAL FLUID DYNAMICS 


WALTER H. MUNK 

Session Chairman 

University of California, San Diego 
La Jolla, California 


teen ee 


The Boussinesq Regime for waves 


in a Gradually Varying Channel 


John Wilder Miles 
University of California 
San Diego, California 


ABSTRACT 


The Boussinesq equations for gravity waves of ampli- 
tude a(x) and characteristic length 2£(x) ina 
gradually varying channel of breath b(x) and depth 
d(x) are derived from Hamilton's principle on the 
assumptions that a/d =a << 1, (a/2) 2 = OCC), io (9) 
= 0(a3/2b/a) and d“(x) = 0(a3/2) (* = d/dx). The 
further assumption of unidirectional propagation 
then leads to the Korteweg-deVries equation for a 
gradually varying channel. It is shown that the 
latter equation admits two integral invariants. 

The second-order (in amplitude) invariant measures 
energy, as expected, but the first-order invariant 
measures mass divided by pid’ ; accordingly, mass 

is conserved only if either the first-order invariant 
vanishes identically or ba is constant, and only 
the former possibility appears to be consistent 

with conservation of energy. An approximate solution 
for a cnoidal wave, which conserves both energy and 
mass, is developed. The corresponding approximation 
for a solitary wave (which may be regarded as a 
limit of a cnoidal wave) does not conserve mass but 
nevertheless provides an approximation to the evolu- 
tion of the amplitude, a « bp 2/3q7l, that is in 
agreement with experiments for gradual decrease of 
depth or increase of breadth but not for decrease 

of breadth. 


1. INTRODUCTION 


The Boussinesq régime for gravity waves of amplitude 
a and characteristic length 2 in water of depth d 
is characterized by 


GCSv“vd<i, Ge GM <« il, BS OG), Gade) 


where a and 8 are measures of nonlinearity and 
dispersion, respectively, and (lc) refers to the 
asymptotic limit a + 0. The assumptions of one- 
dimensional wave motion and uniform depth and the 


523 


neglect of compressibility and viscosity then imply 
Boussinesq's equations for the free-surface displace- 
ment and the depth-averaged velocity, n(x,t) and 
u(x,t). The further assumption of undirectional 
propagation permits the elimination of u to obtain 
the Korteweg-deVries (KdV) equation for n. The 
classical derivations are given by Whitham (1974, 
§13.11). An alternative derivation, starting from 
the Luke-Whitham variational principle and using &, 
the velocity potential at the free surface, and n 
as dependent variables also has been given by Whit- 
ham (1967). 

I consider here the generalization of the 
Boussinesq and KdV equations for a channel of grad- 
ually varying breadth and depth b(x) and d(x) and 
their approximate solution for slowly varying 
cnoidal and solitary waves. I begin (in Section 2) 
by deriving (what may be called) the Boussinesq chan- 
nel equations directly from Hamilton's principle (to 
which the Luke-Whitham variational principle is 
equivalent in the present context) on the basis of 
(1) and the further assumptions (which imply 
gradually varying) 


3 3 
b*(x) = O(a /2b/a) , a7 (x) = O(a /2) (2a,b) 


I then (in Section 3) invoke the hypothesis of uni- 
directional propagation to obtain the KdV channel 
equation, which was developed originally by Shuto 
(1974) through a rather more involved procedure. 

I then go on to consider cnoidal waves in Section 
4 and the solitary wave in Section 5 on the basis 
of the stronger assumptions 

|b*| << 0 3/2 (pa) la*| << a °72 (3a,b) 

A prominent feature of the KdV equation for a 
uniform channel is the existence of an infinite 
number of integral invariants (Whitham, 1974, §17.6). 
The KdV equation for a slowly varying channel admits 
only two such invariants, of first and second order 
in the amplitude; the latter measures energy, as 


524 


expected, but the former measures mass only if ba” 

= constant. This deficiency is presumbly a conse- 
quence of the implicit neglect of the weak reflection 
that accompanies the gradual variation of the channel: 
the reflection coefficient for energy is second 

order in some appropriate measure of the channel 
variation and therefore has no cumulative effect, 
whereas that for mass is first order and does have 

a cumulative effect. The resulting difficulty may 

be avoided for a wave that is either periodic or of 
compact support simply by choosing a horizontal 
reference plane such that the mean value of the 
free-surface displacement vanishes identically (see 
Section 4), but the problem is more subtle for an 
aperiodic disturbance of unlimited extent such as 

a solitary wave (see Section 5) and remains unre- 
solved. 

The primary goal, at least for practical applica- 
tions, of the analysis of waves in a gradually 
varying channel is the prediction of a as a function 
of b and d. Green's law, which neglects both non- 
linearity and dispersion, predicts [Lamb (1932, 
§185)] 

A GI” 5 (4) 
It is often used for practical shoaling calculations, 
and Shuto (1973) finds that a « d “ holds for 
solitary waves on relatively steep slopes for a/d as 
large as 2. On the other hand, the joint assumptions 
of Boussinesq similarity (a/d « d2/22) and conser- 
vation of energy (which is proportional to abo) 
imply [Miles (1977a)] 


ace pb 2/3q-1 . (5) 


Comparison with experiment (see Section 5) suggests 
that (5) should be valid for a shoaling-or laterally 
diverging channel if |6| < 0.1, where 


5 = Ge) 72a) & Sey (6) 


but perhaps not for a laterally converging channel. 
The present results also have implications for 
the approximate treatment of nonlinear wave propa- 
gation along the lines initiated by Whitham (1974, 
Ch. 8) in his treatment of shock-wave propagation 
and since applied to solitary waves [Miles (1977a)]. 


2. BOUSSINESQ CHANNEL EQUATIONS 


The boundary-value problem for gravity waves in an 
ideal, homogeneous liquid may be deduced from 
Hamilton's principle in the form [Broer (1974), 
Miles (1977b) ] 


LP 1 
6[[Lté n}axat = 0, L = En, - Bune esy - 590°, (7a,b) 


where x and y are horizontal and vertical coordinates; 
6 (x,t) and n(x,t) are the velocity potential at, and 
the displacement of, the free surface; dx is an 
element of area in the x space; d(x):is the quiescent 
depth; and the velocity potential $(x,y,t) is 
determined by 


V26= (0)  (-d <y <n) , (8) 


Re ar NKSIONKy = 0) (C7 el) ) StS GZ =) (9a,b) 


The solution of (8) and (9) is given by 


§ = & - y¥-(avey - Sy2v2e + 01826) (10) 


where 8 is defined by (lb) with d and 2 as scales 

of y and x. The corresponding approximation to 

the kinetic energy integral, after invoking n = O(ad), 
aVE = 0(8%E), B = O(a), (2), and V*(AVB) = VA*VB + 
AV2B, is 


n 
f (Vo) 2dy = (atn) (VE)? - $a3 (026)? + 29- [a3 (v2E) VE] 
OGRE) (11) 


Substituting (11) into (7), invoking the further 
approximation that ¢ is independent of the transverse 
coordinate in a channel of slowly varying breadth 
b(x) and depth d(x), and integrating across the 
channel, we obtain 


L 2, 13-2 dL 
-= + =a - = 

offen, 3 (d+n) ES Go ex 7 GIN |bdxdt = 0. (12) 
The corresponding Euler-Lagrange equations, 
Ll g3 
=. + = 
3 (bd See [b(d+n)e), ur bn, 0) (13a) 
and 

E, + 562 + on = 0 (13b) 

t Dex. . 


are counterparts of the Boussinesq equations [cf. 
Whitham (1967)]. 

It is worth noting that the approximations to 
this point are consistent with conservation of both 
mass and energy: 

3, [nbax = Op a, [Jptarm ey - aa3e2 + Sgn? bdx = 0, 
(14a,b) 


where the integrals are over either (-~,~) or a 
periodic interval. The integral (14a) follows 
directly from the integration of (13a) with respect 
to x, subject to appropriate null or periodicity 
conditions at the end points. The integral (14b) 
may be similarly established or may be inferred 
(through Noether's theorem) from the invariance of 
the Lagrangian density in (12) under a translation 
of t;:it is an exact invariant of (13), but it 
would be consistent with the antecedent approxima- 
tions to approximate the specific energy in (14b) 
by 4 (d&2 + gn7). 


3. KORTEWEG-DEVRIES CHANNEL EQUATION 


The Korteweg-deVries (KdV) equation for uni- 
directional wave propagation in a uniform channel 
may be deduced from the Boussinesq equations by 
assuming that & and n are slowly varying functions 
of t in a reference frame moving with the wave speed, 
c. It is expedient in the present context to choose 
x, rather than t, as the slow variable (since b and 
d are prescribed as slowly varying functions of x) 
and to introduce 


- = gd alls) 
aes) E (c gd) (15) 
as a characteristic variable. The direction of 
propagation may be reversed by reversing the sign 
(ope qo) bhigy, ((alS))) 4 


The reduction of (13) on the hypothesis that Ny 
= O(ans) yields 


Qe a j= 

3d Vike ) pe + 3i(cd)) mm. + an, + (Abe)im = 0; (16) 
where 

A( ) = (d/dx)log( ) (17) 
(note that Ac = 4Ad). Equation (16), which appears 


to have been derived originally by Shuto (1974), 
reduces to the KdV equation if b and d are constant. 
The vertically averaged, horizontal velocity is 

given by 
u = (gn/c) [1+0(a)], (18) 
whilst the vertical velocity is O(a %u). The mass, 


Momentum, and energy of the wave therefore are given 
by 


co foe) co 
M = pbc [ nas, M = pba [ uds| ="Me,, & = pgbe [n2ds, 
— co —-o —o 
(lS ay o7C). 


within 1+0(a). The limits of integration may be 
replaced by +4T for a wave of period T. 
Multiplying (16) through by (be) and ben, 


respectively, and integrating over -~ < s < ~ on 
the assumption that n, Ns, and ngg vanish in the 
limits, we obtain the integral invariants 
Cc foe} 
3 2 
I = (bc) nds, J =bc n“ds. (20a,b) 
—oco —oO 


It follows that E = pgJ is conserved. On the other 
hand, 


1 
ai(aa) ema Ma pn@ses)? 


M = (2la,b) 
so that, except for special combinations of b and 
d, M and M are conserved only if S°.nds = 0. Non- 
conservation of momentum is acceptable in consequence 
of the horizontal thrust exerted on the fluid by 
the bottom and walls of the channel, but non- 
conservation of mass is generally unacceptable. 

We remark that the neglect of both dispersion 
and nonlinearity, as represented by the first and 
second terms, respectively, in (16), yields Green's 
law, (be) 2n = f(s), where f is an arbitrary function 
of the characteristic coordinate, s. 


4. SLOWLY VARYING CNOIDAL WAVE 


Theory 


Kinematical and scaling considerations suggest that 
an approximate solution of (16) for a wave of pre- 
scribed period 


We Ann = (t/g)? (22) 
be posited in the form 
n(s,x) = a(x)N(6,x), 68 = ws - x(x), (23a,b) 


where 8 and x are fast and slow variables, a(x) is 
a slowly varying amplitude, and x(x) is a slowly 
varying phase shift. It also is expedient to 
introduce 


525 


y¥ (x) = 2(cd/aw) x7 (x), (24a) 


such that the phase speed of the wave is given by 
. 1 
-8,/8, = c/{1-y(a/a] = [g(dtya)]*. 


Conservation of mass and energy imply the constraints 
(see Section 3) 


ne (24b) 


<N> = (0), a*beT<n2> = 5, (25a,b) 
where < > implies an average over a 2m interval of 

8 and J is the integral invariant obtained through 
the substitution of (23) into (20b). 

A formal, asymptotic development of the descrip- 
tion (23) may be obtained by expanding N(8,x) and 
y(x) in powers of an appropriate measure of the 
slow variation of b and d and invoking (25a) and 
the requirement that the period of 6 be 27. The 
first approximation, which is obtained by substit- 
ing (23) into (16) and then neglecting all 
derivatives with respect to the slow variable x, 
corresponds to that for a cnoidal wave [Lamb (1932, 
§253)]. It may be placed in the form 


N en? [ (K/t) 6|m] - <cn2> F <cn2> = 


l 


[m-1+(E/K)] /m, 
(26a,b) 


il 


y = [2-m-3(E/K)]/m, aL/a% = (16/3)mk2 = U(m), (26c,d) 


where en (u|m) is an elliptic cosine of modulus vn 
and K and E are complete elliptic integrals in the 
notation of Abramowitz and Stegun (1955), and U(m) 
is the local Ursell parameter. Substituting (26) 


into (25b), we obtain 

ah opie 2) = WPaes = Fen), (27a) 
where 

<n2> = <cnt> - <cn2>2 = [2(2-m) (E/K) - 3(E/K)2 

- (1-m)]/(3m?) (27b) 

and 

E = (4°/3°)x2[2(2-m)EK - 3R2 - (1-m) K?]. (27¢c) 

It follows from (27), which determines m(x), that 


m is constant if and only if pad/2 = constant, in 
which special case (23), (26), and (27) constitute 
an exact similarity solution of (16). e 

The results (26a) and (27a) provide a parametric 
relation between aL/d2 and gL3/2 /pa/2 that may be 
graphically represented as a plot of log F vs log U 
[see Miles (1978b)]. The case of constant depth is 
especially simple in that the plot of log F vs log 
U is equivalent to -log b vs log a. The limiting 


relations 

F > au? JAS Ghia cas 2 yo (283i) 
and 

ae an”, ie Saar (U + o) (29a,b) 


intersect at U = 
for U> 150. 

The preceding calculation is a generalization of 
that of Svendsen and Brink-Kjaer (1972), who consider 
the one-dimensional (b = constant) shoaling problem; 
however, they replace 8 + wt in (23b) by the 
equivalent of [1 - 4y(a/da] (x/c), which is clearly 


150 and provide rough approximations 


526 


in error unless both b and d are constant. 
The problem also is attacked by Shuto (1974), 
who allows for the variation of both b and d but 


arrives at a result (which he integrates numerically) 


that appears to be inconsistent with conservation 
of energy. However, his result is consistent with 
(28) in the limit U + 0 and with (29) in the limit 
U + © or, more precisely, with the result obtained 
by neglecting only terms of exponentially small 
order in (27), 


3/2 2 
F~ Gu “- 2G) Ute), (30) 
which is in error by less than 1% for U > 70. It 
therefore appears that Shuto's numerical results 
are not significantly in error (on the scale of his 
plots) over the entire range of U. 


Experiment 
Shuto (1974) compares his results with his own 


experimental observations and with those of Iwagagi 
and Sakai (1969) for shoaling waves periods from 


1.2 to 6 seconds on uniform slopes of 1/20 and 1/70. 


He concludes that linear surface-wave theory (which 
presumably accounts exactly for dispersion) is 
superior to his cnoidal-wave results for U < 30 and 
conversely for U > 30 and that the latter are good 
for a/d as large as 0.8. 


5. SLOWLY VARYING SOLITARY WAVE 
Theory 


The slowly varying solitary wave 


*5 1 
nN = asech? Cra)” (fee t) , C= [g(dta)]°*, 
(3la,b) 
Die OVS =. 
a= a3 %% cia (31c) 


is obtained by letting U + © with KO = 0(1) in (26) 
and (27).* There is, however, a new difficulty: 
none of the integrals I, M, and M [see (20a) and 
(2la,b)], which now are proportional to pl/6g3/4 
b2/3q, and p2/3q3/2 respectively, is conserved ex- 


cept for special variations of b and d. [The failure 
of the condition <N> = 0 in the limit U + ~ is a con- 


sequence of the loss of the displacement -a<cn*> ~ 
a/K, which cancels the mean of acn? (2K8) when inte- 
grated over -K < 2K9 < K.] It follows that, except 
in the special case ba?/2 = constant for which (31) 
is an exact solution of (16) and M and M vary like 
a-? and a3/2, respectively, (31) cannot be a uni- 
formly valid approximation to the solution of the 
KdV channel equation (16); instead, it is the first 
term in an inner expansion, which must be matched 
to an appropriate outer expansion. 

Johnson (1973) obtains the next term in an inner 
expansion for b = constant and finds that it can 
be matched to an appropriate outer expansion if d 
is increasing in the direction of propagation (the 


solitary wave may undergo fission if d is decreasing) ; 


*The prediction that a «= b72/3q71 appears to be due 
originally to Saeki, Takagi, and Ozaki (1971); see 
also Shuto (1973, 1974) and Miles (1977a). 


however, he does not obtain an explicit description 
of the oscillatory tail, nor does he allow for the 
possibility of expanding the slowly varying phase 
X(x) as well as N(6,x) [see (23)]. 

Ko and Kuehl (1978) have criticized Johnson for 
this latter omission and develop a joint expansion 
of (the equivalents of) N and x. They conclude 
that the solitary wave ("soliton") experiences an 
irreversible energy loss in the sense that it does 
not re-establish itself if the channel gradually 
reverts to its initial, uniform breadth and depth. 
This may be, but the proper form of the inner 
expansion is to some extent a matter of expediency, 
and the ultimate validity of any particular expansion 
can be established (albeit heuristically) only 
through matching to a proper outer expansion. Ko 
and Kuehl appear to overlook the crucial role of 
matching, and, at least in this important respect, 
their results must be regarded as incomplete. 

Johnson's results are readily generalized to 
allow for the variation of both b and d and reveal 
that 


§ = 2(3a/d) °/2aA(ba9/2) = (3a)~3/2 (2aAb + 947) 
(32) 

is an appropriate measure of the slow variation of 
the channel (this same measure also is appropriate 
for a cnoidal wave for U > 100). The Boussinesq 
equations (13) and KdV equation (16) are based on 
the restriction 6 = O0(1) asa ¥0 [cf. (2)], whereas 
(26) and (31) are based on the stronger assumption 
|| << 1 [cf. (3)]. Moreover, a consideration of 
the special case of linearly increasing breadth and 
constant depth [Miles (1978a)] suggests that the 
wave ultimately ceases to be solitary and evolves 


0.4 


0.02 


0.01 
10 


FIGURE 1. Decay of a solitary wave in a linearly ex- 

panding channel. The wave is propagating in the posi- 

tive-x direction, where x is measured from the virtual 
origin at which b = 0, and enters the diverging chan- 

nel (from an entry section of uniform width) at 

x/d + 10. The amplitudes at the transition station are 
a/d = 0.05(x), 0.1(+), 0.2(0), and 0.4(:*). The dashed 

lines have slopes of -2/3. 


0.8 


0.4 


0.2 
a 
id 

04 

0.05 
0.025 

Xx /d 
FIGURE 2. Growth of a solitary wave in a linearly con- 


tracting channel. The wave is propagating in the nega- 
tive-x direction (right to left), where x is measured 
from the virtual origin at which b = O, and enters the 
converging channel (from an entry section of uniform 
width) at x/d + 94. The amplitudes at the transition 
station are a/d = 0.05(x), 0.1(+), 0.2(0), and 0.4(:); 
the corresponding slopes of the dashed lines are 
“DoE, “Oo, SO MLE Etvel o)o4i- 


into a dispersive wave train for which the first 

peak closely approximates a solitary wave in shape 
but is followed by successive peaks of only gradually 
diminishing amplitude. There remains, however, the 
difficulty of nonconservation of mass, and the 
general problem of an aperiodic wave (in particular, 
an initially solitary wave) in a gradually varying 
channel is unresolved at this time. 


Experiment 


Shuto (1973) compares Green's law, a = ars, and the 
present prediction a « qd7l, with the experimental 
observations of Camfield and Street (1969) and Ippen 
and Kulin (1954) for shoaling of solitary waves on a 
uniform slope. He concludes that the range of valid- 
ity of the "-l power" law decreases with increasing 
slope and that the "4 power" law holds for slopes in 
excess of 0.045 and a/d as large as 2.0. A more 
precise comparison can be made on the basis of (32), 
which reduces to 


6 = 9(3a/d)~3/2a°  (b = constant) (33) 


for a channel of constant breadth. The estimated 
critical values of 6, such that a « a7! or a-*% pro- 
vide better fits to the data for 6 < 6* or 6 > 6%, 
respectively, are 6* = 0.10, 0.10, and 0.09 for 
slopes of .01, .02, and .03, anda « a! is typically 
within the experimental scatter for 6 < 0.01. 

Chang and Melville (unpublished) have recently 
measured a(x) in linearly diverging and converging 
channels. Their results for a diverging channel 
(Figure 1) tend to confirm the prediction a « b~2/3 
for initial values (at the transition from a uniform 
channel) of 0.05 < a/d < 0.2 [the corresponding 
values of 6 = 2(3a)-3?/2(db“/b) are in the range 
(0.01, 0.07], although the decay ultimately exceeds 
this inviscid prediction--presumably in consequence 
of viscous or other dissipation -- and exceeds it 
after only a rather brief section for an initial 


527 


value of a/d = 0.4. Their results for a converging 
channel (Figure 2) predict a growth that is roughly 
approximated by a = b-9-4, Dissipation in the 
converging channel would tend to decrease the magni- 
tude of the exponent, but why this decrease should 
be so much larger than the corresponding increase 
for the diverging channel is not clear at this time 
(intuition suggests that reflection could be more 
significant in a converging than in a diverging 
channel, but neither analytical nor experimental 
evidence is available to support this conjecture). 


ACKNOWLEDGMENT 


This work was partially supported by the Physical 
Oceanography Division, National Science Foundation, 
NSF Grant OCE74-23791, and by the Office of Naval 
Research under Contract NO0014-76-C-0025. Most of 
the material in Sections 3 and 4 has been published 
elsewhere [Miles (1978b)] in slightly different 
form. 


REFERENCES 


Abramowitz, M., and I. Stegun (1965). Handbook of 
Mathematical Functions, Bureau of Standards, 
Washington, D. C. 

Broer, L. J. F. (1974). On the Hamiltonian theory 
of surface waves. Appl. Sci. Res. 30, 430-446. 

Camfield, F. E., and R. L. Street (1969). Shoaling 
of solitary waves on small slopes. Proc. ASCE, 
Waterways and Harbors Div. 95, 1-22. 

Ippen, A., and G. Kulin (1954). The shoaling and 
breaking of the solitary wave. Proc. 5th Coastal 
Engineering Conference, 27-49. 

Iwagaki, Y., and T. Sakai (1969). Studies on cnoidal 
waves (seventh report) - Experiments on wave 
shoaling. Dis. Pre. Res. Inst. Annals, No. 12B, 
Kyoto Univ., 569-583 [in Japanese; cited by Shuto 
(oT): 

Johnson, R. S. (1973). On the asymptotic solution 
of the Korteweg-deVries equation with slowly 
varying coefficients. J. Fluid Mech. 60, 813-824. 

Ko, K., and H. H. Kuehl (1978). Korteweg-deVries 
soliton in a slowly varying medium. Phys. Rev. 
Lett. 40, 233-236. 

Lamb, H. (1932). Hydrodynamics, Cambridge University 
Press. 

Miles, J. W. (1977a). 
a Slowly varying channel. 
149-152. 

Miles, J. W. (1977b). On Hamilton's principle for 
surface waves. J. Fluid Mech. 83, 153-158. 

Miles, J. W. (1977c). Diffraction of solitary waves. 
ZAMP 28, 889-902. 

Miles, J. W. (1978a). An axisymmetric Boussinesq 
wave. J. Fluid Mech. 84, 181-192. 

Miles, J. W. (1978b). On the Korteweg-deVries 
equation for a gradually varying channel. J. 
Fluid Mech. (sub judice). 

Rayleigh, Lord (1876). On waves, Phil. Mag. 1, 
257-279; Papers 1, 251-271. 

Saeki, H., K. Takagi, and A. Ozaki (1971). Study 
on the transformation of the solitary wave (2). 
Proc. 18th Conf. on Coastal Engg. in Japan, 49- 
53 [in Japanese; cited by Shuto (1974)]. 


Note on a solitary wave in 
J. Fluid Mech. 80, 


Shuto, N. (1973). Shoaling and deformation of non- 
linear long waves. Coastal Engineering in Japan 
HG, alow. 


528 


Shuto, N. (1974). Nonlinear waves in a channel of 
variable section. Coastal Engineering in Japan 
7, dA. 


Svendsen, I. A., and 0. Brink-Kjaer (1972). Shoaling 
of cnoidal waves. Proc. 13th Coastal Engineering 
Conference (Vancouver 1, 365-383. 


Variational methods and 
Proc. Roy. Soc. 


Whitham, G. B. (1967). 
applications to water waves. 


Lond. A 299, 6-25. 
Whitham, G. B. (1974). Linear and nonlinear waves, 


Wiley-Interscience, New York. 


Study on Wind Waves as a 
Strongly Nonlinear Phenomenon 


Yoshiaki Toba 
Tohoku University 
Sendai, Japan 


ABSTRACT 


Recent studies on wind waves in our laboratory, 

from a view point of strong nonlinearity of the 
wind waves, are reviewed. The main items are as 
follows. (1) It has been shown by experiments and 
theoretical analyses that the mechanism of initial 
generation of waves by the wind is the instability 
of shear flows of two-layer viscous fluids, air and 
water. It is a selective amplification of distur- 
bances at the frequency of maximum growth rate. 
However, the transition of the initial wavelets to 
irregular wind waves including turbulence follows 
within several seconds [Kawai (1977)]. (2) Flow 
visualization studies of the internal flow pattern 
of wind waves show that the shearing stress of the 
wind is concentrated at the crest and windward face 
of individual waves, and a special area is formed 
where the surface wind drift, and consequently the 
vorticity is concentrated, causing the forced con- 
vection or the turbulent mode, which is the origin 
of the irregularity of wind waves [e.g., Toba et al. 
(1975); Okuda et al. (1977)]. (3) Statistical 
investigation of instantaneous individual waves in 
a wind-wave tunnel shows clearly the existence of 
similarity in the individual waves [Tokuda and Toba 
(1978)]. Namely, the energy spectrum, which is 
newly defined for the individual waves, is virtually 
equivalent to the traditional energy spectrum at 

the frequency range from 0.7- to 1.5-times the 
frequency of the energy maximum. However, the energy 
peaks which usually appear in the traditional spec- 
trum at the higher harmonics of these dominant waves 
completely disappear. The apparent phase speed of 
individual waves, for each wind and fetch condition, 
is inversely proportional to the square root of 
their frequency, and is much larger than the phase 
speed of linear water waves. For the individual 
waves for each wind and fetch condition, there 
exists statistically a conspicuous relationship of 
the 3/2-power law [cf., Toba (1972, 1978a)] between 
the normalized wave height and period. Consistently 


529 


with this and the phase speed relationships, the 
steepness of the individual waves is statistically 
constant. (4) Discussicn is presented as to the 
possibility of approaching the above-mentioned 
characteristics of the individual waves from the 
similarity hypothesis and dimensional considerations. 
Self-adjustment of the individual waves to the 

local wind drift distribution is postulated to 
explain the 3/2-power relationships, which may be 
the basis of the possibility that the pure wind-wave 
field is represented by a single dimensionless 
parameter [Toba (1978a)]. (5) A new formulation 

is presented for the roughness parameter or the 

drag coefficient over the wind waves, incorporating 
the single dimensionless parameter of the wind-wave 
field. A physical interpretation of the form is 
given from the internal flow pattern of individual 
waves [Toba (1978b)]. 


1. INTRODUCTION 


In a traditional model, the wind waves are treated 
as phenomena, expansible to component free water 
waves having weakly nonlinear interactions among 
waves of different wave numbers. However, detailed 
experimental studies on the actual conditions of 
wind waves produced in wind-wave tunnels, have 
shown that wind waves are much more strongly non- 
linear phenomena, especially in their younger stages. 
This report presents a review of recent studies 
made in our laboratory, giving much emphasis to 

the strong nonlinearities which are inherent in 
wind waves. 


2. INITIAL GENERATION OF WIND WAVES 


The first topic starts with an approach from the 
process of the initial generation. The wind waves 
have long been assumed to be generated from a still 
water surface by the effect of pressure fluctuations. 


FIGURE 1. Flow visualization 
of the initial stage of the 
generation of wind waves by 

use of hydrogen bubble lines 
produced by the electrolysis 

of water. The photographs were 
taken from a viewpoint slightly 
below the air-water interface, 
so images reflected at the in- 
terface are seen in the upper 
1/4 of each picture. The hydro- 
gen bubble lines are produced 
near the left end as pulses of 
0.002-s width at 0.04-s inter- 
vals in a very slow, uniform 
flow of water which was pro- 
duced before the start of 

wind. The wind was 6.2 m/s 
blowing from left to right of 
each picture. The filmed time 
of each picture from the start 
of the wind is shown in sec- 
onds. The out-of-focus areas 
were caused by some fluctuation 
of the mean flow, for very shal- 
low depth of the focus. In (e) 
are seen the initial wavelets, 
and in (f) is seen the onset of 
turbulent mode. [Cited from 
Okuda et al. (1976).] 


| Ma 


(a) 0.40 sec 


4 


(d) 3.29 


A resonance mechanism for the initial generation 
proposed by Phillips (1957) and an instability 
mechanism for further growth proposed first by Miles 
(1957) have been referred to on every occasion. 
Valenzuela (1976) showed that the growth rate of 
waves in the gravity-capillary range, observed by 
Larson and Wright (1957) at the initial stage of 
the generation, agrees with the expected growth 
rate by the instability theory applied to a coupled 
shear flow of the air and the water. 

Kawai (1977 and 1978) of our laboratory has 
arrived at the conclusion, by systematic experiments 
together with theoretical analyses, that the mech- 
anism of generation of the initial wavelets is the 
instability in a two-layer shear flow of viscous 
fluid of air and the water, as a selective amplifi- 
cation of disturbances of the frequency at the 
maximum growth rate. 

The experiments were carried out mainly by use 
of a wind-wave tunnel of 20 m length, 60 cm x 120 cm 
cross-section, containing water of 70 cm depth. 
After the sudden starting of wind on the still 
surface of water, a shear flow first develops in 
the uppermost thin layer of water, and several 
seconds later, regular, long-crested initial wavelets 
appears [Figure l(e)]. His theoretical analysis of 
the shear flow instability of the two-layer viscous 
fluids, using the actual profile of the shear flow 
in water, shows that the system is unstable and 
there exists a frequency at which the growth rate, 
kCj, is maximum (Figure 2). The frequency of kCj- 
maximum does not necessarily coincide with that of 
Cr-minimum, or the minimum phase speed for the 
gravity-capillary wave. Three properties of the 
initial wavelets determined by the experiment, i-e., 
the frequency, the growth rate, and the phase speed 
are all virtually coincident with those of the 
theoretically predicted waves of the maximum growth 
rate as shown in the following. 

Figure 3 shows an evolution of the spectrum 
calculated by the maximum entropy method, which may 
be applicable to nonstationary processes. Each 
spectrum represents an ensemble average of 8 runs. 


- in Figure 7. 


(b) 1.40 (c) 2.36 


(f) 4.20 


(e) 3.78 


Wavelets of a constant frequency of about 15 Hz in 
this case grow as shown in the figure with a smooth 
spectrum. The peak then moves to a lower frequency 
side showing the evolution to irregular wind waves 
having the usual spectral form. In the stage of 
constant frequency, Figure 4 shows the agreement of 
the observed frequency of the initial wavelets with 
the theoretical frequency for the kCj;-maximum, as 

a function of the friction velocity of the air, u,, 
but independent of the fetch. The frequency for the 
Cy-minimum is around 14 to 13 Hz, and does not 
coincide with the observed initial wavelets. Figure 
5 shows the agreement in the phase speed, and Figure 
6 the growth rate between the observed initial 
wavelets and the theoretical initial wavelets for 
kCj-maximum. 

Thus, Kawai's conclusion is that the generation 
of wind waves, whose initial stage is called initial 
wavelets, is caused by the selective amplification 
of small perturbations which inevitably occur in 
the flow by the instability of the two-layer viscous 
shear flow. 

However, the duration of the exponential growth 
of the initial wavelets was limited to from 1 to 8 
seconds in the experiments. The transition from 
the regular, long-crested initial wavelets to short- 
crested, irregular wind waves takes place in a very 
short time. The spectral peak, which has grown up 
with an approximately constant frequency, starts 
wandering at the transition, and then moves toward 
the lower frequency side with the energy increased 
in a general trend as seen in Figure 3, and also 
The transition coincides with the 
onset of turbulence at the water surface as revealed 
in the next section. 


3. INTERNAL FLOW PATTERN OF WIND WAVES — AN 
EXPERIMENTAL SUBSTANTIATION OF THE STRONGLY 
NONLINEAR PROCESSES 


Irregularity is a character inherent in the wind 


waves. This has been demonstrated by detailed 


531 


357 ay Lene UE 35 Ur Taal T ala =] 
(a) (c) 
% % wa 
E 
S 30+ 5 30h 4 
y | ‘ eA r 
r 
Q2- 30 25F 


kCi(8") 


kC) (8) 


3 4 
k (crm) 


studies of the internal flow pattern of wind waves 
by use of flow visualization techniques [Toba 

et al. (1975), Okuda et al. (1976, 1977) and 

Okuda (1977) ]. 

Along the surface of individual undulations, 
hereafter called individual waves, there is a strong 
variation of the tangential stress exerted by the 
wind. The stress value determined locally from the 
distortion of hydrogen bubble lines, is several 
times greater than the average wind stress value at 
the windward face of the crest, and it is negligible 
at the lee side of the crest as shown in Figure 8 
as an example. The concentration of the shearing 
stress results in the development of the local 
surface wind drift forming a special region under 
the crest where the strong vorticity is concentrated. 
The vorticity concentration causes the forced con- 
vection or turbulence, irrespective of whether or 
not the air entrainment, or the breaking in a usual 
sense occurs. As seen in Figure 9, small polystyrene 
particles of 0.99 specific gravity placed just 
beneath the water surface prior to the start of 
the wind, begin to disperse into the interior by 
the forced convection, coincidentally with the 


t(Hz) 


(tz) 


FIGURE 2. Theoretically obtained correlation 
of the amplification factor kC;, the phase 
speed, Cy, and the frequency, f, to the wave 
number, k, in the instability of coupled shear 
flow of the air and the water, for four values 
of the friction velocity of the air uy, of (a) 
13.6, (b) 17.0, (c) 21.4, (da) 24.8 cm/s. 
(Cited from Kawai (1977).] 


transition of the initial wavelets to the irregular 
wind waves. The main stage of the growth of wind 
waves thus seems to proceed as a strongly nonlinear 
processes. 


4. COMPONENT WAVES AND INDIVIDUAL WAVES AS PHYSICAL 
MODEL OF WIND WAVES 


Despite the fact that the wind waves are thus a 
strongly nonlinear phenomenon, they have been 
assumed as expansible to component waves, having 
phase speeds obeying the dispersion relation of 
free water waves, and weak wave-wave interactions 
have been considered. 

Recently there have been some articles reporting 
that the phase speeds of component waves do not 
necessarily satisfy the dispersion relation, notably 
by Ramamonjiarisoa (1974) for the one dimensional 
case and Rikiishi (1978) for two-dimensional com- 
ponent waves. Rikiishi developed an experimental 
technique for the determination of the directional 
structure of the phase speed of component waves 
without pre-assuming the dispersion relation, and 


= 
ft) 
N 
E 
2 
i¥9) 
10 F KK 
L \ 
[ \ 
=) 
rr 
ic t(s) t(s) at(s) 
intta) wavelets 
[ —— 8.00~ 864 .005 128 
Simei 0-64) -019:28) r= 
239) EF= OC oo 
r * 992-1056 - 
— = 1056-1120 - 
developing wind waves 
10° —--- 1024~ 1536 01 
= 1280-1792 - 
— — 15.36 ~ 2048 = 
— — 17.92 ~ 23.04 
2048 ~ 2.60 - 
statronary wand wares 
—— 40.96 - 74.24 MQ 1664 
=9 
10 nn fLennnll 
1 10 100 
f(Hz) 
FIGURE 3. A sequence of spectra for the initial stage 


of the generation of wind waves, showing the growth of 
initial wavelets at a constant frequency of about 15 
Hz, and the transition to irregular wind waves. The 


spectra were calculated by 


the Maximum Entropy Method, 


and each line represents an ensemble average of eight 
cases. The fetch was 8 m and the nominal wind speed 


was 5.1 m/s. 


[Cited from Kawai (1977).] 


Experiment 
e F=3m, NS=1 


1 
1 
8 


0 10 20 30 
ux (cm s") 


FIGURE 4. Observed theoretically predicted frequency 
of the initial wavelets, f., as a function of the 
friction velocity of air, u,- Theoretical values are 
for the condition of the maximum growth rate, where 
U)/u, represents the dimensionless thickness of the 
viscous boundary layer of the air. [Cited from Kawai 


(Le) 01] 
isa | a S| 
Experiment 
8- © F=3m, NS=1 
° 6 17 Ex.! | 
© 9 1 
x 8 8 Ex.0 1 
Tr x 
6 
O U,/ue = 5.0 
A 8.0 
E of i 
/ 
/ 
= BAAN S 
oo 4 noe 4 
cas fe 
Or 7 e 
L Sf | 
Oh 
ore Ky 
7 4 
2r Oo” 6 ef 5 
e Ad ye 
VX of 
hy ay 7 
oe 
ol sas | 
0 10 20 30 
u, (cm s*) 
FIGURE 5. The growth rate of the initial wavelets, 


8, as a function of u,- Theoretical values correspond 
to those for waves of the maximum growth rate. [Cited 
from Kawai (1977).] 


{e) 
{e) 
30,- 2 oe 4 
A S a A 
% 
£ 20+ | 
& 
o 
10/- O Experiment Ex. 1 =| 
O Theory Ui/ux =5.0 
A ” " 8.0 
O|L__ { silt 
0 10 20 30 
Ux (cm Ss") 
FIGURE 6. The phase speed of the initial wavelets C 


as a function of u,. Theoretical values correspond td 
those for waves of the maximum growth rate. The theo- 
retical values were calculated by use of the observed 
velocity profiles in the water at the critical time of 
the first appearance of the initial wavelets. The ob- 
served values for higher three wind speeds were deter- 
mined at the critical time, whereas that for the lowest 
wind speed was determined about 3.5 s after the critical 
time because of the experimental difficulty, and this 
delay may presumably explain the observed higher value 


533 


ea, 
WS ns 


Sp(cm? s ) 


Mi 
E 
' 

Q 
¢ 
1 
' 
8 
iH 


~ 


10 20 


f(Hz) 


FIGURE 7. An example of minute inspection of the ob- 
served time series of the spectral peak for the initial 
stage of the generation of wind waves. The lapse of 

time is indicated in alphabetical order, the interval 
between successive points being 0.32 s. After the growth 
of regular initial wavelets at a constant frequency of 
about 15 Hz, the spectral peak shows an irregular mo- 
tion corresponding to the transition to irregular wind 


than the theoretical ones. [Cited from Kawai (1977).] waves. [Cited from Kawai (1977) .] 
Seen a a a ae an ae ne ne 
zi 
lL 4 
C U= 62 m/sec J 
[ F2285m 
zt 
WIND —> | 
15 2 
Peat : : | : = 
q 9 -180 0 180 
$ Raton 
§ ° oo 
. | 5 aor “og 5 
Per maasp: coca o | FIGURE 8. Observed values of the 
Sa? of, io local shearing stress along the 
oes) VPale) ae ] surface of representative wind 
5 go 9 on 269 | waves. The abscissa is the phase 
P99 9 5] relative to the peak point of the 
cot a e cal crest and the ordinate is ex- 
pressed as the square of the fric- 
pti 88, tion velocity of the water. The 
Oro 5 
ops? aie re fe on'e ih wind speed was 6.2 m/s and the 


Phase 


so 


fetch was 2.85 m. [Cited from 
Okuda et al. (1977).] 


20 150 


FIGURE 9. Flow visualization of the 
initial stage of the generation of wind 
waves by use of polystyrene particles 
which had a 2-mm diameter and the specific 
gravity of about 0.99, and which were 
placed just beneath the water surface 
prior to the start of the wind. The wind 
blows from the left to the right. The 
wind speed in the tunnel section was 

8.6 m/s, and the fetch was 2.85 m. The 
time measured from the start of the wind 
is shown in seconds. In 2.58 s, initial 
wavelets may be recognized by streaks of 
light in the water, and some particles 
have already begun to disperse into the 
water. In 4.78 s, waves are already ir- 
regular wind waves and more particles are 
dispersed. In 13.6 s, particles are dis- 
persed down to more than 10 cm, corre- 
sponding approximately to a half of the 
representative wave length. [Cited from 
Toba et al. (1975) .] 


found that the phase speed was virtually independent 
of the frequency, and had the same value as that 

of the waves of the spectral maximum, at respective 
fetches. These experimental results are interpreted 
as indicating that the assumption of wind waves as 
expansible to component free waves with weak non- 
linearity is not necessarily appropriate for young 
growing wind waves. 

On the other hand, since individual waves as 
instantaneous surface undulations have a specific 
shearing stress distribution, and a specific interval 
flow pattern, they may carry some factors as a phys- 
ical element. We have examined, in a wind-wave 
tunnel of 15-cm width, energy density distributions 
for individual waves, as well as their phase speeds, 
and compared them with those obtained by usual com- 
ponent wave model for the same experimental data 
{Tokuda and Toba (1978)].* 

First, a normalized energy spectrum for individual 
waves has been newly defined and calculated from the 
statistical distribution of two kinds of the individ- 
ual waves: zero-crossing, trough-to-trough and all 
trough-to-trough on our wave records, as illustrated 
in Figure 10. The definition of the normalized 
individual-wave spectral density, O8y, is 


8y (fy) = 6yAE/(AE/E,)E (1) 
where vim 
HABE debh 9) 
are Seatee at ; 
6;Af = > imt G peak: alae, Soon pant ae, db 


and where m; is the number of individual waves of 
the period class, T;, (frequency from f to f + Af), 
Af = 1/(2nAt), where we used At = 0.02 Shen) 00), 
and Af - 0.25 Hz, and also 


*Tokuda, M., and Y. Toba (1978): Component waves 
and individual waves as physical model of wind waves. 
To be published. 


is the frequency of the energy maximum. 
The A-spectrum 


and where f 

Figure 10 shows the comparison. 
is the normalized spectra by the traditional com- 
ponent wave model in which the secondary peak is 


seen at the normalized frequency of 2. The B-spectrum 
is for individual waves of zero-crossing, trough-to- 
trough, and the C-spectrum for all trough-to-trough 
on our wave records. In the main frequency range 
from 0.7 to 1.5, which is the value normalized by 
the peak frequency, the spectra are virtually 
equivalent with one another. The second peak at 
frequency of 2 in the A-spectrum completely dis- 
appears in the individual-wave spectra. The slope 
of these straight lines is £-2 for the high frequency 
side, and £2 for the low frequency, sides ethexe> 
spectrum is considered to give a better represen- 
tation of the high frequency side which is exactly 
on the £79 line, and the B-spectrum represents the 
low frequency side better, which is more similar to 
the traditional A-spectrum. We may infer that much 
energy of the higher frequency part of traditional 
component waves, which is clearly shown as the 
energy at higher harmonics of the spectra, is a 
manifestation of the distorted shape of individual 
waves of the main frequency range, as was already 
suggested by Toba (1973). 

Figure 11 shows the normalized phase speed of 
individual waves determined by two adjacent wave 
gauges. It is inversely proportional to the square 
root of the frequency, in contrast to the phase 
speed of linear waves which is inversely proportional 
to the frequency. In addition, the phase speed of 
the individual waves is much larger than that of 
linear waves as shown later. In the case of the 
phase speed of component waves of one-dimensional 


10' 


535 


INDIVIDUAL WAVES 


ZERO-CROSSING TROUGH-TO-TROUGH 


ALL TROUGH-TO-TROUGH 


10° 
~~ 
z 
= 
~~ 
z 
oe 
{| 
10 
10° 
fr 
COMPONENT WAVES 
a a ee 
+ 
+ 
sls 
FIGURE 10. 


x 


Tj 


Comparison of three kinds of normalized energy spectra from the same wind-wave records in the wind 


wave tunnel. A: Traditional energy spectra by the component-wave model. B: Energy spectra for individual waves 
of zero-crossing trough-to-trough. C: Energy spectra for individual waves of all trough-to-trough. [Cited from 


Tokuda and Toba (1978).] 


spectra, which was obtained from the cross-spectra 
of the records of two adjacent wave gauges (Figure 
12), approximately the same phase speed is obtained 
in the before-mentioned main frequency range, where 
the coherence is close to unity. However, in the 
higher frequency range, it is virtually constant 

in agreement with Ramamonjiarisoa's 1974 measurement. 
The original values are shown in Figure 13, in 
which locations of the spectral peak are shown by 
arrows for the shortest and the longest fetches, 
respectively, and as the peak frequency moves to 
the left, the phase speed of the component waves 
becomes larger. In the figure, the full line shows 


the phase speed of linear waves. Figure 12 is the 
normalization of Figure 13, and Figure 14 shows an 
example of the comparison of phase speeds of com- 
ponent waves and individual waves. It should be 
noted that, as the distance of two wave gauges 
becomes wider, the range of high coherence becomes 
narrower, and the phase speed of component waves 
tends to be more uniform and obscure. However, it 
is at least evident that phase speeds for both com- 
ponent waves and individual waves have the same 
value near the peak frequency, and are inversely 
proportional to the square root of frequency, and 
much higher than the values of linear waves. It 


536 


06 0.8 1-0 1.2 1.4 1.6 
fn 


FIGURE 11. Phase speed distribution of individual 
waves (zero-crossing trough-to-trough), determined 
by a photographic method, and normalized by values 
for waves of maximum energy density. Dispersion re- 
lation for water waves are also entered by the dotted 
line. [Cited from Tokuda and Toba (1978).] 


is caused by the effect of the wind drift, which is 
concentrated near the crests. 

Thus, by using appropriate normalization, we 
may express the energy distribution of physically 
substantial waves by the energy spectra of individual 
waves for some local frequency ranges, excluding 
false energy density. The above mentioned B-spectrum 
and C-spectrum are two examples of these. Further, 
we may reinterprete- the traditional energy spectrum 
for the main frequency range as representing the 
energy distribution of individual waves, rather 
than the usual interpretation of a linear combination 
of small amplitudes of freely travelling component 
waves. In other words, the elementary physical 
substance of wind waves is rather in the individual 
waves, which have a specific distribution of local 
wind stress and flow pattern, and an apparent phase 
speed inversely proportional to the square root of 
the frequency. 

Further, Figure 15 shows that, for the individual 
waves in the main frequency range for each wind and 
fetch condition, there exists a conspicuous statis- 
tical relation between normalized wave height and 
period, for significant waves which Toba (1972) 
proposed as the 3/2 power law: 


H* = prx3/2 (2) 


where H* = gH/u 2 and T* = T/u, represents the 
g * g 


COHERENCE 


FIGURE 12. Phase speed distri- 
bution of one-dimensional compo- 
nent waves, obtained from the 
cross-spectra of records of adja- 
cent two wave gauges, and nor- 
malized by values for waves of 
maximum energy density. The 
coherence of the cross-spectra 
is shown in the upper part. 
[Cited from Tokuda and Toba 
(1978) .] 


Fetch 
m 


(cm s-!) 


Fo05,,9 5.87 
cot ke 
eee .e 


° © 
sob 8 88a g000, eco cog lg 


coe Senet scoce 
eo 
° 
Oper) 
0062250009 
T °ee 
00m 


°° 
wooo 2° 
° 0%, 


PHASE SPEED 


0 2 4 6 8 10 12 
f (Hz ) 


FIGURE 13. Original values of the phase speed distri- 
bution, for eight fetches, before the normalization 
shown in Figure 12. Peak frequencies for the shortest 
and the longest fetches are indicated by arrows, other 
cases being in between of these. The phase speed of 
linear water waves is indicated by the full line. 
[Cited from Tokuda and Toba (1978).] 


dimensionless height and period, respectively, 
normalized by use of the acceleration of gravity 
g and the friction velocity of the air u,. The 
figure shows the data for individual waves for 
various fetches. Except for very short fetches up 
to about 4m, the factor of proportionality B is 
constant of about 0.045. 

It should be noted that although the spectral 
form of wind waves in wind-wave tunnels is different 
from that in the sea as discussed, e.g., by Kawai 
et al. (1977), nevertheless the above power law 
holds for both cases, although the constant, B, is 
slightly different [cf, also Toba (1978a)]. Figure 
16 shows another representation of the same relation: 
between the wave height and the frequency, normalized 
for those waves of maximum energy. The slope of 
the line is -3/2. 

Consistently with this relation and the above- 
mentioned apparent phase speed, the steepness of the 
individual waves determined by a photographic method 
is approximately constant, statistically. It is 


FROM CROSS-SPECTRUM 


FOR INDIVIDUAL WAVES 
AND STANDARD DEVIATION 


c=(9/k)'”? 


(cm s-') 


PHASE SPEED 


1.0 p22? ocoo. 
: 


COHERENCE 


OO 0 2QID BSG YU Ww WA 
FREQUENCY (Hz) 


FIGURE 14. An example of the comparison of one- 
dimensional phase speeds of wind waves, determined 
from cross-spectra of records of two wave gauges, and 
determined for individual waves (zero-crossing trough- 
to-trough) together with the standard deviation, and 
the dispersion relation for water waves. At the bottom 
is shown the coherence of the cross-spectra. The fy, 
represents the frequency at the energy maximum. [Cited 
from Tokuda and Toba (1978) .] 


inferred that these facts strongly indicate the 
existence of similarity in the individual waves or 
in the field of wind waves, presumably as a result 
of the strong nonlinearity. 


Fetch 1-00m 1-70 


10° 10! 


537 


5. APPROACH BY SIMILARITY HYPOTHESIS AND 
DIMENSIONAL CONSIDERATION 


In cases of strongly nonlinear processes, such as 
turbulence, it is hard to approach problems from 
the rigorous way of solving a closed system of 
equations. In these cases, some assumptions based 
on physical considerations are sometimes introduced 
to supplement the system of equations, to arrive at 
useful results. In the case of wind waves, it 
seems that an approach by the traditional model of 
component irrotational free waves with their weak 
interactions is not necessarily realistic as has 
been shown. There is another approach, in which 

a kind of similarity structure in the field of wind 
waves is assumed, and a regularity in gross structure 
is sought by invoking dimensional considerations. 
An example of this line of approach has been 
attempted as partly described in a paper by me 
[Toba (1978a)]. 

Since the local wind stress distribution along 
the surface of individual waves is as shown in 
Figure 7, the local wind drift is forced to be 
stronger near the crest and weaker near the trough. 
Water particles near the surface travel a longer 
distance when they are near the crest than when 
near the trough. On the other hand, water waves of 
finite amplitude cause the wave current, resulting 
from the difference between the foreward and the 
backward movements of the water particles. Some 
self-adjustment should occur for individual waves 
in such a manner than the forward and the backward 
movements by the waves are coincident with the 
difference in the local wind drift as to the phase. 
The wave current ug of the individual waves of 
amplitude, a, and angular frequency, o, is now 
approximated by that of the second order Stokes 
wave: 


ug = a*a3/g 


Number Density 
@ 02 - 


e ONS = O12 

© 010 - O15 

* 005 - 0.10 
0.005 - 0.05 
Significant Wave 


Standard Deviation 


FIGURE 15. Examples showing that the main part of individual waves in the wind-wave tunnel (zero-crossing 
trough-to-trough) satisfies the 3/2 power law between the normalized wave height H* and the period T*. The 


u, was 68 cm/s. [Cited from Tokuda and Toba (1978) .] 


538 


1.5 
NUMBER DENSITY 
1.0 ji 0.005 ~ 0.05 
a 005 - 0.10 
08 0.10 015 
015 0.20 
0.20 
06 
z 
a 
04 
02 4 
05 1.0 2.0 3.0 4.0 
fy 
FIGURE 16. Another representation of the 3/2 power law 


for individual waves (all trough-to-trough). H. and f 
represent the wave height and the frequency, respectively, 
normalized by values for waves of the maximum energy 
density. [Cited from Tokuda and Toba (1978) .] 


Since the difference in the local wind drift is 
caused by the mean wind stress, the self-adjustment 
is expressed by the condition that the wave current 
is proportional to ux, namely, 


a*o3/gu, = constant (3) 
This is transformed immediately to 
H* = B'o*73/2 (4) 


which is equivalent to (2), where o* = u,d/g. 

The condition of constant steepness may arise 
from the similarity requirement. The combination 
of the 3/2-power law relationship and the constant 
steepness condition leads to the apparent phase 
speed proportional to the square root of the 
frequency. These three relationships, which have 
been shown by the experiments to be satisfied by 
the individual waves, are self-consistent with one 
another, and may thus result from the strongly 
nonlinear effects. 

The 3/2-power law makes it possible that the 
wind-wave field is represented by a single dimension- 
less parameter of the frequency at the energy 
maximum as discussed by Toba (1978a). One of the 
consequences of the above paper is that the growth 
of the wind wave field is expressed by the evolution 
of the dimensionless single parameter in a form of 
error function of the parameter itself, in which 
the value of the parameter approaches a final value 
as a simple stochastic process, irrespective of its 
initial conditons, through a rapid self-adjustment 
of the state. 


6. WIND STRESS OVER WIND WAVES 


The final topic of this paper concerns the expression 
of wind stress over wind waves. It has been pointed 
out on Many occasions that the roughness length, or 
equivalently the drag coefficient of the water sur- 


face, depends not only on the wind speed but also 
on the state of the water surface. Various attempts 
have been made to obtain a functional form of the 
roughness length incorporating the state of wind 
waves or the wave breaking. However, in view of 
the complexity of the expressions, together with 
the wide scattering of data points, a simple 
dimensional formula by Charnock (1955) has been 
cited most frequently, but with various values of 

a constant of proportionality, although the formula 
contains only a parameter representing the wind 
field. 


A dimensional consideration leads to an expression: 


Zo* = zo*(ux*, On*) (5) 


where z9* = z9/v is the dimensionless roughness 
parameter, ux* = u,3/gv the dimensionless friction 
velocity representing the overall wind effect, and 
Om* = ux0p/g the single parameter representing the 
wind-wave field as stated in the previous section, 
where Om is the frequency at the energy maximum. 
Charnock's formula 


Zo = Bu,°/g (6) 


zo* = Bu,* (7) 


which is a form of (5) in which o,* is disregarded. 
It is shown that another simple form for zo*, using 
symbols, o and o*, instead of om and o,* hereafter: 


FETCH=13.6m 
TOBA (1972) 


KAWAI et al. (1977) 

KUNISHI (1963) 

MITSUYASU et al. (1971) 
20 = 0.035 ug/g 


UpZo/V 


10! 10? es 108 
ue / gu 

FIGURE 17. Data plots for the relationship (6) in 

a dimensionless form. Data by Toba (1972) and 

Kunishi (1963) are from wind-wave tunnel experiments, 

and data by Kawai et al. (1977) and Mitsuyasu et al. 

(1971) are from tower-station observations. [Cited 

from Toba (1978b).] 


FETCH#13.6m 
TOBA (1972) 


KAWAI et al. (1977) 

KUNISHI (1963) 

MITSUYASU et al. (1971) 
20*0.025 un/o 


u 2 3 4 
10 10 Selo 10 
FIGURE 18. Data plots for the relationship (8) ina 
dimensionless form. The same data with Figure 17 is 
used. [Cited from Toba (1978b).] 


Zo* = au,*/o* = au,2/vo, a = 0.025 (8) 


is a better representation [Toba (1978b)]*. In 
Figures 17 and 18 are shown plots of some available 
data in the forms of (6) and (8) including wind-wave 
tunnel experiments and field observations. It 
should be said that the new formula is better at 
least. It is seen from Figure 19 that the breaking 
of wind waves is also expressed as a function of 
the parameter, u,2/vo, for data from the wind-wave 
tunnel and the sea. The ordinate is the percentage 
of the breaking crests among individual waves 
travelling through a fixed point, and it was deter- 
mined by the same procedure for both cases. The 
breaking of wind waves occurs for the condition 

ux? vo > 103. 

Equation (8) corresponds to an elimination of g 
from the form of (5). In view of the recent 
recognition since Munk (1955) that the waves of 
high frequency components play a major role in the 
transfer of momentum from the wind to the sea, it 
seems rather unreasonable that Eq. (7) contains 
information only of energy containing waves as o 
in the denominator. However, since ug2/V S A = 
du/un represents the magnitude of the average wind 
stress, and o-! « T is a measure of the integration 
time associated with individual waves, u,?/va is 
interpreted as a measure of the accumulation of 
the shearing stress or the concentration of the 
vorticity at each crest of the individual waves, 
conveying the horizontal momentum transferred from 
the air into the interior of the water through 
forced convection, whether or not the waves are 
breaking, as stated in Section 3. As this effect 


*Toba, Y. (1978b). A formula of wind stress over 
wind waves. To be published. 


539 


increases, the total momentum transfer, as well as 
the probability of the occurrence of the breaking 
increases. 

The form of (8) may be transformed to 


Zo = B'u,*/g, (SY = eh Aw, (9) 


which may be interpreted as an extension of Charnock's 
formula (6) to include information of wind waves 

in the form of the wave age, c/u,, where c is the 
phase speed of the dominant waves. Also, the drag 
coefficient, Cp, may be expressed from (8) as 


G. = k*/[ In (z190/au,)]~2 (10) 
where k is the von Karman constant and z;9 the 
reference height of 10 m. According to (10), Cy 
is more sensitive to the wind waves than to the 
wind speed. 


7. SHORT SUMMARY 


We may summarize the review paper as follows. First, 
the initial wavelets are generated by an instability 
of two-layer viscous shear flow of a type of insta- 
bility that immediately transfers to three 
dimensional turbulence. Second, the main phase of 
the growth of wind waves is regarded as the conse- 
quent, strongly nonlinear processes. Third, the 
traditional component wave model is not necessarily 
realistic, and the elementary physical substance 
might better be treated by individual waves, 
especially for younger stages as observed in wind- 
wave tunnels. Fourth, the individual waves 
represent a conspicuous and characteristic similarity 
of structure, presumably as a result of the strong 
nonlinearities, and this may be the basis for the 
pure wind-wave field being represented by a single 
dimensionless parameter. Finally, a new stress 
formula over the wind-wave field is presented. 


40 
Oo FETCH: 13.6m 
10.0 TOBA (1972) 
6.9 


TOBA et al. (1971) 


oa 
(eo) 


i) 
oO 


PERCENTAGE OF BREAKING CRESTS 
9° 


Got tite) 10% 


FIGURE 19. Percentage of breaking crests among indi- 
vidual waves traveling through a fixed point, may be 
expressed as a function of the same parameter with 
Figure 18. Toba et al. (1971) data are from tower sta- 
tion observations, which are common with data of Kawai 
et al. (1977) used in Figure 18. [Cited from Toba 
(1978b) .] 


540 
ACKNOWLEDGMENTS 


The author expresses many thanks to Messrs. M. 
Tokuda, K. Okuda, and Dr. S. Kawai of his laboratory 
for continuous cooperation and discussion, and to 
Professor H. Kunishi of Kyoto University, Professor 
K. Kajiura of University of Tokyo, Professor H. 
Mitsuyasu of Kyushu University, and Dr. N. Iwata of 
National Research Center for Disaster Prevention 

for valuable discussion and comments. He also 
thanks Mrs. F. Ishii for her continuous assistance. 


REFERENCES 


Charnock, H. (1955). Wind-stress on a water surface. 
Quart*. J. Roy). Met. Soc. 81, 639. 
Kawai, S. (1977). Study on the generation of wind 


waves. Ph. D. Dissertation at Tohoku University, 
100 pp. 
Kawai, S. (1978). Generation of initial wavelets 


by instability of a coupled shear flow and their 
evolution to wind waves. Submitted to J. Fluid 
Mech. 

Kawai, S., K. Okada, and Y. Toba (1977). Support 
of the three-halves power law and the gu,o74 
-spectral form for growing wind waves with field 
observational data. J. Oceanogr. Soc. Japan 33, 
WSi7/5 . 

Larson, T. R., and J. W. Wright (1975). Wind- 
generated gravity-capillary waves: Laboratory 
measurements of temporal growth rates using 
microwave backscatter. J. Fluid Mech. 70, 417. 

Miles, J. W. (1957). On the generation of surface 
waves by shear flows. J. Fluid Mech. 3, 185. 

Munk, W. H. (1955). Wind stress on water: an 
hypothesis. Quart. J. Roy. Met. Soc. 81, 320. 

Okuda, J. (1977). Internal flow pattern of wind 
waves, Proc. Nineth Symp. on Turbulence, Inst. 


Space and Aeronautical Sci. Univ. Tokyo, June 
1977, 54. 

Okuda, K., S. Kawai, M. Tokuda, and Y. Toba (1976). 
Detailed observation of the wind-exerted surface 
flow by use of flow visualization methods. J. 
Oceanogr. Soc. Japan 32, 51. 

Okuda, J., S. Kawai, and Y. Toba (1977). Measurement 
of skin friction distribution along the surface 
of wind waves. J. Oceanogr. Soc. Japan 33, 190. 

Phillips, O. M. (1957). On the generation of waves 
by turbulent wind. J. Fluid Mech. 2, 417. 

Ramamonjiarisoa, A. (1974). Contribution a4 1'étude 
de la structure statistique et des mécanismes 
de génération des vagues de vent, Thése a4 
L'Université de Provence Le Grade de Docteur és 
Sciences, 160 pp. 

Rikiishi, K. (1978). A new method for measuring 
the directional wave spectrum. Part II. Measure- 
ment of the directional spectrum and phase 
velocity of laboratory wind waves. J. Phys. 
Oceanogr. 8, 518. 

Toba, Y. (1972). Local balance in the air-sea 
boundary processes, I. On the growth process 
of wind waves. J. Oceanogr. Soc. Japan 28, 109. 

Toba, Y. (1973). Local balance in the air-sea 
boundary processes, III. On the spectrum of 
wind waves. J. Oceanogr. Soc. Japan 29, 209. 

Toba, Y. (1978a). Stochastic form of the growth 
of wind waves in a single-parameter representation 
with physical implications. J. Phys. Oceanogr. 
8, 494. 

Toba, Y., M. Tokuda, K. Okuda, and S. Kawai (1975). 
Forced convection accompanying wind waves. J. 
Oceanogr. Soc. Japan 31, 192. 

Valenzuela, G. R. (1976). The growth of gravity- 
capillary waves in a coupled shear flow. J. 
Fluid Mech. 76, 229. 

Wilson, B. W. (1965). Numerical prediction of 
ocean waves in the North Atlantic for December, 
1959. Deut. Hydrogr. Z. 18, 114. 


An Interaction Mechanism between 
Large and Small Scales for 
Wind-Generated Water Waves 


Marten Landahl, 

Sheila Widnall, 
and 

Lennart Hultgren 


Massachusetts Institute of Technology 
Cambridge, Massachusetts 


ABSTRACT 


By aid of a non-linear two-scale analysis it is 
shown that large-scale water waves can experience 
growth due to spatial non-uniformities in the 

growth rate of the small-scale waves in the non- 
uniform wind field associated with the large-scale 
waves. The growth rate is shown to be proportional 
to the mean-square slope of the small-scale waves 
and their growth rates, but inversely proportional 
to the difference between the phase velocity of the 
large-scale wave and the group velocity of the small- 
scale waves. It is suggested that this mechanism 
can transfer wind energy to short gravity waves at 

a higher rate than the direct linear transfer 
mechanism of Miles (1962). The analysis also 
predicts that a large-scale wave moving against the 
wind will be damped by the action of the small-scale 
waves. 


1. INTRODUCTION 


The mechanism whereby wind generates water waves 
has long proven a difficult and challenging problem 
in theoretical fluid mechanics which has not yet 
been satisfactorily resolved. The simple linear 
mechanism of forcing by pressure fluctuations 
[Phillips (1957)] and by instability induced by 

the mean wind field [Miles (1957), 1962)] have 

been found inadequate to account for the high values 
of energy transfer from wind to waves observed for 
longer waves, both in the laboratory and in the 
open sea. For short waves in the capillary regime, 
laboratory experiments [Larson and Wright (1975) ] 
have given good agreement between observed growth 
rates and Miles' instability theory, particularly 
when the surface drift velocity in the water is 
taken into account [Valenzuela (1976)]. For waves 
in the short gravity range, however, recent experi- 
ments by Plant and Wright (1977) give growth rates 
much in excess of that predicted by the instability 


541 


theory with the discrepancy beginning at a wave 
length of about 10 cm and increasing with wave 
length. Open-sea measurements have also produced 
energy transfer rates for gravity waves which are 
much in excess of the values according to Miles. 
[See, for example, the recent review of Barnett 
and Kenyon (1975)]. 

In view of the failure of linear theory one is 
forced to look for nonlinear mechanisms for energy 
transfer. Nonlinear interaction between waves in 
the gravity range [Phillips (1966)] is a compara-— 
tively weak process (of third order in amplitude) 
which causes redistribution of the energy from 
waves of intermediate wave numbers to waves of lower 
and higher wave numbers. This could be effective 
for the eventual saturation of the spectrum but is 
unlikely to be strong enough to make a large change 
in the initial growth. A more tenable proposition 
is that the modification of the turbulence in the 
air by the wave induced velocity field could change 
the phase shift between surface elevation and the 
pressure so as to alter the energy transfer rate. 
This effect has been investigated by many authors 
[Manton (1972), Davies (1972), and Townsend (1972), 
among others] employing different turbulence models. 
These investigations point to the possibility that 
the modulation of the turbulence by the wind could 
have an important effect, but it is difficult to 
assess the adequacy of the postulated turbulence 
models employed. 

An interesting possibility for transfer of energy 
to gravity waves is through nonlinear interaction 
with capillary waves which can draw energy from 
wind at a much higher rate than the longer waves. 
The interaction between short and long surface 
waves has been subject to a great deal of discussion 
in the literature. A train of short waves riding 
on a long wave becomes modulated by the orbital 
velocity field of the long wave so as to make their 
wave length smaller - and hence their amplitude 
greater - in the region near the crest of the long 
wave. Longuet-Higgins (1969) argued that the 


542 


radiation stress then set up by the short-wave train 
would act to transfer momentum to the long wave. 

In particular, if the short wave were to reach an 
amplitude at the crest of the long wave high enough 
for breaking, it would give up all its momentum to 
the long wave. This maser-like mechanism was 
examined critically by Hasselman (1971) who showed 
that the change in potential energy in the surface 
layer due to Stokes' transport by the short waves 
would give a contribution to the energy transfer 

to the large waves which would exactly cancel that 
arising from Longuet-Higgins' momentum transfer 
term. Hasselman's analysis did not take into 
account any transfer due to modulation of surface 
wind stress or short wave growth rate, however. 
[This effect has been analysed by Valenzuela and 
Wright (1976)]. Also, his analysis concerned 
primarily gravity waves, for which resonant inter- 
action between wave number triads only occurs to 
third order. For capillary-gravity waves, however, 
the dispersion relation allows resonant interaction 
at second order. Valenzuela and Laing (1972) have 
developed a theory for this, and Plant and Wright 
(1977) suggest that part of the measured excess 
growth rate in the low gravity wave range could be 


attributed to capillary-gravity resonant interaction. 


Benny (1976) has also shown that under certain 
conditions, a long gravity wave may grow in the 
presence of small scale capillary waves; the wind 
field was not included in his analysis. 

The present paper reveals yet another possible 
mechanism for the transfer of energy from capillary 
to short gravity waves. The theory presented takes 
into account the effect of shear flow modulation 
on the local growth rate of the capillaries. It 
is found that this variation gives rise to a modu- 
lation of the Stokes' drift which is in phase with 
the long-wave surface slope and therefore makes 
possible an energy interchange with the long wave. 
It is found that the energy transfer rate due to 
this mechanism is positive for capillaries with a 
group velocity higher than the phase velocity of 
the long wave so that it can provide an increase 
in the long-wave growth for waves in the short 
gravity wave regime. For waves running against the 
wind the transfer rate is found to ke negative, so 
that the presence: of the capillaries would always 
increase the decay rate of the long waves. 


2. INTERACTION BETWEEN LONG AND SHORT WAVES 


We shall consider the situation depicted in Figure 
1 with two-dimensional surface wave of small wave 
length, 4', riding on a large-scale wave of wave 
length, A. An asymptotic analysis will be carried 
out under the assumption that 


Ee = A'/d << 1 (1) 


(Prime refers to the short and tilde to the long 
waves). The waves are excited by a wind field 
blowing over the water surface. Only the normal 
stress induced by the wind on the wavy surface is 
considered in this process, the effect of shear 
stresses being neglected. Of particular interest 
is whether the presence of the small-scale waves 
could change the growth rate of small-amplitude 
long waves. 

To arrive at the simplest possible analysis, 
terms that are of higher order than linear in the 


long-wave slope are neglected. For the short waves, 
only quadratic and lower-order terms in the wave 
slope are retained. Further, it will be assumed 
that the flow in the water is irrotational, i.e., 
the effects of surface drift currents are neglected. 
This allows the use of potential-flow theory leading 
to the following boundary-value problem for the 
velocity potential 6 in deep water: 


V4e= ob +6 =0 (2) 
xx ZZ 


with boundary conditions 


Oo eae Oe. (3) 
at z=: 
PW 1 3/2 
Ty ee lee oN bag / EZ) 
(4) 
at z= -™: @=0 (5) 


Here, © = C(x,t) is the surface deflection, P. the 
surface pressure due to the wind, and T the surface 
tension. Since cubic terms are neglected through- 
out, the denominator in the last term of (4) will 
be set equal to unity henceforth. We now separate 
large and small scales by introducing into the 
equations of motion 


ES ie oie (6) 
6=6 + 6! (7) 
Be = Py + Pe (8) 


For the boundary conditions it is useful first to 
transfer them to the surface of the large-scale 
motion, z = t, by a Taylor series expansion. Thus, 


O(x,5) = 0 (x,0) + 5'0,,(x,5) + 
= O,(x,o) + O' (x,o) ae 1G Y (a sn) 
; 3 
+ Or (er) stvemete (9) 
etc. By neglecting terms involving triple and 
higher products one finds from (4) and (5) the 


following boundary conditions to be applied at 
Zi iGes : 


water 


FIGURE 1. Long-wave short-wave interactions in a 
shear flow. 


' ' ' t) ' ' 
Oy OE S ty PER sb (ee ODS, Ge [ise & EA) 
a (10) 
ie ' = = ' & ' "(6 oO! 
lly Pte) = ils 8 ee) = Ws ee Oey Oe he an ee) 


1 b ry 2 6 ry 2 
acy (ee NS se (Oe GANS 


+ To + Oe) EO OD ((aLgb)) 


ibs) (Cabal) P, is the surface pressure in the absence of 
the short waves and Pw the additional surface pres- 
sure added due to the presence of the short waves. 
In deriving (10), partial use has been made of (2), 
which holds for ® and 6! separately. To arrive 

at equations for the long wave, (10) and (11) are 
averaged over the large scales. This is most 
conveniently done by taking the ensemble average 

of a large number of realizations differing only 

as to the phase of the short waves, which is assumed 
to be randomly distributed among the members of 

the ensemble. This procedure yields 


Do & Be tb Ole, ER te oc (12) 
2 a 
We = 2? ss 2 2 
= GS > Os Sale = Oe) eae 
= a= (E94) = AO = OOF) os GS) 
ate 


at Z= co where the tilde denotes the average over 
the large scales and 


aN 


Siw a (14) 


is the Stokes' drift due to the small-scale motion. 
In deriving (13), use has been made of the linearized 
boundary conditions for the small scales, for 
example, 


“—~ SN 
O~j0 = Feuyey 
5 oe Se tt ee 3 
DR ee Pa ao Ce 
ae (BIER) = Ber 063 Se y= oa 
£ 
+ (15) 


The long waves are to be determined as a solution 
of Laplace's equation 


v26 = 0 (16) 


subject to the boundary conditions (12) and (13) 
and the condition that disturbances vanish at large 
depths, i.e., 


iBere 14 = = © (17) 
The corresponding boundary conditions for the short 


waves are obtained by subtracting those for the 
long waves from the full equations. 


543 


'- 6! +77! - 6 6! - 6 oO! 
tt) > ie x6 are a Zz 


F pope (19) 


both to be applied at z = t. The last bracketed 
term of (18) and the last two of (19) will give 
rise to higher harmonics. Their contribution to 
the large-scale motion will be of higher order, 

and they can hence be neglected. In deriving (19), 
use was made of the linearized boundary conditions 
for the long waves. Thus, for example, the term 
eel in (19) arises from replacing Ore, in (11) by 
Tet, which will give a negligible error to within 
the approximation employed. 

Since the major aim of the analysis is to deter- 
mine the lowest-order effect of the short waves on 
the growth rate of the long waves, it is sufficient 
to retain only linear terms. However, all terms 
linear in the large-scale motion which modulate 
the small-scale wave train must be retained. The 
long wave will be taken as a uniform, infinite 
wave train of wave number k = 21/A. Its phase 
velocity differs from the linearized value, 


c= vg/k + kT (20) 


by terms proportional to the square of the small- 
scale wave slope, and by terms due to the wind, 
which are proportional to the density ratio between 
air and water both of which may be expected to be 
small. The short waves driven by the wind may also 
give rise to slow growth, or decay, of the large- 
scale waves. For the subsequent analysis, it is 
convenient to introduce the following nondimensional 
"slow" variables: 


k(x - ct) (21) 


aa! 
i 


t=ket (22) 


The solution for the long wave is sought in the 
form (real part always implied) 


Ze Bi@e” (23) 


ete tkz (24) 


The variation of the surface deflection and potential 
with the "slow" time, t, allows for the effects of 
wind, and the presence of the short waves, to have 
a weak influence on the growth rate, and the phase 
(and consequently also the phase velocity) of the 
long waves. . Without the wind and the short waves 
both f and ® would be constants. 

For the short-waves, on the other hand, both 
the phase velocity and wave number will vary slowly 
along the long wave because of the modulation by 
the latter. We therefore set 


Bo = Lae aye Gr) (25) 


et (x,t) 


Oo = ou(&,zZ" 70) (26) 


544 


where 8 is the phase, 


ie = Oe (27) 


is the wave number, 


TS Oe (28) 
is the frequency (measured in a fixed coordinate 
system), and 

z= k"(z = C). (29) 
The assumption of a slowly varying wave train allows 
one to regard k' and w' as functions of the "slow" 
variables, t and &. 

An approximate asymptotic solution for the short 

waves if sound by expansion in the small quantity 

€ = k/k' (30) 
(That € thus defined is a slowly varying quantity 
causes no special difficulty). Substitution of 
(25) - (29) into (18) and (19) and omission of all 
terms of order lerle, Jerct |, and higher, as well 
as of terms of order e“ and higher, gives the 
following boundary conditions for the small-scale 


motion to be satisfied at z' = 0: 
ef = - itt (c'-a) + cust + ec(Zt - of) + iekt,s 
Saricrs (31) 
Pwo = - (g + k!2n)c' + ik'(c'-u)é" 
p ~ a en a “ n 
+ are ee: 5 g) + 2kuTey + kp Te") 
Se KO 70 Wa (Gra oe) ] ‘Foo ¢ (32) 
where 
ce! = w/k' (33) 
u and w are the perturbation velocities, 6. and 


x 
@, respectively, of the large-scale-flow evaluated 
at z= C and ist is defined by 


So 2 a 80 34) 
a S pla 

0 0 Py ( 
The terms neglected as being of higher order in ¢€ 
include the term Geeon in (19), which expressed in 
the slow coordinates becomes 


Paiva 
ea ra 


and is hence negligible compared to the term k'2Tc'. 
For the long waves one finds similarly 


@, = w= k[e(ct, - it) + 81] +... (35) 
Sn SARA yee Rata a 
0 15 

Sere aa ene 

= 2x1 [id + 6} |2 - 05/7] +... (36) 


where 


ete es 
SS emai Gol a CLE) (37) 


and the star denotes complex conjugate. The velocity 


potential must satisfy Laplace's equation. Substi- 
tution of (26)-(29) into (2) gives 
se (@U ee al) ct iek'[ ko" + 2k'o} 
0m Ss eee 
ae ZA" ke Sek Yee oe 
+ 0(e26') = 0 (38) 


This equation may be solved approximately by series 
expansion in €. One finds in a straight-forward 
manner that 

' 12k; 
as ee at See elt ean ae EA 2 
0) Go Atl = aie e et Itz, Cp) 1A iezA¢ + 0(e*A) } 

(39) 

where A = A(E,T) is to be determined by aid of the 
kinematic boundary condition (31). By substitution 
of (39) into (31) and expanding in powers of € one 
finds 


S(T iaty Se e(Shet)) = => let = 158) } 


+ 0(e20") (40) 


Combination of (40) and (32) yields 


= =(g + kt? = k'(c'=u) 212" + ick" {fe(ey > ae 
+ (et—u)u (ete) (c's 1a) eae 
+ 2c(c'-u)o! 


+ [(e'-a) (c'=2e + u) + 2k"T)eE } (41) 
The induced surface pressure due to the wind may be 
assumed to be related to the small-scale surface 
deflection in a quasilinear manner that takes into 
account the modulation of the wind field by the 
long waves. The following expression is chosen: 


Pe z Shes 
— = k' (eu) (a 2B") (2 = ake)ice 


(42) 
p 


where a' and §' are aerodynamic coefficients (having 
the dimension of velocity) giving the in-phase and 
out-of-phase components, respectively, of the induced 
pressure. The modulation of the wind field due to 
the presence of the long waves is accounted for by 
the factor (1 - akt). For long waves running with 
the wind and having a phase velocity less than the 
wind-speed, the air flow at the crest will slow 

down in the region below the matched layer where 

U= c, and the small-scale growth rate will thus 

be reduced in this region. Conversely, the air 
speed will increase over the troughs leading to an 
increased growth rate there. Hence, the coefficient,- 
a, will be positive for such waves. For waves run- 
ning against the wind, however, or for waves with 

c greater than the wind speed, a will be negative. 
To determine the numerical value of a, one must 
carry out calculations based on the Orr-Sommerfeld 
equation. First the wind field modulation due 

to a long wave of small amplitude is calculated. 
Then, the pressure on the short waves is computed 


on the basis of quasilinear theory, whereafter the 
effect of wind field modulations may be extracted 
from the results. In Section 3, we derive the 
governing equations for the local growth rate for 
short waves in the modulated flow of the long waves. 
Numerical results for a are presented in Section 4. 

Consistency of the two-scale expansion requires 
that the wind-induced growth rate is small, which 
is indeed the case, since it is proportional to 
the air-to-water density ratio. Accordingly, we 
shall set, formally, 


a’ + i8' = e(a+ iB) (43) 


Substituting (42) and (43) into (41) and remembering 
that all the quantities involved are real, we find 
the following pair of relations: 


g + k'?n - k'(c'-u)* + ek'G(c'-u) (1 -akZ) = 0 (44) 


[e(c'-u), + (c'-c) (c'-u) + (c'-a) tg 


o ial = B(c'-u) (1 - akt)]c 


+ Zel(cueu)oe + [(e'=u) (c'=2e + u) 
i 2k'T] oe =0 (45) 


From (44) it thus follows that 


c' =u +t vg/k' + k'T + O(c) = u + v(kK') - O(c) 


(46) 
Inspection of (35) and (36) reveals that the long 
waves receive their growth both directly from wind 
pressure and indirectly from interaction with the 
short waves, the latter effect being proportional 
to the mean-square slope of the short waves. Thus, 
since the variation of long-wave parameters with 
time is small, little error is incurred by taking 
u in (46) to be a function of — alone. Furthermore, 
the frequency of the short waves must then be 
constant in a coordinate system travelling with the 
long waves so that 


ie (P= eS) Sw (47) 


which, together with (46) determines how the wave 
number for the short waves varies along the wave 


train. Differentiation of (47) gives 
k'u 
Ceo 
a ma c!-¢c 3) 
g 
where oe is the group velocity, 
eg =k'y. tv tu (49) 


and where v(k') is defined by (46). With the aid 
of (46)-(49), (45) may thus be written 


2(e'-a) [6b + (cg - E)EL] = {(c'-a)B(1 - ake) 


- (c'-0) Ge + [(e"=e) vy" + T] 
ae /A(Cee) (50) 


This equation may be readily solved by integration 
along the characteristic line 


eS. 8 


7 dé/dt, = (cg - c)/e (51) 


545 


Since only the terms which are linear in the large- 
scale perturbation are to be retained, one may 
ignore the variation of k' with & when carrying 
out this integration. 


- =i : E BCE 2 
oo a @ exp {5 eo a -~— 
c 2 (cg-c) 2(cg-C) 
kha hes 
+ Bw (o,=8)2 [(c “c)v,, ap abl Jy (52) 


where C is a constant to be determined from the 
initial value of ¢. By inserting this expression 
into (37) one finds 


u 


S; = - gi2 any == tee ap GU Pit 
g g 
= = fie? Oey 2 aT 53 
(es) k! ee) 
g 
where 
= iL A 
gi2 = lee |= (54) 
is the mean-square slope of the short waves. (In 


Appendix A an alternative derivation, based on 
kinematic wave theory, is given.) In the second 
bracketed term U may be expressed in terms of f 
by the use of the linearized expression 

W = kez (55) 
Thus, Pe may be written 


st= A'T ae BIZ 


E (56) 


where (ignoring terms which are nonlinear in C) 


At = -s'2 as (57) 
-12~ 
oS = {=c" = ec" + 20 
Sy 
k! ' 4 ' 
dP (ele) [(c -e) vy + T]} (58) 
Gf 


The boundary conditions for the long waves may now 
be written. Substitution of the solution for the 
short waves, and (24), into (35) and (36), gives 


6 = eon = Ae) 4b Ne ee inde (59) 
=- (g + k2r)z - KE (o_ - id) + O(e2) (60) 


For the wave-induced pressure an expression similar 
to (42) is used, namely 
= ke(& + if)c (61) 


Substitution of this and (59) into (60) and separa- 
tion into real and imaginary parts yields 


0 = [g + k2T - ke? - KE(B + G)IZ - kc aoa + REC 


546 
(@ - A')t = (26 - B')Z (63) 


From (63) we find 


18 ns 
Spee B-a' 
GC = S exp ii 5G=nUn at, (64) 


By use of this, ee and Gone may be expressed in 
terms of & and an eigenvalue relation obtained by 
substitution into (62). This then gives 


G = G/RIEEERITD be. (65) 


with correction terms proportional to the mean 
square slope of the short waves and to the air-to- 
water density ratio, both of which are likely to 
be small corrections of little importance. The 
major result of the analysis is that given by (63), 
(64) namely that 

aB'c!' 


= os LBB <1 2 
(2nt) = =e 3 = oP 26 (e!-8) Ss (66) 


ae 
dt 
i.e., the growth of the long-wave amplitude is 
given by the sum of the growth due to direct action 
of pressures in the manner of Miles (1957, 1962) 
and the indirect growth due to the Stokes' transport 
by the growing short waves. The second term may be 
large compared to the first term, if ci is close 
to ¢. However, the analysis presented does not 
hold in the immediate neighborhood of cg = ¢ but a 
separate (and nonlinear) analysis is then required. 
For waves running against the wind, c', cl, anda 
will be negative, so that the presence pf the short 
waves will always increase the decay rate of the 
long wave. 


3. THE WIND-INDUCED GROWTH OF SHORT WAVES IN THE 
PRESENCE OF LONG WAVES 


The perturbation equation governing the modification 
of short waves on the wind-water interface by the 


long-wave field is derived from the momentum equation 


by the procedure used to derive the Orr-Sommerfeld 
equation. Additional effects arise because the 
short waves see not only the mean wind field, U(z), 
but long-wave fluctuations, ti and W. The large- 
scale field is governed by a linear equation, 

the small-scale field by an equation linear in u', 
w' which also contains terms linear in U and w. 

As in 2, we take the water to be inviscid and 
the flow potential but we consider the air to be 
viscous: with no surface current, and continuous 
tangential velocity between air and water, this 
corresponds to the limit un, > ©, vy > 0 with Us 
and v, finite, justified by the large density ratio 
between water and air. Both fluids are taken to 
be incompressible. : 

We begin with the Navier-Stokes equations for 
two-dimensional flow in the air 


du du du opie ily 

Chee ge = ee 

rye us + Woe a 6 + = V“u (67) 
ow dw dw op 1 ik 9) 

DEVEOR Moz = 5 os Ge REG (ee) 


where velocities are scaled to free stream velocity 
outside the boundary layer over the water, and 
lengths scaled to boundary-layer thickness 6 ; R 

is the Reynolds number based on 6. (In Section 2, 
lengths were scaled to k', the short-scale wave 
number) . 

To derive the Orr-Sommerfeld equation, these 
equations are cross-differentiated and subtracted 
to eliminate the pressure. Some use is made of 
the continuity equation and the result is 


32u o2w 


fu 1 y29u 
dzot dxdt 


- uV2w + wV2u - = = (en = 0 


az R 3x 


(69) 


The flow in the air is taken to be a horizontal 
shear flow plus two wave perturbations of disparate 
scales: the fast scale, x and t; and the slow 
scale, % = ex, and t = et where ec = k/k'. The 
variation with z is set by the shear profile and 
viscous effects and will be taken to be the same 
order for both wave fields. The long wave field 
is a function of X,z, and £ only; the short wave 
field is a function of x,z, and t and in addition 
will be influenced by the long scale waves so that 


U(z) + a(X,z,t) + ul (x,z,t;%,t) 


ll 


u 

Wie w(%,z,t) + w! (x,z,t;%,t) (70) 
The surface deflection is taken as 

BEG Cpe) > GY Cesare) 


The major effect of long waves in a parallel 
shear flow on the behavior of the short waves will 
come from changes in the local growth rate and 
convection velocities as well as an unsteady lifting 
of the small scale as the large waves pass. There- 
fore the small scales will be assumed to be of the 
form 


w' (x,Z,t:#,t) = wiz-c(z,t)ler °™ = wiz en” 
Pb) | Meee ee oO = Gene aay 
where z' = z-f and c = c(X,t). Changes in the wave 


number, k, are O(cedU/3X); such terms will be ignored 
in this local analysis. For this assumed form of 
w' and u', the continuity equation becomes 


ee + iku' - ¢€ — =) 10) (72) 


a ae Ds Bo eee Re 

a 1 dw ere Ww i dw wl dw 
=-——+ — C~ —G =F = - ST 73 
O5 cs 8 apt cs Se ae 3) 
since el~ = -w/c. The presence of ia in the assumed 


form for w' and u' introduces several terms into 

the equation for the small scale. In addition, 

the velocity perturbations, U and w, also appear. 
The equation for the large scale is obtained by 

a phase average of (69) written with the assumed 

form (70) and (71). The non-linear coupling of the 

small-scale motions will not be included although 

the corresponding effects in the water are the 

main subject of this paper and are worked out in 

Section 2. We anticipate further work to complete 


the study of non-linear coupling in both the air 
and the water. 
The large-scale motions are taken as 
> 8 aed) 2 8 Ghee 
SPs elk (x ct) em ae) (x-ct) (74) 
soo, & A IRGHSE 
min Bae ests) 
To ease the process of working with products of 
wave perturbations, two distinct complex variables 
i and j are introduced. 
Under these assumptions, the large-scale mo- 
tions are governed by the linear homogeneous Orr- 
Sommerfeld equation 


w'" — 229" + kw - 3kR((U-G) (w" - k2w) - wu") = 0 
(75a) 
and a is related to a through the continuity 
equation 
a = j0'/k (75b) 


where from now on primes will denote derivatives 
with respect to z. 

In the equation for the small scale, we will 
keep all terms linear in the small-scale perturba- 
tions including products of the small-scale and 
large-scale perturbations. 

When the assumed form for the perturbations (71) 
is used in (69) together with the continuity equa- 
tion (73) we obtain the following equation for the 
small scale 


w'" — 2k2W" + k4w - ikR[ (U-c) (w"-k2w)- wU"] 
= Riwa-eSe] (w" - k*w') + ikR{(U'Z + di) (w"-k7w) 
= Spey Oes GEA ew ORS "a! 
ikR (= + U"'S)wt ERs I (2k (U-c) + U"]w 
= U(w"! = k2w') fy U"'w ra U' (w" ne k2w')} 
= ro (wW,w,z) (76) 


where we have introduced the symbol xo (w,Ww,Z) for 
the right-hand side of (76). The various terms in 
(69) are worked out in Appendix C. 

In deriving (76), terms of 0(k2t) have been ig- 
nored, however terms such as 32a/a22 in air have been 
kept since these can be large in a viscous flow. In 
terms of 0(€00/dx) a viscous correction has also been 
neglected since all other terms are proportional to R. 

We are interested in the local equilibrium and 
more specifically the local growth rate of short 
waves in the modified wind-water field. Thus in 
the assumed form of solution for the short waves, 
for a given k the eigenvalue, c, will be a slowly 
varying function of space 

@ = Gy v Gp (eS) (77) 
where Co is the eigenvalue of the short wave field 
in the presence of the wind shear field only; c, (x) 
will be at most O(z), the amplitude of the long 
wave. 

Thus the governing equation for w is the Orr- 
Sommerfeld equation with additional terms arising 
from the long-wave perturbations. Some of these 
terms could be obtained directly by replacing U by 
U + u in the Orr-Sommerfeld equation; additional 


547 


terms come from the unsteady lifting and distortion 
of the small scale flow by the long waves. 

The boundary conditions that are satisfied at 
the free water surface, z =~ + c', will now be 
derived for both the large and small scale motions. 
The first boundary condition is that the tangential 
velocity is continuous at the interface, z' = Z', 


ac C14 
Win Un SM mee > Ww O50 


Expanding the velocities from (70 and 71) ina 
Taylor series about z' = 0, and keeping terms linear 
in the large scale and small scales we obtain for 
the large scale 


Wie Sh, = wh, at z = 0 (78) 


for the small scale [to 0 (kt) ] 


Ui at ea ul =a ae oO at z=0 (79) 
The term 0U,,/dz = Uy has been ignored in deriving 
(79) since it is 0(k*Z) and a*u/dz2 (0) has been 
taken to be zero. 

Conditions (78 and 79) can be expressed in the 
vertical velocity, w' (and w), through use of the 
kinematic boundary condition; that the substantial 
derivative of the surface displacement function, 
S(x,z,t), is zero for both the air and water flow 
at the interface, S = 0. That is, if S(x,z,t) = 
Zou sthenwDS/DtE = Omatnou——OM(zi rat) 

D C) a a 
where a we U(Z) ae + w(T) Bp 

Expanding the velocity field for both air and 
water about z' = 0, and again keeping terms linear 
in the large and small scales, we obtain in the 
long-wave limit for the large scale 


a, WE =O 
for the air aE 5 w= 0 
for the water O8 L. w=0 
at Fi 


and for the small scale 


for the air 


iket - [U't + UjJikz - efu'c' + u'] ae w= 0 
at z' =0 (80) 
and for the water 
ret CU ae oars A Og, 6 : 
OS, > Wheall<ie = eu, re Fe = Oat zi a=a0) 
From (78) to (80) we see that 
ae - 
i at z' =0 
and W = Wy 


From (39) and (40), the velocities in the water at 


z' = 0 are related to the displacement & by the 
expressions 
z : > 1 DE 
Wy = ~ik(e -— u) (1 + tesg)o" 
Ae ee Ba) Ge aoe Set 81 
uy = (c — U,) ( tens (81) 


548 


“ w u .W 

-& ee es 

and t ke (@l sp = iz) 
where € oe i Z 
ox c 


thus Uy = iw. as in fixed coordinates. 
For the large scale motions, 


and 5 (82) 


Ae 


With (75b) and (82), condition (78) for the large 
scale motion becomes 


Uwteéw' -wké=0 at z=0 (83) 


and with (73) and (81) condition (79) for the small 
scale motion [to 0(kt)] becomes 


a R = » Bly thy ~ 
U'w + cw' - wke = - ou -w aul [— + ity 
z Zi é 
Cn WX ES (84) 
= a = 0 
iw a The (w) at z 


where we have introduced the symbol 19 (w) for the 
right hand side of (84). 

The remaining boundary condition to be satisfied 
at z = t and then transferred to z' = 0, is the 
balance of pressure or more precisely of normal 
stress with surface tension 


as (85) 
Ps ~ [nn = Pw Sebel: 9x2 
The viscous normal stress at z = t is given by 
3 
2 Zn 


) 
where hn = > Wee Gls eS ae aes Spe 


Because uwU' (0) HaUd (0) inl the) Damitt uy 2 
there is some cancellation in the stress condition 
and the final result for the large scale is Spy = 
2 w'/R. 

For the small scale, all terms involving the 
large scale perturbations are negligible for k<<k' 
so that Caan = 2 w'/R. 

The pressure in the air at the surface, z=, 


is obtained by expanding the pressure about z' O. 
p(t) = plo) +¢ & (0) (86) 
az 
where p(o) and dp/dz(o) are available from the 
momentum equation (67). 
After considerable manipulation we obtain the 
following formula for pg - Onn 
ikIB, - Opp] = p,[- ¢ G' - wu" + — (w"' - 3k2w') 
SSIS nn’ “a Rk 
du = 0 ~ du ww! Ow 
Cae a ea uae t ie See MaKe OZ 
w a, GE k2w w" 
= (23 — - 22 - ice 87 
z (2ikcw + a 2 = icy )] (87) 
We obtain p_ - k2T~ at z' = ¢' [to 0(kZt)] directly 


from (41) 


By - k2Te = pylk(e - &)? - (g + k*n)]E" (88) 


Using (81) to rewrite (88) in terms of w, introduc- 
ing s = Pa/p,, and using (84) to rewrite the group, 


cw' - U'w, we obtain the final form of the pressure 
boundary condition (85) to be satisfied at z' = 0. 
SiGe & (Gam = = gx - a wl! 
u 
= wlke? - (g+k*T)] ( = + iw/é) - w 2ci, 
ow uy ay 
a 2K D5 piel = 0 (ae iw 
+ s(Gw'e + iw'c aa ke COU ian ( z + 5 
ae: = = Aa Ze 
we Sr Wy a ML “eats Aw k*w) — ~ 
+ awe =) s z (2ikew + ae 2 A ) = Yoo (w) 


We will introduce the variables p,Q,b, and y and 
rewrite (89) as Pw + Qw + bw"' = 


Yoo (w) where 
P = [-k(1-s)c? + (gtk2T)] 
Q= ASK b = -isc/Rk 


and Y59 (w) is the right hand side of (89). 
The corresponding boundary condition for the 
large scale is homogeneous and of the form 


Bw + Qw' + bw"' = 0 (90) 


where 


P = [-k(1-s)¢* + (g+k2T)] 


Q 


jc3sk/R; b= —js/Rk 


To summarize, the long waves satisfy the ordinary 
Orr-Sommerfeld equation (75) with the appropriate 
linear boundary conditions for a free water (83) 
and (90) plus w (©) = w' (2) =o. The resulting 
homogeneous eigenvalue problem is solved numerically 
to determine 4, w, and ¢ for a given long wave 
amplitude, a wave number, k, and R. 

The short waves satisfy a modified Orr-Sommerfeld 
equation, (75) with the effects of the long-wave 
perturbations appearing also in the boundary con- 
ditions (84) and (89). 

To solve this short-wave local-equilibrium 
problem we resort to techniques that by now have 
become standard in stability theory for perturbed 
eigenvalue problems. 

We assume that the short-wave solutions can be 
expanded about the perturbed solution (no long waves 
present) in the form 


aw=uU, +24, 
w=w, +o w, 
BIS ey ee Sp (91) 


“where z is the large wave amplitude. The eigenvalue, 


c, is also expanded 


c=c) +c, (91b) 


For & = 0 the problem of short-wave dynamics 
reduces to the ordinary Orr-Sommerfeld equation 
with free-surface boundary conditions. 


Wo" - 2k2we + kw - GkRL (U-c) (w" —k?w9)-woU"] = 0 
cows + (= - key) Wo = 0 
at 2)" = 0 
Pw) + Q wi +b wi! =) (0) 
and Wo (*) = wi (~) = 0 (92) 


The eigenvalue, cj, which determines the growth 
rate of the short waves in the parallel shear flow, 
is determined by a numerical solution of these 
equations for a given k and R. 

The equations governing the modification to the 
flow due to the long-wave perturbations are derived 
by using (91) in (76) and equating terms of 0(¢). 

In this operation the as-yet-unknown correction 
to the eigenvalue, c cy, will appear multiplying the 
lowest order solution, Wo- The resulting problem 
for Wy is written 


wy -.2k2wt + kYw, - GkR[(U-cg) (WY -k2W,) —w 0") 
= c,r,(w,z) + x, (W,w,z) (93) 


where r, (w,z) and rp (w,W,Z) are known functions of 
the long-wave perturbation, w(z), and the lowest- 


order short-wave perturbation, wy (2). From (84) 
and (91) 
x, (W,Z) = - ikR(w" - k?wy) (94) 


and r»(w,w,z) is defined by (76) with the long 
wave perturbations normalized by ¢. 

The boundary conditions for W, have homogeneous 
operations that are identical to those for w, but 
the equations are non-homogeneous with terms that 
depend on w_, the long-wave perturbations, w, and 
the unknown correction to eigenvalue, c)- 


i du A ate 
' — = = 
cow, + (a kc) Wy Yq c, + Thi (w,w ) 
at z=0 
Bw) iO) wi oe by wh = Ven Ci Va (W,Wo) (95) 


and wy (~) = wy (7) = 0 
where y (w) and Yoo (w) are defined in equations 
(86) and (90) with the long wave perturbations 


normalized by f 


phe eee 
and Yay Wo kw) 


= 2c )k(1-s)wy = ik3s/Rw) + is/kRw) 


Very, 
This problem is similar to that considered by 
Stuart (1960) and many others in later studies. 
In Stuart (1960) we have the problem 


L(w,) = r(z) 


with w, (0) = Ww! (0) =w,(@) =w! (=) =0 (96) 
where L is the Orr-Sommerfeld operator. 
The solvability condition [Ince (1926) pg. 214] 


for this problem is 


549 


where v is the solution to the adjoint problem 
L(v) = 0 


with v (0) = y"'(~) =0 (98) 


i 
< 
S 
Mn 
< 
ay 


Condition (97) is then used to determine the modifi- 
cations of the flow due to non-linearities. 

The present problem differs from that in (96) - 
(98) in that the boundary conditions of (95) involve 
linear combinations of the derivatives of Wy at z 
= O and are non-homogeneous. 

In Appendix B, we show that the adjoint boundary 
conditions that replace (98) in the determination 
of v are 


v" (0) =u" (@) =7 (e) =(0 (99) 
and [B-ikRcU' + o(U'-ke)]v + [U'-kc]v" + cv"' = 09 
where ey ae 
G = - (2k? + ikR(u-e)]ip Bi = =e - 2(u"-ke) 


and the extended solvability condition for non- 
homogeneous boundary conditions is 


co 


J xvae = { [c) ry (z)) + Ly (z) ]vaz 
0 0 


ll 
z 
a 
+ 
\ 
dq 
S 


+ (ye + His On (100) 


The solvability condition (100) is then used to 
determine ¢, and thus the correction to the local 
growth rate due to the presence of the long wave 
perturbations. 


vi {ood Fa NOEs Grae zs 
i eee OE eg EN, ca Js (z) vdz 
Ca 2 
Vv oO : Q a Vv 
eg Oar Beg) 7 (0) I+, Y ¥ (0) - Wes (z) vaz 


(101) 


After c has been determined from (101), the 
normal stress on the small-scale waves in the water 
due to the air flow may be determined. The sim- 
plest approach is to use (88) and infer Pa = Crm 
directly from Poy, using the momentum equation in the 
water (or Bernoulli's equation). 

Retaining the terms linear in the large-scale 
quantities, we have 


Day = Cres = 0h kz (c =H > = (gi TkA)cl (102) 


where c is given by (91b) with cy from (101). The 
correction to the growth rate, Cy, is doubly complex 
in that it has both real and imaginary parts (cee 
and c,) that are in phase and out of phase with @. 


550 


To complete the calculation of Section 2 for 
long-wave growth rate due to the non-uniformity of 
short-wave growth rate, we require the part of c. 
that is in phase with ee For the analysis of short 
waves, Section 2 uses an expression equivalent to 


Duesonry =fosku (cla aa) (user BU) (le ake) cn (42) 


where all quantities are real. This assumes that 

the real and imaginary part of c are modulated by 

the large scale in exactly the same proportions. 
Thus by this assumption, 


<5 (Ze, - uw) 
aky = -2 Gera (103) 
since 
co = V g/k+kT + ACen eee c' = Cy +U 


Should these assumptions not be exactly correct, a 
would have a small imaginary part, which will be 
ignored. 


4. NUMERICAL RESULTS 


We have carried out the calculations described in 
3 using the Orr-Sommerfeld solver developed by 
Gustavsson (1977). This is an implicit method 
which uses an Adam's -integration technique. One 
particularly attractive feature of the program is 
its variable step size. Thus it is possible with 
a reasonable number of points to have a fine mesh 
in the "wall" layer and other regions of high gra- 
dients and to coarsen the mesh as one moves out 
into the boundary layer. The programs and results 
will be more fully described in a subsequent publi- 
cation. Only one set of calculations will be 
reported here. 

The shear flow profile and its derivatives are 
modelled with continuous functions that approximate 
the mean profile of a turbulent boundary layer. 
Calculations were done at a friction velocity, uT, 
of 30 cm/sec; conditions were chosen so that the 
ratio, ut/Uw, was .05, a typical value for wind- 
tunnel experiments. Interaction between long waves 
of 100, 75, 50, 36, 20, and 16.5 cm with short 
waves of 2, 1, 0.75, and 0.6 cm were investigated. 
Although many interesting features of the flow can 
be investigated using this approach (such as the 
distortion of the mean profile as the large wave 
passes, and the variation of the wave speed, local 
‘growth rate, and amplitude of the short waves along 
the large waves), the only systematic investigation 
we have yet performed concerns the energy input to 
the large waves due to the modulation of the short- 
wave Stokes drift. 

The linear temporal growth rate of wind-driven 
waves 2; = ke; is of course a direct output of the 
calculations. Figure 2 shows Ma = sec-l as a 
function of wave number, k ~ cm-l, for we = 30 cm/ 
sec. The growth rates we obtained are slightly 
higher than Miles's viscous calculations [Miles 
(1962)] but when we used his shear-flow profile, we 
obtain close agreement. For Wee = 30 cm/sec, all 
the waves we investigated were viscously dominated, 
that is, their critical layers were sufficiently 
close to the free surface to be essentially merged 
with the surface viscous layer. Thus, little in- 
sight to the behavior of these flows can be obtained 


10. 100 
= 1S) 
1 ® 
1S) 2) 
® Ss 
= 
2 rs) 
Gg LO 10 2 
x = 
" o 
SG oO 
Oui 1 ess ate 1.0 
0.I | 10 
k~cm! 
FIGURE 2. Linear temporal growth rate @ u_ = 30 cm/sec; 
1—-—-—-present calculations; —W— present calcu- 


lations with miles profile U, = 5uU*. 


from an inviscid model of the behavior of shear 
flows. The real part of the wave speed, c,, also 
shown in Figure 2, differs very little from the 
free wave speed of gravity-capillary waves, co. 

The energy input to the large waves from the 
small waves is given by (66). With t = kct, and 
s'2 from (54), the dimensional temporal growth rate 
of the large waves can be written 

ae 


— 1272 
ae i S61] (104) 


= = aB' ~ c! 
= fo [2 1 Tas = 
—C} 
g 


where 2; is the linear growth rate Bk/2. 
the coupling coefficient C as 
-aB'k 


c= 7 (105) 


We define 


where the minus sign is introduced because, contrary 
to our expectations, a turned out to be negative 

for the cases we investigated. Thus the growth of 
the large-wave amplitude is given by 


dé 


— ican 2 
ae Jkt" | 2] (106) 


= Gy lia 


Thus for C positive, an energy input to long waves 
comes from short waves whose group velocity, c!, 

is slightly less than the long wave phase velocity, 
c¢. The theory also predicts that long waves will 
decay if c} > ¢. Since waves satisfying this con- 
dition will be shorter capillary waves which will 
be more strongly damped by viscosity, we expect a 
net energy input to the large waves. Of course 

the theory does not hold at ¢ = c, where non-linear 


“interactions must be considered. 


Numerical values of the coupling coefficient, C, 
are shown in Figure 3 as a function of A for various 


A'. C is certainly 0(1) having a maximum value of 
3 at 4' = 1 cm. It is also a slowly varying function 
of A. It has its maximum value about \’' = 1 cm 


which corresponds to the maximum in the linear 
growth rate for short waves for these conditions. 
It drops off more rapidly with decreasing wave 


length, A', than does the linear growth rate, Qa 
(A'). The long-wave linear growth rate, Qin is 
also shown for comparison; it is much smaller. Of 
course the interaction growth rate also involves 
(k'c') 2 of the short waves which would be typically 
0.01 but the division by cmc would somewhat offset 
the effect of small slope. One calculation for an 
upstream travelling long wave verified that a was 
negative and energy was removed from the long wave 
by interaction with the short wave. We have not 
carried the calculations further to date. 

Some idea of the wavelengths, involved in any 
practical application of these ideas can be seen 
from Figure 4 which shows the group velocity and 
phase velocities for gravity-capillary waves. The 
requirement for strong coupling is ¢ x ci. We 
further note that waves shorter than say 0.3 cm 
are unlikely to be important in a viscous fluid. 
Thus short waves in the range 0.3 cm could interact 
with a 20 cm long wave in the manner we have dis- 
cussed but waves longer than 20 cm would be unlikely 
to be affected. 

Although the effects of surface drift are not 
yet included in our calculations, the range of 
affected long waves can be somewhat broadened by 
considering surface drift. Drift velocities are 
typically 5% of the wind velocity; this is the same 
order as the friction velocity which we have taken 
as ut = 0.05 U.. If we assume that a surface layer 
will advect the short waves [Valuenzuela (1976) ] 
but leave the phase velocity of the long waves 
unaffected (Valenzuela's calculations did not extend 
to long waves), we can consider a broader range of 
interaction possibilities, as sketched in Figure 4. 
For a group velocity augmented by a surface current 
of 30 cm/sec, interactions between a long wave of 
about 50 cm and waves longer than 0.3 cm become 
possible and a 20 cm wave may interact with waves 
of order 1.4 cm. 

Experimental data in the range of wave lengths 
and friction velocities of interest for the inter- 
actions we have investigated here was presented by 
Plant and Wright (1977). Some of their results are 
reproduced in Figure 5, showing the temporal growth 
rate vs. wave number for several values of friction 
velocity. Of particular interest is that while the 
short-wave growth rate is accurately predicted by 
linear theory, there is a departure of theory and 


FIGURE 3. Coupling coefficients for long-wave and 
short-wave interaction; linear temporal growth rate 
Qs. aes 30 cm/sec. 


100 


\~cm 


L4cm 


Cg+ 30cm /sec 


A Se 


LLL) 


“10 100 
C ~ cm/sec 


FIGURE 4. Group velocity and phase velocity for 
gravity-capillary waves. 


(0) sea enn 
O.I 1.0 10 
k~cm! 


FIGURE 5. Measured temporal growth rates for various 
uu. cm/sec; from Plant and Wright (1977). 


55a 


552 


experiment for waves longer than about 10 cm. This 
is close to the first possible long wave that can 
strongly interact with a short wave whose group 
velocity is equal to the long wave phase velocity. 
Thus the results we have obtained to date indi- 
cate that the long waves can receive energy due to 
their interaction with wind driven short waves. 
The interaction mechanism we have investigated 
requires the presence of the wind and the variation 
of the short wave growth rate along the surface of 
the long wave due to changes in the local wind 
field caused by the passage of the long wave. Of 


course further work remains to be done to explore 
the full implications of these results, to complete 
the calculations and to make fuller comparison with 
experiment. 


5. ACKNOWLEDGMENTS 


This work was supported by the National Science 
Foundation under Grant ENG7617265. We acknowledge 
many stimulating discussions with our colleague, 
Professor E. Mollo-Christensen. 


APPENDIX A. 


DERIVATION OF STOKES' DRIFT MODULATION FROM 
KINEMATIC WAVE THEORY 


Kinematic wave theory, modified to allow for small 
dissipation or growth due to energy interchange 
with the wind, gives the following conservation 
equation for the wave action density, A", of the 
train of short waves: 


dA" 3 - 

a= tS UA = 20S A A.l 
at ax (a ) ac ( ) 
where &' is the temporal growth rate. The wave 
action density for waves on a current is given by 
[Bretherton and Garrett (1968) ] 


A' = E'/Q' 2 (A. 2) 


where E' is the energy density and 2' = k'(c' - 0) 
the frequency relative to the fluid at rest. By 
introduction of 


Be = ki (eu a) (a |2 = (c' — ws" (A.3) 


(A.1) may be cast as a conservation equation for 
the Stokes' drift 
as' 


e) tar = 
gee Ucegue 


1 dk u ' ok" ' ' 
gice | Six ) + 2011s 


Il 

7 
fob] Keb) 
x lee 


ap QV) Se (A.4) 
L 


With Qe = k'B" (1-akZ) /2 and expressed in the vari- 
ables T and & this takes the form 


=U 6 2 ey OL Rp ie vi 
[c carat ee c) dE ]2n S' = B(1-akz) 
u 
aus ' eo oe ee eee 
Ste oe +c 20 (re) O) Ven te LN) 


(A.5) 


By neglecting the variation of the left-hand side 
with t one finds from this 


s! set oe 
s (cg-c) (ee -ti) 
k! 


= Cra) [(c'-e)v,, + T)} (A.6) 


which, with S' = (c=) s"2/k", is found to agree 
with (53). 
APPENDIX B. 
THE EXTENDED SOLVABILITY CONDITION 
We first determine the adjoint to the homogeneous 


problem for a shear flow over a water surface. 
This problem is written 


L(w) = 0 
Wi ew (U" - kc) w' =0O 
Wo = Pw + Qw’ + bw"' = 
a ow . o at z=0 
w(e) = w'(~) = 0 (B.1) 


where L is the Orr-Sommerfeld operator. 
The adjoint to the Orr-Sommerfeld equation is 
[Sturart (1960) ] 


L(v) = v"" + ov" - 2ikRU'v' + [k* + ik3R(U-c)]v = 0 


o = -2k2-ikR(U-c) (B.2) 


From the Lagrange identity [Ince (1926) pp. 210, 
214) 


J tenes) - wL(v) }dz = P(w,v) 


0 0 


where P(w,v) is the bilinear concomitant. The 
boundary conditions on v that will complete the 
statement of the adjoint problem are found by the 
requirement that P(w,v) be zero at both end points. 
Since w and its derivatives are zero at Zz = o, this 
leaves the conditions on v to be found for z = 0. 
P(w,v) is written in bilinear form as 


(hepa) UA NAL UA | attastail 1 


(oa Ww 
-o 0-10 w' 
@Q al © © w" 
= v-[U] -w il 0) 0) © w"' (B.3) 


The free surface boundary conditions for w may be 
written 


v'-ke c (0) 0 w 

P Q 0 b w' 
w"' = 0 

ww" (B.4) 


(B.4) is an underdetermined set of equations 
that will yield two solution vectors with arbitrary 


coefficients. They are not unique and any linear 
combination will also be a solution. Two such 
solution vectors are 

Ww, = {0,0,1,0} and Wy = (-c,U'-kc,0,8) (B.5) 
where 

B = - [Q(U'-kc)-cP]/b 


We now enforce the requirement that P(w,v) be 
zero. This requires that certain linear combinations 
of v, v', v", v'"' be zero and these are of course 
the required adjoint boundary conditions. 

Consider the solution vector Wy: For P(w,v) to 
be zero 


P(w,v) =v °[U] - th 2 wo 0 | 


(B.6) 


This requires that 
0 = © (B.7) 


Consider the solution vector We: For P(w,v) to 


be zero, 
aloIsaoj) ap @(WUae)) ap (8 
P(w,v) =v °[U] We = Wo =o 
(Y= Tee 
Cc 
so that (B.8) 
[-icRU'k+o(U'-kc) + B]v - cov' + (U'-kce)v" 
re Ww = © 
Since v' = 0, this term may be eliminated from this 
relationship. Thus given 
WwW, (7) 2 ey 4 (! = ina = 0 | 
at z=0 
W, (w) = Pw + Qw' + bw" = of (B.9) 


for P(w,v) to be zero requires 


(v) = [-ikcRU' + o(U" - ke) + B]v + (U' - ke)v" 


tev = 0) at z= 0 (B.10) 
It can be shown that if (B.9) and (B.10) are used 
to construct P(w,v), the result is identically zero. 
Thus (B.10) are the boundary conditions for the 
adjoint problem. 

The solvability condition for a problem of the 
form [Ince (1926) ] 


553 


L(w) =r 
W, (w) = Va, aly = hyo) (B. 11) 
is that 
[ vrdz = i Von + 15 Wo yaeh + O0000 (B.12) 


where the Von'S are determined such that P(w,v) 
= Wy Wyn Wo Worm +... and v is a solution of 
the adjoint system. 


V, () ="0) i =1,n (B.13) 
For the present problem, only Y, and Yo are 

non-zero. By standard techniques, we have deter- 

mined the additional linear combination of w and 

Vv, 


W3 (w) = - A 

Wy (w) = - w" - ow 

V3 (v) = v/b 

Wh, (CA) = We a (Ce = OAsxe))a7 (B.14) 


such that the bilinear concomitant (B.3) may be 
written in the form 
P(w,v) = W Vy + W> V3 ta Ws Vo + Wy Vv 


1 1 


where W, and W2 are the boundary conditions from 
(B.1) and V, and V, are the adjoint boundary con- 
ditions from (B.10). 

Thus the solvability condition for non-homogeneous 
problem with non-homogeneous free water boundary 
conditions is 


i rvdz = Y, Vv, 1 Foy V3 (B.15) 
0 
where 
Wn = We uz (fe = Ore) 
at z=0 
V3 = v/b 
and v is a solution of the adjoint system 
L(v) = 0 
V, (v(o)) = 0 
Vy (v(0)) = 0 
v(~) = v'(~) = 0 (B.16) 


APPENDIX C. 


In this section we give the expressions for the 
various terms in (69) for the assumed form of the 
small scale (71). The continuity equation (73) 
has been used to express u' in terms of w'. 

The results are as follows 


sin. Oe Oa a Oi OS. dle Dts OS 
dtez  dtoz d22 k 023 dt k 9023 0x 
(Gea) 
o2w i Aw Poe eeOWsO at ow! 
RO 7 © REDE Gy BEG ae HEE BE ae Meca 
Vw = daw Ant alent Be Es + Vw (C.3) 
az2 3 ox i 
She0 > ya Pose 
pvesies Cy Bh, Be oe Otw 
IES ea ae ve om fen vo ae Bee 
+V2R + U" + UNZ (Cc. 4) 
89 _ i atw 32w A, © 3t aw A 33w cs 
ie a eee dz2 «k2 O& Oz5 923 Ox 
(Va) Un! (C.5) 
az 
Die ae open) oy cary SR Be 
ox > dz Ox dz (Ox 
Oo 
+ eaE Vw (C.6) 


The equation for the small scale will contain 
coefficients involving the mean flow expressed as 
a function of z'. 

U(z")) = U(z) + UN(zZ)) Z 
and u"(z") = U"(z) + U"' (z) & (G7) 


so that, for example, the term uV2w, with only 
linear terms retained, becomes : 


2 
uV2w = UV2w + [U + U'E + a] a k2w"] 
Oz 
nee & eon OS, See 
+ u'VCw €2ikU aE Oz (C.8) 
and 
2 ~ i 33w! dw' 
v2 = ao at ' mo4 mt aay RCE e ‘Wo 
wV*u wi w'[{U (0-7) whe aa3 iky ] 
coe (c.9) 


of these, v2w and 32a/ax2 will be ignored. 
The viscous term is manipulated as follows 


1,42 dw 2 du 1 2 ow 2 ou 
eye {co Se a oc 
RL ax v az! R Ny ox az 1 
1 32w! 2 i otw! 32w! 
+ - ") - = i 
plik G72 Ewe) k dzt 7 fe 
tee Oe. 6 Oke | Mh Beh 
+ RI3k 5 a2 = Dee ] (C.10) 


The fifth derivative is obtained from the Orr- 
Sommerfeld equation. Some cancellation occurs 
among these terms to yield the final result (76). 


REFERENCES 


Benny, D. J. (1976). Significant interactions 
between small and large scale surface waves. 
Studies in App. Math. 55, 93. 

Bretherton, F. P., and C.J.R. Garrett (1968). 
Wavetrains in inhomogeneous moving media. Proc. 
ROU ESOC TAy mS ODO 2a. 

Davis, R. E. (1972). On prediction of the turbulent 
flow over a wavy boundary. J. Fluid Mech. 52, 


287. 
Gustavsson, L. H. (1977). Private communication. 
Hasselman, K. (1971). On the mass and momentum 


transfer between short gravity waves and larger 
scale motions. J. Fluid Mech. 50, 189. 

Hinze, J. O. (1975). Turbulence, McGraw-Hill, 
New York. 

Ince, E. L. (1926). Ordinary Differential Equations, 
Dover Publications, New York. 

Larson, T. R., and J. W. Wright (1975). Wind- 
generated gravity-capillary waves: laboratory 
measurements of temporal growth rates using 
microwave backscatter. J. Fluid Mech. 70, 417. 

Longuet-Higgins, M. S. (1969). A nonlinear mech- 
anism for the generation of sea waves. Proc. 
ROUEE SOC -wrAt ae StL maith 

Manton, M. J. (1972). On the generation of sea 
waves by a turbulent wind. Boundary Layer 
Meterology 2, 348. 

Miles, J. W. (1957). On the generation of waves 
by shear flow. J. Fluid Mech. 3, 185. 

Miles, J. W. (1962). On the generation of surface 
waves by shear flow, Part 4. J. Fluid Mech. 13, 
433. 

Phillips, O. M. (1957). On the generation of waves 
by turbulent wind. J. Fluid Mech. 2, 417. 

Phillips, O. M. (1960). On the dynamics of unsteady 
gravity waves of finite amplitude, Part l. J. 
Fluid Mech. 9, 193. 

Phillips, O:. M. (1966). The Dynamics of the Upper 
Ocean, Cambridge University Press, 79-87. 

Plant, W. J. and J. W. Wright (1977). Growth and 
equilibrium of short gravity waves in a wind- 
wave tank. J. Fluid Mech. 82, 767. 

Stuart, J. T. (1960). On the non-linear mechanics 
of wave disturbances in stable and unstable 
parallel flows, Part 1. J. Fluid Mech. 9, 353. 

Townsend, A. A. (1972). Flow in a deep turbulent 
boundary layer over a surface distorted by water 
waves. J. Fluid Mech. 55, 719. 

Valenzuela, G. R. (1976). The growth of gravity- 
capillary waves in’a coupled shear flow. J. 
Fluid Mech. 76, 229. 

Valenzuela, G. R. and M. B. Laing (1972). Nonlinear 
energy transfer in gravity-capillary wave spectra, 
with applications. J. Fluid Mech. 54, 507. 

Valenzuela, G. R. and J. W. Wright (1976). Growth 
of waves by modulated wind stress. J. Geophysical 
Research 81, 5795. 


Preliminary Results of Some 
Stereophotographic Sorties Flown Over 
the Sea Surface 


L. H. HOLTHUIJSEN 


Delft University of Technology 


The Netherlands 


SYNOPSIS 


Preliminary results are presented of a study which 

is concerned with the directional characteristics 

of wind generated waves. The basic approach adopted 
was to measure the actual sea surface elevation as a 
function of horizontal coordinates by means of stereo- 
photogrammetric techniques. The surface representa- 
tions thus obtained were Fourier transformed to 
estimate two-dimensional wave number spectra. 

Basic considerations concerning the photogram- 
metrical process, the transformation rules and the 
statistical significance of the results are described. 
The required stereo photographs were obtained during 
photographic missions carried out in 1973 and 1976 
off the island of Sylt (Germany) and off the coast 
of Holland. So far three two-dimensional spectra, 
each from a different flight, have been calculated. 
The sea and weather conditions during these flights 
are briefly stated. The wind direction in these 
flights was off-shore. 

Frequency spectra computed from the observed wave 
number spectra are compared with an assumed frequency 
spectrum and an observed frequency spectrum. The 
agreement is reasonable but some discrepancy needs 
to be resolved. For two of the three observations 
the directional distribution of the wave energy is 
strongly asymmetrical around the wind direction. 

This asymmetry seems to correspond to asymmetry in 
the up-wind coast line. 

From the observed spectra a directional spreading 
parameter has been computed as a function of wave 
number. The results in normalized form agree well 
with published data. The absolute values of the 
spreading parameter for two spectra are within 30% 
of the anticipated values. For the third spectrum 
the values were almost five times too large but a 
comparison in this case may not be proper. In one 
of the spectra some indications of bi-modality around 
the wind direction have been observed in the direc- 
tional distribution function near the peak of the 
spectrum. 


555 


1. INTRODUCTION 

Observations of the two-dimensional spectrum of wind 
generated waves are relatively few and are mostly 
based on methods with rather poor directional resolu- 
tion. The techniques which are used for the observa- 
tions may be based on such systems as a sparse wave 
gauge array [e.g., Panicker and Borgman (1970)] ora 
buoy capable of detecting directional characteristics 
of the sea surface [e.g., Longuet-Higgins et al. 
(1963)]. The few detailed observations which have 
been published were based on other techniques such 

as high-frequency radio-wave backscatter [e.g., 

Tyler et al. (1974)], analysis of the sea surface 
brightness [e.g., Stilwell (1969), Sugimori (1975) ] 
or stereophotography [e.g., Cote et al. (1960)]. 
These provided information with a high directional 
resolution but the analysis of the results in terms 
of wave characteristics has not been very extensive. 

The Delft University of Technology and the Min- 
istry of Public Works in the Netherlands have devel- 
oped a system based on stereophotography which 
monitors the instantaneous sea surface elevation as 
a function of horizontal coordinates. It has been 
used in this and other studies and it is anticipated 
that it will also be used in future studies of wave 
phenomena such as wave transformation in the surf 
zone or wave patterns around marine structures. The 
present study, which is a joint effort of the Uni- 
versity and the Ministry, is aimed at observing and 
interpreting two-dimensional spectra of wind gener- 
ated waves in a variety of atmospheric conditions. 
The study is primarily directed towards the evalu- 
ation of the shape characteristics of the directional 
energy distribution of the waves. 

For this study a few hundred stereo pictures have 
been taken since 1973 and the analysis has just be- 
gun. The results reported here are preliminary in 
that the number of analyzed pictures is only a frac- 
tion of the total and in that the interpretation of 
these pictures has not as yet been completed. The 
spectra which are presented here were calculated 


556 


from three sets of pictures, each containing ten Cote et al. (1960)] and the present system is es- 
stereo pairs. These sets were chosen on two bases. sentially a revised version of the system used in 
One is the photographic quality which was judged by SWOP. 
photogrammetric experts, the other is the scientific It will suffice here to comment only briefly on 
interest. In this stage of the study it was felt the operational system. Actually two independent 
that wave fields generated by off-shore winds would systems were built. One is based on Hasselblad 
be of most interest because the boundary conditions cameras and has been described in detail elsewhere 
are well defined. Also, results of past investiga- [Holthuijsen et al. (1974)]. The other is an almost 
tions of wave generation [Hasselmann et al. (1973), exact copy of that system except that the Hasselblad 
Hasselmann et al. (1976)] suggests that observations cameras were replaced by UMK cameras of Jenoptik 
in these conditions may be extrapolated to more com- which are superior in optical and metrical aspects. 
plex conditions. The Hasselblad system was used for observations in 
The first set of pictures which was analyzed was the area off Sylt and the UMK system was used in 
taken in September 1973 during almost "ideal" off- the area off the coast of Holland. Synchronization 
shore wind conditions in the area just west of the of the cameras was achieved by using a radio signal 
German island of Sylt. These observations were that triggered a command pulse which was manipulated 
carried out in the framework of an international electronically in such a way that it complied with 
oceanographic project known as the Joint North Sea the timing characteristics of the receiving camera. 
Wave Project (JONSWAP) which is concerned with the The synchronization error for the Hasselblad system 
study of wave generation and prediction. A variety was less than 1 ms for all of the analyzed stereo 
of articles directly related to JONSWAP has been pairs and for the UMK system the synchronization 
published and more are being prepared for publica- error was less than 5 ms. To position the cameras 
tion. Some references are: Hasselmann et al. (1973), two Alouette III helicopters were used. These heli- 
Spiess (1975), Hasse et al. (1977), and Hiihnerfuss copters had a drop-door over which the cameras could 
et al (1978). The two other sets of pictures were be mounted. The distance between the helicopters 
taken in March and November 1976 in the area west of was estimated during the flight through a range 
Holland near the town of Noordwijk, also in off- finder which was imposed on the viewer of a third 
shore wind conditions. Wave observations at sea camera which looked from one helicopter to the other. 
level during the first and last flights are avail- It took a picture of the other helicopter every time 
able and these have been used for comparison with the downward looking cameras were activiated. From 
the stereophotogrammetric results. these photographs the distance between the helicop- 


ters could be computed and the scale of photography 
could be determined. 


2. STEREOPHOTOGRAMMETRY OF THE SEA SURFACE The specification for the helicopter formation 
during a photographic sortie were largely based on 

When an object is photographed from two slightly photogrammetric requirements. Only the altitude 
different positions, the imagery in the two pictures was based on the anticipated sea state since the 
will also be slightly different. The differences noise and resolution in the spectrum are directly 
depend upon the geometry of the object. By measur- related to the altitude of photography. The upper 
ing the differences, the elevation of the surface limit of the altitude was based on noise considera-— 
relative to an arbitrary plane of reference can be tions. The standard deviation of the measurement 
determined. The conventional technique of analysis error is estimated to be 0.03% of the altitude 
requires human interpretation of the pictures and [Holthuijsen et al. (1973)]. Taking a noise to 
complicated stereoscopic viewing devices. More ad- signal variance ratio of 1:10 as an acceptable upper 
vanced procedures, which have only recently been limit, it can be shown that the altitude should be 
developed, use a computer to carry out a correlation less than 1,000 times the standard deviation of the 
between the images to arrive at the same results instantaneous sea surface elevation (or 250 times 
[e.g., Crawley (1975)]. the significant wave height). The lower limit of 

In the conventional geodetic aerial survey the the altitude is directly related to the resolution. 
pictures are taken vertically in sequence from an If a resolution in the spectrum is required equiva- 
airplane and the interval is chosen such that the lent to % of the peak wave number or better, it 
pictures overlap in the area directly under the : appears that for the Hasselblad system the altitude 
line of flight. An obvious condition is that the should be higher than 6.7 times the reciprocal of 
object does not change between exposures. In land the peak wave number. For the UMK system the fac- 
survey this poses no problem since the ground sur- tor is 4.0. For most "young" sea states these upper 
face does not move. The sea surface, however, and lower limits are not in conflict. The final 
changes very rapidly. To limit the distortions be- choice of the altitude was confined to multiples of 
tweeen two successive pictures to an acceptable 250 ft for the pilot's convenience. 
level, they should be taken within an interval of The size of the sea surface covered in stereo in 
1-5 ms. The airplane cannot possibly fly from one stereo pair is usually too small to produce suf- 
one required point of photography to the other within ficient data for a reliable estimate of the two- 
this time lapse. The consequence is that not one but . dimensional spectrum. To increase the amount of 
two cameras are needed which take the pictures "simul- data more pictures were taken in sequence with a 
taneously," that is, within an interval of 1 - 5 ms space interval sufficiently large to ensure photog- 
and that two aircraft are needed to position the two raphy of non-overlapping sea areas. The correspond- 
cameras. Apart from these technical differences in ing time interval between the exposures would be 
obtaining the stereo pairs, the methods and pro- typically between 4 s and 20 s (depending on camera 
cedures used in this study are standard in geodetic type, ground speed, and altitude). The photographic 
survey and they have been used in the past by various operation to obtain this sequence is called a sortie. 
Oceanographic investigators. A well publicized ef- In principle, the pictures can be analyzed with 


fort is the Stereo Wave Observation Project [SWOP, recently developed, fully automated processes. The 


facilities, however, were not available for the pres- 
ent study and the conventional technique was used. 

In the three-dimensional space which is reproduced 

in the stereoscopic viewing devices a right-handed 
system of coordinates was defined with the y-axis 

in the direction of flight and the z-axis upward. 
During the analysis the sea surface was read at a 
square grid with spacing Ax = Ay, which was chosen 
such that aliasing in the spectrum would be limited 
to only a fraction of the total wave variance. For 
each stereo pair the analysis was carried out ina 
square field as large as possible and the elevations 
were determined relative to an arbitrary plane of 
reference. In the subsequent numerical analysis the 
linear trend was removed through a least-squares 
analysis. The fields obtained from a series of 
stereo pairs were initially arbitrary in shape but 
fairly close to a rectangle. Later they were clipped 
or extended to a square of one common size of Ly.*Ly 
as required in the spectral analysis. Sections where 
no stereo information was available (mainly in the 
areas of extension) were filled with zeros. 


3. TRANSFORMATION AND STATISTICAL SIGNIFICANCE 


The sea surface data from the stereophotogrammetric 
analysis were Fourier transformed to estimate the 
two-dimensional wave number spectrum (k-spectrum) . 
To inspect the directional characteristics as a 
function of wave number, the K-spectrum was trans— 
formed to the wave-number, direction space to pro- 
duce the k,9-spectrum. The k-spectrum was also 
transformed to the frequency domain. 


The k-Spectrum 
The definition adopted here for the two-dimensional 


wavenumber spectrum E(k) is given by Eqs. 1, 2, and 
30 


re 
E(k) = lim < ee > (1) 
Aoo 
where 
Ba > -i2tk:x 12 
H(k) = |SS h(x) e ax (2) 
> 
R 
A = Sf dx (3) 
> 
R 


and <> denotes ensemble averaging. Observations of 


E is estimate of E(k) 


5 


a 
k-plane 


grid in k-plane 


E. linearly interpolated between EB and E 
Ee linearly interpolated between E 


Eo linearly interpolated between E 


557, 


h(x) were available from the stereo analysis ina 
number of square fields and these fields were con- 
sidered to be realizations of the ensemble. They 
were Fourier transformed with a multi-dimensional, 
multi-radix FFT procedure [Singleton (1969)] and 
the final estimates were obtained by averaging the 
results over the available realizations. The sea 
surface data were not tapered and the spectral 
estimates were not convolved; consequently the 
spectral estimates are "raw" estimates. In analogy 
with time series analysis [e.g., Bendat and Piersol 
(1971)] the reliability is represented by a y2= 
distribution with 2n degrees of freedom, where n 

is the number of fields. The resolution denoted 

by Ak, Aky is) on the order of (Ly ° Ly) ae 


The k,9-Spectrum 


ze 
The transformation of the k-spectrum to the k, 
6-spectrum is formally given by Eq. 4. 


nm 
E(k,6) = E(k) |J;| (4) 


where k = magnitude of ik, 8 = orientation of k and 
where the Jacobian Jj = k. Computing the values of 
E(k,8) at a regular grid in the k,8-plane requires 
the estimation of E(K) at corresponding values of 
k. This was done by bi-linear interpolation of 
E(K) at the proper values of K (see Figure 1). 

The directional resolution can be estimated by 
considering the angular distance between two 
neighbouring, independent estimates of E(K) ona 
circle in the ¥-plane centred in k = 0. On this 
circle with arbitrary radius, k, approximately 
27Tk/Ak independent estimates of E(K) are available 
and the directional increment between these estimates 
in radians is Ak/k. This would be a fair approxima- 
tion of the directional resolution if all pictures 
were oriented in the same direction. But actually 
the orientation is a random variable due to the heli- 
copter motion during the sortie. The directional 
bandwidth to be added will be on the order of twice 
the standard deviation (dg) of the helicopter yaw. 
The final expression for the directional resolution 
(A8) is given in Eq. 5. 


A® = Ak/k + 20, (5) 


The resolution in k will be on the order of the 
increment between estimates of E(k) in the k-plane 
which is LZ! = Lo 

The reliability of the estimates of E(k,6) can 
again be expressed in terms of a y2-distribution but 
the number of degrees of freedom is not uniformly dis- 
tributed over the k,@-plane. It constitutes an un- 


2 
3 and E, 


5 and EG 


+ gridpoint in k,@ plane transformed to 


FIGURE 1. Bi-linear interpola- 
tion in the k-plane. 


558 


FIGURE 2. Sites of the field operations. The areas in 
the boxes are shown enlarged in Figures 3 and 4. 


dulating function due to the fact that the estimated 


value of E(k,@) is based on four values of E(k) which 


are usually not equally weighted in the given in- 
terpolation technique. They are equally weighted 
only when a transformed gridpoint in the k,8-plane 
coincides with the centre of a mesh in the ¥-plane. 
In that case the number of degrees of freedom for 
E(k,6) is four times the number of degrees of free- 
dom for each individual estimate of E(K). This is 
the upper extreme of the undulating function. The 
lower extreme occurs when a transformed k,® grid- 
point coincides with a gridpoint in the K-plane. 
Then the number of degrees of freedom of the esti- 
mate of E(k,8) is equal to the number of.degrees of 
freedom of an individual estimate of E(k). The 


values of the two extremes are 8n and 2n respectively. 


The £-Spectrum 


The f-spectrum is determined by integrating the f, 
8-spectrum over the range (0,7) and multiplying the 
result by two. The operation is given by Eq. 6. 


T 
E(f) = 2 f E(£,6)d0 (6) 
0 


The £,8-spectrum has been computed from the k- 
spectrum. The relationship to transform from wave 
number vector to frequency is based on the linear 
dispersion relation for deep water corrected for 
currents. This expression and the transformation 
are given in Eqs. 7, 8, and 9. 


L. >> 
f = (gk/2m) * + k.V (7) 


E(£,0) = E(k)|g.| (8) 


By 
Jo = Ds (g/2m) 3/2 + vk-! cos (0, - e)17? (9) 


a 
V is the current vector and V and 8. are its magni- 
tude and orientation. To determine the values of 

E(k) the same procedure as described above was used. 


The resulting spectrum is the frequency spectrum as 


Norway 


\ 
Germany | 


Great Britain | 


Belgium 


observed in a point stationary with respect to the 
sea bottom. This was done so as to be able to com-— 
pare the results with measurements carried out with 
anchored buoys. Expressions for the approximate 
resolution (Af) and number of degrees of freedom 
(N) are given by Eqs. 10 and ll. 


> Use el 
KE ak Ak = on eis (10) 
Phe) 
ee ues ab 
N= 8 G AK (11) 


4. DESCRIPTION OF THE SITES AND THE WEATHER 
CONDITIONS 


Maps of the areas off Sylt and off Noordwijk and 

two bottom profiles are given in Figures 2, 3, 4 

and 5. It may be noted that both areas are similar 
in general appearance but an important difference 
seems to be that the coast near Sylt recedes sharply 
North and South of the island and is strongly asym- 
metric with respect to the off-shore direction, 


25km 


+ observation tower 


FIGURE 3. The area of observation off Noordwijk. Lo- 
cations of observations indicated by dots. 


» Denmark 


554 


7e \ o 
= EOE | 
FIGURE 4. The area of observation off Sylt. Active 


wave monitoring stations and station of observation 
indicated by dots. Wind direction indicated by arrow. 


whereas the coast near Noordwijk is more continuous 
and symmetric. For both sites the water is effec- 
tively deep for waves generated by an off-shore wind. 

The sortie in the area west of Sylt was carried 
out during the field operations of JONSWAP in 1973, 
on September 18th, at 17:30 hr (local). Britimmer et 
al. (1974) describe the large scale weather features 
during the JONSWAP operations of 1973 and also give 
results of meteorological observations from ships, 
buoys, and balloons in the area. According to this 
information the windspeed and direction prior to the 
flight had been fairly constant for one day. Since 
the wind was almost perfectly off-shore the situa- 
tion was classified as an "ideal" generation case. 

In the two hours prior to the flight the windspeed 
and direction at station 8 (see Figure 4), at 10 m 
elevation was approximately 13 m/s and 110° respec- 
tively. The direction is only a few degrees off the 
"ideal" off-shore direction of 107°. 

The weather during this flight was poor for photo- 
graphic operations and all pictures which were taken 
were under-exposed, in spite of the best possible 
photographic measures. Pictures were taken over 
six stations of JONSWAP, including active wave mon- 
itoring stations 5, 7 and 9 (see Figure 4). The 
frequency spectra observed at these stations are 
given in Figure 6 and they may be used for a direct 
comparison with the results of stereo observations 
over these stations. But in selecting the pictures 
for preliminary investigation preference was given 
to photographic quality rather than availability of 
ground-true information and it appeared that the best 
pictures were taken over station 10, which was other- 
wise inactive during the flight. 


TTT, ON 
———— Noordwijk 


FIGURE 5. Bottom profiles off Sylt (direction 287°) 
and off Noordwijk (direction 300°). 


559 


The frequency spectrum at station 10 was estimated 
with a "hindcast" procedure based on the JONSWAP pa- 
rameter relationships [Hasselmann et al. (1973)]. 

The "hindcast" was attempted for stations 5, 7, and 

9 with the observed windspeed of 13 m/s but the re- 
sults (Figure 6) were rather poor, although they 
seemed consistent with the statistical variation in 
the observations of JONSWAP. The agreement improved 
when a windspeed of 15 m/s was used (Figure 7). This 
was the windspeed estimated just prior to the flight. 
Since this fictitious windspeed produced more real- 
istic results, in particular for station 9 which was 
the nearest to station 10, it was used for the "hind- 
cast" at station 10. The resulting spectrum is given 
in Figure 8, the comparison with the stereophoto- 
graphic results will be discussed in Section 5. 

The second and third set of pictures to be ana- 
lyzed were chosen from the pictures obtained in the 
area off Noordwijk. The main reason for selecting 
these pictures rather than the pictures taken off 
Sylt was that the results of the sortie just de- 
scribed indicated that the data were influenced by 
the asymmetry of the coastline of Sylt. The coast 
near Noordwijk is more symmetric for off-shore wind 
directions. The information on the atmospheric con- 
ditions during these flights was based on standard 
synoptical observations which were received through 
the office of the Royal Netherlands Meteorological 
Institute. In addition a cup-anemometer and a wind- 
cone were available at an observation tower located 
9.5 km off-shore from Noordwijk (see Figure 3). 

The second sortie (the sequence refers to the 
sequence of analysis, not the time sequence of the 
flights) was flown in off-shore wind conditions on 
November 12, 1976, at 13:05 hr (local). From the 
synoptical observations it was found that the wind 
was rather weak over the entire North Sea and the 
wind in the area of observation was mainly caused 


Sylt 730918 


ast JONSWAP spectrum _— 


observed spectrum 
17:51 start of record (duration 25 min.) 

N stat.9 
30+ 17:25hr 


257 


energy density (m*/Hz) 


0.30 0.35 0.40 


010 015 0.20 


FIGURE 6. Observed frequency spectra at stations 5, 7, 
and 9 and corresponding JONSWAP spectra for U = 13 m/s. 


560 


Sylt 730918 


35} 
stat.9 JONSWAP spectrum 

30 | 17.25 hr 
r | observed spectrum 


| \ 17:51 start of record (duration 25 min.) 


20 


wn 


i=) 


spectral density ( m?/Hz) 


os 


Qo ——, 
0.10 a15 020 025 030 035 040 
frequency (Hz) 


FIGURE 7. Observed frequency spectra at stations 5, 7, 
and 9 and corresponding JONSWAP spectra for U = 15 m/s. 


by a weak and fairly large low pressure area over 
central France. Synoptical observations in the 
coastal region 25 km North and 8 km South of Noord- 
wijk indicated windspeeds of 4.5 m/s and 4.0 m/s 
respectively and the wind directions of 100° and 
160° respectively. The wind observation at the 
platform was carried out at 23 m above mean sea 
level. Averaged over the duration of the photo- 
graphic operations (about 40 min.), the observed 
windspeed was 6.4 m/s and the directions just prior 
and just after the flight were approximately 140°. 
The “ideal" off-shore direction would have been 120°. 
To estimate the windspeed at 10 m elevation, the 
observed value was corrected. The correction for 
the bulk of the tower, is known from wind-tunnel 


5.0 
Sylt 730918 


JONSWAP spectrum ------ 
Observed spectrum 


40 
(from stereo dota) 


w 
o 


spectral density (m?/Hz) 
Nn 
i=) 


is) 


01 015 0.20 0.25 030 a35 f (Hz) 


FIGURE 8. Spectrum inferred from stereo data and cor- 
responding JONSWAP spectrum for U = 15 m/s. 


tests, and the windspeed was extrapolated using a 
logarithmic wind-profile with a drag coefficient, 
cj9 = 1-5 x 10-3. The resulting windspeed is 6.0 
m/s. The corrections for the wind direction are 
marginal and well within the error of observation. 
During this flight pictures were taken over the 
observation tower and at locations 30 km and 50 km 
from the coast (see Figure 3). The pictures taken 
30 km off-shore seemed to contain sufficient stereo 
information to obtain a relatively high directional 
resolution and these were chosen for preliminary 
investigation. 

Wave observations at sea level were available 
from a wave gauge at the observation tower and from 
an accelerometer buoy at the location 30 km off- 
shore. The spectrum of the buoy is given in Figure 
9. It will be used for comparison with the stereo- 
photographic results. During the flight some swell 
coming from south-westerly directions was observed 
from the helicopters. 

The third sortie was flown off Noordwijk on 23 
March 1976 at 12:20 hr (local). The wind was rather 
weak over the entire North Sea and the direction 
varied from ENE off the Dutch coast to SSW off the 
Norwegian coast. This windfield was caused mainly 
by a fairly weak high pressure ridge over the North 
Sea and a low pressure area over central France. 
Synoptical observations at the same coastal stations 
as mentioned above indicated windspeeds of 11.0 m/s 
and 8.0 m/s respectively and wind directions of 80° 
and 70° respectively. The corrected wind speed and 
direction at the observation tower (averaged over 
20 min.) were 8.3 m/s and 70°. Since the "ideal" 
off-shore wind direction would have been 120° the 
wind is slanting across the coast line at an angle 
of approximately 50°. Obviously this implies a 
strong asymmetry of the coast line with respect to 
the wind direction. Pictures were taken over the 
observation tower and at locations 17 km and 30 km 
off-shore. Since the pictures taken 17 km off-shore 
seemed to be the best, they were analyzed. Unfor- 
tunately no simultaneous wave observations in the 
area were available. 


0.50 Noordwijk 761112 
buoy spectrum = = ———-—- 
observed spectrum 

0.40 (from stereo data) 


030- 


spectral density(m*/Hz) 


010 0.20 030 0.40 0.50 
frequency (Hz) 


FIGURE 9. Spectrum inferred from stereo data and 
spectrum from buoy measurement. 


561 


TABLE 1 Photogrammetric Sylt Noordwijk | Noordwijk 
parameters Sept. 1973 | Nov. 1976 March 1976 
altitude of photography 1500 ft 


orientation of helicopters 
elative to true North 


percentage of zeros added 
in stereo areas 


umber of pictures 
accepted for stereo analysis 


2 
stereo area per picture 220x220 m™ 156x156 ma 170x170 an” 


es ey 2 2 2. 
grid in x-plane 5), 33 5) in 3} big Sh am Dose~aos) iu 


5. RESULTS resolution, Og was estimated at 0.06 [c-.f., 
Holthiujsen et al. (1974) ]. 

The values of a number of parameters relevant to the On closer inspection of the contour-line plot of 

photogrammetric process are given in Table 1. In the spectrum of Sylt two wave fields can be identi- 

view of the preceding paragraphs this table is fied: one coming from approximately 110° and one 

largely self-explanatory but a few parameters will from approximately 155°. This is rather surprising 

be discussed briefly. because neither the wind conditions nor the ground- 
The altitudes of photography are based on antic-— true information gave such indication. The swell 

ipated significant wave heights and peak wave num- in the second spectrum (off Noordwijk) coming from 

bers. These were estimated by substituting the south-westerly directions was observed during the 

windspeed and fetch in the JONSWAP parameter rela- flight. It is well separated from the locally 

tionships [Hasselmann et al. (1973)]. For the sortie generated wind sea and it will be largely ignored 

off Sylt the wind information was fairly good as it in the following discussion. The peak of the third 

was based on ship observations in the area but for spectrum is, surprisingly, coming from Northerly 

the sorties off Noordwijk this information was poorer, directions rather than from Easterly directions, as 

partly because no observations prior to the flights may be antitipated from the wind direction. 

were available. The helicopters were flying directly Instead of the k, -spectra, the normalized direc-— 

into the wind during the sortie off Sylt. During tional distribution functions have been plotted in 

the second and third sortie they were flying with Figures 13, 14 and 15. The definition of these 

the wind in the left respectively right rear quarter functions is given by Eqs. 12 and 13. 

with 11° drift. Ten stereo pairs were taken in each 

sortie but one pair was rejected from the set taken E(k,8) 

off Sylt because it covered too small an area. Us— DO BES) ° FOO ies (te) 

ing the sea surface information from the stereophoto- Halen eek 

grammetric analysis the three K&-spectra were computed 

according to the procedures described in Section 3. 

The results are presented in the form of countour- D(8;k) = 0 mone Wy SO) Ss Ay (13) 

line plots in Figures 10, 11, and 12. Some isolated 

regions in the k-plane have been indicated where the This seemed to be more illustrative than a contour- 

spectra are thought to be seriously affected by line plot of the k,§-spectra, the normalized direc-— 

noise. This noise is dealt with in the Appendix. primarily for the directional characteristics. An 

Values of relevant spectral parameters are given in evaluation of these functions will be given in the 

Table 2. For the determination of the directional next paragraph. 


TABLE 2 Spectral 
parameters 


Sylt 
Sept. 1973 


Noordwijk | Noordwijk 
Nov. 1976 | March, 1976 


] 


resolution in k-plane imal (220x220) 1 (156x156) (170x170)! 


number of degrees of 20 


freedom 


1 


] 0.0641 &*) 


peak wave number (Gc) [m 


directional resolution 
atke=s ok 
k = 2k" 
k = 3k” 
m 


as 
10 


562 


FIGURE 10. Contour-line plot of K-spectrum off Sylt, 
Sept. 18th, 1973. Contour-line interval equivalent 
to factor 2. Minor variations are dashed, shaded 
areas seriously affected by noise. Orientation of 
positive ky-axis 110°, k_-axis 200° from true North. 
Wind direction 110°. 


FIGURE 11. Contour-line plot of K-spectrum off 
Noordwijk, Nov. 12th, 1976. Contour-line interval 
equivalent to factor 2. Minor variations are dashed, 
shaded areas seriously affected by noise. Orientation 
of ky-axis 275°, negative k -axis 185° from true 
North. Wind direction 140°. 


FIGURE 12. Contour-line plot of k-spectrum off 
Noordwijk, March 23rd, 1976. Contour-line interval 
equivalent to factor 2. Minor variations are dashed, 
shaded area seriously affected by noise. Orientation 
of ky-axis 310°, k,-axis 40° from true North. Wind 
direction 70°. 


mt 
is 
Ber 
B)e 
els Sylt 730918 
5 ae 
(J resolution area Aky=Ak,4220m) 3 5 
0.5 spectral density(m4) 0 
. 1 10 


ee 4 


—_—_, 
Q03 Q04 Q05 Q06 007 Q08 008 0.10 


001 002 


wavenumber component ky[m7"] 


North {i 
Wind 
al Noordwijk 761112 
een es 
ol 
v4 
A ot 

C] resolution area AK, =AKy=(156m) © 

0S spectral density (m4) F a 
3|E a 
3/8 = G 

‘F015 = 


Li 
002 006 006 g08 O10 O12 014 016 
wavenumber component k[m"] 
North Wind 
Ss Noordwijk 760323 
heh Ot 
re 
oe 
O resolution area AKx=AKy=(170m)~! e|s 
05 spectral density (m4) i 8 
Oren: 
FD) g BS 


01 015 02 
wavenumber component - ky [m? ] 


North 


1.0 
SCALE 


20° 110° 200° k =1.90 km 
k =0.71km k=2.14km 
k =0.95km k =238km 
k=119 km k=2.62km 
k =1.43 km k =2.86 Km 
k =157km k=3.10 km 


KeTetSSixt Om min, 


The f-spectrum has been computed from the ie 
spectrum according to the procedures described in 
Section 3. The result for the spectrum off Sylt 
is given in Figure 8 along with the corresponding 
JONSWAP spectrum. The resolution is about 0.02 Hz 
near the peakfrequency, which is 0.165 Hz, and 0.01 
Hz at twice the peak frequency. The number of de- 
grees of freedom for frequencies greater than 0.13 
Hz is 250 or more. Considering the scatter in the 
original data set of JONSWAP and taking into account 
the resolution, it is concluded that the agreement 
between the two spectra is fair. 

The frequency spectrum computed from the observed 
k-spectrum of the second sortie is plotted in Fig- 
ure 9 along with the frequency spectrum of the buoy. 
The resolution of the spectrum based on the stereo 
data is on the order of 0.02 Hz near the peak of 
the swell and 0.015 Hz near the peak of the locally 
generated wind sea. The number of degrees of free- 
dom is 125 or more for frequencies greater than 
0.10 Hz. For the spectrum of the buoy the resolu- 
tion is about 0.02 Hz and the number of degrees of 
freedom is about 48. 

The spectrum based on the stereo data seems to 
be shifted in energy density. This may have been 
caused by noise and to appreciate this influence 
the R-spectrum was corrected. The noise was assumed 


563 


Sylt 730918 


k =3.33 Km 
k =3.57km 
k =3.80 km 


k=4.28km 


FIGURE 13. Normalized directional 
distribution functions of the k- 
spectrum off Sylt, Sept. 18th, 
1973. Directions are relative to 
true North. 


to be uniformly distributed over the K-plane and the 
variance was estimated at 0.002 m2 (based on the 
anticipated measurement error of 0.03% of the alti- 
tude of photography, see Section 2). Accordingly 
a uniform noise level of 0.018 m* was subtracted 
from the ¥-spectrum and the transformation was 
carried out again. The differences were marginal 
compared with the earlier results and the shift 
cannot be explained with the anticipated noise uni- 
formly distributed over the K-plane. Further in- 
vestigation is needed to resolve the remaining 
discrepancy. 

The frequency spectrum of the third sortie is 
given in Figure 16 but no attempt has been made 
to compare this spectrum with a "hindcasted" spec— 
trum because the relatively simple relationships 
for off-shore wind situations cannot be applied. 


6. DISCUSSION OF THE RESULTS 


In the area off Sylt, where the wind was almost 
perfectly off-shore and fairly homogeneous and 
stationary, one would expect to find a frequency 
spectrum with a shape similar to the shape found 
earlier in JONSWAP. Finding a JONSWAP-type spectrum 
in the conditions off Noordwijk seems to be less 


564 


k=0.2 km 


k=0.3km 


k=0.4 km 


k=0.7 kp, 


FIGURE 14. Normalized directional distribution 
functions of the k-spectrum off Noordwijk, 

Nov. 12th, 1976. Directions are relative to 
true North. The peak wave number kn is related 
to the locally generated wind sea. 


likely because the differences between the wind ob- 
servations at the coast and at the tower are fairly 
large and the wind may have varied between the point 
of observation and the coast. In particular for the 
slanting wind conditions it is obvious that a JONSWAP-— 
type spectrum would not be found, due to the asym- 
metry in the coastline around the wind direction. On 
the other hand, non-linear interactions in the spec— 
trum may produce a JONSWAP-type spectrum, in spite 

of the asymmetry and the variations in the windfield 
[Hasselmann, et al. (1976)]. From an inspection of 
Figure 8 it can be concluded that the frequency 
spectrum in the sortie off Sylt is indeed JONSWAP- 
like. The correspondence of the frequency spectra 
off Noordwijk with a JONSWAP-type spectrum has not 
yet been investigated. 

For the k-spectra of the first two sorties one 
would expect to find directional distribution func- 
tions having some kind of standard shape, symmetrical 
about the mean direction although some skewness may 
be expected in the observation off Noordwijk because 
the wind direction was not perfectly off-shore. For 


SCALE 


Noordwijk 761112 


oo 
= 
3 


= 
W 
is) 
= 
3 
= 
i 
nN 
= 
3 


k=26 km 


km=6.41x 10-2 m_~! 
k=18k rp 


the third spectrum strong skewness may be anticipated 
due to the slanting position of the coastline. 

These expectations seem to be far from reality 
in the k-spectrum off Sylt. The directional distri- 
bution near the peak of the spectrum (see Figure 13) 
is distinctly asymmetric with respect to the wind 
direction with the highest peak at + 45° off the 
wind direction (155° from true North). It is highly 
improbable that the wave generation mechanism would 
build a directional distribution as strongly asym- 
metrical as this. An explanation for this unexpected 
observation can perhaps be found through a detailed 
study of the wind and wave fields, possibly using 
"hindcasting" procedures. But in the context of this 
paper one can only speculate on some possible causes. 
The source function is symmetrical, as is the radia- 
tive energy transfer, since bottom and current re- 
fraction is virtually non-existent. It seems then 
that the asymmetry stems from asymmetry in the wind 
field or in the boundary conditions. As for the 
wind field, a cursory inspection of the large scale 
weather maps revealed no asymmetry. As for the 


boundary conditions, the coast of Sylt, rather than 
the main-land coast was deemed to be relevant as up- 
wind boundary. This was based on the expectation 
that the wave energy is propagating in a narrow 
angular sector around the wind direction [e.g., 
Hasselman et al, (1973)] and since the coast of 

Sylt is rather symmetric it should not cause asym— 
metry in the wave field. But the coast to the North 
and South of Sylt is strongly asymmetric. In fact, 
the distance to shore in the direction of 155° (the 
direction of the highest peak) is almost 2.5 times 
the distance to shore in the direction of 65° (the 
"symmetrical" direction, see Figure 4). If this 
asymmetry in the windward boundary is indeed the 
cause, then it seems that the "ideal" generation 
cases of JONSWAP may be contaminated to some degree 
by asymmetric boundary conditions. Still, relating 
this conclusion to the observed K-spectrum is largely 
speculative as long as it is not substantiated with 
more data. In particular the shapes of the k- 


1.07 
SCALE / ; 


———————er 
310° 40° 130° k=2.00km 

eee ey 
k= 022 km k=2.22km 
a 
k=0.44 km k=2.44Km 

| ele 
ed (Be 
k=0.67k 4 k=2.67 km 


k=1.56km k=355 km 
ale Bea 
k= 1.78 km k=3.78 kp 

FIGURE 15. 


565 


spectra at locations closer to shore may give some 
clues. 

The expectations regarding the directional dis- 
tributions for the locally generated wind sea off 
Noordwijk in the second sortie seem to be more 
realistic, at least in an overall sense (Figure 14, 
for k > k,). Any skewness is hard to identify 
through visual inspection of the plots due to the 
small scale variations in the functions. These 
probably stem from the statistical variability of 
the estimates. The swell peak (k = 0.3 con 
0.6 k,) is unimodal and covers a narrow angular 
sector with a half power width of about 35°. 

The directional distribution functions of the 
spectrum in the third sortie seem to be strongly 
skewed for the lower wave numbers (Figure 15, 
mS 2 kp, Say) but for higher wave numbers skewness 
is hard to identify visually. As for the main di- 
rection of the energy distribution, it varies almost 
monotonously from approximately 80° at higher wave 


Noordwijk 760323 


=s00K =a ean eae k=600 Km 

I ll 
AN aie PIS AEN 
eos = Nihewrsy 
k=4.44 km k=6.44km 
ee weir 


(he 
k=489 km 


k=5.11 km k=711 km 
ab 
[jira ee eee 
k=533m k=733 km 


k=7.55km 


k= 5:88 x1073m_! 


k=5.78 km 


as 
Normalized directional distribution functions of the k-spectrum off Noordwijk, 


March 23rd, 1976. Directions are relative to true North. 


566 


numbers to about O° for the lowest wave numbers (see 
also Figure 17). The energy of the higher wave num- 
bers travels more or less in the wind direction but 
the main direction of the peak of the spectrum ap- 
pears to be about 10° relative to true North; that 
is about 60° from the wind direction and almost 
parallel to the coast. This seems to be the most 
remarkable feature of this spectrum as one would 
expect to find a uniform main direction of 70°, con- 
sidering the wind direction and the effects of non- 
linear interactions [Hasselmann et al. (1976)]. 
Again, as with the spectrum off Sylt, it is felt 
that the observed phenomenon is due to the asym- 
metry of the coastline around the wind direction. 

To substantiate this preliminary conclusion 
qualitatively, a simplified "hindcasting" model was 
implemented for homogeneous, stationary wind fields, 
arbitrary coastlines, and deep water. In this model, 
which is basically the same as suggested by Seymour 
(1977), the wave components from different direc-— 
tions are decoupled. In this version the parameter 
relationships from JONSWAP [Hasselmann et al. (1973) ] 
were taken and the suggestions of Mitsuyasu et al. 
(1975) were used for the directional distribution 
function. When applied to the situation of the 
first and third sortie it did produce two-dimensional 
£,6-spectra which at least qualitatively agreed with 
the so far unexpected main directions in the observed 
k-spectra. 

This seems to be.in contradiction with the con- 
clusions of Hasselmann et al. (1976) that the shape 
of the spectrum is fairly insensitive to variations 
in the wind field due to the non-linear interactions 
in the spectrum. It should be noted however that 
the distance to the coast, in terms of wave lengths, 
seems to be rather short for the lower wave numbers 
in the two spectra so that non-linear interactions 
May not have been sufficiently effective to over- 


0.40 
Noordwijk 760323 
frequency spectrum 
| from stereo data 
0.30 


spectral density (m*/Hz) 


0.10 0.20 0.30 0.40 050 
frequency (Hz) 


FIGURE 16. Spectrum inferred from stereo data of ob- 
servation off Noordwijk, March 23rd, 1976. 


+ Noordwijk, 761112, Ak=1/156m~" 
1 


8 


270 © Sylt, 730918, Ak = 1/220 m7! 
rm 


26a 8 Noordwijk, 760323, Ak=1/170 m7 


210 F net 


180 


0 - —e — Ll Uhre ts =} 
LS 6 NT 8 IO ee 2618 ez O Nn 22 226 26 SONS AES 
k=mAk 
39 
a 
FIGURE 17. The mean direction of the waves relative 


to true North, as function of wave number. 


come the influence of the geometry of the coastline. 
For the higher wave numbers the distance to shore 

is relatively long and the non-linear interactions 
may have produced the observed directional distribu- 
tion functions which indeed seem to be hardly af- 
fected by the asymmetry of the coastline. The ob- 
servations therefore may still be consistent with 
the theory of non-linear interactions and the con- 
clusions of Hasselmann et al. (1976) if the relevant 
space and time scales are considered. 

In an "ideal" generation case the directional 
distribution of the wave energy is often approximated 
with a simple unimodal function. The observed situa- 
tions are distinctly multi-modal, but one such func-— 
tion, given in Eq. 14, has been fitted to the data. 
This was done mainly to compare the results with the 
published data. 


lL WG = A) a= 8 


Di(e) = Se We ay cos 2s (Come) (14) 


In this expression s is the spreading parameter and 
8m is the mean direction, both of which may vary 
with k. The values of 8,, and s have been computed 
using a least-squares technique. The results for 
8m as a function of wave number are given in Figure 
15. Noise in the spectra (see Appendix) did influ- 
ence these results and outliers had to be identified. 
As a criterion for acceptation, the rate of change 
of 6, along the wave number axis has been chosen. 
An accepted value of Om should be within 30° of its 
neighboring values on the wave number axis.* This 
is equivalent to a rate of change of approximately 
0.0024 m for the first sortie, 0.0033 m for the 
second sortie, and 0.0031 m for the third sortie. 
This allows for slow but significant variations in 
6m which is required, for instance, in the spectrum 
of the third sortie. The resulting set of accepted 
values of 6,, is also indicated in Figure 17. The 
values of s at the corresponding values of the wave 
number have been plotted in Figure 18 in a format 


* 
The value of 30° was chosen arbitrarily. 


suitable for a comparison with data published by 
Mitsuyasu et al. (1975). 

Mitsuyasu et al. (1975) presented results of a 
number of measurements (five) which were carried 
out with a cloverleaf buoy at several locations 
around the Japanese islands. The observed wave 
fields were generated by various types of wind 
fields, including on-shore and off-shore winds. It 
appears from the ratio of the wind speed to the phase 
speed of the peak frequency of these observations, 
that the state of development of the wave fields was 
rather advanced (the ratios ranging from 0.75 to 
1.25). Based on the observed values of s, relation- 
ships in the frequency domain were suggested. The 
relevant expressions have been transformed here to 
the wave number domain to produce Eqs. 15 and 16. 


s = k7}-25 for k 2 
S| = k259 for k <1 (15) 
= 268) 
s 1.5 (u/c) (16) 


where § = S/S and k = k/ky, a is the maximum value 
of s, ky is the peak wave number, cm is the phase 
speed of the peak wave number, and U is the wind 
speed. The data of Mitsuyasu et al. (1975) are 
probably obtained in situations where tidal currents 
were negligible and in the above transformation the 
deep water linear relationship between frequency and 
wave number was used. 

Equations 15 and 16 are also plotted in Figure 18 
and the agreement is fair, the scatter being on the 
same order of magnitude as the scatter in the data 
of Mitsuyasu et al. (1975). The values of sp com- 
puted from the stereo data are 6.0 for the spectrum 
off Sylt, 5.0 for the first spectrum off Noordwijk. 
These are also in fair agreement with the values 
suggested by Mitsuyasu et al. (1975) which are 4.6 
and 6.1 respectively. However, for the second spec-— 
trum off Noordwijk the observed value of s is 27.4 
whereas the value following from expression 16 is 
5.9. This is a very large discrepancy which is 
possibly due to the rather extreme asymmetry of the 
coastline around the wind direction where the sug- 


© Sylt 730518 
A Noordwijk 760323 
+ Noordwijk 761112 


Sf=15)/Sim 
R= K/Kym 


(74 


OIF 


FIGURE 18. The normalized spreading parameter as a 
function of the normalized wave number. 


567 


gestions of Mitsuyasu et al (1975) may not be ap- 
plicable. 

The above discussion concerned rather overall- 
characteristics of the directional distributions. 
It is planned to investigate these functions more 
in detail. For instance, in the &-spectrum off 
Sylt one aspect which will require closer study is 
the shape of the directional distribution near the 
peak of the spectrum in a sector around the wind 
direction. Two peaks at + and - 15° relative to 
the wind direction can be identified and this phe- 
nomenon seems to be "real" in the sense that the 
directional resolution seems sufficiently high (20°) 
to resolve these peaks in terms of statistical sig- 
nificance. The resonance theory of Phillips (1957) 
predicts a bimodal distribution for frequencies in 
the initial stage of development, but the components 
around the peak have passed that stage and there is 
no relation with this theory. More relevant seem 
to be the theory and calculations of Hasselmann 
(1963), Longuet-Higgins (1976), and Fox (1976) which 
produce a non-linear energy transfer in wave number 
space with two lobes towards the lower wave numbers 
and two lobes towards the higher wave numbers. Fox 
(1976) noted that this function resembles a "butter- 
fly." Also the results of Tyler et al. (1976), who 
observed directional distributions of wind generated 
waves with high-frequency radio-wave backscatter, 
may be of interest since some of the distributions 
have a bimodal character around the mean direction. 


7. CONCLUSIONS 


Three, two-dimensional, wave number spectra have 
been computed from stereophotographic data obtained 
in off-shore wind conditions. The agreement with 
ground-true information is reasonable but some dis- 
crepancy needs to be resolved. 

The directional distribution of the wave energy 
near the peak of the first spectrum is strongly 
asymmetric. In the third spectrum the main direc-— 
tion of the waves differs appreciably from the wind 
direction. It is speculated that these phenomena 
are due to asymmetry in the up-wind coastline. The 
directional distribution functions of the second 
spectrum are more symmetric and unimodal, at least 
in an overall sense. 

A bimodality in a sector around the wind direction 
is observed near the peak of the first spectrum. 
This bimodality may be related to a multi-modal non- 
linear interaction in the spectrum. 

The observed normalized directional spreading 
parameter as function of a normalized wave number 
is in fair agreement with published data. The ab- 
solute values are about 30% larger for the first 
spectrum and about 20% lower for the second spectrum. 
The values for the third spectrum are almost five 
times too large. This may be due to the rather 
extreme asymmetry of the coastline where a compari- 
son with the published data may not be proper. 

The results reported herein are preliminary. 
Additional analysis of available data is being 
carried out. 


ACKNOWLEDGMENTS 


The helicopters were provided by the Royal Nether- 
lands Air Force and they were flown by the Search 
and Rescue team of Soesterberg airbase (the Nether— 


568 


lands). This is gratefully acknowledged. Consider- 
able support in terms of logistics, groundtruth 
data, meteorological observations, etc. was received 
from colleagues in the framework of JONSWAP and this 
is greatly appreciated. 


NOTATION 


A area of spatial integration 
Cm phase speed of component fp 
Cg group velocity 
D(8) standard directional distribution function 
E(k) spectral density in k-space 
spectral density in k,8-space 
spectral density in f,8-space 
E(£) spectral density in f-space 
£ frequency 
g acceleration due to gravity 
ES instantaneous surface elevation 
H (k) Fourier transform of surface elevation 
J Jacobian 3) 
k wavenumber vector k = (kx, Ky) 
k wavenumber, modulus of wavenumber vector 
km wavenumber at peak of wavenumber vector 
spectrum of locally generated wind sea 
dimension of area of analysis in x-direction 
dimension of area of analysis in y-direction 
number of degrees of freedom 
number of transformations 
boundary of spatial integration 
directional spreading parameter 
maximum value of s 
dimensionless spreading parameters s/sp 
windspeed at 10 m elevation 
tidal current vector 
magnitude of V 
place vector x= (x,y) 
ASP EB spatial coordinates 
increment 
direction, orientation of wavenumber vector 
orientation of tidal current 
mean direction 
standard deviation of helicopter yaw 


i 
* 


2 Q YS et J 


TOEPXK ¥I am 


QD 
=] 


REFERENCES 


Bendat, J. S., and A. G. Piersol (1971). Random 
Data: Analysis and Measurement Procedures, 
Wiley-Interscience, New York. 

Brummer, B., D. Heinrich, L. Kriigermeyer, and D. 
Prim (1974). The Large-Scale Weather Features 
over the North Sea during the JONSWAP II Experi- 
ment. Berichte des Instituts fur Radiometeor- 
ologie und Maritime Meteorologie, Universitat 
Hamburg, Institut der Frauenhofer Gesellschaft, 
24. 

Cote, L. J., J. O. Davis, W. Marks, R. J. McCough, 
E. Mehr, W. J. Pierson, J. F. Ropek, G. Stephen- 
son, and R. Vetter (1960). The Directional 
Spectrum of a Wind-generated Sea as determined 
from Data obtained by the Stereo Wave Observation 
Project. Meteorological Papers, 2, No. 6, New 
York University. 

Crawley, G. B. (1975). Automatic contouring on the 
Gestalt photomapper, testing and evaluation. 
American Society of Photogrammetry, Workshop IIT, 
San Antonio, Texas, U.S.A. 


Fox, M. J. H. (1976). On the non-linear transfer of 
energy in the peak of a gravity-wave spectrum. 
II, Proceedings Royal Society of London, A. 348, 
467. 

Hasse, L., M. Griinewald, and D. E. Hasselmann (1977). 
Field observations of flow above the waves. Pre- 
print from the Proceedings of the NATO-Symposium 
on "Turbulent Fluxes through the Sea Surface, 
Wave Dynamics and Prediction," Bendol, to be 
published by Plenum Press (New York, London) . 

Hasselmann, K. (1963). On the non-linear energy 
transfer in a gravity-wave spectrum. Part 3. 
Evaluation of the energy flux and swell-sea in- 
teraction for a Neumann spectrum. Journal of 
Fluid Mechanics, 15, 385. 

Hasselmann, K., R. P. Barnett, E. Bouws, H. Carlson, 


D. E. Cartwright, K. Enke, J. A. Ewing, H. Gienapp, 


D. E. Hasselmann, P. Krusemann, A. Meerburg, P. 
Muller, D. J. Olbers, K. Richter, W. Sell, and 
H. Walden (1973). Measurements of Wind-Wave 
Growth and Swell Decay during the Joint North 
Sea Wave Project (JONSWAP). Ergd&nzungsheft zur 
Deutschen Hydrographischen Zeitschrift, Reihe 

A (EO), Wis 1A 

Hasselmann, K., D. B. Ross, P. Muller, and W. Sell 
(1976). A parametric Wave Prediction Model. 
Journal of Physical Oceanography, 6, 2; 200. 

Holthuijsen, L. H., M. Tienstra, and G. J. v.d. 
Vliet (1974). Stereophotography of the Sea Sur- 
face, an Experiment, Proceedings of the Inter- 
national Symposium on Ocean Wave Measurement and 
Analysis, American Society of Civil Engineers, 
alsyako 

Huhnerfuss, J., W. Alpers, and L. Jones (1978). 
Measurements at 13.9 GHz of the radar backscat- 
tering cross section of the North Sea covered 
with an artifical surface film to be published 
in Radio Science. 

Longuet-Higgins, M. S., D. E. Cartwright, and N. D. 
Smith (1963). Observations of the directional 
spectrum of sea waves using the motions of a 
floating buoy. Ocean Wave Spectra, 111-132, 
Prentice Hall, Inc., New Jersey. 

Longuet-Higgins, M. S. (1976). On the non-linear 
transfer of energy in the peak of a gravity-wave 
spectrum: a simplified model. Proceedings 
Royal Society of London, A. 347, 311. 

Mitsuyasu, H., F. Tasai, T. Suhara, S. Mizuno, M. 
Ohkusu, T. Honda, and K. Rikiishi (1975). Ob- 
servations of the Directional Spectrum of Ocean 
Waves Using a Cloverleaf Buoy. Journal of 
Physical Oceanography, 5, 750. 

Panicker, N. N., and L. E. Borgman (1970). Direc— 
tional Spectra from Wave Gage Arrays. Proceed- 
ings of the 12th International Conference on 
Coastal Engineering, Washington, D.C., p. 117. 

Phillips, O. M. (1957). On the generation of waves 
by turbulent wind. Journal of Fluid Mechanics, 
2 aie 

Singleton, R. C. (1960). An algorithm for computing 
the Mixed Radix Fast Fourier Transform, IEEE 
Transactions on Audio and Electroacoustics, AU-17, 
An \S)sh< 

Spiess, F. N. (1975). Joint North Sea Wave Project 
(JONSWAP) progress - an observer's report. Re- 
port ONRL-C-8-75, Office of Naval Research, 
London. 

Seymour, R. J. (1977). 
on restricted fetches. 
ican Society of Civil Engineers. 


Estimating wave generation 
Proceedings of the Amer- 
Journal of the 


Waterway, Port, Coastal and Ocean Division, WW2, 
paper 12924, p. 251. 

Stilwell, D. (1969). Directional Energy Spectra of 
the Sea from Photographs. Journal of Geophysical 
Research, 74, 8; 1974. 

Sugimori, Y. (1975). A study of the application of 
the holographic method to the determination of 


APPENDIX 


NOISE 


Inspection of contour-line maps of the sea surface 
obtained from the observation off Sylt revealed a 
dome-shaped distortion. This distortion is probably 
caused by the fact that the pictures could not be 
positioned in the stereoscopic viewing devices with 
the accuracy normally obtained with high grade pic- 
tures. When this positioning is not optimal, a 
dome-shaped distortion is to be expected. Unfor- 
tunately the exact distortion cannot be determined, 
but in the k-plane it seems to be well separated 
from the wave information (area No. 1 in Figure 19) 
and the data in this area was removed in the sub- 
sequent analysis. 

The other noise-affected areas are related to a 
phenomenon introduced by the manner of scanning 
the pictures during the photogrammetrical process: 
the sea surface elevation at even-numbered lines 


wavenumber [m~] 
t ky 


© componen 


o, L 


ao 


569 


the directional spectrum of ocean waves. 
Sea Research, 22, 339. 

Dyllew,, GCG. Lie, Cu (Ce. Teague, Ro H. Stewart), A. Ms 
Peterson, W. H. Munk, and J. W. Joy (1974). Wave 
directional spectra from synthetic aperture ob- 
servations of radio scatter. Deep-Sea Research, 
21, 989. 


Deep- 


(scanned in positive y-direction) is systematically 
slightly too low, while the elevation at odd- 
numbered lines (scanned in negative y-direction) 

are systematically slightly too high. This effect 
has been observed earlier in the analysis of stereo 
photos of regular waves generated in a hydraulic 
laboratory. The principal wave length and direction 
of this distortion correspond with the location of 
area No. 2 in Figure 19, which is the location of 
the Nyquist wavenumber in x-direction. This spectral 
information was removed from the spectra in the sub- 
sequent analysis. The noise in areas No. 3, 4, and 
5 was labeled as such mainly because of the delta- 
type behavior of the directional distribution func- 
tions in these regions. It is probably due to 
variations in the error introduced by the scanning 
and possibly also by "leakage" from area No. 2. In 
the k-spectrum off Sylt this noise was not removed. 
In the k-spectrum off Noordwijk the noise in the 
indicated region in Figure 11 has been removed. 


4 


areaS (ee 


—1____1 
ay | GS | 0.01 002 003 004 005 006 007 008 ao9 010 


wavenumber component k,[m"] 


IGURE 19. Location of noise in 


F 
> 
k-plane. 


Gerstner Edge Waves in a Stratified Fluid 
Rotating about a Vertical Axis 


Erik Molo-Christensen 
Massachusetts Institute of Technology 
Cambridge, Massachusetts 


ABSTRACT 


An exact solution is obtained for edge waves along 
one inclined planar boundary in a fluid rotating 
about a vertical axis. The solution is based on a 
modification of Gerstner's rotational waves, and 
includes the effect of mean drift. The solution re- 
duces to Yih's edge wave solution for zero rotation 
and to Pollard's rotational deep water Gerstner waves 
in rotating flow. Satellite observations of sea sur- 
face are shown which reveal patterns similar to those 
which would be generated by Gerstner edge waves. 


1. INTRODUCTION 


The early, exact solution by Gerstner (1802, see 
1932, p. 419) was rediscovered by Rankine (1863), 
discussed by Lamb (1932), found to be valid for free 
surface waves in an arbitrarily stratified flow by 
Dubreil-Jacotin (1932), further modified to describe 
edge waves by Yih (1966), and free surface waves in 

a rotating flow by Pollard (1970). However, there 
has been a tendency to dismiss Gerstner waves as of 
limited applicability to phenomena in nature. As 
Lamb (1932) has pointed out, the generation of 
Gerstner, free surface waves by the application of 
surface stresses requires a certain mean vorticity 
distribution to exist in the fluid. It can be argued 
that in a nonrotating fluid of uniform density it is 
difficult to conceive how the required vorticity dis- 
tribution can be established. However, in a strati- 


fied and rotating fluid, there are mechanisms capable 


of generating vorticity without viscous diffusion. 
In a stratified fluid, the baroclinic term, express- 
ing the action of a pressure gradient normal to a 
density gradient in generating vorticity will be 
capable of establishing a horizontal vorticity field. 
In a rotating fluid, the effects of vortex stretch- 
ing and compression can establish distributed vertical 
vorticity. 

There, in a rotating stratified flow, waves simi- 


570 


lar to Gerstner waves are more likely to be encoun- 
tered. In fact, the uniform flow, usually assumed 

as the mean flow on which small perturbation waves 
may ride, would be less likely to occur in a rotating 
stratified fluid. But the small perturbation solu- 
tions for waves, as well as exact, finite amplitude 
solutions, are all useful as approximate descriptions 
of real phenomena and actual observations. 

If such solutions do not fit the exact circum— 
stances, they can possibly serve as starting points 
for perturbation expansions. Furthermore, we may 
learn about some of the special features of finite 
amplitude exact wave solutions; there is a tendency 
to forget some of these facts when preoccupied with 
linear wave solutions. 

In the following, I shall present a Lagrangian 
description of an edge wave field, point out where 
it differs from previous solutions, and develop the 
dispersion relation for the waves. 


2. COORDINATE SYSTEMS AND DISPLACEMENT FIELD 
Coordinate System 


The waves propagate in the x - direction, normal to 
the plane of Figure 1. In the planes normal to the 
x - direction we define the oyZ- coordinates, with 
o0Z vertical and the oyz-coordinates, with oy in the 
plane of the inclined boundary, inclined at an angle 
a with the vertical. The particle motion will be in 
planes parallel to xy. 

While Yih (1966) could let the amplitude of 
particle motion decay with negative y-distance, and 
Pollard (1970), for deep water waves away from a 
side boundary, made the obvious and correct choice 
of letting the particle motion decay with decreasing 
vertical position; here I have to make a different 
choice. The amplitude of particle motion will decay 
along a direction - or, shown in Figure 1 as another 
coordinate system, ors. 


—> n> DO] 


FIGURE 1. Coordinate system, looking along the direc- 
tion of wave propagation, ox, and along the labeling 
coordinate direction, oq. 


Displacement Field 


Using labeling variables, q, r, s, to identify fluid 
particles, define the field of particle positions in 
terms of 1, r, s and time, t, as follows: 


x = q + Ut - a (exp mr) sin (kq - ot) (1) 
y =x cos 8 - s sin B 
+ a (exp mr) cos (kq - ot) (2) 
BS 1 Sali 5} a S Cos § (3) 
for m SIRs O 


U is a constant mean particle velocity in the x- 
direction, a is an oscillation amplitude parameter, 
m is an inverse decay distance measure, K is wave- 
number and o is the frequency of particle motion. 

First consider the kinematics of wave motion, 
next find the condition for incompressibility before 
proceeding to apply dynamics to give the dispersion 
relation. A surface defined by letting r be a func- 
tion of s will have waves that proceed in the x- 
direction. For example, a string (line) of particles 
defined by fixed values of r and s will have maxima 
in y-displacement at 


kq - ot = 2ntT (4) 


From Eq. 1, substituting for q from Eq. 4 gives the 
x-positions of crests to be at 


Ke aee = [2nt + (o + Uk)t]/k (5) 


The crests move at a speed of 


c = (o + Uk)k = w/k (6) 


571 


w is the wave encounter frequency, and differs from 
the particle oscillation frequency by the Doppler 
SHEsEe Uke 
Mass Conservation 
The displacement field defined by Eqs. 1, 2, and 3 
can be made to satisfy the requirement that the 
density of a fluid particle is independent of time 
by requiring that the Jacobian: 
d(x,y,2)/9(q,xr,S) 
= 1 - a@km (exp2mr) cos 8 
+ (m cos 8 - k) a(exp mr) cos (kq - ot) (7) 
is independent of time. This requires 
k =m cos £ (8) 
Now proceed to apply the momentum equations to cal- 
culate the pressure, which in turn will be set con- 
stant at the free surface. 
3. PRESSURE FLUCTUATIONS 
The momentum equation in Lagrangian variables gives, 
for the derivative of pressure with respect to the 


labeling variable q: 


“Pg/P = (% + z £ sina - y f cos Oe 
+ (¥ + x £ cos ON, + (2 - x f sin Ne 


(9) 


N> 


ar Cj g 
The equations for the r and s-derivatives are 
similar. f = 22 is the angular velocity of rotation 
of the coordinate system, the angular velocity being 
vertical as mentioned before. Substituting for x, 
Vin and ez etromeEqs..15,, 27) sand 3) into) Eq= 197) onesob= 
tains: 


-Pg/P [o2 - £ cos a(o + Uk) 


-gk sin a] a (exp mr) sin 6 (10) 
-p,/p = - [0% - f 0 cos aja? exp2mr 
+ [-o2 cos 8 + £ o cos(a + 8) 
+ fUm cos a+ gm sin aJa(exp mr) cos 6 
+ fU cos (a + 8) + g sin (a + 8) (11) 


-ps/0 = [o2 sin B - £ o sin (a + 8)]a(exp mr) cos 6 


= £Ul san) (0) +18) +g) cosh (a7 298) (12) 


where § = kg - ot is the phase of particle oscilla- 
tion. 

At the free surface, which consists of particles 
with a specified relation between r and s, and with 
values of labeling variable, q, from - ~ to + ~, the 
pressure must be independent of q and t. This is 
satisfied, as can be seen from Eqs. 10, 11, and 12, 


572 


if the pressure is independent of phase 6, and Pq 
Py, and Pp, are independent of 6. 
From Eq. 12, pz is independent of 6 when 


cot 6 = an = (elon (3 (13) 
f sina 


Since a is given by the slope of the boundary, Eq. 
13 gives 8 for a given o and a. Equation 10 shows 
Pq to be independent of 8 when 


o* - £ cos a(o + Uk) - g k sina = 0 (14) 


For a given value of o, Eq. 14 yields k, and m is 
then found from Eqs. 8 and 13. 

This leaves Eq. 11 unused, but it can be shown 
that the requirement that Py be independent of 6 is 
not independent of Eqs. 13 and 14. Equation 11 also 
shows that there will be a mean pressure gradient 
across the wave propagation direction, proportional 
to a*. This is a nonlinear effect of the presence 
of waves. 


4. DISCUSSION 


The equivalent to a linear dispersion relation con- 
sists of Eqs. 8, 13, and 14, relating particle fre- 
quency, 0, decay direction angle, 8, horizontal 
wavenumber, k, and decay parameter, m, with f, a, 
and U as parameters. 

Note that the introduction of a mean drift veloc- 
ity, U, has a now-trivial effect on dispersion, as 
can be seen from Eq. 14, where the effect is not a 
simple Doppler shift in frequency. The equations 
of rotating fluids are not invariant to Galilean 
transformations. Also note that the dispersion is 
independent of the amplitude parameter, a; this is 
an unexpected result for non-linear waves. But the 
amplitude of particle motion parallel to oy is really 
a exp[2mR(s)], where R is the value of r at the sur- 
face. Since m is found from the equations involved 
in determining dispersion, one cannot really claim 
that dispersion is independent of amplitude. 

With the dependence on phase, 9, eliminated in 
Eqs. 10, 11, and 12 by satisfying the dispersion 
relations, one can see that the mean surface slope 
across the wave propagation direction will vary with 
wave amplitude and with y- position. 

As pointed out by Dubreil-Jacotin (1932), and 
later by Yih (1966) the results are valid for a 
fluid of arbitrary stable density stratification. 

The solutions given here can be further extended 
to replace the free surface by an interface between 
the given flow field and a homogeneous wave trapped 
fluid, giving the gravitational billows described 
elsewhere [Mollo-Christensen (1978)]. This will re- 
place the acceleration of gravity, g, by g' = 
g(Ap/p), where Apis the density difference between 
the two fluids and p the density of the lower fluid 
at the interface. 

Similarly, the flow field at the off-shore or 
inside end may be bounded by a field of geostrophic 
billows or a combination of gravitational and geo- 
strophic billows [see Mollo-Christensen (1978)]. 


FIGURE 2. High-passed and contrast enhanced satellite 
infrared images from January 27, 1975, at 1600, 1700, 
and 1800 hrs., GMT. Florida on the right side, Gulf 
Coast on top. 


5. SOME EXAMPLES OF OBSERVATIONS OF FINITE AMPLITUDE 
WAVES ALONG A SLOPING BOUNDARY 


By processing satellite data on sea surface infrared 
emission one can see moving patterns of sea surface 
temperature in the Gulf of Mexico between the con- 
tinental shelf edge and the coast. 

A sequence of processed satellite images taken 
one hour apart is shown in Figure 2. Because the 
mean current, U, at the time of observation is not 
known, one cannot say whether these waves satisfy 
the dispersion relations for the kind of edge waves 
discussed here. All one can say at this point is 
that it appears possible to satisfy the dispersion 
relations given with wavelengths, bottom slopes, 
and currents of reasonable orders of magnitude, but 
one needs to refine the observations further before 
one can reach any definite conclusions. 


6. CONCLUSIONS 


Nonlinear edge waves of finite amplitude can have 
dispersion relations defined by a set of equations 
relating particle oscillation frequency, encounter 
frequency, wave number, and other parameters in a 
way that can be solved systematically if one starts 
by specifying a suitable wave variable, in the pres- 
ent case, frequency. 

The observations which inspired the present anal- 
ysis show Gerstner edge waves or possibly waves of 
a different kind; one cannot tell with the evidence 
now at hand. 


573 


ACKNOWLEDGMENT 


The research reported here was supported by the 
Office of Naval Research under Contract No. NOO014- 
76-C-0413. The observations cited were made with 
support from the Office of Naval Research under 
Contract No. NO0014-75-C-0291. The satellite images 
were processed using the facilities of Air Force 
Geophysics Laboratory, Lincoln, Mass. 


REFERENCES 


Dubreil-Jacotin (1932). Sur les oudes de type 
permanente dans les liquides heterogenes. Atti. 
Accad. Lincei. Rend. Cl. Sci. Fis. Mat. Nat., 6, 
15; 814-819. 

Gerstner, F. (1802). Theorie der Wellen. Abh. d. 
Koénigl. Boéhmische Ges. d. Wissenschaften zu Prag 
fur das Jahr 1802. 

Lamb, H. (1932). Hydrodynamics, 738 pp. 
New York, 1945. 

Mollo-Christensen, E. (1978). Gravitational and geo- 
strophic billows, some exact solutions. J. Atmos. 
Sci. To be published. 

Pollard, R. T. (1970). Surface waves with rotation: 
an exact solution. J. Geophys. Res., 75, 5895- 
5898. 

Rankine, W. J. M. (1863). On the exact form of 
waves near the surface of deep water. Phil. 
Trans., 127-138. 

Yann Co S35 (SG). 
fied fluid. 


Dover, 


Note on edge waves in a strati- 
J. Fluid Mech., 24, 765-767. 


The Origin of the 
Oceanic Microstructure 


Gia abo 


Barenblatt and A. S. Monin 


P. P. Shirshov Institute of Oceanology 


Moscow, USSR 


ABSTRACT 


Microstructure of hydrodynamical fields, a well- 
known phenomenon in the ocean, is attributed to the 
formation and development of turbulent spots gener- 
ated due to the loss of stability or breaking of 
internal waves. Under some general assumptions the 
relations are obtained governing the development of 
turbulent spots at various 'stages of their evolution 


It is shown that the longest and slowest stage of the 


extension of a turbulent spot is the final, viscous 
one. Simple self-similar laws of the extension of 
turbulent spots are obtained for this stage and com- 
pared with experiment. Long-standing turbulent 
layers of the "blini" shape, sharply bound by am- 
bient non-turbulent stratified fluid, are identified 
with turbulent spots of the above-mentioned origin 
which are in the final viscous stage of their evolu- 
tion. The relations are also obtained governing 
viscous intrusion of the bottom seawater into the 
body of the ocean. 


1. INTRODUCTION 


Under strongly stable stratification, turbulent mix- 


ing is inhibited due to large losses of the turbulent 


energy for the work against the buoyancy forces. 
der natural conditions, therefore, turbulence cannot 
be present in the whole body of the fluid during 
rather long periods of time [Woods (1968), Monin et 
al. (1977), Federov (1976)]. In fact, it is concen- 
trated only in separate turbulent layers having the 
shape of "blini," vertically quasi-homogeneous due 
to mixing, and separated by thin streaks with micro- 
jumps of temperature, electrical conductivity, sound 
velocity, salinity, density, refraction index, and 
other thermodynamic parameters of sea water some- 
times accompanied by microjumps of flow velocity. 
Such thin-layered vertical structure, which is ap- 
parent from inhomogeneities ("steps") on the verti- 
cal profiles of density and other thermodynamic 


Un- 


574 


parameters (see schematic drawing in Figure 1) or 
even more sharply from multiple peaks on the pro- 
files of vertical gradients of these parameters, is 
called microstructure or fine structure of hydro- 
dynamical fields. Numerous measurements performed 
using the method of continuous vertical sounding in 
the cruises of the research vessels of the Institute 
of Oceanology, USSR Academy of Sciences, and re- 
search vessels of other countries showed that the 
microstructure exists always and everywhere in the 
World Ocean (the lack of microstructure may be ex- 
pected only for the regions of macroconvection which 
occur rather seldom in the ocean, at least in the 
low and temperate latitudes). 

Smoothing over the microstructural "steps" on 
the profile of a thermodynamic parameter, e.g., 
density or temperature, we obtain a smooth curve 
characterizing large-scale stratification of the 
ocean (gross-stratification). We have to emphasize 
that from the point of view of the Richardson cri- 
terion gross-stratification is nearly always stable 
- the Richardson number computed for it, Ri(z), as 
a rule, is essentially larger than its critical 
value, 1/4. How can the turbulence be generated 
under such conditions? Graphs of Ri(z), taking 
into account the "steps" of microstructure, show 
values of Ri < 1/4 in several layers of the micro- 
structure - apprently in these very layers, at the 
momeht of sounding, the generation of small-scale 
turbulence took place (in other layers where Ri > 
1/4 turbulence decayed with time). The appropriate 
conditions for local generation of turbulence at 
stable gross-stratification may be created by in- 
ternal waves. Indeed, in the field of internal 
waves in the regions near their crests and hollows 
the local values of the Richardson number can be 
reduced lower than the critical value, 1/4, and the 
turbulence spots would then be formed there. The 
internal waves can also break. For the turbulent 
spots formed after the breaking of internal waves, 
the formation is characteristic of continuous spec- 
trum, i.e., of developed turbulence immediately 


FIGURE 1. 
chronous vertical distribution of 
density and shear in the ocean. The 
dashed line shows the shear distri- 
bution for intrusions. 


Schematic form of syn- 


after the breaking [Belyaev et al. (1975)]. 

The evolution of a newly-formed turbulent spot 
appears to be the following. The turbulent mixing 
makes the spot vertically quasi-homogeneous, there- 
fore, within the spot the density of the water be- 
comes uniform. For stable stratification, when the 
density grows with depth, the density in the upper 
half of the mixed spot is higher and in the lower 
half of the spot lower than at the same levels in 
ambient fluid. Therefore, under the action of the 
buoyancy forces, the upper half of the spot should 
go down and the lower half of the spot should rise 
to its middle level. Therefore, the spot should 
"collapse," simultaneously spreading and transform- 
ing itself into a thin "blin." The intrusion of 
such a "blin" into the body of surrounding strati- 
fied fluid creates in it a new layer of microstruc-— 
ture. 

If the initial internal wave has a long period 
and wave length (e.g., internal waves with tide 
periods may be generated by tide forming forces 
and tides themselves) turbulent spots formed by 
this wave are large and corresponding turbulent 
layers are very thick. Internal waves of smaller 
periods and lengths may develop on these layers 
forming turbulent spots of smaller sizes and layers 
of microstructure of smaller thicknesses, etc.; 
internal waves of minimum periods and lengths, 
turbulent spots of minimum sizes and layers of 
microstructure of minimum thicknesses. Thus, the 
answer to the question "which came first, the chicken 
or the egg?" consists for this case in the indica- 
tion of a cascade process "internal waves > turbulent 
spots > layers of microstructure +> internal waves 
etc." This cascade process may lead to the forma- 
tion of a quasi-steady spectrum of internal waves, 
intermittent turbulence, and layers of microstructure 
(although in real nature the action of some other 
processes influencing real spectra is possible, in- 
cluding storms and quasi-steady horizontal inhomo- 
geneities of geographic and dynamic origin). The 
turbulent spots also take part in a rising cascade 
generated by local instabilities of available shear 
flows, breaking of surface waves, sinking of cooled 


575 


fluid from the turbulized surface layer, etc. As 
distinct from the classical Kolmogorov cascade in 
non-stratified fluid, here, in passing from a larger 
scale to a smaller one, the energy is not preserved, 
being left in turbulent spots in the final stage of 
their evolution where internal waves do not gener- 
ate. Thus, in stratified fluid turbulent spots of 
various scales are continuously generated and the 
process of their evolution is of considerable 
interest. 

The first stages of the evolution of turbulent 
spots* where the radiation of internal waves takes 
place are rather short: by estimates of J. Wu 
(1969) and T. W. Kao (1976) they come to an end in 
a time interval of the order of several tens of 
n7! (N is the Brunt-Vaisdla frequency) after the 
beginning of the process. The final stage of the 
evolution of turbulent spots is much longer. This 
stage is much less known: in the paper of J. Wu 
(1969) concerning this stage it is mentioned only 
that viscosity is of significance at this stage and 
it is noted that the profile of the spot is pre- 
served during this stage. The analysis presented 
here shows that the velocity of the extension of 
turbulent spots at the viscous stage is essentially 
lower than at the initial stages. It is our opin- 
ion that the "blini"-shaped turbulent structures 
are the intrusions of the turbulent spots of various 
scales into surrounding stratified fluid which are 
mainly at the final stage of their evolution. 

Thus, let a turbulent spot (Figure 1) be formed in 
a stable continuously density-stratified (linearly 
for definiteness) fluid due to some reason (breaking 
of internal waves, local loss of stability of shear 
flow, penetration of denser fluid from the turbulent 
surface layer, etc.). The density of fluid within 
the turbulent spot due to mixing is uniform in con- 
trast to an ambient continuously stratified fluid 
being in a state of rest or laminar motion. Certain 
potential energy is stored due to mixing in the tur- 
bulent spot, so the state of the mixed fluid- 
stratified environment system ceases to be in 
equilibrium. Mixed turbulent fluid starts to strike 
(Figure 2) into stratified non-turbulent fluid by 
tongues - "intrusions" which are formed at the 
level, z = 2), (z is the vertical coordinate) where 
the density of stratified fluid is equal to the 
density of mixed fluid. 

Potential energy, stored by the fluid at initial 
turbulization and mixing in the spot, dissipates 
during the intrusion of mixed fluid into stratified 
non-turbulent fluid. It is natural to consider 
three stages of the evolution of the spot: 

(1) Initial stage of free intrusion. The motive 
force of the intrusion at this stage exceeds greatly 
the drag forces. The turbulent spot extends slightly 
but the internal waves are intensively formed by the 
spot. 

(2) Intermediate steady state. The motive force 
at this stage is balanced mainly by form drag and 
wave drag due to radiation of internal waves by an 
extending turbulent spot. The acceleration of the 
tongue is negligible. 


*the classification of stages of the evolution of the spot of 
mixed fluid in the continuously density-stratified fluid goes 
back to the fundamental work of J. Wu (1969) where the ex- 
perimental investigation of the initial stages of this process 
was performed for the wake of circular initial cross-section. 
T. W. Kao (1976) performed semi-empirical theoretical investi- 
gation for the initial stages of the evolution of such wakes. 


576 


FIGURE 2. The intrusion of a turbulent spot into con- 
tinuously stratified fluid. 


(3) Final viscous stage. The motive force is 
balanced at this stage mainly by viscous drag. 

Of course, between the first and second and the 
second and third stages there exist intermediate 
transitional periods. When the third stage comes 
to the end the spot is mixed due to diffusion with 
ambient fluid and disappears. 

The turbulent motion inside of the intrusion 
tongue is supported by general shear stress together 
with eddy motions inside of the intrusion due to the 
difference of the velocities of the tongue and en- 
vironmental non-turbulent fluid. The boundary of 
turbulent and non-turbulent fluid is sharp and if 
the thickness of the intrusion is not too small, 
the shear required for supporting the turbulence 
within the intrusion is not large. 

Indeed, let us consider the equation of the 
balance of turbulent energy in a shear flow of 
stratified fluid neglecting, as usually, the viscous 
transfer term [Monin and Yaglom (1971)-] 


3,5 + 3 {w'E' + p'w'} 


= - pw'g - pe - p u'w' au (1) 


Here t is the time, E the turbulent energy of 
unit mass, € the dissipation rate per unit mass, 
u the longitudinal and w the vertical velocity 
components, p the pressure. The flow is considered, 
for the estimates we need, as horizontally homogen- 
eous and the Boussinesque approximation is accepted, 
i.e., the density variation is taken into account 
only if it is multiplied by very large factor - 
gravity acceleration g. 

Let us accept for the terms of the equation of 
balance of turbulent energy, the Kolmogorov approxi- 
mations [Monin and Yaglom (1971) ] 


wwii} ap joy eS pvp. 38 


u'w' = - 2vB aa, é = y'*p3/272 (2) 


Here 8 = E/p is the mean turbulent energy per 
unit mass, 2 the external turbulent scale. Thus, 
the equation of balance of turbulent energy takes 
the form 


3,8 = a LB 3,8 - pw'g/p 
+ 278 (2 a)? - y'tB3/272 (3) 


The mathematical nature of sharp interface between 


the turbulent and the non-turbulent regions becomes 
completely transparent from this equation. In fact, 
Eq. (3) is a non-linear equation of heat conductivity 
type with heat inflow where the coefficient of trans- 
fer of turbulent energy equal to ave tends to zero 
with turbulent energy itself. For such equations 
under zero initial conditions the disturbed region, 
in contrast to the linear heat conductivity equation, 
is always finite; this explains (cf. below) mathe- 
matically the existence of a sharp interface between 
the turbulent and the non-turbulent regions. 

It is important that, due to mixing following the 
generation of a turbulent spot, the losses of turbu- 
lent energy for the work of suspending a stratified 
fluid [the second term of the right-hand side of the 
Eq. (3)] disappear because the density within the 
spot becomes uniform. Furthermore, the first term 
of the right-hand side of (3) governs the diffusional 
transfer of turbulent energy within the mixed region 
and does not influence the averaged, through the 
spot, value of turbulent energy. Therefore, the 
decay of turbulent energy within the spot is governed 
by the balance of the two last terms of the right- 
hand side of the equation (3) representing genera- 
tion and dissipation of turbulent energy, 
respectively. 

It seems natural to accept that the external scale 
of turbulence 2, within a factor of the order of 
unity, coincides with the transverse size of the 
tongue of intrusion h; the constant y by estimates 
has a value of about 0.5. Thus, the shear d,u ~ 
VB/h is sufficient to support the turbulence within 
the spot at a steady level together with the state 
of mixing within the spot. If h has the value of 
tens of centimeters - one meter or more, then for 
the value VB ~ 1 cm/sec, characteristic of oceanic 
turbulence, the shear required for supporting steady 
turbulence is small. In thin layers it is large; 
therefore, the turbulence in thin layers decays 
rather quickly and the spot of mixed fluid exists 
during the time interval required only for the dif- 
fusional mixing of the spot with the ambient strati- 
fied fluid. 

Furthermore, available experimental data show 
(J. Wu (1969)] that turbulent entrainment and the 
erosion of a turbulent spot may be neglected, start- 
ing from a very early stage of the evolution till 
rather late stages of this process. Therefore, we 
shall take the volume of turbulent spot constant at 
all stages of its collapse to be described. 

For simplicity we shall further suppose that the 
initial form of a turbulent spot is symmetric in 
respect to the equilibrium plane where the densities 
of stratified fluid and mixed fluid coincide. 


2. INITIAL STAGES OF THE EVOLUTION OF THE SPOT OF 
MIXED FLUID 


At the first stage, free fall (lifting from below) 
of the particles of mixed fluid to the equilibrium 
plane takes place, followed by the spreading of 
fluid particles along this plane. Therefore, the 
rate of change of the area of horizontal projection, 
S, of a turbulent spot is proportional at this stage 
to the product of the actual area by the rate of 
fluid influx to the equilibrium plane. The latter 
quantity is equal to the product of the acceleration 
of free fall proportional to N2 and time t. Thus, 
we obtain for the initial stage 


dS/dt ~ sN*t (4) 
For small Nt we obtain by integration 


= = ee 
(S S,)/S, Nft (5) 


(Sg is the initial area of horizontal projection of 
the spot). Thus, at the first stage the character- 
istic size of the plan form of the turbulent spot, 
L, changes proportionally to the square of time 


(L-L)/L. ~ n2t2 aL/dt ~ L N2t (6) 
fo) fo} {o) 


[for the wake, S ~ L, and the relation (6) follows 
from (5) in an elementary way; for the spot of the 
circular plan form, S ~ L2, but at (L = Lo) SS) op 
we = Tae ~ 2 (L - Lo)Lo and (6) follows again from 
(5) Ic 

The relations of the type of (6) were obtained 
by J. Wu (1969) from the experimental investigation 
for a spot having the form of a cylinder with a 
horizontal axis; they were confirmed by some nu- 
merical investigations [see Kao (1976)]. Actually 
they were confirmed to be valid to Nt ~ 2.5. 

At the intermediate stage the motive force of 
the intrusion is balanced by form drag and wave 
drag, thus, the velocity of the propagation of the 
intrusion tongue is governed by the parameter of 
stratification - Brunt-Vaisala frequency N - to- 
gether with the actual height of the tongue, h, 
whence by dimensional considerations we obtain 


aL/dt ~ Nh (7) 


We see that at this stage the dependence of the 
velocity of the extension of the intrusion tongue 
is different for various geometries of the problem. 
In fact, the volume of the turbulent spot V is 
constant; for the cylindrical spot h ~ V/LH (H is 
the longitudinal size of the spot) and h ~ V/L2 for 
a spot of the circular plane form. Therefore, we 
obtain for the cylindrical spot 


dL2/at ~ NV/H , L ~ VYNV(t —- to) (8) 


(to is a conditional time moment of the beginning 
of the second stage), whereas for the spot of the 
circular plane form 


3 
aL3/dt ~ NV , L~ YNV(t — to) (9) 


The relations of the type (8) were obtained by 
J. Wu (1969) from the experimental data for collapse 
of a turbulent wake of initial circular cross- 


FIGURE 3. Elementary particle of the diffusion tongue. 


577 


section. They were confirmed to be valid for 
3S ie S BS, 


3. FINAL, VISCOUS STAGE OF THE INTRUSION 


Under accepted assumptions the equation of mass con- 
servation for a mixed fluid takes the following form 
in hydraulic approximation. 


a,h + div (hy) = 0 (10) 


Here h(x,y,t) is the height of the intrusion 
tongue; x,y are the spatial horizontal coordinates, 
t is the time, v is the velocity of fluid displace- 
ment averaged through the height of the tongue. 

For the determination of the velocity, v, let us 
consider the system of forces acting on the cylin- 
drical particle of the intrusion tongue leaning upon 
the area 56S (Figure 3). The motive force of this 
particle is caused by the action of the gradient of 
redundant pressure, P, and spatial variation of the 
height of the tongue of intrusion 


Fm = - grad(ph) 6s (11) 

Furthermore, the drag force per unit area of a 
particle surface due to the viscous character of 
the drag at the final stage of the intrusion under 
consideration is governed by the velocity, v, of 
the particle relative to ambient fluid, viscosity 
of the fluid, u, and particle height, h. The di- 
mensional considerations give the viscous drag 
force per unit area of particle surface proportional 
to uv/h. Therefore, the viscous drag force acting 
on the particle leaning upon the area, 6S, is equal 
to 


Fr = CuvéS/h (12) 


where C is a constant, under given assumptions - a 
universal one. For estimating the constant, C, the 
well-known solution of the problem of viscous flow 
between flat plates may be used. This solution 

gives for the viscous drag the value 12uvéS/h, whence 
C = 12. Equaling drag force to motive force (the 
inertia force, as at the second stage, is supposed 
to be a negligible one) we find 


v = - hgrad(ph) /Cu (13) 


To complete the statement of the problem we have 
to find the redundant pressure in the mixed fluid. 
In stratified fluids the density varies linearly with 
height. The intrusion tongue propagates symmetri- 
cally, thus, the equilibrium plane divides the 
height of the tongue in half. Let us denote by pj 
and ~,, correspondingly, the pressure and the density 
in stratified fluid at the level, 2 = zy. ‘Then, 
evidently, the pressure in the stratified fluid 
varies with depth following the relation 


Pp = pi ~— 919(2 - 2) 
+ p{N2(z - 2)) 2/2 (14) 


Here, as before, N is the Brunt-Vdisdla frequency 
N2 = ag, g is the gravity acceleration, a = (dp/dz)p . 
Thus the pressure at the upper and the lower points of 
a vertical section of the tongue z = z; + h/2 are 
equal, respectively, to 


578 
Ur = 2,2 
P = Pp] — Pigh/2 + p\N°h*/8 
Pp = pj + pigh/2 + p\N2h2/8 (15) 


because at the upper and the lower points the pres- 
sure in the tongue coincides with the pressure in 
ambient stratified fluid. Hence, the pressure 
within the tongue is distributed according to the 
hydrostatic law 


P = Pp) — p1g(z - 21) + p1N*h?/8 (16) 


The pressure averaged over the section of the tongue 
is equal to 


Ki 2y2 
Pog 7 Pl + ei h“/8 (17) 


The pressure averaged in the same way in the strati- 
fied fluid due to (14) is equal to 


22 

= ot 4 18 
Be Pl p1Nch*/2 (18) 
Thus, the redundant pressure entering the expres- 


sion of motive force of the intrusion tongue at a 
given vertical line is 


= = = 2n2/12 
2) See ee = ho/, (19) 


The relations (13) and (19) give 


p,NA aN 


1 1 
= 3) eo 
M 12Cu lore telte) 4cyu 


hegrad(h) (20) 


Putting this expression into the equation of 
mass conservation of mixed fluid (10) we obtain 
for h a non-linear equation of the heat’ conductivity 
type 


a,h - nAh®> = 0 , n = p,N2/20Cy = N2/20Cv = (211) 


Here A is the Laplace operator, v the kinematic 
viscosity of the fluid. In particular, for one- 
dimensional motions Eq. (21) takes the form 


d.h - nd*2 h°2 = 0 (22) 
iS xx 


= 5 = 
dh n(1/r) 0x0 h 0) (23) 


for the plane and the axisymmetrical cases, respec- 
tively. Here x is the horizontal Cartesian co- 
ordinate, r the horizontal polar radius. 


4. SELF-SIMILAR ASYMPTOTIC LAWS OF TURBULENT SPOT 
EXTENSION AT THE VISCOUS STAGE 


We neglected turbulent entrainment and the erosion 
of a turbulent region; therefore, the volume of the 
turbulent mixed region is considered to be constant 
and equal to the initial volume of the turbulent 
spot. It stands to reason that this assumption at 
the viscous stage is valid for sufficiently high 
stratification only. If the characteristic dimen- 
sions of the plane form of a turbulent spot are 
nearly equal, it is natural to expect that the ex- 
tension of the intrusion starts already to be axi- 
symmetric at the end of the intermediate stage and 


deliberately is axisymmetric at the viscous stage. 
Hence, Eq. (23) may be applied for its description. 
Thus, the condition of conservation of the volume 
of a turbulent spot takes the form 


co 


27 ff rh(r,t)dr = V = Const (24) 
fo) 


The asymptotic stage of the spreading of the spot 
is of primary interest when the plane size of the 
intrusion exceeds the corresponding initial size 
of the turbulent spot. At this stage the details 
of the initial distribution h(r,0) cease to be 
essential and for an asymptotic description or the 
viscous stage of the intrusion the initial distribu- 
tion may be represented in the form of an instantan- 
eous point source 


h(r,t;) = 0 (r #0), 2m f rh(x,t))dr = V (25) 
(0) 


Here, t; is the conditional time moment of the be- 
ginning of the viscous stage. 

The solutions of such type for non-linear heat 
conductivity equations with the power-type non- 
linearity to which Eqs. (22, 23) belong were con- 
sidered in the papers of Ya. B. Zel'dovich, A. S. 
Kompaneets, and one of the present authors [see 
Barenblatt et al. (1972)]. In our case the solu- 
tion depends on the quantities t - tj, n, V, r. 
The dimensional considerations show that it is a 
self similar one: 


@ 1/5 
a Anais > 1851) £(o) 


-1/10 
t= r[vin(t - t))/l6n*] (26) 


Putting (26) into Eq. (23) and integrating the 
ordinary differential equation obtained for the 
function, f(t), we find 


(aol ye A r2 1/4 
6 ( ae ) gOSES So 


0,520.5 103/572 = 2 (27) 


(GG) = 


Thus, at each moment of time the intrusion tongue 
stretches for a finite distance: this is (cf. Sec- 
tion 1) the peculiar feature of non-linearity dis- 
tinguishing the equation of intrusion from the 
linear equation of heat conductivity. The edge of 
the intrusion propagates following the law 


ro(t) = 2(vin(t = ey fen 2/2 (28) 


The form of the intrusion tongue represented by 
the curve 1 in Figure 4 also is peculiar: the 
thickness of the tongue changes slowly to the very 
edge where it comes abruptly to naught. The maxi- 
mum spot thickness, ho (t) = h(o,t), also changes 
very slowly with time 


0 0.5 70 


FIGURE 4. The distribution of thickness along an 
intrusion. 
anne (2 1/4 ( Vv ) 1/5 (29) 
fe) 6 Cra (te = te) 


Equation (28) seems very simple and accessible 
for experimental confirmation: confirmation of 
this equation will give some confidence in the 
validity of the model proposed here. The experi- 
mental checking of Eq. (28) was performed by A. G. 
Zatsepin, K. N. Federov, S. I. Voropaev, and A. M. 
Pavlov. They used the following scheme for the ex- 
periment (Figure 5). An open plexiglass tank having 
the form of a rectangular parallelepiped contained a 
stable, temperature-stratified fluid. A hollow 
cylindrical tube was introduced from above under 
the surface of the fluid. The fluid in the tube was 
mixed and then the tube was raised, leaving in its 
place a spot of mixed fluid which immediately started 
penetrating the ambient stratified fluid. The ob- 
servations, photo- and movie camera, were performed 
using a shadow device. The experiment allowed one 
to observe clearly the two last stages of spot evolu- 
tion; the spot extension at the viscous stage is 
represented in Figure 6. The mixed fluid volume in 
the spot was fixed for all experiments, as well as 
the kinematic viscosity of the fluid and the diameter 
of the tube. Therefore, if Eq. (28) is correct, the 
experimental data in the coordinates Lg[2ro(t)/D], 
&gIN (t - t))] had to fall on a single straight line 
with the slope 0.1. This is confirmed by the graph 
of Figure 6 where the slope of the solid straight 
line is 0.1 and t; = -10 sec. Thus, the law of one 
tenth Eq. (28) for the viscous extension of a spot 
was confirmed by the experiments of A. G. Zatsepin, 
K. N. Federov, S. I. Voropaev, and A. M. Pavlov with 
a Satisfactory accuracy. 

Analogously, in the case when the form of the 
turbulent spot is close to the cylinder with a 
horizontal axis Eq. (22) for the height of the in- 
trusion tongue will hold, where x is the horizontal 
coordinate normal to the axis of the spot. The con- 
dition of conservation of the volume of the spot of 
mixed fluid takes, for this case, the form 


H ff h(x,t)dx = V = Const (30) 


where H is the longitudinal size of the cylindrical 
spot. The initial conditions corresponding to the 


579 


asymptotic solution of the instantaneous point 
source type may be written in the form 


Die = ONG 70) { h(x,t))dx = Vv (31) 


a) 


and the asymptotic solution itself due to the same 
reasons, as before, may be represented in the form 


v2 1/6 
h 4n(t - t,)H2 =i) 


1/6 


5 = x[Vin(t - t,)/l6Ht] 


MG. = Be Aye ae Se 
(0) 10) 


2/3 

1/6 

2 =U 0, 62%. = Us)” om 2 3).(6 
Ae (5, 27s) 4? 2 0.97 (32) 


so that the leading edge of the intrusion, x = X(t), 
propagates according to the law 


ee? 


(2) = co lvin(t - t,)/16H* (33) 


o} 


while the maximum thickness of the intrusion, ho(t) 
= h(o,t), decays with time according to 


/6 


hot(t) = 0.97(v2/4H2n(t - ty))? (34) 


Thus, in both cases a strong deceleration of the 
extension of intrusion was characteristic for a 
turbulent spot in the transition to the viscous 
stage. Indeed, at the free intrusion stage the ex- 
tension of a turbulent spot is proportional to the 


The scheme of the experimental checking of 


FIGURE 5. 
the law of viscous extension of a spot of mixed fluid. 
1) The tank, 2) Point light source with collimator, 

3) Lens, 4) Vertical elevator with electromotor, 5) 
Mixer, 6) Tube, 7) Screen, 8) Movie camera. 


eee | 


Hi. 1 


Bes 6 6 20 


FIGURE 6. The one tenth law as confirmed by laboratory 
experiments of A. G. Zatsepin, K. N. Fedorov, S. I. 
Voropaev, and A. M. Pavlov. 


square of time; at the intermediate stage it is 
proportional to the square root of time for a cylin- 
drical spot and to the cube root of time for an axi- 
symmetric spot. At the viscous stage the extension 
is proportional to time; to one sixth in the case of 
a cylindrical spot and to one tenth in the case of 
an axisymmetric spot. Thus the extension of the 
spot is sharply decelerated at the viscous stage in 
comparison with the initial stages. 

It seems plausible to us that the "blini" shaped 
regions of constant density and temperature observed 
in the ocean are turbulent spots of various scales 
generated by the loss of stability or breaking of 
internal waves, local instability of shear flows, 
penetration of cooled turbulized fluid from the 
curbulized surface layer, etc. which are mainly in 
the last, viscous stage of their evolution. Note 
that along with the states in which turbulence is 
preserved within the spot, the states are possible 
and apparently rather frequent, especially for spots 
of small scales, in which turbulence within the spot 
has disappeared but the fluid remains mixed and homog- 
eneous. This assumption is supported qualitatively 
by some data of simultaneous measurements of vertical 
distributions of density and velocity gradient 
[Federov (1976)]. These distributions have the form 
presented by solid lines in Figure 1. Indeed, if 
the regions of constant density are intrusions, then 
the shear should increase near their boundaries com- 
pared to ambient fluid (cf., Figure 2). However, in 
this case the shear should be reduced near the cen- 
tral line of intrusion (dashed line in Figure l). 

It is plausible that the resolution in these mea- 
surements was not sufficient to observe this shear 
reduction. 


5. THE INTRUSION OF BOTTOM SEA WATER INTO THE 
BODY OF THE OCEAN 


The intrusion of mixed fluid into a continuously 
stratified medium is widely distributed in nature; 


IO) 100 


it is of interest from the point of view of the 
evolution of turbulent spots in stratified fluid. 
A characteristic example - the intrusion of the 
bottom Mediterranean water into the body of the 
Atlantic (Figure 7). The bottom water descends 
through the Straits of Gibraltar down the contin- 
ental slope and enters the body of the ocean in an 
intermediate layer where the density of the ocean 
water is equal to its own density. The intrusion 
of the bottom water of the Red Sea into the body of 
the Indian Ocean is completely analogous. The in- 
trusion of bottom water is a slow process and we 
may assume that for its description, Eq. (22), 
corresponding to a pure viscous mechanism of the 
intrusion drag, is valid. 

The intrusion of bottom sea water into the body 
of the ocean goes by separate portions [Federov 
(1976) ] and it is possible to assume that, at the 
beginning of the intrusion of a new portion, the 
bottom fluid that intruded earlier is carried suf- 
ficiently far away so that the initial condition 
holds 
(<= (0) (35) 

Here h, as before, is the height of the intrusion 
tongue, x the horizontal coordinate in the direction 
of intrusion from its origin. Let us suppose that 


FIGURE 7. The intrusion of sea bottom water into the 
body of the ocean. 


the height of the bottom water layer at the origin 
of intrusion does not depend on time: 


IM(Opi) = hy = (Const (36) 


The solution of Eq. (22) under conditions (35) 
and (36) is also self-similar and has the form 


= = Yoh + 
h hot, (a) + t x/ nh t (37) 
where the function f9(t) which satisfies the equa- 
tion 


a2£5° df> 


A a SER = @ (38) 
ac? dt 


under the conditions 


10) (39) 


(0) Si, sy) 


is continuous and has a continuous derivative 
df°/at (the last requirement follows from the 
continuity of the flow of bottom fluid). The solu- 
tion, f(t), is represented in Figure 4 (curve 2). 
It is also different from zero only in a finite 
interval 0 = G = To 6 «61.66, so that the leading 
edge of the intrusion x,(t) propagates as 


REFERENCES 


Woods, J. D. (1968). Wave-induced shear instability 
in the summer thermocline. J. Fluid Mech. 32, 
791. 

Monin, A. S., V. M. Kamenkovich, and V. G. Kort 
(1977). Variability of the ocean. J. Wiley. 
Federov, K. N. (1976). Fine thermohaline structure 
of oceanic waters. Gidrometeoizdat, Leningrad. 
Belyaev, V. S., I. D. Lozovatsky, and R. V. Ozmidov 

(1975). On the relation between the small-scale 
turbulence parameters and the local stratifica- 
tion conditions in the ocean. JIzv. AN SSR, Ser. 

Physics of Atmosphere and Ocean II, 718. 

Wu Jin (1969). Mixed region collapse with internal 
wave generation in a density stratiied medium. 
J. Fluid Mech. 35, 531. 

Kao, T. W. (1976). Principal stage of wake collapse 
in a stratified fluid: two-dimensional theory. 
Physics of Fluids 19, 1071. 

Monin, A. S., A. M. Yaglom (1971). Statistical 
hydromechanics. Part I. The MIT Press. 

Barenblatt, G. I., V. M. Entov, and V. M. Ryzhik 
(1972). Theory of non-steady filtration of liquid 
and gaS. Nedra, Moscow. 


v 4 
che 
a 
en 
may a 
ar 2 
i i ex AS 
I 
u 1 
i = Ve : 
ca i i 
ey I Ai 
= i ; — 
; i 
ry 
1 L 1 ‘ 
he { zs 
; 
‘ . b 
| 
~ 


1 


“1% 


eld OP, Raw fees tated marta Me™ 


ne 


Z 


eee: Ey pain’ 2 


Session VITT 


GEOPHYSICAL FLUID DYNAMICS 


LOUIS N. HOWARD 

Session Chairman 

Massachusetts Institute of Technology 
Cambridge, Massachusetts 


The Rise of a Strong Inversion 
Caused by Heating at the Ground 


Robert R. Long and Lakshmi H. Kantha 
The Johns Hopkins University 


Baltimore, Maryland 


ABSTRACT 


A theory is offered for the rise of a strong inver- 
sion in the atmosphere caused by heating at the 
ground. The heating, specified by the buoyancy 
flux, q,;, near the ground, causes turbulence in a 
growing layer of depth, D, above the ground with an 
inversion or interfacial layer of thickness, h, 
separating the mixed layer from the non-turbulent 
air above. There is a buoyancy jump, Ab, across 
the interfacial layer and the air above the inver- 
sion has a buoyancy gradient, No: 

The lower surface of the inversion layer rises 
(at a speed, Us = dD/dt) because of two processes. 
One is related to the mean temperature rise of the 
mixed layer which, in the present model, leaves h + 
D unaffected but which causes the interfacial thick- 
ness, h, to decrease and therefore D to increase at 
a rate proportional to eee where Ri = DAb/w% is 
the Richardson number and wx = (q,D) ? is the con- 
vective velocity typical of the rms velocities in 
the main portion of the mixed layer. The second 
process, increasing both h and D, is the erosion of 
the stable fluid by the turbulence in the mixed 
layer and the intermittent turbulence in the inter- 
facial layer. This causes D to increase at a rate 
proportional to Rea a The total effect is con- 
tained in the equation 


Ye 
— = aRi7! + cRi-7/* 
Wy 


where a and c are universal constants. Other re- 
sults are presented, notably the ratio, lqo/ay|, where 
qo is the (negative) buoyancy flux near the level 
Z=D. This ratio decreases with increase of sta- 
bility as observed in experiments of Willis and Dear- 
dorff. |qo/q,| ~ Ri-3/7*. 


1. INTRODUCTION 


When the sun rises and begins to heat the ground, 
the atmosphere is normally in a stable state (po- 


585 


tential temperature increases with height). If we 
neglect the effect of mean wind for the moment, the 
heating creates instability and turbulence near the 
ground and a mixed layer of depth, D, appears, capped 
by an inversion. This phenomenon is called penetra- 
tive convection. The potential temperature of the 
mixed layer is nearly constant with height except 
very close to the ground, where a superadiabatic 
lapse rate exists in a thin layer, and just below 
the inversion base where there is weak stability. 
The inversion base rises because of two processes. 
The first is heating alone which tends to decrease 
the thickness, h, of the inversion layer, (IL), and 
so increase D. The second is the entrainment effect 
of the turbulent eddies just below the inversion 
base. We do not have a detailed understanding of 
this erosion process but laboratory experiments with 
mechanical stirring [Moore and Long (1971), Linden 
(1973) ] suggest that the eddies in the mixed layer 
deflect the IL upward storing potential energy. When 
this is released by downward motion, a portion of 
the lighter fluid in the IL is ejected into the 
homogeneous layer where it is carried away by the 
turbulent eddies, leaving the lower surface of the 
IL sharp again. 

If there is no mean wind, the energy for the tur- 
bulence comes from the energy flux divergence term 
and from the buoyancy flux term in the energy equa- 
tion, where q = -w'b" is the buoyancy flux*. When 
there is a mean wind, as is usual in the atmosphere, 
the shear yields another energy source. This serves 
to increase the turbulence energy and thus to in- 
crease the entrainment effect through greater agita- 
tion of the IL. In addition, the shear may cause 
Kelvin-Helmholtz instability and consequent wave 
breaking at the interface and thereby enhance ero- 
sion. 

On the other hand, the effect of shear should be 


“SHOVES in an incompressible fluid is defined as b = 
g(p - P9)/p9 where g is gravity, p is density and pg is a 
representative density. In the atmosphere, p and pg are 
potential densities. We may also write b = g(@ -89)/8o 
where 6 is a potential temperature. 


586 


negligible if the mixed layer depth is much greater 
than the Monin-Obukhov length, L = -u3/q), [Monin 
and Yaglom (1971, p. 427)] where ux is the friction 
velocity. Thus (-L/D) 173 is proportional to the 
ratio, ux/wy, of the turbulent velocity in the mixed 
layer associated with shear to the turbulent veloc- 
ity associated with convection, wx = (q,D) 173. The 
shear effect becomes less important as this ratio 
decreases. Lenschow (1970, 1974) presents aircraft 
measurements, which appear to confirm the unimpor- 
tance of energy production by the shear for the 
turbulence near the inversion if |L/D| is small 
enough. 

The purpose of this paper is to construct a 
theory for the rise of an inversion in the atmo- 
sphere neglecting the effect of shear. The analysis 
is similar in some respects to that in a recent paper 
by the first author, [Long (1977b), hereinafter 
referred to as MISF] in which a theory is developed 
for turbulence in a stably stratified liquid, as 
for example in the experiments of Rouse and Dodu 


(1955), Turner (1968), Wolanski (1972), Linden (1973), 
Crapper and Linden (1974), Linden (1975), Thompson and 


Turner (1975), Wolanski and Brush (1975), and Hop- 
finger and Toly (1976). In these experiments a 
stably stratified fluid is agitated by a grid oscil- 
lating up and down near the bottom of the vessel 
(Figure 1). A growing mixed layer of depth, D, 
appears in the lower portion of the fluid separated 
from the non-turbulent fluid above, in which the 
buoyancy gradient is given, by an IL of thickness, 
h. Observations indicate that the lower mixed layer 
has a very weak mean buoyancy gradient. The buoyancy 
difference across the IL is relatively large and is 
denoted by Ab. 

As indicated by the experiments of Thompson and 
Turner and Hopfinger and Toly, and derived by the 
first author in a recent paper [Long (1977a)], the 
turbulence generated by the grid in a homogeneous 
fluid is nearly isotropic, and if u is the rms veloc- 
ity and 2 is the integral length scale, the quantity, 
uZ(proportional to eddy viscosity), is constant with 
height. When there is stratification, the mixed 
layer is nearly homogeneous and us = K is again con- 
stant near the grid [Hopfinger and Toly (1976)]. 
Since 2 is proportional to the depth, D, the veloc— 
ity, u, = K/D, is characteristic of the turbulent 
velocities in the mixed layer. The quantity, K, 
may be taken to be characteristic of the "action" 
of the energy source (grid). 

On the basis of observations, experimenters have 


FIGURE 1. 
the grid.) 


Oscillating grid experiment. (S = stroke of 


proposed that the entrainment velocity u, = dD/dt 
is expressible in the form 


Ue «3/2 _* DAb 
fs Ri , Rl = Feg2 (1) 


where Ri* is the overall Richardson number, f is the 
frequency, and S is the stroke of the grid. The 
measurements correspond to large values of Ri* so 
that attention is confined to the usual situation 

in nature in which the Richardson number is large. 
In terms of the "action" K of the grid, another 
Richardson number is 


Mee | msi (2) 
ea 


This is very similar to the number Ri = 2Ab/u2 
proposed by Turner (1973), where 2 and u are the 
integral length scale and rms velocity measured at 
the level z = D in a homogeneous fluid agitated by 
the same grid at the same grid frequency and stroke. 

In MISF and in the present paper, the role of the 
IL separating the mixed layer from the non-turbulent 
fluid above is essential. This contrasts with ear- 
lier theories in which h is neglected despite ex- 
perimental evidence [Linden (1975)] that h is 
proportional to D and is not particularly small 
(h/D = 1/4). Observations [for example, Wolanski 
and Brush (1975)] indicate that the IL with its 
large density gradient is typified by wave motion. 
Wolanski and Brush found that the frequency of dis- 
turbances in this layer was proportional to the 
Brunt-Vaisdla frequency (Ab/h)? although numerically 
one order of magnitude smaller. Certainly turbulence 
of some kind exists in the IL and since the density 
gradient there is strong rather than weak as in the 
mixed layer, it is reasonable to assume that the 
turbulence in the IL is intermittent and that this 
intermittent, weak turbuience transfers the buoyancy 
in the layer. In MISF the intermittency factor de- 
creases with increase of stability so that for the 
large Richardson numbers of the asymptotic theory 
the layer is, for the most part, in laminar wave 
motion with occasional breaking waves in the interior 
and at the lower surface of the interface. 

Similar ideas may be applied to the present prob- 
lem in: which the turbulence in the mixed layer is 
caused by heating at the lower surface. The princi- 
pal differences are the effect of heating in causing 
h to decrease and D to increase, and the differences 
in the sources of turbulence kinetic energy. The 
energy equation is 


' Dae 12 12 ey 

+ 

Open w(2 +% a ) Vu & (3) 
C4 Po 2 

where the first term is the energy flux divergence; 

u', v', w' are the instantaneous velocities, p' is 

the pressure, Po is a reference density, q = -w'b" 


-is the buoyancy flux, and € is the energy dissipa- 


tion. In the present problem the buoyancy flux 
term, -w'b', is of basic importance and corresponds 
to the conversion of potential energy to kinetic 
energy. This effect is missing of course, in the 
case of mechanical stirring in a homogeneous fluid. 
Equation (3) omits the local time rate-of-change 
of kinetic energy although, in fact, the inversion 
is rising and conditions are therefore unsteady. 
With respect to the mixed layer, the kinetic energy 


Heated Surface 


FIGURE 2. Model of entrainment at an interface by 
heating from below. The curve on the left is the mean 
buoyancy, b, with an assumed linear profile above the 
interfacial layer. The curve for buoyancy flux, q, is 
on the right. The superadiabatic layer near z = 0 is 
not shown. 


is proportional to the square of the convective 
velocity, (q,D) 273, so that the ratio of the time 
rate-of-change term to the other terms in Eq. (3) is 
Ue/Wx. This ratio is of order one if the convective 
motions are spreading upward at a speed, ug, in 
initial conditions of neutral stability. Evena 
fairly weak inversion will cause a great slowdown 
and ue/wy will be small. Similar remarks apply to 
the IL and we are assured that the time dependence 
is negligible in the stable conditions of the paper, 
although it has received some attention in considera- 
tions of the real atmosphere [Zilitinkevich (1975)]. 

We may conclude this introduction with reference 
to work on penetrative convection in the atmosphere 
and oceans including atmospheric observations: 
Lettau and Davidson (1957), Ball (1960), Veronis 
(1963), Izumi (1964), Summers (1965), Deardorff 
(1967), Kraus and Turner (1967), Lilly (1968), Dear- 
domki (972) Betts) (197/3)",, Carson) (197/s)),) Stull 
(1973), Tennekes (1973a,b, and 1975), Adrian (1975), 
Farmer (1975), Zilitinkevich (1975), Kuo & Sun (1976), 
Stull (1976a,b,c), and Zeman and Tennekes (1977). 
Related experiments have been run by Deardorff, 
Willis, and Lilly (1969), Willis and Deardorff (1974), 
and Hedit (1977). A second-order closure model has 
been given by Zeman and Lumley (1977). More recent 
field observations have been made by Kaimal, et al. 
(1976). Mixed layer deepening in the upper layers 
of the ocean, which is almost always associated with 
wind stirring has been discussed by Niiler and Kraus 
(ID 77) 6 


2. RELATION OF FLUXES TO THE BUOYANCY JUMP AND TO 
MIXING LAYER AND INTERFACIAL LAYER THICKNESSES 


In the theory of the paper we ignore rotation, radia- 
tive heating, water vapor, and horizontal variations 
of mean quantities. The model is shown in Figure 2 
which contains curves for the mean buoyancy and buoy- 


587 


ancy flux. The mean buoyancy curve above the IL is 
assumed to be linear with buoyancy gradient N2. In 
one case we assume that N* = 0 so that the inversion 
rises and weakens, eventually disappearing. When 

n2 # O we assume that the air was at rest with uni- 
form buoyancy gradient when heating began. Then the 
inversion strength increases with time. Since the 
theory of this paper is concerned with very stable 
conditions, the solutions hold for large values of 
the Richardson number. 

The buoyancy flux curve is derived below from the 
assumed buoyancy distribution. The latter is assumed 
to be linear in the IL (region R3). This is an ex- 
cellent approximation* in certain circumstances at 
least, for example in the mechanical stirring ex- 
periments of Wolanski and Brush (1975). Observa- 
tions in the mixed layer [Willis and Deardorff 
(1974) ] indicate that there is very little mean 
buoyancy variation in this layer except for some 
indication of a stable mean gradient near the heated 
plate. If we ignore these gradients for the moment, 
the equation 


db aq 

at Oz i) 
indicates that q is a linear function of z. In fact, 
experiments show that q is nearly linear [Willis and 


Deardorff (1974)] so that the neglect of mean buoy- 
ancy variations in the mixed layer in the model of 
Figure 2 seems reasonable. The lower surface is 
heated and the buoyancy flux q = -w'b' (proportional 
to the heat flux) is held constant at the lower sur- 
face where it is denoted by q,. The mean buoyancy 
in the mixed layer is 


b =b.. = N2(D + h) + Ab (5) 
m 00 


where Ab is the buoyancy jump across the interfacial 
layer and bog is constant equal to the buoyancy at 
the surface 1f the linear gradient above is extra- 
polated down to the surface. Integrating (4) over 
the mixed layer, we get the flux, qo, just below the 
IL. IRS Als) 


Se ky GD 


De Ch 
N2D S= (D + h) (6) 


On physical grounds qo must be negative (Figure 2) 
and this is confirmed by laboratory measurements 
(Willis and Deardorff (1974)]. In the IL, the mean 
buoyancy is 


- N2(D = h) (7) 


= Ab 
Ney) Ne) ee (ES 1D) Ie 1a 


Integrating (4), we get the flux at a given level 
in the interfacial layer 


2 2 
dAb t 2 dh z ap) 
Bou? G+ a) @ ESS Ee 
aD . dh 
peter sem (lola tet 8 
wee (2+ B) (3) 


where T= z- D. At z=D+h, the buoyancy flux 


is zero so that 


* 
Even when the approximation is only fair, the error in as- 


suming a linear profile is small. We discuss this in Section 
6. 


Using (6) we get 


ei = = x [(D + sh) Ab - 4N?(D + h) 2] (10) 


The integral of (10) is 


(D + sh) Ab - 4N2(D + h)? = Vo - q,t (11) 


where 


We = (oe ‘sh ) Abg - 4N?(D, + ny)? (12) 


and the zero subscript denotes values at t = 0. 
Tennekes (1973b) obtained (11) and (12) with h and 
ho missing. As we have indicated, the interfacial 
layer thickness h plays an important role in the 
theory of this paper. The time to = vp /a1 is the 
time for an initial buoyancy difference to disappear 
when the upper air has a uniform potential tempera- 
ture [Tennekes (1973b) ].- 


3. THE INTERFACIAL LAYER (REGION R3) 


According to the discussion in Section 1, the IL 

in our model is turbulent with intermittency factor, 
I3, defined here as the ratio of the volume in tur- 
bulent motion to the whole volume*. Much of the 
layer is in wave motion in which all of the compo- 
nents of the fluid velocity are of the same order, 
i.e., the ratios w3/u3, w3/v3 are independent of 

the Richardson number. The intermittent turbulence 
is caused by the intermittent breaking of these 
waves. Since the wave amplitude is of the order of 
the wave length when the wave breaks, we should have 
u3 ~ V3 ~ w3 initially in the breaking waves as well 
and we assume this. Of course the "homogeneous" 
fluid in the breaking patch will tend to flatten 

out and the vertical velocities in the patch will 
decrease relatively as time goes on. In our model 
we ignore the patch after a time of order (h/Ab)% 
and consider that the local heat transfer has al- 
ready been accomplished. In actual fact this trans- 
fer is accomplished by the spreading of the patch 
over a larger time interval and the ultimate trans- 
fer by molecular processes. Since buoyancy flux 
occurs only in the turbulent portions of this layer, 
we get, at any level in the IL, 


a3 = ~ Byu3b313 (13) 


where b3 is the rms buoyancy fluctuation in the 
interfacial layer. B , is a universal constant? but 


“the introduction of intermittency may result in confusion 

if one inadvertently thinks of the IL as a’surface or even 

as a layer with thickness of the order of the amplitude of 
the wave disturbances. The latter is not excluded as a 
possibility in this section but, in fact, as we see in 

Eq. (26) the wave amplitude is much smaller than the thick- 
ness of the IL so that I is not the ratio of the times that 

a fixed point is in the upper (non-turbulent) and lower (tur- 
bulent) fluid. 


twe use symbols B), Byj,... to denote universal constants. 
Later, "constants" arise which, at first glance at least, 
may be functions of s = N2/(Ab/h) , i.e., the ratio of the 
stabilities of the upper "quiescent" layer and the inter- 
facial layer. We denote these "constants" by Aj,Ap,..-- 


b3, u3, and I3 may vary with height. The turbulence 
is certainly strongly influenced by buoyancy in this 
layer so that kinetic and available potential ener- 
gies [Long (1977d)] are of the same order not only 
in the waves but in the turbulent patches, i-.e., 


eS BS, S B05 a (14) 


where 63 is the order of the size of the disturbances 

and because of the tendency for conservation of buoy- 
ancy, we assume b3 is proportional to 63(Ab/h). Us- 

ing (14), Eq. (13) becomes 


B2B 5 
coils erase a ab) 
43 aa (4 1 (15) 


Let us now find the dissipation. This occurs only 
in the turbulent patches and we assume that the 


local dissipation Gia = £(u3,63, b3). Since us ~ 
b353, we get Ep ~ 3/63 and 
Bu? ls Ab 45 
a 
€3 = 13 4 = B,Bju3 (2) T3 (16) 
63 h 
Equations (15) and (16) show that €3 ~ q3. Since 


these are both dissipative, it follows that they are 
of the order of the energy flux divergence. At the 
upper boundary of the IL, the kinetic energy of the 
waves has been so reduced by losses to potential 
energy and dissipation, that there can no longer be 
wave breaking and turbulence. Thus h is the depth 
of penetration of the turbulence. At the height z 

= D+h, the energy flux is too weak to support tur- 
bulence so that it has apparently decreased to a 
value well below that at the bottom of the IL. 
Therefore, the increment in energy flux over the IL 
is proportional to the value at the bottom of the 
IL. Integrating Eq. (3) between levels in the layer 
near the upper and lower surface, we find that q3h 
is of the order of the energy flux just below the 
inversion where q3 is the average buoyancy flux in 
R3-. Since the interface is being distorted by the 
vertical motions (inducing pressure fluctuations) , 
the energy flux should be proportional to W5P5/00 


~ we in Ro. We may write 


= 3 
q3h SO Aowo (17) 


Equation (17) has a form superficially similar to 
that proposed by others in a number of papers [for 
example Long (1975), Zeman and Tennekes (1977)] on 
the basis of assumptions about the size of terms in 
the mixed layer. In present notation, these authors 
propose qoD ~ we and this leads rather directly to 
the Ri-! law for the entrainment. Equation (17) is 
really quite different. If the upper fluid is ho- 
mogeneous, Ay should be a universal constant. How- 


“ever, when the upper layer is stratified, losses of 


energy may occur by wave radiation and Ap may then 
be a function of s = N2/(Ab/h). 
Using (6), (8), (14), (15), (17), we get 


3 
Agw? dip nh dAb) Ab dh aD 
iy See nO EO Be Sn, Se 
h ae) cles 9G Ge DS Ge 
2 1 d 
+ q; - N*“(D + %h) ae (D + h) (18) 


1, 
=e ma \ 2 
= eff, 
So = B3 (2) wo (19) 
BB 1s 
od Ab dAb d 
yaa) = ceo Dl a = anes 


where the subscript "2" denotes values at a level 
just above z = D. Equation (19), which follows from 
Eq. (18), is consistent with the assumption that the 
pressure fluctuations in eddies in region Rp of fre- 
quency wo/59 of order of the natural frequency 
(Ab/h)2 are generating the breaking waves by reso- 
nance. 


4. TURBULENCE IN THE MIXED LAYER 


According to (17) the vertical turbulence velocity 
in Ro is related to the average buoyancy flux in 
the interfacial layer. The latter is related to 
the entrainment velocity so that it is essential to 
relate w 2 to turbulence in the main portion of the 
mixing layer, or to w, = (qb) 173 This is often 
called the convective velocity. A great deal of 
confusion has arisen regarding this problem because 
of two explicit or implicit assumptions often made: 
(1) that the turbulence near the interface is quasi- 
isotropic, i.e., ug ~ vo ~ wo, and (2) that wo ~ wy. 
We will try to show that both of these assumptions 
are incorrect*. 

In laboratory experiments with mechanical mixing, 
measurements indicate that the mean buoyancy gradi- 
ent in the mixed layer is very weak and, in fact, 
approaches zero as the Richardson number increases 
(Wolanski (1972)]. Instantaneously, the lower sur- 
face of the interfacial layer is very sharp (perhaps 
a discontinuity for infinite Reynolds numbers!) . 
This surface is agitated by the disturbances of the 
mixed layer so that the mean buoyancy curve varies 
continuously, although rapidly in the region, R». 

It seems quite safe, however, to neglect effects of 
buoyancy on the turbulence of the instantaneous mixed 
layer. Let us do this tentatively although we will 
return to this point later. Since, for the highly 
stable conditions of this paper, the interface dis- 
turbances will be very small, the inversion will act 
like a 'rigid lid" with slipt and the turbulence will 
be similar to turbulence between a rigid heated plate 
at z = 0 and a rigid plate az=D. The first ques- 
tion to face, then, is the nature of the turbulence 
at some level € = D - z near the upper "plate." To 
do this, we first consider the findings in two recent 
papers by Hunt (1977) and Hunt and Graham (1977) re- 
garding the distorting effect of a rigid plane on 
homogeneous turbulence. The corresponding labora- 
tory experiment is produced by passing air through a 
grid in a wind tunnel. The rigid plane is a moving 
belt along one wall of the wind tunnel with speed 
equal to the mean wind. This serves to eliminate 

the shear near the wall and the corresponding energy 
source. The wall causes two boundary layers (Fig- 
ure 3). One is a very thin viscous layer of thick- 
ness dy near the wall in which all three components 
of velocity go to zero, and the other, called a 
source layer of thickness 6,, extends from the vis- 


* 
We mean by A ~ B that A/B is finite and non-zero in the limit 


as ixul $7 dp 
"This is the opinion also of Zeman and Temnekes (1977). 


589 


FIGURE 3. Turbulence near a wall. 


cous layer to a level at which the disturbing ef- 
fects of the wall are negligible. The vertical 
velocity must decrease throughout the source layer 
because it is very small at the top of the viscous 
layer, but there is no obvious reason for a decrease 
of the horizontal velocity components in the source 
layer. This is confirmed by experiment and by the 
mathematical analysis by Hunt and Graham who derive 
the following results of interest in the present 
problem: The rms vertical velocity in the lower 
portions of the source layer is wo = B(et) 1/3, where 
B is a universal constant and € is the dissipation 
function far from the wall, and the rms horizontal 
velocities are of the same order as those far from 
the wall although somewhat larger. It is useful to 
obtain these and other results more intuitively. 
In a recent paper, the first author [Long (1977c) ] 
has shown that turbulence at high Reynolds number in 
a wind tunnel far from a wall is determined com- 
pletely by two quantities, K and u/x, where K is a 
quantity of dimensions L?T-! characteristic of the 
grid and proportional to ul. u is the mean velocity 
and x is distance downstream from the grid (or more 
accurately from a virtual energy source replacing 
the grid). For example, the dissipation function 
far from the wall is e€ ~ Ku2/x*, the rms velocity 
is u ~ (Ku/x)%, and the integral length scale is 
2 ~ (Kx/u)*s. 

Obviously the source layer thickness is 6, ~ 2 
[Hunt (1977)] and the dissipation in the source 
layer is 


u 

e, = ef a (21) 
K ey 2 

Just outside of the viscous layer, Es is Eso or 

rt 
bya? 

ein 68 \| Soae (22) 
K°x? 


INS Ws Op Sy > 0O, and, since ¢€ must be independent 
of viscosity for high Reynolds number turbulence, 
aq @ Co 

At small ¢, eddies of length much less than f 
will not feel the distorting effect of the surface 
and will be isotropic. Eddies of length much greater 
than ¢ will feel the surface very strongly and will 
be strongly flattened. Eddies of length of order 
& << & will feel the surface but will remain quasi- 
isotropic. From the equation of continuity the 
large flattened eddies of horizontal dimensions D 
yield vertical velocities of order ujZ/D ~ KE/D2. 
The quasi-isotropic eddies are much smaller and for 


590 


high Reynolds numbers will lie in the inertial sub- 


range. They will have a spectrum function 
2 5 
Sg ie ae wd Spall (23) 
E 
1 (k) oe k Aids (6 


where k is the wave number so that ,the contribution 
to the vertical velocity is et6°cl/3, This is much 
larger than the contribution from the flattened 
eddies so that w, ~ elf3zi/3 or we ~el/3 ti/3, as 
derived rigorously by Hunt and Graham (1977) . 

In the mixing experiments the surface at z = D 
is not rigid but is agitated by disturbances of 
amplitude 65. Assuming that eddies of this size 
are in the inertial subrange, we get vertical veloc- 
ities of order BMSINiee and again these, rather than 
the eddies of size D, contribute most to the rms. 
Then wo ~ el/35,1/3, Since € ~ K3/p4, we get, as 
in MISF, 


=B (24) 


The problem of the present paper is somewhat more 
complicated but the distorting effect of the inter- 
face should be the same since the buoyancy varia- 
tions in the mixed layer are very small. The air 

in the main portion of the mixed layer has velocities 
of order (q)D)!/3 rather than K/D and in Ro the 
buoyancy flux is similar to that in the case of 
mechanical stirring. Equation (24) takes the form 


3 
WwW 
Ee BCH (25) 
85 : 


This result, together with (19), implies wo ~ 
W*Ri-4(h/D)% , where Ri = DAb/w2, and differs 
fundamentally from that of Tennekes (1973b) who 
assumed wo ~ w, by arbitrarily equating the buoy- 
ancy flux and the energy flux divergence. Tennekes 
has acknowledged [Zeman and Tennekes (1977) ] the 
inadequacy of this assumption. 

The drop-off of w as the interface is approached 
is revealed in the data of Willis and Deardorff 
(1974). As shown by Hunt and Graham (1977), the 
total kinetic energy is the same near the distorting 
surface as it is far away so that the horizontal com- 
ponents of rms velocity should increase toward the 
interface. There is an indication of this also in 
the data of Willis and Deardorff. 

It is also interesting that we may predict the 
same type of behavior near the lower heated surface. 
In fact, earlier data of Deardorff and Willis (1967) 
as well as the more recent data of Willis and Dear- 
dorff (1974) show that the vertical velocity near 
the heated plate increases with height, roughly in 
accordance with similarity theory [Prandtl (1932)], 
but that the horizontal velocity decreases with 
height. Thus, it is possible to apply similarity 
theory to obtain the vertical component, w, but not 
to obtain the horizontal components, u and v. The 
dimensional analysis for the horizontal components 
at large Rayleigh number must include D as well as 
q, and z no matter how small the ratio, z/D! There 
are experimental indications that the classical 
arguments of "localness" are also incorrect in prob- 
lems of turbulent shear flow [Tritton (1977, p. 

Using (25), the relations in (18)-(20) and the 
expression for wy are 


283)]. 


oat oD 3 
Sy a iain is 
LONER PS 
=p 2a. 8 A) 27 
pe techy cel (2 bean) 
mies ie oy an = @= hb) = (28) 
Bp 119162 = at at a 
3 
-a,q)2 h\dAb Ab dh 1 aD 
2th 
=> — —_— + —_- — — — 
Spe gus ea ae ( Sig ) cae | 2S ae 
ht (Ab) 4 
Se = hy ESL ae ey) (29) 
BPD sa aie) as 
Seanes 
3 


q 
where Q5 = A,B) 1/33 5 


5. DIFFERENTIAL EQUATIONS 


Equation (29) is a single differential equation in 
three unknowns, D, h, Ab. Let us now seek additional 
information. The quantity, 03/53, is the dissipa- 
tion in the turbulent patches in the interfacial 
layer. We have seen that it is independent of Ri 
in the lower portions of the layer. Obviously it 
will vary continuously with < (now defined as z - D) 
in the layer and, to the first order, will remain 
independent of Ri although it may vary with the 
quantity s = N2h/Ab when the upper fluid has a 
linear buoyancy field. We may therefore write 


U3 t 
oe = Ci 4? (é, s) (30) 


or using (14) 


4 nye 
4 h ic 
u3 = B3 Ae) ¥7 (, s) (31) 


We may obtain another expression for us by in- 
tegrating the energy equation over the interfacial 
layer. We have already seen that Je3|~|a3| and 
assuming that the energy flux is proportional to 
u3 in this layer*, we have from the energy equation 


Ww 


3 
3 = Brigg (32) 


Using (8) and integrating, we get 


dab (72 3 
ud = wi + Bio [ ans + MB(E- =) 


(DE) (33) 


* 

We have seen that the energy flux at the bottom of the layer 

1s proporticnal to ugae To the first order it should be pro- 

portional to us in the rest of the layer, i.e., independent of 


Ri. 


Comparing (31) and (33) and using (27), we get 


3 
ga aes 2 ne a tr eC Ee 
Yy (Es) =e ge Tr usp + 25 (34) 
where? A3, Ay, and As may depend on s. Equating 


coefficients in (34) we get (29) again and the 
following 


ied 
2 4 
dAb a Gl ie aor 
Ghy WPMD) are" 7 NED) a (D + h) = -a3 3 (35) 
D(Ab) # 
33 
2 rane 
Dddb , db aD _ N*D a 7 1 
5 Ge Do as a ae (O° 2) = ch 3 (36) 
D(Ab) * 
32 
24 
D2 aAb | D2Ab dh q, h 
- = + 5 = a5 (37) 
6h dat 6h2 at 3 
D(Ab) * 
my BYAR na ; 
where Oj] = A;/B)2B3 (i = 3,4,5). Equations (35)- 


(37), (29), and (11) are five equations in the three 
unknowns. They determine the solution to the first 
order for large Ri, although we must make sure that 
all equations are satisfied to that order. In this 
regard, if we use (35)-(37), (29), and the deriva- 
tive of (11), i.e., (10), we may consider these as 
five homogeneous linear, algebraic cquat tous in five 
EET SA FA dAb/dt, dD/dt, dh/dt, q,, and qi? 24374 7p 
(Ab) 3 +. The determinant of these eguations vanishes 
and we satisfy compatibility. 


6. HOMOGENEOUS CASE (N = 0) 


If N = O, the upper fluid is homogeneous and (11) 
becomes 


1 2 
(D + 5h) Ab = Vp - ayt (38) 


2 9 Aad 
where Vp is a constant related to initial conditions. 
We use (35) and (38) to eliminate Ab in (29), (36), 


and (37). We get 
3 3 3 
9.4 4 4 
d a, °h (D sr 2 ) 
at (Ab) = - ea a3 52 ViEgaeene i! (39) 
(V5 - q,t) 
dh qih (D+ oh) 
dt D 
(vi = Cite) 
Si o 
q, 2h" ines tn) 
+ Cy = 6a5 D = 7 = (0) (40) 
ON = qyt)* 


+ 5 be PD: A 

.For arbitrary Ri, the quantities A,; may depend on Ri. As 
Ri > ©, however, A, will approach "constants" which may, of 
course, be zero. 


591 


1 
dap h (D+ gh) 
SG 
at 1 >) 4 
(Vo - q,t) 
a7 As 
aon (D + 5h) 4 
= (a3 + 204) =0 (41) 
D2 2 Z 
(Vo - a, t) 
1 
Bc e anya penap ny aie 
dt dt 2) De 
(Vo - q,t) 
307 7 
ae 
D2 D q,°h! (D + zh)" 
sue 6a9 oe 603 a 203 = (0) (1)) 
h 2 f 


Two effects occur in (40) and (41). We may separate 


them by adding the two equations. We get 
3 7 
24 1 4 
él h\ 41 h }D) Sr oh 
—_ + = + _ _ 
dt (D h) (2u, 6a5 ae 2 (43) 
Vo ~— qt 


The term on the right of (43) expresses the upward 
motion of the boundary between (intermittently) 
turbulent and non-turbulent fluid due to turbulence 
in the interfacial layer causing entrainment of the 
upper, non-turbulent fluid. On the other hand, the 
second terms in (40) and (41) express the upward 
motion of the boundary between fully turbulent and 
intermittently turbulent fluid (and the consequent 
decrease of h) due to heating alone. This contribu- 
tion to the entrainment velocity is proportional to 
the interfacial thickness, h, and disappears when 
the common approximation is made that h = 0. 

Let us find an approximate solution to (40)-(42). 
If we let Dp and ho be the values of D and h at t = 
0, we make the following definitions: 


ho h D 
S'S =) = h = = jp 
Do a , Do 1 , Do i 
2 2 1 
3 3 a J. 
q 
a8 pet (44) 
Vo Do3 


Then equations (40)-(42) may be written 


1 
+ 5 hj) 
dh i By, Uo 
dt Dy (1 - 6t) 
L, ilies 
4 ii yoy 
2, ye SOE Ze =0 (45) 
+ a3 > 645 Dy 2 Z 
: Dy (AL > Ox) 
1 
aD, hy (D, ar zhy) ; 
dt Dy (lL = Or) 
ih 1 Ho 
GD, & Bin, ) B08 
2 (Gg > 2p) 7 0 (46) 


Do (2 - 6t)* 


592 


D Dj hi 
(602 ne 6a3 hy + 604 + 605 7 


72 Le 
q 1 rar 
ot SS ST ee, (47) 
2 £ 
Dy (1 - at)4 
Solutions are of the form 
h i q,t 
—_ = _ + = —_ 
mh aL (al 74) ve 
3 7 sa 
= < 2n2 
4 Pace fay Das 
= A(Cley = Seley) Gi: (al + $a) at 
v2 
0 
Qe 
mate 
+5 a(2t+ayo 1+... (48) 
Vo 
Gast 
Zen aG so) = 
Do 2 V6 
atl 
fr il Dee Dye 
ae (ie), ae Aton) ery (al ae 58) aetricn 
Vo2 
Gace 
- 5 a2(2 + a)? ae (49) 
Vo 
Pee 2 Se 
vy 2+a ve 
0 
Sel 
iL 1 c= 
7 oF Pip ne 
O13 ) a a y qy Do 
G 204 3asa}a{l+ 5 z 
Vo 
2 D2 
a(2 + a) qjt 
ee aC SOO (50) 


4 
4 Vo 

where hg and Dp are related by the equation 
6a9 - 6a3a + 6ay4a2 a 6as5a2 = 0 (51) 


The entrainment velocity, ug = dD/dt, may be ex- 
pressed in terms of the Richardson number, Ri = 
DAb/w? , by using 


6(1 + Sa) = fal a 25) 


Y 2 


np Ce (Ox 4 2a4)a* (52) 


The first term is of the same form as the non- 
dimensional entrainment velocity of Tennekes (1973) 
but, as already pointed out, the derivation and 
physical mechanism are very different. It is easy 
to trace the error in (52) arising from the simpli- 
fication of Section 2 that the IL has a linear 
buoyancy field. The error is proportional to 
(u,/w,) ab./Ab where b, is the maximum difference 
between the actual buoyancy in the IL and the as- 
sumed buoyancy. Since a is 1/6 or so and be/Ab is 
fairly small, this error is negligible. Notice also 
that the theory concerns strongly stable conditions 
so that (52) does not apply in the limit as Ru, == Op 
As Ri tends to order one ue becomes of order wy as 
one would expect. 


The ratio go/q, is of interest. Using (6), we get 


a2 D_ dAb 
— si ¢ = S— (53) 
qi qi dat 
Using (50) and (52), we get 
Z 
3 £ 
a2 a TR = a 
|| = yRi * , y = — (3+ 2ay + asa) (54) 
ql (ea) 
2 
The expressions (26)-(28) are 
1 
Td 5 3 
Cy aguas Op BHae: Se 
—_—= at atri # = 8 ua Rea ' 
We im iB 3.3 
B3! rears 
B3 alt 
3 
I SeE Rae (55) 
2 ak ae ee 
BiB) * 


These relations are identical to those in MISF. The 
result that the disturbances in the IL are small 
compared to the thickness of the IL is contrary to 
speculation [Stull (1973) and Zeman and Tennekes 
(1977)] that h is the depth of penetration of the 
eddies into the stable region. 


7. LINEAR BUOYANCY FIELD IN THE UPPER LAYER (N # 0). 


We consider initial conditions in which the fluid is 
at rest initially with a linear buoyancy field, so 
that D, h, Ab are zero at t = 0. Equation (11) be- 
comes 


n2 


(D + sh)ab = (D+ ny? = - gt (56) 


Equations (56), (29), and (35)-(37) determine the 
problem. The approximate solutions* are 


* 

The solutions, as throughout the paper, are for strong 
stability, which implies here that Nt is large. Thens +1, 
and %9,043,4,45 are independent of s. 


i L 
2 
(2q)t) (2q)) “by 
= , 
N Ne 
1 
l 2 
= (2q,)“b 
2a 1 2 
h = (2q)t) z + 3 
n2 
at ee 
Ab = (2q,t)2aN + (2q)) 2b3N? (57) 
where 
1 
2: b a 
a=-1+(Q+2 Pyles ME Se) \ See L 
Oy ay i ay 3 
22 
1 
ao 2 ide} = lok) = 
—=a*(1 +a) , ———= 22a (58) 
(omn 3) Oy 
Using the relationship 
3 3 3 
42 uf b b aS 
Ri 3a 1 sl Ban oh 
Nt = 3 1 eee Ri (59) 
2a? 22 
we obtain for the entrainment velocity 
u x Tek math 
— = ari7! + 22b,a'Ri * (60) 
We 
The ratio of fluxes is 
3 ees 
|*| = 22p aki 4 (61) 
qi 


Notice that Ab/h > N2 as t > © so that the IL be- 
comes indistinguishable from the upper layer as the 


turbulence in it weakens (becomes more intermittent). 


This contrasts with MISF in which the stability in 
the IL is several times larger than the stability 
in the upper fluid. Notice also that s + 1 implies 
09...d5 are universal constants. More accurately, 


Nie 


[Nd 2. 
ae = 1+ ay (<2) (62) 


We see from (32) that a, > 0 so that the buoyancy 
gradient in the IL is more stable than in the air 
above. These results suggest that an interfacial 
layer will be difficult to identify when there is 
a stable buoyancy gradient aloft. This is certainly 
the case in the experiments of Deardorff, Willis, 
and Lilly (1969) and Willis and Deardorff (1974). 


8. DISCUSSION 


We have already contrasted the theory of this paper 
with that of Tennekes (1973). He obtains 


593 


Ue OY 
Beanie (63) 
* 


which has the same form as the first term in (52) 
or (60). The present theory should not, however, 
be regarded as an extension or modification of the 
Tennekes' theory because, as we have noted in sev- 
eral places, the two theories differ fundamentally. 
This is also evident in the difference in the nature 
of the two constants of proportionality for the 
Ri71 term in the two theories. The a, in (63) may 
be identified physically as the ratio |qo/q | which 
is a universal constant in the Tennekes' theory. 
The constant, a, in (52) or (60), however, is a 
universal constant equal to the asymptotic value 
of the ratio of the inversion layer thickness to 
the thickness of the mixed layer. Tennekes assumed 
a value of 0.2 or so for a; and it is a coincidence 
that this is also a reasonable choice for a. 

We may attempt to estimate the constants in the 
expressions 


7 3 
u a air qo a) oe 
e ates a. 4 0 
— = aRi7! + cRi * f Asal 4 (64) 
Wy ql 


using the data of Willis and Deardorff (1974)*. 
Approximate estimates for the two cases: 


Sily iD) = Se) Guy, in = S) on, Ae S 157%, 

0.39 cm/sec, Qo = 0.18°C cm/sec, 
WwW 1.3 cm/sec, Ri = U5 5,5 4 = Ostler 
¢ = 1.09, y = 1.61 
B48) te) S FSeidp iy SS 55) vem, (aw = Be, 

A 


He the 


0.69 cm/sec?, Qo = 0.22°C cm/sec, 
1.4 cm/sec, Ri = 20, a = 0.15, 
@ = 1,05, 7 = 1.05 


Ee top 
te ak 


We may also attempt to compare with atmospheric data. 
For example, using the 1200 observation on Day 33 
for the Wangara data, [Zeman and Tennekes (1977)], 
we obtain 


2 S10" Guy AO S 2°C;, 
20°C cm/sec, 
3, 4 = O58 C= 22 


D = 1.1 10°cm, h = 
13 cm/sec’, Qo 
194 cm/sec, Ri 


Ile 


Cc 
o 
Idk 


These computations indicate that the two terms in 
the expression for u, in Eq. (64) are roughly 
similar in magnitude for atmospheric and laboratory 
conditions. 

It is interesting to compare the theory of the 
erosion of a linear buoyancy field with a numerical 
experiment of Zeman and Lumley (1977) using a 
second-order closure model. The numerical calcula- 
tion began from an initial instant, ty, at which 
eshte) IY S Dip Wa = Wag) = (qaDo)it7 2. The present 
theory sat timelt) =) tS tp ais 


D T 9) 
— = te SP ooo Sn) = WDE AW 
Do i So 7 SH) Uf 


where we have assumed that (tgN)’s is large. The 
numerical’ curves [Figure 1 of the paper of Zeman 
and Lumley] are nearly linear after tT exceeds 2 or 
so although, as (57) would indicate, D/Dop increases 
somewhat more slowly after considerable time. The 


i . 
Supplemented by information in a personal communication from 
Dr. Willis. 


594 


FIGURE 4. Comparison of present theory and numerical 
experiment of Zeman and Tennekes (1977). The curves 
correspond to values of S, in (65). 


most important comparison, however, is that the 
curves of Zeman and Lumley for various Sg collapse 
rather well when plotted against t/Sg instead of Tt 
as in Figure 1 of Zeman and Lumley. Conversely, we 
may reproduce Figure 1 of Zeman and Lumley together 
with plots of D/Do in (65) for the same values of 
Sg chosen by Zeman and Lumley. This is shown in 
Figure 4. The agreement is good, especially at 
large stabilities where the approximation in (65) 
should be best. This indicates that the two models 
have some similar features. 


ACKNOWLEDGMENT 


This research was supported by the National Science 
Foundation under Grant Nos. ATM/6-22284, ATM/6-04050, 
and OCE76-18887, and by the Office of Naval Research, 
Fluid Dynamics, under Contract No. NO0014-75-C-0805. 


REFERENCES. 


Adrian, R. J. (1975). Turbulent convection in water 
over ice. J. Fluid Mech. 69, 753. 

Ball, F. K. (1960). Control of inversion height by 
surface heating. Quart. J. Roy. Met. Soc. 86, 
483. 

Betts, A. K. (1973). Non-precipitating cumulus con- 
vection and its parameterization. Quart. J. Roy. 
Met. Soc. 99, 178. 

Carson, D. J. (1973). The development of a dry, 
inversion-capped convectively unstable boundary 
layer. Quart. J. Roy. Met. Soc. 99, 450. 

Crapper, P. F., and P. F. Linden (1974). The struc- 
ture of turbulent density interfaces. J. Fluid 
Mech. 65, 45. 

Deardorff, J. W. (1967). Empirical dependence of 
the eddy coefficient for heat upon stability 
above the lowest 50 m. J. Appl. Meteor. 6, 631. 

Deardorff, J. W. (1972). Parameterization of the 
planetary boundary layer for use in general cir- 
culation models. Mon. Wea. Rev. 100, 73. 

Deardorff, J. W., G. E. Willis, and D. K. Lilly 
(1969). Laboratory investigation of non-steady 
penetrative convection. J. Fluid Mech. 35, 7. 


Farmer, D. M. (1975). Penetrative convection in 
the absence of mean shear. Quart. J. Roy. Met. 
Soc... LOD, 869) 

Heidt, F. -D. (1977). Comparison of laboratory ex- 
periments on penetrative convection with measure- 
ments in nature. Heat Transfer and Turbulent 
Buoyant Convection, 1. D. Spalding and N. Afgan, 
ede), 199-2010; 

Hopfinger, E. J. and J. A. Toly (1976). Spatially 
decaying turbulence and its relation to mixing 
across density interfaces. J. Fluid Mech. 78, 
I1G)5}< 

Hunt, J. C. R. (1977). A review of the theory of 
rapidly distorted turbulent flows and its appli- 
cations. XII Biennial Fluid Dynamics Symposium, 
Poland. 

Hunt, J. C. R., and J. M. R. Graham (1977). Free 
stream turbulence near plane boundaries. J. Fluid 
Mech. (In press). 

Izumi, T. (1964). Evolution of temperature and 
velocity profiles during a breakdown of a noctur- 
nal inversion and a low level jet. JG. Appl. 
Meteor. 3, 70. 

Kaimal, J. C., J. C. Wyngaard, D. A. Haugen, O. R. 
Cote, Y. Izumi, S. J. Caughey, and C. J. Readings 
(1976). Turbulence structure in the convective 
boundary layer. J. Atmos. Sci. 33, 2152. 

Kraus, E. B., and J. S. Turner (1967). A one- 
dimensional model of the seasonal thermocline. 
Part II. The general theory and its consequences. 
Tellus) 19), 98. 

Kuo, H. ie, and Ww. Y. Sun (1976). Convection win 
the lower atmosphere and its effects. J. Atmos. 
Seals Bp Zils 

Lenschow, D. H. (1970). Airplane measurements of 
planetary boundary layer structure. J. Appl. 
Meteor. 9, 874. 

Lenschow, D. H. (1974). Model of the height varia- 
tion of the turbulence kinetic energy budget in 


the unstable planetary boundary layer. J. Atmos. 
Sexo sel, AXS5\5 
Lettau, H. H., and B. Davidson (1957). Exploring 


the Atmosphere's First Mile., Vols. 1 and 2. 
Pergamon Press, New York. 

Lilly, D. K. (1968). Models of cloud-topped mixed 
layers under a strong inversion. Quart. J. Roy. 
Meteor. Soc. 94, 292. 

Linden, BP. 'F.' (1973). The interaction o£ jay vortex 
ring with a sharp density interface: a model 
for turbulent entrainment. J. Fluid Mech. 60, 
467. 

Linden, P. F. (1975). The deepening of a mixed 
layer in a stratified fluid. J. Fluid Mech. 71, 
385. 

Long, R. R. (1975). The influence of shear on mix- 
ing across density interfaces. J. Fluid Mech. 70, 
305. 

Long, R. R. (1977a). A theory of turbulence in a 
homogeneous fluid induced by an oscillating grid. 
Submitted to Physics of Fluids. 

Long, R. R. (1977b). A theory of mixing in a stably 

; stratified fluid. Submitted to Physics of Fluids. 

Long, R. R. (1977c). The decay of turbulence. Sub- 
mitted to Physics of Fluids. 

Long, R. R. (1977d). Some aspects of turbulence in 
geophysical systems. Advances in Applied Me- 
eChaniiicsil/a melts 

Monin, A. S. and A. M. Yaglom (1971). Statistical 
Fluid Mechanics: Mechanics of Turbulence 1. The 
MIT press., Mass. 


Moore, M. J., and R. R. Long (1971). An experimental 
investigation of turbulent stratified shearing 
flow. J. Fluid Mech. 49, 635. 

Nawdiern, P. Pi. .and E., Bi. Kraus) (9/77). \One= 
dimensional models of the upper ocean. Modelling 
and Prediction of the Upper Layers of the Ocean, 
E. B. Kraus, ed., Pergamon Press, Oxford, England, 
pp. 143-172. 

Prandtl, L. (1932). Meteorlogische Andwendungen der 
Strémungslehre. Beitr. Phys. fr. Atmos. 19, 188. 

Rouse, H. and J. Dodu (1955). Turbulent diffusion 
across a density discontinuity. Houille Blanche 
10, 405. 

Stull, R. B.(1973). Inversion rise model based on 
penetrative convection. J. Atmos. Sci. 30, 1092. 

Stull, R. B. (1976a). The energetics of entrainment 
across a density interface. J. Atmos. Sci. 33, 
1260. 

Stull, R. B. (1976b). Mixed-layer depth model based 
on turbulent energetics. J. Atmos. Sci. 33, 1268. 

Stull, R. B. (1976c). Internal gravity waves gen- 
erated by penetrative convection. J. Atmos. Sci. 
33, IAT). 

Summers, P. W. (1965). An urban heat island model: 
Its role in air pollution problems with special 
application to Montreal. Proc. First Canadian 
Conf. Micrometeorology, Toronto, Atmos. Environ. 
Service. 

Tennekes, H. (1973a). Similarity laws and scale 
relations in planetary boundary layers. Workshop 
on Micrometeorology, D. A. Haugen, ed., AMS, 
Boston. 

Tennekes, H. (1973b). A model for the dynamics of 
inversion above a convective boundary layer. J. 
AEMOS a SCL SO) DD Sl. 


595 


Tennekes, H. (1975). Reply (to Zilitinkevich, S. 
Soo Mo Aewess Metis st, hq 

Thompson, S. M., and J. S. Turner (1975). Mixing 
across an interface due to turbulence generated 
by an oscillating grid. J. Fluid Mech. 67, 349. 

Tritton, D. J. (1977). Physical Fluid Dynamics, 
Van Nostrand Reinhold, New York. 

Turner, J. S. (1968). The influence of molecular 
diffusivity on turbulent entrainment across a 
density interface. J. Fluid Mech. 33, 639. 

Turner, J. S. (1973). Buoyancy Effects in Fluids, 
Chapter 9, Cambridge Univ. Press. 


Veronis, G. (1963). Penetrative convection. Astro- 
phys. J. 137, 641. 
Willis, G. F., and J. W. Deardorff (1974). A labora- 


tory model of the unstable planetary boundary 
layer. J. Atmos. Sci. 31, 1297. 

Wolanski, E. J. (1972). Turbulent entrainment across 
stable density-stratified liquids and suspensions. 
Ph.D. Thesis, The Johns Hopkins Univ., Baltimore, 
Md. 

Wolanski, E. J., and L. M. Brush (1975). Turbulent 
entrainment across stable density step structures. 
RELI) Ap PBa)c 

Zeman, O. and J. L. Lumley (1977). A second-order 
model for buoyancy driven mixed layers. Heat 
Transfer and Turbulent Buoyant Convection, Vol. l, 
D. Spalding and N. Afgan, ed., pp. 65-76. 

Zeman, O., and H. Tennekes (1977). Parameterization 
of the turbulent energy budget at the top of the 
daytime atmospheric boundary layer. J. Atmos. 
seit, S44, dali, 

Zilitinkevich, S. S. (1975). Comments on 'A Model 
for the Dynamics of the Inversion above a Con- 
vective Boundary Layer.' J. Atmos Sci. 32, 991. 


Laboratory Models of Double-Diffusive 
Processes in the Ocean 


J. 


Stewart Turner 


Australian National University 


Canberra, 


ABSTRACT 


There is now good observational evidence to support 
the ideas that double-diffusive processes, i.e., 
those for which the differential diffusion of heat 
and salt are important, can affect the rates of 
vertical transport of these properties in the ocean, 
and are responsible for the formation of certain 
types of microstructure. Much of our detailed 
understanding of these effects has come from related 
laboratory experiments, but new phenomena are still 
being discovered which are as yet untested by direct 
measurements in the ocean. It is the purpose of 
this paper first to review the background to this 


Australia 


subject, and then to describe the more recent experi- 


ments which suggest further double-diffusive effects 
likely to be significant in various oceanographic 
contexts. 

A convenient laboratory technique has been to 
use two solutes (commonly salt and sugar) to model 
the T-S variations; some of these experiments with 
closer diffusivities are in fact directly relevant 
to the ocean. When more than two diffusing compo- 
nents are present it has been shown that even small 
differences in molecular diffusivity can signifi- 
cantly affect the relative rates of transfer of 
solutes through an interface, and this should be 
considered more carefully in geochemical studies. 
Strong double-diffusive layering is often associated 
with large horizontal gradients of T and S, and 
related effects have been studied in our laboratory 
in three different geometries: the circulation 
produced by a block of ice in a salinity gradient; 
a line source of one fluid intruding at its own 
density level into a gradient with different prop- 
erties; and the spreading across a frontal surface 
separating two fluids having the same vertical 
density but different T-S structures. 


596 


1. INTRODUCTION 


It is past the stage when the relevance of double- 
diffusive effects has to be justified ab initio to 
an audience of oceanographers. Over the last few 
years, there have been many observations of fine- 
structure and microstructure in the deep ocean 
which can only be explained in these terms. Wherever 
there is a systematic association between T and S 
variations, with both properties increasing or 
decreasing together (so that their effect on the 
density is in opposite senses), then it is clear 
that the difference in molecular diffusivities for 
heat and salt can affect the vertical structure 
and the transports of the two properties. It is 
not then sufficient to base predictions of mixing 
on the net density distribution alone. 

Our understanding of these processes has been 
greatly influenced by related laboratory experiments 
[see Turner (1973, 1974)]. Much of the detailed 
work has concentrated on the properties of sharp in- 
terfaces separating relatively well-mixed layers: 
it has been shown that when there are compensating 
T-S gradients, a smoothly stratified water column 
typically breaks up into a series of steps, and 
molecular processes must be more important across 
such interfaces. Once layers have formed there 
remains little doubt that the coupled transports 
can be estimated using the laboratory results. It 
is much less certain, however, that the processes 
of formation of layers have always been adequately 
modelled in the laboratory, where most of the experi- 
ments have been one-dimensional in form. 

More recent experiments [Turner and Chen (1974), 
Huppert and Turner (1978), Turner (1978)] have begun 
to explore a variety of two-dimensional effects, and 
it is these which will be given most attention in 
the verbal presentation of this paper. It should be 


admitted right at the beginning that these experi- 
ments are still largely qualitative, and that much 
more remains to be done, but already they suggest 
new explanations of some existing observations in 
the ocean, and allow us to predict what might be 
measurable in future work. 


2. ONE-DIMENSIONAL PROCESSES 
Formation of Layers from a Gradient 


For completeness, the fundamental physics of the 
double-diffusive convection will be outlined briefly 
by referring to the simpler early experiments. The 
review of one-dimensional experiments will then be 
brought up to date and specific oceanographic 
examples of these processes will also be described. 
The necessary conditions for double-diffusive 
convection to occur in a fluid are firstly that 
there should be two or more components having 
different molecular diffusivities, and secondly 
that these components should make compensating 
contributions to the density. It is remarkable 
that under these conditions strong convective 
motions can arise even when the net density distri- 
bution increases downwards. The overall density is 
"statically stable' in this sense in all the cases 
described here. Motions are nevertheless generated 
since the action of molecular diffusion, at different 
rates for the two components, makes it possible to 
release the potential energy in the component which 
is heavy at the top. This can drive convection in 
relatively well-mixed layers, while the second 
(stably distributed) component preserves the density 
difference across the interfaces separating them. 
There are two cases to be considered, depending 
on the relation between the diffusivities and the 
density gradients, i.e., on whether the driving 
energy comes from the component having the higher 


FIGURE 1. 


Layering produced from an initially 
smooth salinity gradient by heating from below. 
Three well-mixed layers are marked by fluorescein 
dye, lit from the top. (Tank diameter, 300mm. ) 


597 


or lower diffusivity. The simplest example of 

the former is a linear stable salinity gradient, 
heated from below. An unstratified tank would over- 
turn from top to bottom, but because of the stabi- 
lizing salinity gradient only a thin temperature 
boundary layer is formed at first, which breaks 

down through an overstable oscillation [Shirtcliffe 
(1967)] to form a shallow convecting layer. This 
layer grows by incorporating fluid from the gradient 
above it, in such a way that the steps of S and T 
are nearly compensating, and there is no disconti- 
nuity of density, only of density gradient. 

When the thermal boundary layer ahead of the 
convecting region reaches a critical Rayleigh number, 
it too becomes unstable. A second layer then forms 
above, and eventually many other layers form in 
succession (See Figure 1). The vertical scale of 
these layers increases as the heating rate is 
increased, and decreases with larger salinity gra- 
dients. Turner (1968) has shown that the first 
layer stops growing when 


Je =D Nee (1) 


Here d, is the critical depth, D is a dimensional 
constant which depends on the critical Rayleigh 
number and the molecular properties, B = -gaFm/pC 
is the imposed buoyancy flux corresponding to a 
heat flux Fp (a being the coefficient of expansion 
and C the specific heat), and Ng = [(g/p) (dp/dz]% 
is the initial buoyancy frequency of the stabilizing 
salinity distribution. The criterion for the for- 
mation of further layers is currently being studied 
by Huppert and Linden (personal communication). 

A device which has proved very helpful in elim- 
inating uncontrolled sidewall heat losses (as well 
as providing results directly relevant to the ocean) 
is to carry out experiments with two solutes, say 
sucrose and sodium chloride solutions, instead of 
salt and heat. Essentially the same phenomena can 
be observed, although the diffusivities are much 
more nearly equal (the ratio T = Ksg/Kz, where Km 
denotes the larger and ks the smaller diffusivity 
in each case, is about 1/3 for sugar and salt, 
compared with = 1072 for salt and heat). 

Linden (1976) has in this way extended the 
"heated gradient" experiments to study the case 
where there is a destabilizing salt (T) gradient 
partially compensating the stabilizing sugar (S) 
gradient in the interior. He has shown, both 
theoretically and experimentally, that during the 
formation of layers the relative contributions of 
the energy provided by the boundary flux, and that 
released in the interior, change systematically 
with the ratio of the vertical T and S gradients. 
In the limit where these gradients become equal, 
all the energy comes from the destabilizing compo- 
nent in the interior, and the ultimate layer depth 
is finite and proportional to N,-% (where N, is the 
buoyancy frequency corresponding to the stabilizing 
component) . 

Once layers and interfaces have formed, it is 
important to understand what governs the fluxes of 
S and T across them. For this purpose two or more 
layers can be set up directly, and the interfaces 
examined using a variety of optical techniques. 

For example, Figure 2 is a shadowgraph picture of 
a very sharp interface formed between a layer of 
salt solution above a layer of sugar solution, which 
is equivalent to colder fresh water above hot salty 


water. Note that salt is here the analogue of heat, 


FIGURE 2. Shadowgraph picture of a sharp "diffusive" 
interface, formed between a layer of salt solution above 
a denser sugar solution. Note the convective plumes each 
side of the interface, evidence of strong interfacial 
transports. (Scale: the tank is 150mm. wide.) 


and sugar the analogue of salt, since in each case 
the convection is maintained, and the interface 

kept sharp, by the more rapid vertical transfer of 
the faster diffusing component. Such interfaces 
have been called "diffusive interfaces", for reasons 
which will become clearer in the following section. 


Fluxes through Diffusive Interfaces 
Quantitative laboratory measurements have been made 
of the S and T fluxes across the interface between 
a hot salty layer below a cold fresh layer, and 
they have been interpreted in terms of an extension 
of well-known results for simple thermal convection 
at high Rayleigh number. Explicitly, Turner (1965), 
Crapper (1975), and Marmorino and Caldwell (1976) 
have shown that the heat flux oF», (in density units) 
is described by 
473 

OF = AL (aAT) (2) 
where Aj has the dimensions of velocity. Fora 
specified pair of diffusing substances, A, is a 
function of the density ratio Ro=BAS/aAT, where 8 
is the corresponding "coefficient of expansion" 
relating salinity to density differences. The 
deviation of A, from the constant A obtained using 
solid boundaries, with a heat flux but no salt flux, 
is a measure of the effect of AS on Fm. When Rp 
is less than about 2, A, > A due to the increased 
mobility of the interface, and when Rp > 2, Ay 
falls progressively below A as R, increases and 
more energy is used to transport salt across the 
interface. The empirical form 


A,/A = 3.8 (BAS/aAT) 7 (3) 


{Huppert (1971)] provides a good fit to the obser- 
vations over the whole of the measured range 1.3< 
Rp <7. 

The salt flux also depends systematically on 
Rp, and has the same dependence on AT as does the 
heat flux. Thus the ratio of salt to heat fluxes 


(both expressed in density units) should be a 
function of Rp alone for given diffusing substances: 


BF./oF, = £, (8AS/aAT) (4) 


The results reproduced in Figure 3 support this 
relation, and they also reveal the striking feature 
that the flux ratio is substantially constant (=0-.15) 
for 2<R, <7. [The more recent experiments of 
Marmorino and Caldwell (1976) suggest that the flux 
ratio can be as high as 0.4 with much smaller heat 
fluxes, but the reason for this discrepancy is not 
yet resolved]. Experiments by Shirtcliffe (1973), 
using a layer of salt solution above sugar solution, 
have shown a much stronger dependence of Fm on Rp 
than (3), but again a constant flux ratio, the 
measured value (for NaCl and sucrose) being 

BF</OF p x 0.60. Note that the flux ratio must 
always be <l, for energetic reasons: the increase 
in potential energy of the driven component must 
always be less than that released by the component 
providing the energy. This implies that the density 
difference between two layers must always increase 
as a result of a double-diffusive transport between 
them. 

Direct measurements through the interface in 
Shirtcliffe's experiment suggest that this has a 
diffusive core, in which the transport is entirely 
molecular, and which is bounded above and below by 
unstable boundary layers. The "thermal burst" 
model of Howard (1964) has recently been extended 
to this two-component case by Linden and Shirtcliffe 
(1978), to predict both the fluxes and flux ratios. 
The constant range of flux ratio can be explained 
in the following way. Boundary layers of both T 
and S grow by diffusion to thicknesses proportional 
to Ken and Keer and then both break away intermit-— 
tently. If only the statically unstable part at 
the edge of the double boundary layer is removed 
(such that aAT=8AS), then the fluxes will be in 
the ratio tz, in reasonable agreement with the 
laboratory results for the two values of T used. 
Linden (1974a) has given a mechanistic argument to 
explain the increase of flux ratio at lower values 
of Ry, which he attributes to the direct entrainment 
of both properties across the interface. 

It is worth noting in passing that Huppert (1971) 


10 


0-8 


FIGURE 3. The ratio of the fluxes of salt and heat (in 
density units) across an interface between a layer of 
hot, salty water below colder, fresh water, plotted as 
a function of the density ratio R.. [From Turner 

(1965) .] " 


has shown theoretically that an intermediate layer, 
or series of layers, is stable if the overall S 
and T differences lie in the range where the flux 
ratio is constant, and unstable if the flux ratio 
varies with Rj. Observations of stable layers in 
the ocean seem to be consistent with this criterion. 
The merging of layers by this and other mechanisms 
has been studied experimentally by Linden (1976). 
Some measurements have also been made in the 
case where several solutes with different diffusivi- 
ties, Ky, are driven across an interface by heating 
from below. Turner, Shirtcliffe, and Brewer (1970) 
showed that the individual eddy-transport coeffi- 
cients can be different, and suggested that they 
are proportional to yee More recent work by 
Griffiths (personal communication) predicts theo- 
retically that the ratios of transports of pairs 
of solutes should be proportional to T2 at low 
solute-heat density ratios, and to Tt at higher 
ratios. His much more accurate and extensive 
experiments show an even larger variation, for 
reasons which are still unexplained. These results 
are potentially of great importance for the inter- 
pretation of geochemical data, as will be discussed 
further below. 


Observations of Diffusive Interfaces 


There are now many observations of layering in the 
ocean which can unambiguously be associated with 
"diffusive" interfaces, and where a one-dimensional 
interpretation seems appropriate. The regularity 
of the steps and the systematic increase of both 

S and T with depth serves to distinguish these from 
layers produced in other ways (by internal wave 
breaking, for example). Neal et al. (1969) and 
Neshyba et al. (1971) have observed layers about 

5 m thick, underneath a drifting ice island in the 
Arctic where cold fresh melt water overlies warm 
salty water. A common observation in Norwegian 
fjords is that cold fresh water, formed by melting 
snow, can often form a thin layer on top of warmer 
seawater, with an interface which remains extremely 
sharp, and thickens much less rapidly than expected. 
This is due to double-diffusive convection driven 
by the heat flux from below, which will stir the 
layers on each side of the interface (independently 
of any wind stirring at the surface) and thus keep 
the interface sharpened. 

There are also fresh-water lakes in various parts 
of the world which have become stratified in the 
past by the intrusion of sea water. Some of these 
are heated at the bottom by solar radiation, and 
convectively mixed layers separated by diffusive in- 
terfaces are formed. A particularly well-documented 
example is Lake Vanda in the Antarctic [Hoare (1968), 
Shirtcliffe and Calhaem (1968)]. Since these lakes 
are not complicated by horizontal advection pro- 
cesses, Huppert and Turner (1972) were able to 
use the Lake Vanda data to show that the one- 
dimensional laboratory result (3) can be applied 
quantitatively to comparable large-scale motions. 

Other striking examples are the multiple steps 
observed in a lake in the East African Rift zone, 
which is heated geothermally by the injection of 
hot saline water at the bottom [Newman (1976)], 
and the layers of hot salty water found at the 
bottom of various Deeps in the Red Sea [Degens and 
Ross (1969)]. These layers are nearly saturated 
with salts of geothermal origin, including a high 


599 


proportion of heavy metals, and are of special 
interest because of the potential commerical value 
of the associated thick sediments. [Another related 
application, to the genesis of ore deposits on the 
sea floor, has recently been proposed by Turner and 
Gustafson (1978)]. 

The existence of many components in these layers 
raises another question which should be explored 
more systematically in the oceanic context. 
Griffiths' laboratory measurements mentioned above 
indicate that different solutes are transferred 
across diffusive interfaces at different rates, 
depending on their molecular diffusivities. The 
"Mixing rate" for a tracer is thus not necessarily 
a good indicator of the transport of a major com- 
ponent if interfaces are important. In the absence 
of definite knowledge of the mixing mechanisms 
which have operated between the sources and the 
sampling point, the assumption that all components 
are mixed simultaneously (i.e., that a single "eddy 
diffusivity" should be used) seems likely to lead 
to large errors, and even to gross misinterpretations 
of geochemical data. 

Double-diffusive processes can also be important 
in other systems besides aqueous solutions. A 
situation of oceanographic interest arises if liquid 
natural gas (LNG) or some other liquid gas spills 
(following a tanker accident for instance) onto 
the sea surface [Fay and MacKenzie (1972)]. The 
liquid quickly evaporates to form a layer of cold 
gas, predominantly methane, which would be lighter 
than the air above it except that it is much colder. 
Since methane, and also water vapour picked up from 
the sea surface, have larger diffusivities than heat 
in air, double-diffusive effects can again be 
important in this gaseous system. The driving 
energy comes from the distribution of methane and 
water vapour, so the interface is "diffusive". 

The limited observations available suggest that 

the top of such a layer is very sharp, and its rate 
of spread vertically small, which is consistent 
with a self-stabilizing double-diffusive transport 
across the interface. Another application, to 
explain the phenomenon of "rollover" in LNG storage 
tanks, will not be described in detail here, but it 
too depends on double-diffusive effects, this time 
in the liquified gas [see Sarsten (1972)]. 


Salt Fingers and Related Phenomena 


We now turn to the second type of double-diffusive 
convection, that for which the driving energy is 
derived from the component having the lower molecular 
diffusivity. Though this is associated with the 
very different phenomenon of "salt fingers", there 
are many similarities between it and the "diffusive" 
case already presented, and these will be emphasized 
in the following discussion. 

When a small amount of hot salty water is poured 
on top of cooler fresh water, long narrow convection 
cells or "salt fingers" rapidly form. These motions 
were first predicted by Stern (1960) [and see Stern 
(1975) for a more up to date account of the theoret-— 
ical work]. They are sustained by the slower 
horizontal diffusion of salt relative to heat, which 
permits the release of the potential energy in the 
salt field. Again, fingers may be produced using 
two solutes with much closer diffusivities, and 
when there are strong contrasts of properties, the 
fingers are confined to an interface. Figure 4 


600 


FIGURE 4. 


Shadowgraph of a thickened "finger" inter- 
face, formed between a layer of sugar solution on top 
of salt solution. (Scale: the tank is 150mm. wide.) 


shows a shadowgraph picture of such an interface 
between a layer of sugar solution (S) above heavier 
salt solution (T). This is bounded by sharp edges, 
where the fingers break down and feed an unstable 
buoyancy flux into the convecting layers on either 
side. 

Finger interfaces between two such layers have 
been shown to thicken linearly in time [Stern and 
Turner (1969), Linden (1973)]. They have also been 
observed in plan by Shirtcliffe and Turner (1970) 
who showed that the convection cells have a square 
cross section, with upward and downward motions 
alternating in a close-packed array. The initial 
stability of an interface has been examined quanti- 
tatively by Huppert and Manins (1973). Whena 
layer of S is placed on a layer of T, the sharp 
boundary thickens by diffusion; the condition for 
formation of fingers depends on the magnitude of 
the gradients and the ratio of diffusivities, T, 
and is related to the overall differences by 


BAS/aAT > tv¥t . (5) 


These results can be extended to three components, 
as can the earlier linear stability theories 
[Griffiths (1978)]. For heat and salt, (5) shows 
that fingers should form with very small destabiliz- 
ing salinity differences, and suggests that they 
will be ubiquitous phenomena in the ocean. 

Our confidence in applying these results ona 
geophysical scale has recently been increased greatly 
by the direct observations of fingers (using an 
optical method) by Williams (1974, 1975) under 
conditions close to those predicted by Linden (1973) 
on the basis of laboratory results. Magnell (1976) 
has also measured horizontal conductivity variations 
with the right scale (=2cm.) to support this inter- 
pretation. 

As mentioned above, there is not as big a differ- 
ence between the "diffusive" and "finger" cases as 
there appears to be when we simply compare the 
interfaces illustrated in Figures 2 and 4. Layers 
can be produced from a smooth gradient in the latter 
case too, by supplying a flux of S at the edge of 
a gradient of T; this was first demonstrated, using 
a sugar flux above a salt gradient, by Stern and 


Turner (1969). When viewed on the scale of the 
convecting layers, there is in fact a close corres- 
pondence between the two systems. The inequality 
of diffusivities results in an unstable buoyancy 
flux across statically stable interfaces in both. 
cases, and this maintains convection above and below. 
Only the mechanism of interfacial transport differs, 
and it is here that the detailed structure of the 
interface enters. Across a finger interface the 
buoyancy flux is dominated by the destabilizing 
component, S, and salt is transported faster than 
heat, whereas the opposite holds a diffusive inter- 
face. 

Corresponding laboratory measurements of the 
two coupled fluxes have been made for finger inter- 
faces. Again there is a strong dependence on the 
density ratio across the interface, and the ratio 
of heat to salt fluxes is constant over a consider- 
able range. Turner (1967) has shown in the heat 
salt case that the salt flux is about 50 times as 
large at R)* = aAT/BAS + 1 as it would be if the 
same salinity difference were maintained at solid 
boundaries, and falls slowly as R,* increases. 

He also obtained a value for the flux ratio aFp/8F, 

= 0.56 over the range 2 < Ro* < 10. Linden (1973) 
has made direct observations of the structure of 

salt fingers and the velocity in them that support 
these estimates of the salt flux. His estimate of 
the flux ratio was much lower, but recent experiments 
in our laboratory have supported the earlier value. 
These new experiments have concentrated on achieving 
as small a value of Rp* as possible, but measurements 
in the "variable" range of flux ratio are still 
elusive. This range could, however, be of great 
importance in the ocean, where Ro* is often close 

to unity. 

It is also of interest to mention the experiments 
of Linden (1974b) who applied a shear across a 
salt-finger interface. He showed that a steady 
shear has little effect on the fluxes, though it 
changes the fingers into two-dimensional sheets | 
aligned down shear. Unsteady shears (i-.e., stirring 
on both sides of the interface) can, on the other 
hand, rapidly disrupt the interface, and actually 
decrease the salt flux. 

There are now many examples of layering in the 
ocean which are consistent with the "fingering" 
process. These are observed in situations where 
both the mean salinity and the temperature decrease 
with increasing depth, and often occur under warm 
salt intrusions of one water mass into another. 

The first observations were made by Tait and Howe 
(1968, 1971) under the Mediterranean outflow, and 

a good summary of other measurements is to be found 
in Fedorov (1976). For reasons which will be dis- 
cussed more fully in later sections, it is difficult 
to find cases where one can be sure that the forma- 
tion of layers bounded by finger interfaces has 

been the result of one-dimensional processes, 
strictly analogous to those studied in the labora- 
tory. Once layers have formed, however, the effects 


‘of the fluxes through the finger interfaces between 


them can properly be discussed in these terms, and 
two practical examples will be given. 

The first arises in the context of sewage disposal 
in the sea. Fischer (1971) has discussed the case 
where effluent, which can be regarded for this 
purpose as nearly fresh (though polluted) water, is 
ejected from a pipe laid along the bottom, and 
rises as a line plume into sea water which is strat-— 
ified in temperature. Careful design of the outfall 


ensures that the effluent, diluted with many times 
its volume of cold sea water, will spread out ina 
layer below the thermocline. But this layer will 
remain colder and fresher than the water above it, 
so the salt finger mechanism can cause it to thicken 
vertically, and even extend to the surface. A 
related case, in which the environmental effects 
could be even more serious, arises in the disposal 
of effluent from a desalination plant. Suppose 

that the brine from which water has been evaporated, 
and the heated water from the cooling plant, are 


mixed together to be disposed of as a single effluent. 


This hot, salty water will have about the same 
density as the original sea water - according to 
the precise design conditions, it can be slightly 
heavier or slightly lighter. If it is made heavier, 
and forms a layer along the bottom, a diffusive 
interface will be formed, and the coupled transports 
will tend to increase the density difference and 
thus keep the layer distinct. If it is put in at 
the surface, or at an intermediate level ina 
gradient, fingers will form, and there will be more 
rapid vertical mixing. One thing is certain: the 
rate of mixing cannot be determined using only the 
net density distribution and leaving out of account 
the double-diffusive effects. 


3. TWO-DIMENSIONAL EFFECTS 
Side-wall Heating and Related Processes 


It became clear in early laboratory experiments on 
double-diffusive convection that layers will readily 
form from a salt gradient in another way, if it is 
heated from the side. This effect was studied 
systematically by Thorpe, Hutt, and Soulsby (1969) 
and by Chen, Briggs, and Wirtz (1971), and their 
results can be summarized as follows. The thermal 
boundary layer at a heated vertical wall grows by 
conduction and begins to rise. Salt is lifted to 
a level where the net density is close to that in 
the interior; then fluid flows out away from the 
wall, producing a series of layers that form 
simultaneously at all levels and grow inwards from 
the boundary. The layer thickness is close to the 
length-scale 
aAT 

= Bas/da &) 
which is the height to which a fluid element with 
temperature difference AT would rise in the initial 
salinity gradient. 

The stability problem corresponding to sidewall 
heating of a wide container has not been solved, 
though Stern (1967) has shown theoretically how 
lateral gradients could lead to the generation of 
layers. Thorpe, Hutt, and Soulsby (1969) have 
analyzed the simpler case of a fluid containing 
compensating linear horizontal gradients of S and 
T, contained in a narrow vertical slot and Hart 
(1971) improved their analysis; both theories 
predict slightly inclined cells extending right 
across the gap, with a spacing in fair agreement 
with the measurements. 

Similar layers are formed when the salinity as 
well as the temperature of the vertical boundary 
does not match that in the interior, for example 
when a block of ice is inserted into a salinity 
gradient and allowed to melt. A qualitative experi- 
ment of this kind was reported by Turner (1975), 


601 


FIGURE 5. Showing the tilted layers formed by insert- 
ing a block of ice into salt-stratified water at room 
temperature. Fluorescein was frozen into the ice, and 
was illuminated from the side, so that the spread of 
the dye indicates the distribution of the melt water. 
(Negative print.) 


but interest in the process has increased recently, 
because of the application to melting icebergs. 
Huppert and Turner (1978) have carried out a more 
extensive set of experiments with this problem in 
mind. 

An understanding of the melting of icebergs 
could be important in various contexts. Several 
groups are currently examining the feasibility of 
towing icebergs to their coasts and melting them 
to provide fresh water, but there are many unsolved 
scientific and engineering problems [see, for 
example, Bader (1977)]. It has been proposed that 
fresh water could be obtained by building a shallow 
pen round a grounded iceberg, allowing the melt 
water to collect in this, and siphoning it off the 
surface. On the other hand Neshyba (1977) has 
suggested that the melt water produced by icebergs 
would mix with the surrounding sea water, and could 
thus be effective in lifting water and nutrients 
from deeper layers to the surface, where it would 
increase biological production. 

Huppert and Turner's (1978) experiments have 
shown, however, that neither idea is likely to be 
valid, because of the neglect of the stable salinity 
gradient which exists in the upper layers of the 
oceans where icebergs are found. As demonstrated 
in Figure 5, the presence of horizontal S and T 
differences then produces a regular series of tilted 
convecting layers, which feed most of the meltwater 
into the interior; very little rises to the surface. 
A more detailed analysis of the experiments is 
continuing. At present it appears that for a 
cooled sidewall the layer depths are similar whether 
melting is occurring or not, and that they are not 
described simply by (6) but depend more weakly on 
the initial salinity gradient. Another phenomenon 
which deserves more careful study is the series of 
grooves and ridges produced by non-uniform melting 
associated with the circulation in the layers (see 
Figure 6). 


Sloping Boundaries 


Phenomena analogous to those described above can 
be observed in systems containing smooth gradients 
of more slowly diffusing solutes. The essential 
physical feature of the heated sidewall process is 


FIGURE 6. 


Shadowgraph photograph of a melting ice- 
block in a salinity gradient. Note the regularly spaced 
scallops and ridges, caused by uneven melting asso- 
ciated with the convection in layers. 


that the boundary conditions (on temperature or 
salinity or both) do not match the conditions in 

the interior. In tanks containing opposing gradients 
of two components, with say a maximum salt concen- 
tration at the top falling linearly to zero at the 
bottom, and a maximum (slightly larger) sugar 
concentration at the bottom falling to zero at the 
top, the same kind of instability can be produced 

in another way. With vertical side walls, the 
surfaces of constant concentration are normal to 

the boundaries, and the no-flux boundary condition 

is automatically satisfied. But when an inclined 
boundary is inserted, diffusion will distort the 
surfaces of constant concentration away from the 
horizontal, so that they become normal to the 
boundary. Density anomalies are produced which 

tend to drive flows along the wall; these cannot 
remain steady, but instead turn out into the interior 
and produce a series of layers. 

This process was first investigated experimentally 
by Turner and Chen (1974), with the initial strati- 
fication in the "diffusive" sense. A prominent 
feature of the intruding layers is the local reversal 
of gradients in the extending "noses", where fingers 
are prominent. In the later stages of that experiment, 
the advancing noses have become independent of the 
mechanism which produced them, and this suggested 
the systematic study of double-diffusive sources 
in various environments which is pursued below. 
Linden and Weber (1977) have investigated layer 
formation in the "finger" case; they have also 
discussed the instability of the boundary layer 
at the sloping wall, and the criteria determining 
the layer depths. In the limit where the opposing 
gradients are nearly equal, the characteristic 
vertical lengthscale depends mostly on the initial 
vertical distributions of S and T, and little on 
the mechanism triggering the instability. 

A different effect of sloping boundaries should 


be mentioned here. In a two-layer system, in which 
the layer depths vary because one wall of the 
containing vessel is inclined, large-scale quasi- 
horizontal motions can be set up even when the 
buoyancy flux across the horizontal interface is 
uniform. This effect is a purely geometrical 
consequence of the sloping boundary. The net result 
of the double-diffusive transports across the 
interface is to provide an unstable buoyancy flux 
which makes the bottom layer heavier. A given flux 
produces more rapid density changes in shallower 
regions where there is less dilution, and this sets 
up a circulation in the sense which includes a flow 
down the slope. Gill and Turner (1969) have shown 
that this flow can reverse the relative gradients 

of the two components, for example, giving rise to 
salt fingers at the bottom of a tank originally 
stratified in the diffusive sense. They have also 
suggested an application to the formation of bottom 
water near the Antarctic continent. Similar effects 
have been observed Ly Turner and Chen (1974) when 

a sloping interface, rather than a solid sloping 
boundary, produces the non-uniformity of depth, 

and this too can have implications for the formation 
of bottom water in deeper water. 


Double-diffusive Intrusions 


The experiments described in the two preceding 
sections have recognized the importance of horizontal 
gradients, but they still have not dealt with the 
common situation where fluid with one set of T-S 
properties intrudes into another having different 
properties. This question has recently been 
addressed by Turner (1978), using sources of sugar 
and salt solutions released into gradients of 
various kinds. 

The basic intrusion process with which other 
phenomena can be compared is the two-dimensional 
flow of a uniform fluid at its own density level 
into a linear gradient set up using the same property. 
Figure 7 shows the behaviour of a (dyed) source of 
salt solution released into a salinity gradient. 
This is what we might intuitively expect: the 
intruding fluid just displaces its surroundings 
upwards and downwards, and is kept confined to a 
horizontal layer by the denisty gradient. Detailed 
studies of this process have been reported by 
Maxworthy (1972), Manins (1976), and Imberger, 
Thompson, and Fandry (1976). Note praticularly the 
"upstream wake" effect, leading to a considerable 
disturbance of the environment ahead of the advancing 
nose. 

When the source of salt is replaced by sugar 
solution (S), while the same salinity gradient (T) 
is retained in the environment, the behaviour is 
very different. (It is worth keeping in mind 
throughout the following, the analogous situation 
with temperature and salinity: this corresponds 
to the intrusions of a layer of warmer, saltier 
water into a stable temperature gradient). As 
shown in Figure 8, there is strong vertical convec- 
tion near the source: this is produced by a 
mechanism which also occurs with a uniform ambient 
fluid close to the same density as that injected. 
The more rapid diffusion of T relative to S across 
the plume boundary causes it to become heavier, 
and its immediate surroundings lighter, than the 
fluid at the level of the source. The vertical 
spread is limited by the stratification, and "noses" 
begin to spread out at levels above and below the 


source. The process of vertical convection continues, 
and further layers appear as the layers first formed 
extend away from the source. The total volume of 
fluid affected by mixing is many times that of the 
input, showing that the intrusions are overtaking 
and incorporating the environment, rather than 
just displacing it as in the experiment of Figure 
7. The implication for the ocean is, of course, 
that large scale intrusions will tend to break up 
into thinner noses and layers, as is indeed observed. 

Each individual nose as it spreads contains an 
excess of S relative to its environment, so that 
conditions are favourable for the formation of a 
diffusive interface above and fingers below, as can 
be seen in Figure 8. This also implies that there 
will be a local decrease with depth or an inversion 
of T through each layer, and that the density 
gradient above such an intrusion will be greater 
than that below. These features have been demon- 
strated in oceanic data by Howe and Tait (1972), 
Gregg (1975), and Gargett (1976). 

Note too the slight upward tilt of each layer 
as it extends, which can be interpreted as follows. 
Above and below an intrusion, the net density 
differences are small and the double-diffusive 
fluxes therefore large. The one-dimensional labo- 
ratory observations indicate that the transports 
across a finger interface (both in the sugar-salt 


FIGURE 8. The flow produced by releasing 
Sugar solution at its own density level into 
a salinity gradient. Strong vertical convec- 
tion occurs, followed by intrusion at several 
levels. The density gradient and flow-rate, 
and the scale of the photograph, are approxi- 
mately the same as for Figure 7. 


603 


FIGURE 7. The intrusion of dyed salt solu- 
tion into a salinity gradient at its own 
density level. The distorted dye streaks 
show that the fluid in the environment be- 
gins to flow well ahead of the advancing 
fluid. (The region shown is about 400mm. 
wide.) 


and salt-heat case) are larger than those across 

a comparable diffusive interface. Thus the flux 

of positive buoyance through the fingers from below 
can exceed the negative flux from above, so a layer 
becomes lighter and rises across isopycnals as 

it advances away from the source. There is also 

a systematic shear flow associated with the inclined 
layers, and both these features would seem worth 
looking for when observations are made of oceanic 
finestructure in the future. 

The interpretation of the layer slope in terms 
of the differences in fluxes across the two inter- 
faces is supported by experiments carried out in 
the inverse sense. With a source of salt solution 
(T) flowing at its own density level into a gradient 
of sugar solution (S), the behaviour is as shown 
in Figure 9. Vertical convection near the source 
is again followed by the spread of noses at various 
levels, but now with diffusive interfaces below 
and fingers above, corresponding to the excess of 
T in the noses relative to their S environment. 
There is a systematic downward tilt as the noses 
advance, due again to the dominance of the buoyancy 
flux at the finger interfaces, which now causes 
the layers to become heavier as they extend. The 
sense of the internal shear is also consistent 
with this picture: the motion is inclined slightly 
down and away from the source at the bottom of the 


604 


FIGURE 9. The flow produced by releasing salt 
solution into a gradient of sugar solution, 
using conditions comparable with, but the in- 
verse of those shown in Figure 8. 


fingers and above the diffusive interfaces, indica- 
ting again that there is an increase in density due 
to the continuing flux in the fingers. 

Two other features of the laboratory observations 
which have important implications for the ocean 
should also be mentioned. The most rapid formation 
of layers in the series of experiments reported by 
Turner (1978) occurred when the tank was stratified 
in the "finger" sense, and the fingers were allowed 
to run down towards a marginally stable state. 

When source fluid was introduced, layers formed 
more rapidly and regularly than before, because of 
the potential energy already available in the 
ambient fluid. This implies that "reactivation" 

of layers in a region where they have previously 
formed will proceed more quickly than the original 
layering process. It suggests that the patches of 
strong layers, under the Mediterranean outflow 

into the Atlantic for example, are associated with 
the arrival of a fresh pulse of intruding fluid. 

The second related observation is that the further 
stage of overturning to produce nearly uniformly- 
mixed layers, bounded above and below by finger 
interfaces is also more likely to be reached near 
the source of the intruding water. The relationship 
between the two types of layering has been demon- 
strated directly in the measurements of Gregg (1975), 
which show that inversions of intrusive origin can 
in the course of time break down to form well-mixed 
layers. 


Layer Formation at Fronts 


An important geometry which merits separate study 
is a discontinuity of T-S properties over a vertical 
or inclined surface, i.e., a front. The motions 
produced when an inclined boundary is inserted into 
a fluid stratified with opposing gradients (Section 
3) have some of the required features, but the 
presence of the solid wall is clearly undesirable. 
Fronts can be set up in the laboratory in several 
ways. Large horizontal T and S gradients can, for 
example, be produced just by pouring fluid with 
contrasting properties into one end of a stratified 
tank at several levels, or by stirring it in 
throughout the depth. A somewhat more controllable 
method is to insert a vertical barrier in a previ- 
ously stratified tank, to introduce the extra fluid 
on one side of it, and allow the disturbances to 


This tech- 


die away before removing the barrier. 
nique has been used in the experiment shown in 


Figure 10. It is difficult to get the two vertical 
gradients exactly matched, and so when the barrier 
is removed internal waves are set up, which soon 
die away, leaving the isopycnals horizontal but 

the front distorted. The initial state illustrated 
in Figure 10a is completely determined by the 
readjustment of the density field, but note that 
diffusive interfaces have already developed in the 
sense to be expected with an excess of S on the 
left. At a later stage (Figure 10b) the frontal 
surface is spread out horizontally by the inter- 
leaving of inclined layers, the scale of which is 
unrelated to that of the initial adjustment, and 
which are driven entirely by the local density 
anomalies produced by double-diffusive transports. 

A more sophisticated version of this experiment 
is currently being studied by Ruddick (personal 
communication). He has set up identical vertical 
density distributions on two sides of a barrier, 
using sugar (S) in one half and salt (T) in the 
other. When the barrier is withdrawn, there is 
some small scale mixing, but virtually no larger 
scale distortion. A series of regular, interleaving 
layers then develops, with a spacing and speed of 
advance which are systematically related to the 
horizontal property differences. 

There are now many. measurements which support 
the view that the prominence and strength of 
layering in the ocean are related to the magnitude 
of the horizontal gradients of properties. To 
cite just two examples: profiles across the Antarc- 
tic polar front [Gordon et al. (1977)] reveal 
inversions which decrease in strength with increasing 
distance away from the front. Coastal fronts 
between colder fresh water on a continental shelf 
and warmer salty water offshore also exhibit strong 
interleaving [Voorhis, Webb, and Millard (1976)]. 

A general conclusion which can already be drawn 
from the laboratory experiments described in this 
section is that the formation and propagation of 
interleaving double-diffusive layers is a self- 
driven process, sustained by local density anomalies 
due to double-diffusive transports. Once a series 
of noses and layers has formed, the changes of T 
and S within them can be described in terms of the 
one-dimensional (vertical) transport processes 
previously studied. It should eventually be possible 


[Joyce (1977)] to parameterize the effective 
increase in the horizontal diffusion of T and S, 
produced by interleaving, in terms of the horizontal 
gradients and these quasi-vertical fluxes. 


ACKNOWLEDGMENT 


I am grateful to R. Wylde-Browne for his assistance 
with many of the recent experiments described here, 
and particularly with the photography. Discussions 
with R. W. Griffiths and B. R. Ruddick about their 
current work have been very stimulating. 


REFERENCES 


Bader, H. (1977). A critical look at the iceberg 
project. First International Conference on 
Iceberg Utilization, Iowa State University. 

Chen CoE DAmGHEBralggsmand Re Al Wastzan(lhOv/as)ir 
Stability of thermal convection in a salinity 
gradient due to lateral heating. Int. J. Heat 
and Mass Trans. 14, 57-65. 

Crapper, P. F. (1975). Measurements across a 
diffusive interface. Deep-Sea Res. 22, 537-545. 

Degens, E. T, and D. A. Ross (eds.) (1969). Hot 
brines and recent heavy metal deposits in the 
Red Sea. Berlin: Springer-Verlag. 


605 


FIGURE 10. Showing the spread of a front, 

formed by withdrawing a barrier from the center 

of an experimental tank. The initial state cor- 
responded to "run down" fingers, with extra 

(dyed) sugar solution mixed in at the left. 

Upper: 2-1/2 mins. after barrier was removed, 
showing intrusions formed by the den- 
sity adjustment. 

Lower: 65 mins. after barrier removed, showing 
the smaller scale motions driven by 
double-diffusive processes across the 
frontal region. 


Fay, J. A., and J. J. MacKenzie (1972). Cold cargo. 
Environment 14, 21-29. 

Fedorov, K. N. (1976). The thermohaline finestruc-— 
ture of the ocean. Gidrometeoizdat, Leningrad. 

Fischer, H. B. (1971). The dilution of an undersea 
sewage cloud by salt fingers. Water Research 
57 DOP—Gi5. 

Gargett, A. E. (1976). An investigation of the 
occurrence of oceanic turbulence with respect to 
finestructure. J. Phys. Oceanog. 6, 139-156. 

Gilly A Heeand. Jie Se Turner (969) Some mnew 
ideas about the formation of Antarctic bottom 
water. Nature 224, 1287-1288. 

Gregg, M. C. (1975). Microstructure and intrusions 
in the California current. J. Phys. Oceanog. 5, 
253=2 15 

Griffiths, R. W. (1978). The influence of a third 
diffusing component upon the onset of convection. 
J. Fluid Mech. (in the press) 

Gordon, A. L., Ds i. Georgi, and! H.) We Taylor (977) 
Antarctic polar front zone in the western Scotia 
Sea - summer 1975. J. Phys. Oceanog. 7, 309-328. 

Hart, J. E. (1971). On sideways diffusive instabil- 
ity. J. Fluid Mech. 49, 279-288. 

Hoare, R. A. (1968). Thermohaline convection in 
Lake Vanda, Antarctica. J. Geophys. Res. 73, 
607-612. 

Howard, L. N. (1964). Convection at high Rayleigh 
number. Proc. eleventh Int. Congress Applied 


606 


Mechanics, Munich (ed. H. Gértler), 1109-1115. 
Berlin: Springer-Verlag. 
Howe, M. R., and R. I. Tait (1972). The role of 
temperature inversions in the mixing processes 
of the deep ocean. Deep-Sea Res. 19, 781-791. 
Huppert, H. E. (1971). On the stability of a 


series of double-diffusive layers. Deep-Sea 
Res. 18, 1005-1021. 
Huppert, H. E., and P. C. Manins (1973). Limiting 


conditions for salt-fingering at an interface. 
Deep-Sea Res. 20, 315-323. 

Huppert, H. E., and J. S. Turner (1972). Double- 
diffusive convection and its implications for 
the temperature and salinity structure of the 
ocean and Lake Vanda. J. Phys. Oceanog. 2, 
456-461. 

Huppert, H. E., and J. S. Turner (1978). 
icebergs. Nature 271, 46-48. 

Imberger, J., R. Thompson, and C. Fandry (1976). 
Selective withdrawal from a finite rectangular 
tank. J. Fluid Mech. 78, 489-512. 

Joyce, T. M. (1977). A note on the lateral mixing 
of water masses. J. Phys. Oceanog. 7, 626-629. 

Linden, P. F. (1973). On the structure of salt 
fingers. Deep-Sea Res. 20, 325-340. 

Linden, P. F. (1974a). A note on the transport 
across a diffusive interface. Deep-Sea Res. 21, 
283-287. 

Linden, P. F. 


On melting 


(1974b). Salt fingers in a steady 
shear flow. Geophys. Fluid Dynam. 6, 1-27. 

Linden, P. F. (1976). The formation and destruction 
of fine-structure by double-diffusive processes. 
Deep-Sea Res. 23, 895-908. 

Linden, P. F., and T. G. L. Shirtcliffe (1978). The 
diffusive interface in double-diffusive convection. 
J. Fluid Mech. (in the press) 

Linden, P. F., and J. E. Weber (1977). ‘The formation 
of layers in a double-diffusive system with a 
sloping boundary. J. Fluid Mech. 81, 757-774. 

Magnell, B. (1976). Salt fingers observed in the 
Mediterranean outflow region (34°N, 11°W) using 
a towed sensor. J. Phys. Oceanog. 6, 511-523. 

Manins, P. C. (1976). Intrusion into a stratified 
fluid. J. Fluid Mech. 74, 547-560. 

Marmorino, G. O., and D. R. Caldwell (1976). 
librium heat and salt transport through a 
diffusive thermohaline interface. Deep-Sea Res. 
23, 59-68. 

Neal, V. T., S. Neshyba, and W. Denner (1969). 
Thermal stratification in the Arctic Ocean. 
Science 166, 373-374. 

Neshyba, S. (1977). Upwelling by icebergs. 
267, 507-508. 

Neshyba, S., V. T. Neal, and W. Denner (1971). 
Temperature and conductivity measurements under 
Ice Island T-3. J. Geophys. Res. 76, 8107-8120. 

Newman, F. C. (1976). Temperature steps in Lake 
Kivu: a bottom heated saline lake. J. Phys. 
Oceanog. 6, 157-163. 

Sarsten, J. A. (1972). 


Equi- 


Nature 


LNG stratification and 


rollover. Pipeline and Gas Journal. Sept. 
1972, 37-39. : 

Shirtcliffe, T. G. L. (1967). Thermosolutal convec- 
tion: observation of an overstable mode. Nature, 


213, 489-490. 
Shirtcliffe, T. G. L. (1973). Transport and profile 
measurements of the diffusive interface in 


double-diffusive convection with similar diffusi- 
vities. J. Fluid Mech. 57, 27-43. 

Shirtcliffe, T. G. L., and I. M. Calhaem (1968). 
Measurements of temperature and electrical 
conductivity in Lake Vanda, Victoria Land, 
Antarctica. WN. Z. J. Geol. Geophys. 11, 976-981. 

Shirtcliffe, T. G. L., and J. S. Turner (1970). 
Observations of the cell structure of salt 
fingers. J. Fluid Mech. 41, 707-719. 

Stern, M. E. (1960). The ‘salt fountain' and 
thermohaline convection. Tellus 12, 172-175. 

Stern, M. E. (1967). Lateral mixing of water 
Masses. Deep-Sea Res. 14, 747-753. 

Stern, M. E. (1969). Collective instability of 
salt fingers. J. Fluid Mech. 35, 209-218. 

Stern, M. E. (1975). Ocean circulation physics. 
Academic Press, pp. 246. 

Stern, M. E., and J. S. Turner (1969). Salt fingers 
and convecting layers. Deep-Sea Res. 16, 
497-511. 

Tait, R. I., and M. R. Howe (1968). Some observa- 
tions of thermohaline stratification in the deep 
ocean. Deep-Sea Res. 15, 275-280. 

Tait, R. I., and M. R. Howe (1971). 
staircase. Nature, 231, 178-179. 

Thorpe, S. A., P. K. Hutt, and R. Soulsby (1969). 

The effect of horizontal gradients on thermohaline 
convection. J. Fluid Mech. 38, 375-400. 

Turner, J. S. (1965). The coupled turbulent trans- 

ports of salt and heat across a sharp density 


Thermohaline 


interace. Int. J. Heat and Mass Trans. 8, 
759-767. 

Turner, J. S. (1967). Salt fingers across a density 
interface. Deep-Sea Res. 14, 599-611. 


The behaviour of a stable 
J. Fluid 


Turney ist rol LOOS is 
salinity gradient heated from below. 
Mech. 33, 183-200. 

Turner, J. S. (1973). Buoyancy effects in fluids. 
Cambridge University Press, 367 pp. 

Turner, J. S. (1974). Double-diffusive phenomena. 
Ann. Rev. Fluid Mech. 6, 37-56. 

Turner, J. S. (1975). Laboratory experiments on 
double-diffusive instabilities. Adv. Chem. 
Phys. 32, 135-149. 

Turner, J. S. (1978). 
into a density gradient. 
the press) 

Turner, J. S., and C. F. Chen (1974). ‘Two-= 
dimensional effects in double-diffusive convec-— 
tion. J. Fluid Mech. 63, 577-592. 

Turner, J. S., and L. B. Gustafson (1978). The 
flow of hot, saline solutions from vents in the 
sea floor - some implications for exhalative 
sulfide deposits. Econ. Geol. (in the press) 

Turners, Ji. Si, Le Ga Le. Shirtclitre\andmebesGe 
Brewer (1970). Elemental variations of trans- 
port coefficients across density interfaces in 
multiple-diffusive systems. Nature 228, 1083- 
1084, 

Voorhis, A. D., D. C. Webb, and R. C. Millard (1976). 
Current structure and mixing in the shelf-slope 
water front south of New England. J. Geophys. 
Res. 81, 3695-3708. 

Williams, A. J. (1974). 
the Mediterranean out-flow. 


Double-diffusive intrusions 
J. Geophys. Res. (in 


Salt fingers observed in 
Science 185, 


941-943. 
Williams, A. J. (1975). Images of ocean micro- 
structure. Deep-Sea Res. 22, 811-829. 


Buoyant Plumes in a Transverse Wind 


Chia-Shun Yih 


The University of Michigan 


Ann Arbor, Michigan 


ABSTRACT 


With the rise in energy needs and the consequent 
proliferation of cooling towers (not to mention 
smoke stacks) on the one hand, and society's 
enchanced concern with the environment on the other, 
the study of buoyant plumes caused by heat sources 
in a transverse wind has become important. Buoyant 
plumes may also occur in the ocean, such as when 

a deeply submerged heat source moves horizontally 
in it. The fluid mechanics involved in buoyant 
plumes is very nearly the same, be they atmospheric 
or submarine. 

In this paper a similarity solution for turbulent 
buoyant plumes due to a point heat source in a 
transverse wind is presented. By a set of trans- 
formations the mathematical dimension of the 
problem is reduced from 3 to 2. Analytical solutions 
for the first and second approximations are obtained 
for the temperature and velocity fields. The 
solution exhibits the often observed pair of longi- 
tudinal counter-rotating vortices. As a result of 
buoyancy, the point of highest temperature and the 
"eyes" of the vortices at any section normal to 
the wind direction continuously rise as the longi- 
tudinal distance from the heat source increases. 


1. INTRODUCTION 


As industry expands and energy needs rise, the 
buoyant plumes caused by ever-increasing cooling 
towers and smoke stacks have become an important 
concern for societies anxious to protect their 
environment. Much effort has been expanded on the 
so-called numerical modeling of the phenomenon 

of plumes both in the United States and in Europe. 
In most of the numerical studies, the eddy viscosity 
is assumed constant, and its value is chosen to 
make the results agree with whatever gross observa- 
tions are available. The power of modern computers 
has made it possible to obtain numerical solutions 


607 


for partial differential equations with very 
irregular data, such as wind and temperature profiles 
in the atmosphere. On the other hand, one can 

only carry out a number of these special solutions, 
and while the power of the computer makes computa- 
tion possible it also makes the intermediate steps 
so opaque that one can only have faith in the 
accuracy of the results and the correctness of the 
programing; and one can attempt to interpret the 
results and understand the phenomenon only at the 
very end, when numerical results are available. 

One can hardly see, for example, the effects of 
changing one single parameter of the problem, without 
giving that parameter several values and going to 
the computer again and again. It is in view of 

this condition that even people most concerned with 
the immediate applicability of calculated results 
desire a certain measure of transparency in the 
analysis of the phenomenon. 

At the same time systematic and detailed experi- 
ments on buoyant plumes in transverse winds, with 
temperature and velocity measurements, are lacking. 
This being so, it seems that an analytical solution 
of the problem is most desirable and timely, even 
if it must of necessity be constructed by assuming 
certain quantities (such as the turbulence level 
in the plume) on the basis of whatever related 
experimental results are available. The assumed 
quantities (or quantity) will appear in the analysis 
as unspecified coefficients (or coefficient, as in 
this analysis), to be determined by experiments 
later. In the present work only one coefficient 
related to the turbulence level is left unspecified, 
to be determined by future experiments. But the 
probable range in which it lies is given. 

The solution is based on a set of transformations 
that reduces the mathematical dimension of the 
phenomenon from 3 to 2 is thus characterized by the 
striking feature of similarity between cross sections 
normal to the wind direction. The laws of decay 
of the temperature and velocity fields are given 
in simple, explicit terms. Thus, apart from the 


608 


quantitative predictions that this analysis is 
intended to furnish, I hope that the general features 
of the solution will be found especially useful. 


2. THE DIFFERENTIAL EQUATIONS 


The two basic assumptions underlying the analysis 
are that the longitudinal velocity component in 
the direction of the wind is constant and that an 
eddy viscosity, €, is constant in any cross section 
normal to the wind direction. It can be shown that 
the first assumption ceases to be true only at 
stages of approximations later than those arrived 
at in the present analysis, and its violation is 
therefore not very important. The second assumption 
mentioned above has been made in all analytic 
solutions for turbulent jets and plumes, according 
to Prandtl's simplified theory. These solutions 
are well known. See, for example, the paper by 
Yih (1977) on turbulent plumes for the latest 
application of that theory. One feels reassured 
that for a calculation of the mean temperature and 
velocity fields, this theory can again be used. 

We shall take the direction of the wind to be 
the direction of increasing x, and the z direction 
to be vertically upward. The y direction will then 
be a horizontal direction transverse to the x direc- 
tion. In general ¢ depends on x, y, and z. But it 
has been repeatedly shown before in other studies 
of jets and plumes that in their core, € can be 
taken as constant at a constant value of x, and 
that only at their outer edges does the nonuniformity 
in the y-z plane introduce some errors in the 
calculated mean quantities. (Very far away from 
the jets and plumes the value of € is immaterial 
for the determination of the temperature and velocity 
distributions). Accepting these outer-edge errors, 
which are fairly small, we shall take € to be a 
function of x only, apart from the parameters of 
the problem to be defined later. We note that if 
an eddy viscosity is used to determine the velocity 
distribution in turbulent flow in a circular pipe, 
Laufer's (1953) measurements show that in the core, 
that is, away from the narrow region near the pipe 
wall, € is nearly constant. 

The equations of motion are then, with subscripts 
denoting partial differentiations, 


Wii ar WAI SP ni = O 


al 
x y Zz p Py + E(Vyy a7 View (1) 


), (2) 


Uw, + vw, + wwo = - e = 60) ap eX te 
m8 y A 0 Pz GI e Wy Woz 


in which U is the wind velocity, assumed constant, 

v and w are the velocity components in the directions 
of increasing y and z, respectively, p is the 
density, p is the pressure, and g is the gravita- 
tional acceleration. The variable 8 is defined by 


Ap 


a=", 
- (3) 


where Ap is the variation of the density from the 
ambient density p, assumed constant. Thus the 
Boussinesq approximation has been used in Eqs. (1) 
and (2). Since § is small and the pressure vari- 
ation in the plume, though important for determina- 
tion of the flow field, is unimportant in the deter- 
mination of Ap from the temperature variation by 

the equation of state, 8 can also be written, by 
virtue of the equation of state of ideal gases. 


oa 

T 
where AT is the temperature variation and T the 
ambient temperature. For a liquid, the relationship 
between Ap and AT is still linear if 8 is small, 
and the constant. of proportionality is determined 
by the property of the liquid. 

We shall assume the eddy viscoity for heat 
diffusion to be the same as that for momentum 
diffusion. This may not be strictly true, for the 
turbulent Prandtl number may be slightly different 
from 1. The effect of this difference, if any, is 
not of great importance in our attempt to determine 
the mean temperature and velocity fields. The 
equation for heat diffusion can then be written in 
the form 


+ wO, = EN + 6 (4) 


x y y zz)- 


Longitudinal diffusion of heat or of momentum is 
ignored in Eq. (1), (2), and (4). This is justified 
in the same way as in other works that use the 
boundary-layer theory. 

The equation of continuity is, since the longi- 
tudinal velocity component is assumed constant, 


Yoo WS Oe (5) 


The heat source, located at the origin, is 
measured by the quantity 


UOdydz. (6) 


Note that solid boundaries are assumed to be far 
away from the source, so that their effects are 
negligible. Equations (1), (2), (4), (5), and (6), 
with appropriate boundary conditions, govern the 
phenomenon under investigation. 

The equation of continuity (5) allows the use 
of a stream function | in terms of which v and w 
can be expressed: 


V =z, w = -Wy. (7) 


By cross-differentiation of Eqs. 
obtain the vorticity equation 


(1) and (2), we 


Wiese or vey + wi, = € (Evy AP So) = 980 (8) 
in which é is the x component of the vorticity and 
is given by 


E=w 


ne te Wy S Vado (9) 


3. THE FORM OF THE EDDY VISCOSITY 


We assume the terms in Eqs. (1) or (2) or (4) to 

be of the same order of magnitude. In particular, 
this means that the diffusive and the convective 
terms are of the same order of magnitude in any of 
these equations. It also means that in Eq. (2) the 
buoyancy term is of the same order of magnitude as 
the convective and diffusive terms for w. This 
assumption underlies all existing analytical studies 


of jets and plumes and can be regarded as amply 
justified. 

Comparing the first and last terms in Eq. (2), 
then, we have 


(10) 


in which 2, and 2, are the length scales for the 
x and z directions. Comparing the first term in 
Eq. (2) with the term g@, we have 


Uw 


§@ ~ ——, (11) 
gh, 


where 9 and w stand for the magnitudes of 9 and w, 
rather than 8 and w rigorously, as they do also 
in the following proportionalities. Equation (6) 
gives, further, 


8 Nea w 8, (12) 
2 
UL, 
if we take 2, and 2, to be equal. From proportion- 
alities (11) and (12) we have, after some rearrange- 
ment, 


gGr 
Wy = (13) 
UL, 
But surely 
Se ~ Woo (14) 
Hence 
gGe 
x 
Coy : (15) 
UL, 


From proportionalities (10) and (15) we have 


G G 
Oi ta ar (16) 
U U 
since the doer the scale of x, is just x. Thus (12), 
(13), (14), and (16) give 
Le a x2/3, G0 xl/3, w ~ alae 8 ~ ee 


These results are unaffected when other comparisons 
are made between terms in either Eq. (1), (2), (4), 
@xe ((5)) o 

From porportionalities (15) and (16) we have 


a= Se? : (17) 


where a is a dimensionless constant to be determined 
experimentally or estimated from known values of 

€ in similar phenomena. We shall leave it free 
throughout our analysis. Equation (17) gives the 
form of € to be used in this paper. 

It seems strange at first sight that € should 
vary inversely as U. I believe that the interpre- 
tation of € ~ U7l is that € increases with the 
time that is required for the wind to travel a unit 
distance in the x direction, because turbulence 
needs time to develop. 


609 


4. THE TRANSFORMATIONS AND THE DIFFERENTIAL 
SYSTEM TO BE SOLVED 


The transformations to be used to obtain similarity 
solutions are already suggested by (12), (13), and 
(16) and are 


U Wis) 
Oia alae (5) Sin Un 7iB)) 7 (18) 
g 
1/3 
(vyw) = = (3) (V,W), (19) 


S/S} 


U 
(n,) = Gayi7z (69x) (y,z)- (20) 


Then the equation of continuity (5) becomes 


and Eqs. (7) become 


Va Yrs Wil (21) 
in which ¥ is the dimensionless stream function 
related to yp by 

me all 
DS eae (Cae) 7 ne) o (22) 
Equation (9) now takes the form 
= = = ap bd 
E we Ve Ce aaa (23) 


where € is the dimensionless vorticity component in 
the x direction. 

With the transformations (18), (19), and (20), 
Eq. (4) becomes 


Lh = A(Vh_ + Wh_), (24) 
n 6 


where L is the linear operator defined by 


92 92 3 3 
= =——— —— — 5 
an2 + 02 + 2n55 + 2057 + 4, (25) 
and 
VS ee (26) 


Equation (8) now has the form 
(> DE S S05 A(VE, + WE). (27) 


Equations (23), (24), and (27) are the final equa- 
tions governing the dynamics of the plume in a 
transverse wind. They are to be solved with the 
boundary conditions 


(ij) by =O, G20, YaO, aU = 0 ae i = O- 


(atat))) 394 0 at n = to or G = to. 


lI 
[e) 
wy 

i 
jo) 
E 

tl 


Boundary conditions (i) correspond to symmetry with 
respect to the ¢ axis, and conditions (ii) ensure 
that there is no temperature variation and no 


610 


velocity components v and w at infinity. The 
integral relation (6) now takes the form 


hdndg = 1, (28) 


The mathematical problem is now completely specified. 


5. THE METHOD OF SOLUTION 


The mathematical problem just formulated can be 
solved numerically once \ is known. But consider- 
able effort is required for this solution, since 
there are three second-order partial differential 
equations to be solved, two of which are nonlinear. 
It is true that computers can deal with nonlinear- 
ities, but the domain is infinite, and some estimate 
has to be made of how far to go in the numerical 
computation. Furthermore the integral condition 
(28) can only be imposed after the computations 

are done for h, and this makes the computation very 
cumbersome. 

For arbitrarily large values of \ an analytical 
solution is extremely difficult because the non- 
linearities present formidable difficulties. We 
shall attempt a power-series solution of the form 


h = hy + Ah, + Near ie Daetiel hk 
= Z2 

BS Ba PAS 3 Be ae ool Gg (29) 
a 2 

Des, OMG AW ole eo 


The success or failure of this approach depends 

not only on the value of A, but also on the magni- 
tudes of h,/ho, ho/hy, etc. Thus we need to make 
an estimate of the range of X, and we have to find 
out how fast hy, &,, and Y, decrease as n increases. 
Furthermore, even the estimate of A cannot be made 
without knowing the magnitudes of Yg. It turns 

out that a reasonable estimate of i is 


SOME OF 


Using Eq. (29), we shall show in the following 
sections that h,/ho, E1/eor and ¥1/¥ are all of 
the order of 1072. Thus, if i = 30, stopping at 
the second approximation, that is, at the terms 
with the first power in X, would introduce an error 
of about 10%, if we assume, as we evidently can, 
that the ratio 1072 would apply to (An+1)/hpy etc. 
for n equal and greater than 1. If X = 50 this 
error would be about 25 to 30 percent, and it would 
be necessary to go to at least the 042 terms to 
reduce the error to less than 15%. 

We shall delay the presentation of the estimate 
of A until later and shall proceed with the solution 
according to the approach in Eq. (29). In awaiting 
the experimental determination of i, we shall carry 
out the solution to the second approximation. 


6. THE FIRST APPROXIMATION 
The first approximation is governed by the equations 


Lh, = 0, (30) 


(L - 1) &9= -ho,, (31) 


WY ar Md = SS, 1 (32) 
0 Orr 0 


with the boundary conditions (i) and (ii) stated 
before, which we need not repeat here. 
The solution of Eq. (30) is 
2 2 
- + 
Mis ee oo) 

and application of the integral condition (28) on 
ho gives the value 1/m for C, so that 


2 
al ie 
ho = =e F (33) 
where 
me! Ge 4b 62. 
Then the solution of Eq. (31) is 
2 2 
2 75 2 =e 
Eo =a a, 1e = ee Gels 8 re 0 


where 


Oe= tern & 4 
n 


Given Eo. Eq. (32) can be easily integrated by 
separation of the variables r and 8. The result is 


= 
Y= ee (So he (34) 

The isotherms given by Eq. (32) are just concen- 
tric circles. But the streamlines given by Eq. 
(34) are already interesting. They are shown in 
Figure 1, which shows two very prominent vortices, 
with the vorticity pointing in the x direction. 
Thus the first approximation already shows the 
prominent features of the flow pattern in any plane 
normal to the x axis. Note that both the flow 
pattern and the temperature field are symmetric 


Cic 


FIGURE 1. Flow pattern from the first approximation. 
The horizontal axis is the n axis, the vertical axis 
the © axis, and the arrow indicates the direction of 
the gravitational acceleration. The value of 679 is 
zero on the t axis. It increases toward the left and 
decreases toward the right. The increments (or decre- 
ments) are all 0.1. 


with respect to the € axis, and that both ho and 
Yo vanish at infinity, as desired. 

The maximum vorticity is 0.09 and is at the 
point 


i) = WD, 6 = Op 


at which both V and W are zero. The maximum vertical 


velocity is 1/6m and is at the origin. The maximum 
absolute value of Yo is 0.63817/61, which occurs 
a © 2 © 2 wy te = alos. 


7. THE SECOND APPROXIMATION 
The equation for h, is 


Lh, = Vghon + Woh (35) 


1 Omi Ole” 


where Vj) and Wo are the velocity components from 
the first approximation. The right-hand side of 
Eq. (35) can be written in polar coordinates as 


KIFR 


Cle) #9 dr 
Hence Eq. (35) can be written as 
A ae apd 
iy, = = Sea P (oe? Yep 
3r@r 


where L, in its polar-coordinate form, is 


32 i 1 92 3 
b= Sl “= te’ at eZ te’ 
Writing 
in 6 
bh sets mm ©), (36) 
1 2 1 
31 
we have 
Ve 5 
e Aig 
L,H, = - (l-e Yip (37) 


if we write L, for L with the operator 32/902 in 
L replaced by -n2. 
To solve Eq. (37), we let 


sl SS ie ep (38) 


so that Eq. (37) becomes 


2 2 


= (Loe Jo (a) 


cel) see (2 = 3) £" + 2£ = =e 


Then we approximate the right-hand side of this 
equation by 


2 2 
2.-r x x ie x 
r“e ( 5 B ) (40) 


The greatest error occurs at r = 1.8, but it is 
less than 6.5% of the maximum value of the quantity 
approximated. Up to r = 1.2 the approximation is 
excellent. It is expected that the local errors 
around r = 1.8 will be diffused out when Eq. (39) 


611 


is integrated and will introduce negligible errors 
in the result. After (40) is substituted into 
Eq. (39), the latter is solved by repeated use of 
the following formula for various values of n: 


=r ZIn(n = 2) - 2(n - 1)r2]Jer¥ 


The result for f is put into Eq. (38), and we have 


1S | Gi ya 
H, = == - —— —— rt - —— 
ie 3 G2 1G » Iya) Bona 


r®) eons (41) 


The function Hy is tabulated in Table 1. A look 
at h, given by Eq. (36) then reveals that the 
temperature is increased in the upper half of the 
n-G plane and decreased in the lower-half plane, 
making the isotherms more widely spaced in the 
upper-half plane and more crowded in the lower-half 
plane. 

The tabulated values of H, show a very smooth 
variation of H, with r, verifying the expectation 
that the local irregular variation of (40) is 
diffused away when Eq. (39) is solved with (40) 
replacing its right-hand side. 

The next step is to solve 


Gb = ais, 2 Sa + VoEon + Woboc- (42) 


1n 


A simplification is possible before we attempt to 
solve Eq. (42). Differentiating Eq. (35), we have 


(L + 2) = Vohonn + WoPonz + VonPon 
+ WonPoc: (43) 
Let 
hy 
f= aa + q. (44) 
Then Eq. (42) becomes 
BT il 
(L- lq + (L + 2) S~ = FZ Wohgnn + Wohonz) » (45) 
since 
1 
S09 = 3 Aon 


By virtue of (43), Eq. (45) becomes 


aL 
(ieee) qa 3 Yonon ap WonPoc) - (46) 
But 
Von = (¥o)on, Wor eS -(¥9)nn, 
so that ¥ is a stream function for the fictitious 


velocity field (Vont Won)» and we can write Eq. (46) 
as 


612 


TABLE 1 Values of H,, S, and F; for r > 4, -100F = 5.04 r72 


r 0.1 0.2 0.3 0.4 OS 0.6 0.7 0.8 0.9 1.0 
100H), 3.03 5.85 8.28 10.19 11.48 12.14 12.20 11.75 10.89 9.76 
-100S 0.25 0.97 2.06 39 - 80 12 UoP3 8.00 8.41 8.43 
-100F 0.03 0.11 0.24 0.41 0.60 0.80 1.00 iL AL 7/ igs 1.44 

r itoal eo? tgs} 1.4 eo) 1.6 ed 1.8 ike) 2.0 
100H, 8.47 PoU3} 5.84 4.65 3.62 2.74 2.04 1.48 1.06 0.75 
-100S 8.11 7.51 6.70 5 7/8) 4.83 Soe)al 3.07 2.34 Leys} 1675) 
-100F 1D 1.56 1G S)7/ 1.55 iLook 1.45 1.38 1.30 eer 212} iligals} 

r PAL Deed, od) 2.4 Bod) 2.6 Bet 2.8 oe) 3.0 
100H, 0.53 0.38 0.27 0.19 0.14 0.11 0.08 0.06 0.05 0.03 
-100S 0.88 0.61 0.42 0.29 0.20 0.14 0.11 0.08 0.06 0.05 
-100F 1.05 0.98 0.90 0.84 0.78 On75 0.68 0.63 O59 0.56 

r Shoal Sie SSS) 3.4 355) 3.6 a7 3.8 3.9 4.0 
100H, 0.02 0.02 0.01 0.01 0.01 0.00 
-100S 0.04 . 0.03 0.02 0.02 0.01 0.01 0.01 0.00 
-100F OF 2 0.49 0.46 0.44 0.41 OF39 0.37 0.35 0233 0.32 

(L-Da=- = Z%y - ut Aa eer (2? és ) -2 Ses = art (elt Eee) 
Remembering that for 
p) sin 6 9 L GG oo one E) = en (22 a z) 
aa aor e cos 8 We 2 8 24 8 
aie 
and with ¥o and ho given by Eqs. (34) and (32), we oy ‘ 


have, finally, and the last member of (49) can be approximated by 


one eighth of (40). By repeated use of the formula 


re al 

(@ = 1)q =—— sin 26 |e Q f =) Pare Bans) a8 

on BD inGe~ee, )) = & [n(n - 4) - (2n - 3)r?Je 

7 2 for various values of n, we can then find the 

- — enn (47) solution for (48), and the final result for q is 
r2 
2 f 
To solve this, let q = —z sin 20 + Q, (50) 
on 
) : 
qa= — epi 2G) 0 ae i, wanes 
on 2 2) 2 
al pee, Cre a al 116 5 
= = - -=+ 
2 Bie fc Wa ( 8° 9945 * 
Then Eq. (47) becomes e r 
— Sh 6 
A 3 =P 95472 42432 
Te = ki (2x - =) IRD Sik BS © [e (eS) = a), 
(48) Lgl sia) (51) 
32640 


where L is the linear operator defined by (48). 
It is advantageous to write the right-hand side of With hy given by (36) and (41) and therefore with 
(48) as hy, known, (44), (50), and (51) give 


f= sin 26 + S(r), (52) 
oT 
where 
2 2, 
19 1 = aL 1OS283 a2 
Ss = —— = += = 
Se 2 fe DB) oT SAD 
4r 
114713 i, SSALz/ ¥ 
1670760 247520 


x 181 ome aL id 
68544 4320 
==) 

(53) 


The values of S(r) are tabulated in Table 1, from 
which it can be seen that the maximum absolute 
value of S occurs at about r = 0.95 and is about 


0.847. Since S is negative throughout, inspection 
of Eq. (52) shows that the maximum value of &) is 
at 


= 095), 6 = 


The effect of S is to reduce the strenths of the 
vorticity for the lower-half plane, but to augment 
them in the upper-half plane, thus to raise the 
eyes of the vortices. 

Finally, ¥; is to be found from 


9 i, 6 iy Oe 
Ty Yee, = A eee 2 Te) Yn = Ene 
or 1G 08 
Let 
vy, z sin 2 F (r) (54) 
9 
Then 
po Bop os Sop S este), 
re 2 


Two integrations by the method of variation of 
parameters (since a complimentary solution of F is 
simply rv?) gives, with due regard for the boundary 
conditions, 


a2 i2 
=5 
F = -r? Yr r3Sdr dr 
0 a 
Yr 12 
a) 
al -2 r3sdar - r2 nG l car 2 (55) 
= a @ 


6 


613 
which is given in Table 1 also. The calculations 


for the second approximation have now been accom- 
plished. 


8. ESTIMATE OF i 


The terms involving € in Eq. 
in the Reynolds stress terms 


(2) have their origin 


o) Ure Q 12 

By (v'w') and De (KE) 5 
where the primes indicate turbulent quantities. 
The terms were originally on the left-hand side of 
Eq. (2). The nonlinear terms on the left-hand 
side of Eq. (2) can be written as 


oh OE @ 
ay (vw) + ae (wo). 


Thus the ratio of 


2. G2 wo). 12 

ae (w*) and aS (w'<) 
is the ratio of 

3 2 

a5 (w-) and “SW 0 


and this ratio has the magnitude of 
-\W2/W . 
/ c 


The magnitude of Wy is 1/67, and the magnitude of 
Wo, is 0.267/31, which is the maximum value of Wor 
along the n axis. Thus, approximately, 


Ar AL 
0.267(12m) 5 me 
where s is the square of w'/w. The convection in 
the bent plume is like the convection in a two- 
dimensional plume, since the plume is bent by the 
wind to a nearly horizontal position. The measure- 
ments of Kotsovinos (1977) for the plane plume 
give the value 0.2 to s. This is considered by 
some people to be too high. But for the problem 
under investigation s may be even higher, because 
any swaying or deformation of the vortices would 
contribute a good deal to turbulence. Thus using 
0.2 for s in Eq. (56) would overestimate A. Using 
0.2 for s, we obtain from Eq. (2) 


dX = 48.5. 


This is probably too high. My estimate of i is 
that it is somewhere in the range 


50) S AS AIO), 


The value 30 for A} corresponds to a value of 0.34 
LOSI: 

Let us now see what errors would be committed 
for h, ¥, and & by stopping at the second approxi- 
Mation. For A = 30, the errors (in ratio of the 


614 


FIGURE 2. Flow pattern from the second approximation. 
The n axis is horizontal and the f axis vertical. The 
arrow indicates the direction of the gravitational ac- 
celeration. The value of 67¥ is, starting from the ¢ 
axis and going to the curves on the right, respectively, 
Op Obi, “Os4p “Ossi “Wot, =—Oo5p —OaGg Einel —MoGEin Gus 
values of 67¥ on the curves to the left of the f axis 
have corresponding absolute values but are positive. 


estimated* maximum value of the terms neglected to 
the maximum value of the computed quantity) are, 
respectively, less than 15%, 3%, and 10%. For A 
= 40 these percentage errors are, respectively, 
25%, 5%, and 18%. The most interesting thing to 
note is that ¥ is the most accurately calculated 
quantity. Figure 2 shows the flow pattern ina 
plane normal to the x axis, and Figure 3 shows the 
isotherms therein, all for } = 30. The flow pattern 
in Figure 2 can be regarded as sufficiently accurate 
to be representative of the actual flow pattern in 
a plane normal to the x axis. As expected, the 
hottest point and the "eyes" of the vortices occur 
at positive values of f. That is to say, the plume 
rises according to the x°/3 law. After the present 
work was done, I found that this law had recently 
been verified experimentally by Wright (1977), 
although he did not measure the detailed velocity 
and temperature distributions in the plume. 

If later measurements show i is larger than 30, 
higher approximations would be necessary. 


9. DISCUSSION 


It is perhaps surprising that the analysis shows 
that the results in dimensionless terms are indepen- 
dent of the parameter Gg2/u°. The explanation is 
that the velocity (v,w) far downwind from the heat 
source becomes vanishingly small, and whatever the 
value of U, the transverse wind is asymptotically 
always strong. 

Near the heat source the flow indeed depends 
very much on the magnitude of U. The plume may 


*On the basis that hj/hp and (hn+)) /Ay are of the same order 
of magnitude and that the same is true for ¥ and &. 


FIGURE 3. Isotherms from the second approximation. 

The n axis is horizontal and the ¢ axis vertical. The 
arrow indicates the direction of the gravitational ac- 
celeration. The value of th is 1.1 on the smallest 
closed curve and 0.3 on the outermost curve. The incre- 
ments are 0.1. 


rise high in a weak wind before being bent suffici- 
ently for the present theory to apply. In using 
the present theory it is always necessary to 
determine a virtual position for the heat source, 
which for small value of U can be considerably 
higher than its actual position. 


ACKNOWLEDGMENT 


This work was partially supported by the Office of 
Naval Research. The subject of this work was 


suggested to me by my friend Dr. Michel Hug, Director 


of the Department of Equipment, Electricity of 
France, through Mr. F. Boulot of the National 
Hydraulics Laboratory at Chatou, France, during my 
brief sojourn there in the summer of 1977. Their 
interest in this work, as well as the interest of 
Dr. A. Daubert, director of that laboratory, is 
very much appreciated. The work, begun at Chatou, 
was substantially improved and finished during the 
tenure of my Humboldt Award, at the University of 
Karlsruhe. To the Humboldt Foundation and my 
Karlsruhe hosts I should like to express my sincere 
appreciation. 


REFERENCES 


Kotsovinos, N. E. (1977). Plane turbulent buoyant 
jets. Part 2. Turbulence structure. J. Fluid 
Mech. 81, 45-62 (see P. 52, Figure 7). 

Laufer, J. (1953). The structure of turbulence in 
fully developed pipe flow. NACA Tech. Note 
2954. 

Wright, S. J. (1977). Report KH-R-36, Keck Labor- 
atory, California Institute of Technology. 

Yih, C.-S. (1977). Turbulent buoyant plumes. 
Phys. Fluids 20, 1234-1237. 


APPENDIX: 
THE EFFECT OF NEGLECTING THE PRESSURE 
GRADIENT IN CALCULATIONS FOR THE CONVECTION 
PLUME IN A TRANSVERSE WIND 


J. P. Benqué 
Electricité de France 
Chatou, France 


In many previous studies on jets and plumes, the 
pressure distribution in the jets or plumes is 
assumed hydrostatic, so that if the body-force term 
in the equation of motion is written in the form 
-gAp, where Ap is the difference between the local 
density and the ambient density, the pressure gradi- 
ent can be neglected in the equations of motion. 

If, further, the flow is two dimensional or axisym- 
metric, only the equation of motion for the vertical 
velocity component is then needed. After that 
velocity component is determined, the equation of 
continuity can be used to determine the other veloc- 
ity component. 

In the preceding paper by Yih, the assumption 
that the x component of the velocity is constant 
leaves only two other velocity components to be 
determined, and it is tempting to adopt the usual 
procedure of neglecting the pressure gradient. Yih 
has resisted that temptation. But it is useful to 
see what effects such a neglect would have on the 
flow and to determine whether in the problem treated 
by Yih such a neglect is allowable. This Appendix 
is devoted to this question. 

If the pressure distribution is assumed hydro- 
static and the usual procedure is followed, one will 
drop Eq. (1) and retain Eq. (2), with the first 
term on its right-hand side dropped. [Equation 
numbers in Yih's paper are retained.] Equations 
(3) to (7) will remain but (8) and (9) will not be 
needed. 

Following Yih's development and using his nota- 
tion, then, we have, as the dimensionless equations 
to solve, (24) and 


(L - 3)W = -h + A(VW, + WW). (A.1) 


Using the A-series (29), we have again (33) for the 
solution of hy. The equation for Wor obtained from 
(A.1), however, is now 


(L - 3)Wy = hy: (A. 2) 
The solution of this equation, satisfying all the 
boundary conditions for W stated in Yih's paper, is 


1 =y2 1 -n2-72 
0 3 - aC s (A. 3) 
Although it can be readily verified that Eq. 
(A.3) satisfies Eq. (A.2), it is not obvious that 
Eq. (A.3) is the unique solution. We shall show 

in the following that it indeed is the unique 
solution. The complementary solution Wo¢ of Eq. 
(A.2) satisfies 


(L - 3)Woc = O (A.4) 
and must be even in both n and ft. Let 


Woo = E(Mg(Z), 


615 


where the f is in no manner the same as the f in 
Eq. (38) of Yih's paper, we have 


£" 4) 2n£! Gat =" 0}, (A.5) 
g' + 20g' + bg = 0, (A.6) 
where 
a+b = 1. (A.7) 
Now let 
-n2/ 
fi) =e 28 (h) - (A.8) 


Then Eq. (A.5) becomes 
BM = (n2 + b)R = O- (A.9) 


Similarly, if we let 


=P 
g(t) =e b /2(c) . 


Then 


y" - (t2 + a)y = 0. (Al10) 


Because of Eq. (A.7), a or b must be positive. Let 
b be positive. (The argument is strictly similar 
if a is positive.) Because of the symmetry with 
respect to the f axis, 


B' (O)| = 0. 


Then Eq. (A.9) shows that 8 will approach infinity 
as n° approaches infinity, if 8(0) is not zero. 

[If 8(0) = 0 then 8B = O throughout.] To see how 
£(n) behaves at infinity, it is necessary to see 
how 8(n) behaves asymptotically. A simple calcula- 
tion shows that the two solutions of Eq. (A.9) 
behave, for large values of n2, like 


1 
exp |- (n2 - a - 2)%dn| and exp (n2 - a) 2an|. 


As we have seen, 8 must contain the second solution 
since 8 approaches infinity as n2 + ©, Using the 
second solution as the dominant term (a constant 
multiplier being understood), and recalling that 


Df en di ai part io 
(n a) n mn + 0 ( 2 , 


we see from Eq. (A.8) that for large 2 


aqy.< lal 2, (A.11) 


which can be seen to satisfy Eq. (A.5) asymptotically. 
If a is negative, (A.11) shows that f(n), and there- 
fore Wo, cannot satisfy the condition on Wp at 
infinity. If a is positive, it must be less than 

1, because of (A.7) and because b is positive. 

Then if Wo contains Woor 


616 


ol which shows that at In| = ©, Vj does not vanish. 

We must then, if we adopt the procedure of neglecting 
the pressure gradient, not demand that Vo vanish 

at infinity, but instead demand 


SS 0 


This boundary condition for V, must, for consistency, 
be demanded of V, i.e., of Vj, V2, etc., as we 
proceed to higher and higher approximations. 

In this connection we can also see that it is 
not possible to add a multiple of Woo to the Wy 
given by Eq. (A.3) to make V, vanish at infinity. 
For, in order to make Vo vanish at infinity, the 
only possibility is to add to the Wo given by Eq. 
(A.3) a multiple of 


FIGURE A.1. Flow pattern for (V,W)). Aém¥, = 0.2. 


2 


Worse) a(n) (A.15) 


co co 
where f satisfies 
I= W )dndz = oe, (A.12) £" + 2nf' - f = 0. 


That means 


a=-l, 
But this cannot be true, because integration of 
(A.2), by parts if necessary, gives and Eq. (A.11) gives 
1 
c} i) £(n) ~ [In| 2, 
=spl SS = hydnds —tiLy, which makes Woar and therefore Wo if it contains 
Woc, infinite at |n| = ©, Any other dependence of 


Wo on © than exp (-67) would, of course, not make 


rae rc a4 Vo vanish at infinity, for the part of Vo that 
so that arises from Wo, would not be able to cancel out Eq. 
cael (AQ13) at |=. 
3) Hence Wy and V) are uniquely given by Eqs. (A.3) 


and (A.13). Using them in 
Hence W cannot contain a multiple of Woar and Eq. 
(A.3) is the unique solution. Lh, = Voho, + Wohors (A.16) 
Then the equation of continuity gives 


and 
: 2 2 
2 — + - — - 
V2 f te (n G Fane (A.13) (L - 3)W, hy + VoWo, + WoWors (A.17) 
g we find that 
> hy 
0,10 W) = 5 (A.18) 


I have computed h, numerically from Eq. (A.16) 
and the boundary conditions, and therefore W}). 
The velocity component, V,, is then found from the 
by equation of continuity. The flow pattern corres- 
ponding to (V,,W)) is given in Figure A.1, where 


250 
¢ the streamlines are shown, with ¥; = 0 on the ¢ 
: axis and Aém¥, = 0.2. 


Then the flow field for 


oF V=vV 


ee o + AV, and W = Wy + AW) 


is shown in Figure A.2, with X = 30, where the 
AetY = 0.2. 

It is clear that the "streamlines" do not close 
to form closed eddies, as in the figures of Yih's 
paper. Thus the effect of the pressure gradient 
Flow pattern for (Vj + AV, Wo + AW))- cannot be neglected in the problem studied by Yih. 

In past studies of jets and plumes, where the 


FIGURE 
A6rY = 


Oop 
OS) 


pressure gradient has been successfully neglected, 
the velocity component other than the one retained 
is one order of magnitude smaller than the one 
retained. Thus the equation of motion for it can 
be neglected together with the gradient of (the 
dynamic part of) the pressure, and the flow pattern 


617 


can be determined from the equation of continuity 
once the principal component of the velocity is 
determined. Such is not the case in the problem 
under discussion here, and therefore for this prob- 


lem it is necessary to retain the pressure gradient, 
as Yih has done. 


Internal Waves 


OF Me SPheslaaip's 


Johns Hopkins University 


Baltimore, Maryland 


ABSTRACT 


It has become evident in the past few years that 
the wave-number, frequency spectrum of deep ocean 
oscillations has a remarkably consistent form close 
to that which would be expected for statistical 
equilibrium among the modes under wave-wave resonant 
interactions. The energy sources that maintain deep 
oceanic internal waves are, however, not well under- 
stood. - 

In the vicinity of the thermocline, the energy 
density (per unit mass) of internal wave activity 
is generally much greater than in the ocean depths. 
Relatively high frequency internal waves, generated 
in a variety of ways, are to a first approximation, 
trapped in this region. Disturbances whose fre- 
quencies are less than Ng, the deep stability fre- 
quency, do however radiate downwards effectively. 
Also, groups of high frequency, low mode waves 
generate second order mean perturbations to the 
thermocline structure, and if the group frequency 
is less than Ng, again energy radiates down. The 
flux of energy into the deep ocean is illustrated 
first in a simple model in which a sharp pycnocline 
lies over uniformly weakly stratified water. The 
more general problem involving an arbitrary strati- 
fication is formulated and some preliminary asymp- 
totic solutions are presented. 


1. INTRODUCTION 


During the last 10 years or so, a variety of new 
and ingeneous oceanographic observations has been 
made on the structure of internal waves fluctua- 
tions in the ocean. Twelve years ago, in the first 
edition of The Dynamics of the Upper Ocean, I was 
forced to write that in view of the difficulty and 
expense involved in the systematic study of oceanic 
internal waves, "those (measurements) that do exist 
are correspondingly rare and valuable." The present 
situation is gratifyingly different. Deep oceanic 


618 


observations of internal waves are no longer rare, 
but they remain valuable; Cairns (1975), Katz (1975), 
Gould, Simmons, and Wunsch (1974), and a number of 
others have provided different kinds of observations 
from which a consistent pattern is emerging. It ap- 
pears that the deep oceanic internal wave spectrum 
has a remarkably universal form close to that speci- 
fied by the Garrett-Munk (1975) spectrum, though why 
this is so cannot yet, I think, be asserted with con- 
fidence. McComas' (1975) calculations on resonant 
wave-wave interactions indicate that the Garrett- 
Munk spectrum is close to what one would expect in 

a state of statistical equilibrium under the balance 
of these interactions. On the other hand, there are 
indications, such as the occurrence of sporadic, 
isolated patches of turbulence in the stably strati- 
fied regions of the ocean which suggest that local 
instabilities may be limiting the wave spectral 
density. 

Soviet investigations, such as those of 
Brekhovskikh et al. (1975) have concentrated on the 
low mode structure in the thermocline region whose 
energy density (per unit mass) exceeds, usually by 
an order of magnitude, that of the deep oceanic in- 
ternal waves. The characteristic frequencies are 
also about an order of magnitude higher. The cal- 
culations of Watson, West, and Cohen (1975) among 
others indicate that the lowest modes are generated 
quite rapidly by interactions among surface wave 
components; a number of studies along these lines 
are described in the useful review by Thorpe (1975) 
and by the present author (1977). The upper ocean 
is certainly the site of considerable dynamical 
activity, but how much of it is radiated downwards 


-to provide a source for those motions encountered 


in the deeper, less strongly stratified region be- 
low? According to the usual linear analysis, the 
low mode, relatively high frequency waves are trapped 
to the strongly stratified thermocline region; only 
the low frequency high modes have structure that can 
penetrate great depth. 

Yet the description of deep oceanic motions as a 


linear superposition of high modes may make little 
sense. A linear mode can itself be considered the 
superposition of two disturbance trains, one propa- 
gating downwards and the other upwards with reflec- 
tions either at the bottom or at a region where the 
buoyancy (or stability) frequency N drops below the 
wave frequency. McComas! calculations indicate that 
the non-linear interaction time of such components 
at the spectral densities found in the deep ocean, 
is remarkably short, only a few wave periods in 
many wave cases. Accordingly, a train of waves 
generated, say, near the thermocline will in actu- 
ality have little opportunity to travel to the 
bottom, reflect upwards, and combine with a down- 
wards travelling wave to produce a 'mode' as usually 
conceived. More realistic would be the view of dis- 
turbances generated in the more active thermocline 
region, radiated downwards but being 'scrambled' by 
wave-wave interactions into a more diffuse spectral 
background. 

This contribution is concerned with some aspects 
of the energy flux downwards from high frequency, 
low mode internal waves at the thermocline. If the 


internal wave frequency is greater than the stability 


frequency Ng below the thermocline, the waves are 

of course trapped to the thermocline region. How- 
ever, as their frequency decreases below Ng, they 
become 'leaky' and their energy radiates rapidly 
downwards as the simple analysis of the next section 
will demonstrate. Yet, if Brekhofskikh et al. (1975) 
measurements are at all typical, most of the energy 
of the low mode internal waves in the thermocline 
region is at frequencies considerably above Nagi 
indeed, in view of the efficiency with which such 
low frequency energy is propagated downwards, we 
would not expect to find much energy at these fre- 
quencies in the main thermocline. However, one 
possible link is suggested by the work of McIntyre 
(1973) who showed that groups of internal waves in 

a fluid of constant frequency N, confined between 
horizontal boundaries, produce second order 'mean' 
motions, modulated as are the wave groups. There 

is no reason to believe that these second order dis- 
turbances are confined only to the particularly 
simple case that he considered, and indeed in Sec- 
tion 3 it is shown that they are not. 
internal waves, occurring in groups and trapped 
within the main thermocline, produce second order 
low frequency disturbances; if the group frequency 
is less than Ng, their energy is radiated downwards 
at the group frequency. 

The results presented here are preliminary but 
intended to provoke consideration of this mechanism 
as a source of oceanic internal waves. The simplest 
case of a sharp thermocline overlying a deep, uni- 


formly stratified region is described in some detail. 


The more realistic (and complicated) case with a 

general distribution of N(z) can be considered by 
asymptotic methods and these results will be de- 

scribed elsewhere. 


2. RADIATION DOWNWARDS--A "LEAKY MODE" 


Consider the following experiment: a laboratory 
tank (Figure 1) is stratified with a layer of uni- 
form density lying over a density jump 6p below 
which the fluid is uniformly stratified, with N? = 
(-p7lg d0/dz) = constant. A wave-maker at the end 
of the tank generates a periodic disturbance with 
(real) frequency n. What are the characteristics 
of the motion induced? 


High frequency 


619 


It is, I think intuitively evident that if n > N 
an interfacial wave mode will propagate. The struc- 
ture of the mode below the pycnocline will be in- 
fluenced by the stratification but at these high 
frequencies, no internal waves can be supported in 
the lower layer and the interfacial wave will propa- 
gate without loss. If, however, n < N, internal 
waves induced in the lower region by the interfacial 
disturbance can carry energy downwards so that the 
interfacial wave will attenuate. The question is: 
how rapidly does this occur? 

A linear analyses suffices. Suppose the pycno- 
cline displacement is represented by the real part 
of © = a exp i(kx - nt), where n is real and k may 
be real or complex. Above the pycnocline at z = 0, 
the motion is irrotational with u = Vd and V2 = 0. 
In the uniformly stratified region below, the vert- 
ical velocity component, w, obeys the internal wave 
equation 


a2 2 Dep We 
cen VA Row 2 © , (1) 


where Vine is the horizontal Laplacian operator, 
32 /ax2 in this two-dimensional problem. At the 
upper free surface at z = d, w = O to sufficient 
accuracy; at the pycnocline the vertical displace- 
ment and the pressure must both be continuous and 
as z>- ©”, the disturbance must either die away 
or represent internal waves with an energy flux 
downwards. 

In the upper region, the solution for 9 is 
readily found to be 


_ ina cosh k(z - dq) F 2 
> = Te sinh ka CxPi (kx Me) p (2) 


while in the lower layer, if 
w = - ina exp [kz + i(kx - nt)] , (3) 


(which satisfies the condition of continuity of w 
at z = 0), then substitution into (1) requires that 


(Pje silo Gye c (4) 


Note that since n is real, «/k is either purely real 
(Gin 2 N) or purely imaginary (if n < N). 

The dispersional relation is obtained from the 
condition that the pressure be continuous at z = f. 
In the upper region of density p, 


Pp, = - pg = 22 | 


- palg + (n*/k) coth kd] exp i(kx -nt) , (5) 


Wave 
Absorbers 


FIGURE 1. Tank stratified with a layer of uniform 
density over a density jump below which the fluid is 
uniformly stratified. 


620 


to the first order in the wave amplitude. In the 
lower region, where the density is p + 6p - oN2z/g, 
the horizontal pressure gradient 


DS 2 ou (ond Woy Se 


to the lowest order, 
ax ot 


i(p + 6p)an2(k/k) exp i(kx - nt) 


at z = 0 from (3), so that 


Po = (9 + 6p) an2(k/k2) exp i(kx - nt) 
At z = T, below the pycnocline, 
Py =- (p + dp)a(g - n2K/k?) exp i(kx - nt) (6) 


From (5) and (6) it follows that 


2 _ __(6p/p) gk Bie ek ae (7) 
coth kd + (k/k) coth kd + (K/k) ’ 


to the Boussinesq approximation, when é6p/p << 1, and 
where b is the contrast in buoyancy across the pycno- 
cline. 

For high frequency oscillations, when n 2 N, 
equation (4) shows that k/k is real and less than 
unity; from (7) k is real and the waves propagate 
without attenuation. The additional restoring 
forces provided by the stratification below do how- 
ever increase the wave frequency for given k and b 
above the value for an unstratified lower layer by 
the ratio 


[coth kd + 1] / [coth kd + (1 - N2/n2)*%] 


The case when n < N is algebraically simplest when 
|ka| +e, In view of the upper boundary conditions, 
the real part of k > 0, while from (4) 


2\}5 
K N - 
mm = 2 i(a - a) =e ta stan 6), (8) 
where n = N cos @. From (7) 
Pe = (Gai) (il Hh ee Ch) G (9) 


Since the interfacial waves attenuate in the posi- 
tive x-direction as energy leaks downwards, the 
positive sign in (9) is relevant and the vertical 
wave-number 


ray 
i] 


iktan 6, 


(n2/b) ( - tan26 + i tan 6) . (10) 


The motion of the pycnocline is therefore repre- 
sented by ; 


2 2 \ 
= _ nx fetes hotel 
C=a exp ( Spe tan 8) exp 1 (a ne) , 
n?x 
= a(x) exp i SS nt p (11) 
where 


2 
a(x) = a exp (- n* tan e) : 


The ratio of the spatial attenuation rate to the 
wave-number is simply tan 6 = (N2/n2 = 1)”; when n 
is significantly less than N the attenuation dis-— 
tance is short as the energy leaks downwards very 
effectively. 

Expressions for the motion in the upper and 
lower regions can be written down simply. In the 
lower layer energy flows along the characteristics 
—& = x cos @ + z sin 8 = const., and the distribu- 
tion of vertical velocity is 


wes fm e@ (- nieint . (25 aoe t) 
*P bcos26 ©XP + \beosé u 
2 
mete aff rte! 
=)/=no!7)(&))) exp i( ST oe ) fF (12) 
Zz 
g 
6 x 
S 
IS q 
Ne SS 
N ~ 
SS SS 
S 


where a)(&) is the amplitude of the interfacial 
wave at the point where the characteristic inter- 
sects the pycnocline. The horizontal component 
of the velocity field in the lower layer is u = 
- w tan 6, since the motion here consists of alter- 
nate layers sliding relative to one another along 
the characteristic surface inclined at an angle 98 
to the vertical. The pressure fluctuation can be 
found most simply from the horizontal momentum 
equation: 

n2E 


p = - a1(&)bsin6(sin® + icos§)exp i eaeeat = 


The vertical energy flux is therefore 


2 
ES = -4noa ](&)b sin@cosé , 


and the total energy flux, directed downwards along 
the characteristics — = const is 


E = mnaj(—) bsine . (14) 


In the upper region, the fluctuations in pressure 
are found from (2): 


pe DD 2 es 
DRE) seca neces, [-kz + i (kx - nt)] , 


when kd >> 1, whose real part, in view of (9) and 
(12) reduces to 


IS. = & ei(Ge op zcot6) b(cos*8cosy + cos@sin@sinyx) , (15) 


ae 


where 
n2 
xX = pb (* - ztané@) - nt P 


The real part of the horizontal velocity field is 
likewise 


UL. = (0$/3x) = - na(x + zcot@) cosy ; 


so that the horizontal energy flux in the upper 
layer 


the horizontal divergence of which 


ab = - kna?(x)b sin@cosé (17) 
provides for the radiative flux in the lower layer. 
This simple example illustrates the way that 

energy can be radiated downwards by the low fre- 
quency perturbations produced by groups of high 
frequency waves, but they have a deeper theoretical 
interest. Gaster (1977) has pointed out that if 
the dispersion relation for waves involves complex 
wave-numbers or frequencies, the usual kinematic 
definition of group velocity may not be correct, 
and a simple calculation shows that the solution 

is an example of this failure. Here the wave- 
numbers are complex as the energy leaks into the 
lower layer, but the energy flux is not at the rate 
represented by the local energy density, n2a}2/2 
cos*6, times the ordinary group velocity Vw = 

c tan 6 = (b/n) sin @. The correct interpretation 
of these situations will be considered elsewhere. 


3. ENERGY RADIATION DOWNWARDS FROM GROUPS OF 
INTERFACTAL WAVES 


To illustrate the way in which groups of internal 
waves produce 'mean,' second order disturbances locked 
to the wave group, let us consider the same basic 
stratification as in the previous section, with 

fluid of depth d and constant density lying over 

a buoyancy jump b below which the stability fre- 
quency N is constant. Suppose that interfacial 

waves with frequency n > N are maintained by high 
frequency forcing £ from the upper layer, perhaps 

by the surface wave-wave interactions described by 
Watson, West, and Cohen (1976). If the internal 

wave amplitude is characterised by a and the wave- 
number by k, then, to order E2 = (ak)*, the condition 
or continuity of pressure across the interface can 

be expressed as 


du pY4 ( 2u ) 
S)\)-bp = =- tu: V ; 8 
(2) b Aen +f I Wie azct u u (18) 
ae B= O, wae M( )} = Cn = ( Yag the difference 
across the density jumps. Since 
aie 2 - pO harness a GS 
C= we uy VG =wo +f a2 I, we ae 0 
= = oe (u ) (19) 
0 3x 0S , 


to order E2, where the suffixes, tT and 0, represent 
quantities measured at z = t,0, then the condition 
that tf be continuous across the interface assumes 
the form 


621 


(sie) 


Aw = A ox 


at z=0O . (20) 


Finally, in the lower layer, 


nae Wo a Ne ae SO oy (21) 


93 
2 axdzot go wey 
- Cee * Vb + oa + Vw) ) (22) 
ox = dt . 


Variations in energy density of the primary waves 
will propagate with the group velocity, c,; let us 
therefore average these equations at aeslnres fixed 
with respect to the wave groups but over random 
phases of the waves themselves, a process repre- 
sented by brackets [ ]. The averaged interfacial 


conditions are then, to order E% = (ak)2, 
a, | oe S95 22d Nel eee gon gyi (23) 
at ox azot ~ U 
a 
AS ev} = Nee (x6) ' (24) 
ox | 


both at z = 0, and 
(2) = fe) = & Teel (25) 
ox 4 


also at z= 0. The averaged field equation for the 
lower layer follows similarly from (21) and (22). 

The linear fluctuating internal wave motion is 
as given in the previous section when n > N; through 
the non-linear terms on the right of (23)-(25), this 
forces a second order mean disturbance [zc], [wl], 
etc., that moves with the velocity of the wave groups. 
The pycnocline disturbance can be represented as 


t = ka{cos(k'x - n't) + cos(k"x - n"t)} . 
The form of the forcing functions is simplest when 


the pycnocline depth is such that kd >> 1, and it 
is found that (23) reduces to 


ou eC) 
ae | Bes [c] 


2 
= % a2 uipts 
4a es {2 («) | 


(53S i ‘8) 
g 


ips 


(c_ + 4e)Sin k 
g g 


{1 + O(ak, n? fn?) 3 0 


i} 


c 
- k a2k NZ (<2 + 1) Salina Ik (5 © C38) 5 (26) 
g c 2 g g 


Wont Be = et ca jel Ng = n' - n", and cy, represent 
the wave-number, frequency, and velocity of the 
groups. Similarly, from (24) 


622 


1 2 f K Q 
Mw = q an ae + «) sin xX) 7 (27) 
where X = kg(x - cgt) and from (25) 


A al : 
(Es) [wil g ank, Same’ 


i] 


an 
| 


aly igs ; 
[w]_ 3 2 is e/a) sini! 7 (28) 


where [ ], and [ ]J_ represent averages taken just 
above and below the discontinuity in density. 

These matching conditions to be applied as z = 0 
involve the non-linear forcing provided by the wave 
groups. The field equations are, however, linear 
to this order. 
we have Laplace's equation for the averaged velocity 


lols O. | (29) 


while in the averaged internal wave equation (21) 
for z < 0, the non-linear terms are smaller by at 
least (ng/n) 2 << 1 than those in the matching con- 
ditions, since they involve two horizontal deriva- 
tives (or one x and one t derivative) of averaged 
second order quantities. Accordingly, to sufficient 


accuracy, 
Oe ea 2 22 = 
DED Vi[w] + N aes Kal SO 5 1a <6 © «6 (30) 


Since the length of the wave groups is large 
compared with the wavelength of the interfacial 
waves, k. << k and it is consistent to assume that 
kgd << 1, even though kd >> 1. Furthermore n/N << 
1 while NgN = O(1). Under these conditions the 
solutions for the mean pycnocline displacement and 
the low frequency internal waves radiated downwards 
are found to be 


a2nc 


aed 
[co] =- ba COs ka (x - ea) g (31) 


[w] 2» - 5 arnk, (2 + a cos OSes ar Sa = ye) nS) 


where 


1 

2 cy) 
Baral paga are (33) 
g 


is the vertical wave-number of the radiated field. 
The horizontal velocity component in the internal 
wave motion below the pycnocline 


{u] = [w] tanyp , 


where cos j = ng/N and the energy density (twice 
the kinetic energy density) is 


E =» (Tae + Twi), 


patkg2n2n2 eN2 
~ 126n2 | ( + £) G 


4 2n2 
an. 2pa n“N 


ea ; (34) 
g 


Above the pycnocline, when d > z > O, 


since n/N >> 1 and kK/k = 1. The vertical component 
of the group velocity of the radiated waves is fg 
cos ~ sin ~ where c, is the group of the inter- 
facial waves, so that the vertical energy flux is 


m7 


to 


n2 * 
= (9/128) a'tn2Nk ( = el ) : (35) 
g N2 


Although this representation of the density dis- 
tribution by a discontinuity at the pycnocline, 
followed by a uniform stratification below, is a 
gross simplification of typical oceanic conditions, 
it is of interest to examine the order of magnitude 
of the vertical energy flux that might be generated 
in this way. If the interfacial wave amplitude is 
10 m at a frequency of 5 c.p.h., having groups 1 km 
in length and if N = 2 c.p.h., the downwards energy 
flux is about 2 erg/cm* sec., which is of the same 
order as the 5 erg/cm* sec. estimated by Garrett 
and Munk (1972) for the rate of energy dissipation 
from internal waves by sheer instability. This 
correspondence is sufficiently close to encourage 
a more detailed study with N(z) arbitrary, the re- 
sults of which will be presented elsewhere. 


ACKNOWLEDGMENT 


This work was supported by the Fluid Dynamics 
Branch of the Office of Naval Research under con- 
tract NR 062-245. 


REFERENCES 


Brekhovskikhk, L. M., K. V. Konjaev, K. D. Sabinin, 
and A. N. Serikov (1975). Short period internal 
waves in the sea. J. Geophys. Res., 80, 856-64. 

Cairns, J. L. (1975). Internal wave measurements 
from a midwater float. J. Geophys. Res., 80, 
299-306. 

Garrett, C., and W. H. Munk (1972). Space-time 
scales of internal waves. J. Geophys. Fluid 
Dyn., 3), 225-64. 

Garrett, C., and W. H. Munk (1975). Space-time 
scales of internal waves: a progress report. 

J. Geophys. Res., 80, 291-7. 

Gaster, M. (1977). On the application of ray math- 
ematics to nonconservative systems. Geofluid- 
dynamical wave mathematics, Appl. Math. Gp., 

U. Washington, 61-6. 

Gould, W. J., W. J. Schmitz, and C. Wunsch (1974). 
Preliminary field results of a mid-ocean dynamics 
experiment (MODE-0). Deep-sea Res., 21, 911-32. 

Katz, E. J. (1975). Tow spectra from MODE. J. 
Geophys. Res., 80, 1163-7. 

McComas, C. H., and F. P. Bretherton (1977). Reso- 
nant interaction of oceanic internal waves. J. 
Geophys. Res., 82, 1397-1412. 

McIntyre, M. E. (1973). Mean motions and impulse 
of a guided internal wave packet. J. Fluid 
Mech., 60, 801-11. 

Phillips, O. M. (1977). Dynamics of the Upper Ocean 
2nd ed., Cambridge University Press. 

Thorpe, S. A. (1975). The excitation, dissipation 
and interaction of internal waves in the deep 
ocean. J. Geophys. Res., 80, 328-38. 

Watson, K. M., B. West, and B. I. Cohen (1976). 
Coupling of surface and internal gravity waves: 
a Hamiltonian model. J. Fluid Mech., 77, 185- 
208. 


Breaking Internal Waves in Shear Flow 


Si iS whorpe 


Institute of Oceanographic Sciences, 
Wormley, United Kingdom 


ABSTRACT 


During and following periods of strong winds, the 
Richardson number (the square of the ratio of the 
Brunt-Vdisdla frequency to the shear) in the 
thermocline is of order unity, and the shear becomes 
an important factor in determining the properties 

of internal gravity waves. These properties are 
discussed and the shape and breaking of waves ina 
shear flow is investigated in laboratory experiments. 
These experiments show that the waves may break at 
their crests or their troughs depending on the sign 
of a certain vector scalar product. An analogy 
between surface waves and interfacial waves is 
invoked to account for this behaviour. Breaking 

is observed to occur by particles of fluid moving 
forward more rapidly than the wave crest advances, 
leading to gravitational instability. The effect 

of breaking in the ocean will not only enhance 
diffusion rates, but it will modify the directional 
spectrum of the internal waves. 

Although many acoustic backscatter observations 
from ships reveal clearly the presence of internal 
waves in the ocean seasonal thermocline, very few 
have been published which appear to show signs of 
their breaking. This is surprising in view of the 
clear and not infrequent evidence of 'breaking 
events' in the equivalent acoustic or Doppler radar 
Measurements in the atmosphere. Our knowledge of 
internal wave breaking in the ocean still rests 
almost entirely on the direct observations by divers 
using dye in the Mediterranean thermocline [Woods 
(1968)]. The present towed, moored, or dropped 
instruments give inadequate information on the 
nature or structure of the intermittent mixing events 
in the ocean to be certain of their cause, or even 
of the scales of motion which contribute most to 
diffusion across density surfaces in spite of its 
great importance to the prediction of the thermo- 
cline structure of the upper ocean. 

It is against this background of poorly known 
dynamical structures that this paper is presented. 


623 


One aim is to describe the patterns which accompany 
wave breaking, for without a knowledge of such 
patterns it is difficult to design the appropriate 
experiment to detect wave breaking or, conversely, 
to correctly identify the processes involved once 
observations are available. 

It would be naive to ignore the effect of wind 
in a description of breaking waves on the surface 
of the sea in deep water [see, for example, Phillips 
and Banner (1974)]. (Wave breaking on a beach is 
a different matter). It is similarly inappropriate 
to ignore the effect of mean shear on internal waves 
in the seasonal thermocline, since the Richardson 
number there is low, especially during, and follow- 
ing, storms [Halpern (1974)]. Internal gravity 
waves can exist and propagate in a shear flow just 
as they can when a mean flow is absent. These waves 
belong to a group which Banks, Drazin, and Zaturska 
(1976) have classified as 'modified' (-by shear) 
"internal gravity waves'. They may sometimes coexist 
with a set of wavelike disturbances which grow in 
amplitude (the 'unstable wave solutions' of the 
Taylor-Goldstein equation) and which may eventually 
lead to turbulence (Figure 1). It is known however 
that (for steady mean flows) the latter solution cor- 
responding to Kelvin-Helmholtz instability (K-H.I) 
only exists if the Richardson number, Ri, in the flow 
is somewhere less than a quarter [Miles (1961), 
Howard (1961)] and even then in some flows an un- 
stable solution may not exist. One way in which 
internal gravity waves may break is by themselves 
causing or augmenting a mean shear to induce regions 
of such low Ri that small-scale disturbances may 
grow as K-H.I and generate turbulence. It appears 
that Woods' (1968) billows were generated in this 
way, and similar structures in Loch Ness [Thorpe, 
Hall, Taylor, and Allen (1976)] may have a like cause. 
It is however known that internal waves may break in 
quite a different way, by what has been termed 'con- 
vective instability' [Orlanski and Bryan (1969) ]. 
This form of instability becomes much more likely in 
the presence of a mean shear. 


FIGURE 1. The development of Kelvin Helmholtz In- 
stability (K-H.I) in a stratified shear flow [from 
Thorpe (1971)]. 


Shear affects internal gravity waves in several 
ways. Perhaps the most important concern the wave 
speed. Bell (1974) has shown that for any wave 
mode, the phase speed, c, is a decreasing function 
of wavenumber, k, which, for waves moving faster 
than the mean flow at any level, tends to k71Nnax 


+ U, as k increases indefinitely, where Nna 


max x 
is of the Brunt-Vaisdld frequency, N, and Uneee che 
maximum mean flow. (A similar result holds for 


waves travelling more slowly than the mean flow.) 
This result reduces to the well-known property, 
OS Rep Cie internal waves in the absence of shear 
[Groen (1948)] where o = ck is the wave frequency 
relative to the mean flow. It implies that even 
in a shear flow the wave frequency is less than 
Nmax Provided the waves are viewed in frame of 
reference which moves forward at the speed, Upax- 
Banks et al. showed further that, at least for 
simple mean flow profiles, the speed of waves of 

a given mode and wavenumber tends to Where (from 
above) as Ri decreases. We see a consequence of 
this result later. 

The vertical structure of internal waves is also 
changed by shear. Figure 2 shows how the distri- 
bution of the amplitude of a small amplitude wave 
of given k varies with z as the shear increases 
for (a) plane Couette flow of a fluid with constant 
N and (b) hyperbolic tangent profiles of mean speed 
and density. The profiles are distorted as Ri 
decreases with the largest amplitudes displaced 
towards the level at which the mean speed in the 
direction of wave propagation is greatest. We shall 
find it convenient to distinguish these cases by 
the sign of x = c.g X 2 where 2 is the mean flow 
vorticity and c the phase speed of the waves in a 
frame of reference in which the depth averaged mean 
flow is zero. Positive U,) in Figure 2 corresponds 
to x > 0, and conversely. 

The shape of waves in a fluid with density and 
velocity distributed as tanh z (corresponding to 
Figure 2b) is shown in Figure 3 for (a) backward 
relative motion in the upper layer, x < 0, (b) no 
shear, (c) forward motion in the upper layer, x > 0. 
The waves in (b) and (c) have narrower crests than 
troughs, whilst the waves in (a) have wide crests 
and narrow troughs. 

This second-order effect is not unexpected. It 
may easily be shown [Thorpe (1974, Appendix C)] 
that interfacial waves (see Figure 4) which move 
forward with the speed of the upper layer (the 
limit, as we have seen, towards which the phase 
speed of the internal waves tends as Ri decreases) 
have exactly the same shape as have surface waves 
on a fluid of depth equal to the lower layer. Con- 
versely those moving at the speed of the lower layer 
have the shape of surface waves on a fluid of depth 
equal to the upper layer, but inverted. This is 
just the trend shown in Figure 3. The limiting 
form of the surface wave is one with a sharp apex 
of 120°. Such an angle can exist in a two-layer 
flow only in the cases we have considered where 
the wave speed is the same as the flow in one of 
the two layers. Otherwise there is a relative flow 
around the apex in the upper (or lower) fluid 
leading to a singularity of infinite flow in the 
irrotational fluid. In general, some other limiting 
profile must appear, although it is likely to tend 
in a continuous way towards the limiting sharp apex 
profile. Recent work on breaking surface waves 
[Cokelet (1977)] cannot be applied even in the 
special case for the analogy is valid only for 
steady waves. 

Experiments, however, [Thorpe (1968)] demonstrate 
how internal waves break in a shear flow. Figure 
5 shows wave breaking for x > 0. A jet of fluid 
moves forward (that is faster than the waves advance) 
from the wave crest above the level of the mean 


interface where we saw in Figure 2 that the dis- 
placement was concentrated, and, in Figure 3, where 


the curvature was greatest. The fluid particles 
move forward (at speed C,) more rapidly than the 
wave advances and this leads to a layered structure 
with a region of slightly denser fluid overlying 
less dense fluid with the potential consequence of 
gravitational instability. Similar 'forward' 
breaking occurs at the wave troughs when X < 0. 

The experiments demonstrate clearly the difference 
between K-H.I of the mean flow (seen in Figure 5}j) 
and the convective instability of the waves. In 
the former the wave-like disturbances grow, extract- 
ing energy from the mean flow, whilst in the latter 
the waves do not grow in amplitude and lose energy 
as a consequence of instability. 

The condition for convective instability to 
occur (C, = c) has been used in a calculation to 
produce the stability diagrams of Figure 6. These 
are appropriate only to a particular wavelength 
and show the wave slope at which instability will 
occur for a given Ri. The Couette flow (Figure 6a) 
is stable in the absence of waves for all Ri > 0, 
but the hyperbolic tangent profile (Figure 6b) is 
unstable at Ri = 0.25 and the dashed lines show 
the value Ri = 0.25 at the interface marking the 
boundary at which K-H.I will occur in a quasi steady 
flow. These diagrams demonstrate how shear greatly 
reduces the wave slope at which convective instabil- 
ity sets in, a partial consequence of the trend of 
the phase speed toward Umax and hence a reduction 
of the wave particle speed necessary to promote net 
speeds, Cpr which exceed the phase speed. The non- 
linear terms are also very important however, the 
finite amplitude change in the phase speed being 
as important as other non-linear effects. 


625 


FIGURE 2. The amplitude of the displace- 
ment of lines of constant density in 
internal waves of the first mode with wave 
number k = m/H calculated from linear theory 
(i.e., from the Taylor-Goldstein equation) 
at various Richardson numbers (as labelled) 
in 
(a) Couette flow, U = Ug(2z/H - 1), with 
constant density gradient. Ug is posi- 
tive for the left hand set of curves 
and negative for the right hand set. 


(b) Hyperbolic tangent profiles, U = 
Uptanh y and density p = pg(1 - 
Atanh y) where y = 20z/H - 15. 
Up is positive for the first three 
curves at the left, zero for Ri = = 
and negative for the three curves on 
the right. The value of Ri marked on 
these curves is the minimum mean flow 
value at z = 3H/4. 


We may press the analogy between interfacial 
internal waves in a shear flow and surface waves 
further. The shape of surface gravity waves 
(narrower crests than troughs) and their habit of 
breaking forwards at the crests seems universal, 
in that it is independent of water depth, being 
observed and (where theory is available) predicted 
for both shallow and deep water waves. The internal 
waves observed in the experiments have similar prop- 
erties, accepting that the profile is inverted with 
respect to the surface waves if x < 0, even though 
they are not strictly interfacial waves or moving 
at the speed of one of the layers. This suggests 
that the shape and breaking, by convective overturn, 
of long first mode internal waves on a relatively 
narrow interface between two uniform layers follow 
the pattern observed in the experiments, independent 
of the depths of the layers, provided that the 
Richardson number of the mean flow in the interfacial 
region is small. 

Figure 6b is not symmetrical, a consequence of 
the asymmetry introduced by having unequal layer 
thicknesses above and below the interface. Trans-— 
lated to a situation in which wind is driving a 
flow above a shallow thermocline, the diagram 
implies that internal waves travelling with the 
wind (x > 0) will break at a greater amplitude (or 
later if the shear flow is increasing) then waves 
of the same length travelling against the wind. 

This result also follows from our analogy with 
surface waves since, for a given wavelength, surface 
waves of limiting (120° apex) amplitude in deep 
water (corresponding to the forward moving, x > 0, 
internal gravity waves) are higher than waves in 
shallow water (which correspond to the backward 
moving waves). Waves moving across the flow will 


TAAL AAAT NNN RAT IAAAL  IA T 


FIGURE 3. Internal waves in a shear flow with profiles of U and p similar to those of Figure 
2(b), except that the interface is at z = H/4 and the mean, depth averaged, flow is zero. The 
waves propagate to the left and in (a) the mean flow in the upper layer is to the right, lower 
to the left (y<0), in (b) there is no mean flow, whilst in (c) the mean flow in the upper layer 
is to the right and in the lower layer to the left (y > 0). 


h FIGURE 4. Interfacial waves in a two-layer fluid. In 
(a) the phase speed of the waves, c, is equal to the 
speed of the lower layer, Up. The wave shape is identi- 
cal to that of surface waves on a layer of depth hj, 

h, but inverted. (This corresponds to xy < 0). In (b), 

c = U;, and the wave shape is identical to that of 


surface waves on a layer of depth ho. 


——= 
—— 


627 


FIGURE 5. The onset of wave breaking for xy > 0. The waves are moving to the left. The mean Richardson number at the in- 


terface in the accelerating flow is approximately (a) 2.5 (b) 


(c) 0.36 (d) 0.25 (e) 0.18 (f) 0.14 (g) 0.11 (h) 0.09 


(i) 0.07 (j) 0.06 [from Thorpe (1968)]. Convective overturn is seen to begin at (c) and K-H.I at (i). The instability is 
not seen at the critical value of Ri because of the time needed for growth in the accelerating flow. 


not be unaffected by it. This process may be 
important in producing asymmetric directional wave 
spectra in the seasonal thermocline. 

In practice of course unidirectional flows and 
long trains of internal waves do not occur in the 
ocean. The component of the mean flow velocity 
normal to the direction of wave propagation appears 
to play no part in the breaking or dynamics of the 
waves, and the results should be valid for long 
crested waves even in (Ekman) spiral flows. A 
periodic shear flow applied to a wave, as when one 
internal wave moves through another, may produce 
locally the conditions for convective overturn of 
the kind we have described. The final stages of 
the experiments of Keulegan and Carpenter (1961) 
or Davis and Acrivos (1967) illustrate this process. 
In these experiments a short second mode wave is 
driven by resonant interaction from a long first 
mode wave, itself generated by a wavemaker. The 
shorter wave eventually breaks in the shear field 
of the longer first mode wave. 

Flow acceleration accompanies both the periodic 
flows in a wave field and the motion of the upper 
layers of the ocean during periods of wind forcing. 
In the experiments shown here breaking was induced 
by allowing the flow to accelerate uniformly. It 
was discovered that the energy of the fluctuating 
wave components was reduced very rapidly as a 
result of this acceleration. The consequent Rey— 
nolds stress working on the mean velocity gradient 
transferred energy to the mean flow. This inter- 
action may have important consequences on the 
development of the seasonal thermocline during 


1:0 
Wave 
slope 
0-5 
UNSTABLE 
0:5 1 5 (a) 
Rig 
0-4 
Wave 1 
slope 
UNSTABLE UNSTABLE 


Limit for 
convective overturn 


o 

nN 
nN 
Ww 


ail at 
XEON Rie XCHIONR NS 


FIGURE 6. Stability diagrams corresponding 
to the waves described in Figure 2, based 
on a calculation extended to third order 
(Thorpe, 1968). (a) Couette flow (b) Hyper- 
bolic tangent profiles. 


628 


periods of wind forcing and the acceleration of the 
mixing layer, but they are beyond the scope of this 
paper. 

It seems likely that in the seasonal thermocline 
short internal waves may break predominantly by 
convective overturn whilst the longer are more 
prone to K-H.I, but the balance of effects is not 
known. The importance of non-linearities in 
determining the condition of convective overturn 
and the unknown structure of the density and veloc-— 
ity fields make the problem difficult to resolve 
theoretically, and some effort is being directed 
towards an observational, and hence empirical, 
solution using small arrays of thermistors with 
rapid response times, and sensitive CTDs. 


REFERENCES 


Banks, W. H. H., P. G. Drazin, and M. B. Zaturska 
(1976). On the normal modes of parallel flow 
of inviscid stratified fluid. J. Fluid Mech. 
75, 149. 

Bell, T. H. (1974). Effects of shear on the prop- 
erties of internal gravity wave modes. Dt. 
jeitlehgooies 45 Bip Silo 

Cokelet, E. D. (1977). Breaking waves. Nature 
2677, 169% 

Davis, R. E, and A. Acrivos (1967). The stability 
of oscillating intérnal waves. J. Fluid Mech. 
30, 723. 

Groen, P. (1948). Two fundamental theorems on 
gravity waves in inhomogeneous incompressible 
fluids. Physica 14, 294. 


Halpern, D. (1974). Observations of the deepening 
of the wind-mixed layer in the Northeast Pacific 
Ocean. J. Phys. Oceanog. 4, 454. 

Howard, L. N. (1961). Note on a paper by John W. 
Miles. J. Fluid Mech. 10, 509. 

Keulegan, G. H., and L. H. Carpenter (1961). An 
experimental study of internal progressive 
oscillatory waves. Wat. Bur. Stand. Rep. No. 
7319. 

Miles, J. W. (1961). On the stability of hetero- 
geneous shear flows. J. Fluid Mech. 10, 496. 

Orlanski, I., and K. Bryan (1969). Formation of 
the thermocline step structure by large amplitude 
internal gravity waves. J. Geophys. Res. 74, 
6975Se 

Phillips, O. M., and M. L. Banner (1974). Wave 
breaking in the presence of wind drift and swell. 
J. Fluid Mech. 66, 625. 

Thorpe, S. A. (1971). Experiments on the instabil- 
ity of stratified shear flows: miscible fluids. 
J. Fluid Mech. 46, 299. 

Thorpe, S. A. (1974). Near-resonant forcing in a 
shallow two-layer fluid: a model for the 
internal surge in Loch Ness? J. Fluid Mech. 

O37, SOL)- 

Thorpe, S. As, A. J. Hall; es Taylor, and ai Alden 
(1977). Billows in Loch Ness. Deep-Sea Res. 
Zaye Sills 

Thorpe, S. A. (1978). On the shape and breaking 
of finite amplitude internal gravity waves in 
a shear flow. J. Fluid Mech. 85, 7. 

Woods, J. D. (1968). Wave-induced shear instability 
in the summer thermocline. J. Fluid Mech. 32, 
Tbe 


List of Participants 


Allan J. Acosta, California Institute of Technology, 
Pasadena, USA 

Bruce H. Adee, University of Washington, Seattle, 
USA 

Jose A. Alaez, Canal de Experiencias Hidrodinamicas, 
Madrid, Spain 

Klaus Albrecht, Institut fur Hydroakustik, Ottobrunn, 
Federal Republic of Germany 

Vladimir K. Ankudinov, Hydronautics, Inc., Laurel, 
USA 

Robert E. Apfel, Yale University, New Haven, USA 

Roger E. A. Arndt, University of Minnesota, 
Minneapolis, USA 

Glenn M. Ashe, U. S. Coast Guard, Washington, USA 

Daniel G. Bagnell, U. S. Coast Guard, Washington, 
USA 

Kwang-June Bai, David Taylor Naval Ship R & D 
Center, Bethesda, USA 

Ignacio Baquerizo Briones, Spanish Society of 
Naval Architects, Madrid, Spain 

Goran B. R. Bark, Swedish State Shipbuilding 
Experimental Tank, Goteborg, Sweden 

Steven J. Barker, University of California, 
Los Angeles, USA 

Franco C. Bau, Cantieri Navali Riuniti, Genoa, 
Italy 

Robert F. Beck, University of Michigan, Ann Arbor, 
USA 

Michael L. Billet, Pennsylvania State University, 
State College, USA 

William K. Blake, David Taylor Naval Ship R & D 
Center, Bethesda, USA 

Christian Bratu, Institut Francais Du Petrole, 
Rueil-Malmaison, France 

John P. Breslin, Stevens Institute of Technology, 
Hoboken, USA 

Neal G. Brower, Johns Hopkins University, Baltimore, 
USA 

Samuel H. Brown, David Taylor Naval Ship R & D 
Center, Bethesda, USA 

Donald R. Burklew, Operations Research, Inc., 
Silver Spring, USA 


629 


Otto Bussemaker, Schottel-Nederland B. V., The 
Hague, Netherlands 

Ben J. Cagle, Office of Naval Research, Pasadena, 
USA 

Nicholas Caracostas, Advanced Marine Enterprises, 
Inc., Washington, USA 

George F. Carrier, Harvard University, Cambridge, 
USA 

F. Sherman Cauldwell, Naval Ship Engineering 
Center, Washington, USA 

Tuncer Cebeci, Douglas Aircraft Company, Long 
Beach, USA 

Georges L. Chahine, Ecole Nationale Superieure de 
Techniques Avancees, Paris, France 

Ming-Shun Chang, David Taylor Naval Ship R & D 
Center, Bethesda, USA 

Richard B. Chapman, Science Applications, Inc., 
San Diego, USA 

Howard A: Chatterton, U. S. Coast Guard, Washington, 
USA 

Michael A. Chaszeyka, Office of Naval Research, 
Chicago, USA 

Henry M. Cheng, Office of the Chief of Naval 
Operations, Washington, USA 

Teresa Chereskin, Massachusetts Institute of 
Technology, Cambridge, USA 

George H. Christoph, Sun Shipbuilding & Dry Dock 
Company, Chester, USA 

Allen T. Chwang, California Institute of Technology, 
Pasadena, USA 

David W. Coder, David Taylor Naval Ship R & D 
Center, Bethesda, USA 

E. N. Comstock, Naval Ship Engineering Center, 
Washington, USA 

Genevieve Comte-Bellot, Ecole Centrale de Lyon, 
Ecully, France 

Reilley E. Conrad, Naval Ship Engineering Center, 
Washington, USA 

Ralph D. Cooper, Office of Naval Research, 
Washington, USA 

Bruce D. Cox, David Taylor Naval Ship R & D Center, 
Bethesda, USA 


630 


William E. Cummins, David Taylor Naval Ship R & D 
Center, Bethesda, USA 

Douglas J. Dahmer, David Taylor Naval Ship R & D 
Center, Bethesda, USA 

Tore G. Dalvag, AB Karlstads Mekaniska Werkstad, 
Kristinehamn, Sweden 

Stephen H. Davis, John Hopkins University, 
Baltimore, USA 

Charles W. Dawson, David Taylor Naval Ship R & D 
Center, Bethesda, USA 

William G. Day, Jr., David Taylor Naval Ship R & D 
Center, Bethesda, USA 

Jean-Francois M. Demanche, Bassin d'Essais des 
Carenes, Paris, France 

Jean-Claude Dern, Bassin d'Essais des Carenes, 
Paris, France 

William K. Dewar, Woods Hole Oceanographic Institu- 
tion, Woods Hole, USA 

Warren C. Dietz, U. S. Coast Guard, Washington, USA 

Richard C. DiPrima, Rensselaer Polytechnic 
Institute, Troy, USA 

Jan M. Dirkzwager, Ministry of Defence, The Hague, 
Netherlands 

Stanley W. Doroff, Office of Naval Research, 
Washington, USA 

Phillip Eisenberg, Hydronautics, Inc., Laurel, 
USA 

N. M. El-Hady, Virginia Polytechnic Institute, 
Blacksburg, USA 

J. W. English, National Maritime Institute, 
Feltham, England 

Robert Falls, Maritime Administration, Washington, 
USA 


Hermann F. Fasel, University of Stuttgart, Stuttgart, 


Federal Republic of Germany 

Archibald M. Ferguson, University of Glasgow, 
Glasgow, Scotland ; 

Peter D. Fitzgerald, Exxon International, Florham 
Park, USA 

Francois N. Frenkiel, David Taylor Naval Ship R & D 
Center, Bethesda, USA 

Daniel H. Fruman, Laboratoire d'Aerodynamique, 
Orsay, France 

Donald Fuhs, David Taylor Naval Ship R & D Center, 
Bethesda, USA 

Michael Gaster, National Maritime Institute, 
Middlesex, England 

Edward M. Gates, University of Alberta, Edmonton, 
Canada 

Carl Gazley, Jr., The Rand Corporation, Santa 
Monica, USA 

William K. George, State University of New York, 
Buffalo, USA 

Robert K. Geiger, Office of Naval Research, 
Washington, USA 

Douglas L. Gile, Boulder, USA 

Alex Goodman, Hydronautics, Inc., Laurel, USA 

Stephan At. Goranov, Bulgarian Ship Hydrodynamics 
Center, Varna, Bulgaria 

Paul S. Granville, David Taylor Naval Ship R & D 
Center, Bethesda, USA 


Richard A. Griffiths, U. S. Coast Guard, Washington, 


USA 

William L. Haberman, Rockville, USA 

Jacques B. Halder, David Taylor Naval Ship R & D 
Center, Bethesda, USA 

Francis R. Hama, Princeton University, Princeton, 
USA 

Henry J. Haussling, David Taylor Naval Ship R & D 
Center, Bethesda, USA 


Grant E. Hearn, British Ship Research Association, 
Wailsend, England 

Harold I. Heaton, Johns Hopkins University, Applied 
Physics Laboratory, Laurel, USA 

Isom H. Herron, Howard University, Washington, USA 

Leo H. Holthuijsen, Delft University of Technology, 
Delft, Netherlands 

Max G. A. Honkanen, Engineering Company M. G. 
Honkanen, Helsinki, Finland 

Louis N. Howard, Massachusetts Institute of 
Technology, Cambridge, USA 

Chun-Che Hsu, Hydronautics, Inc., Laurel, USA 

Thomas T. Huang, David Taylor Naval Ship R & D 
Center, Bethesda, USA 

Lee M. Hunt, National Academy of Sciences-National 
Research Council, Washington, USA 

Stephen J. Hunter, Admiralty Marine Technology 
Establishment, Haslar, England 

Erling Huse, Ship Research Institute of Norway, 
Trondheim, Norway 

Takao Inui, University of Tokyo, Tokyo, Japan 

Shunichi Ishida, Ishikawajima-Harima Heavy 
Industries Co., Ltd., Yokohama, Japan 

Gerald S. Janowitz, North Carolina State University, 
Raleigh, USA 

Stuart D. Jessup, David Taylor Naval Ship R & D 
Center, Bethesda, USA 

Bruce Johnson, U. S. Naval Academy, Annapolis, 
USA 

Virgil E. Johnson, Hydronautics, Inc., Laurel, USA 

Francois J. Jouaillec, French Ministry of Defence, 
Paris, France 

Peter Numa Joubert, University of Melbourne, 
Melbourne, Australia 

Vijay K. Jyoti, Dominion Engineering Works, Ltd., 
Montreal, Canada 

Lakshmi H. Kantha, Johns Hopkins University, 
Baltimore, USA 

Paul Kaplan, Oceanics, Inc., Plainview, USA 

George M. Kapsilis, M. Rosenblatt & Son, Inc., 
Gaithersburg, USA 

Hiroharu Kato, University of Tokyo, Tokyo, Japan 

R. G. Keane, Jr., Naval Ship Engineering Center, 
Washington, USA 

Andreas P. Keller, Technical University of Munich, 
Munich, Federal Republic of Germany 

Colen G. Kennell, Naval Ship Engineering Center, 
Washington, USA 

Philip S. Klebanoff, National Bureau of Standards 
Washington, USA 

Leslie S. G. Kovasznay, Johns Hopkins University, 
Baltimore, USA 

Ruby E. Krishnamurti, Florida State University, 
Tallahassee, USA 

Gert Kuiper, Netherlands Ship Model Basin, 
Wageningen, Netherlands 

Jurgen H. Kux, University of Hamburg, Hamburg, 
West Germany 

Louis Landweber, University of Iowa, Iowa City, USA 

Arie J. W. Lap, Royal Netherlands Naval College, 
Dan Helder, Netherlands 


“Jochen Lauden, Hamburgische Shiffbau-Versuchsanstalt, 


Hamburg, Federal Republic of Germany 

George K. Lea, National Science Foundation, 
Washington, USA 

Yves Lecoffre, Alsthom Atlantique, Grenoble Cedex, 
France 

Choung M. Lee, David Taylor Naval Ship R & D 
Center, Bethesda, USA 

Yu-Tai Lee, University of Iowa, Iowa City, USA 


Lennox J. Leggat, Defence Research Establishment 
Atlantic, Nova Scotia, Canada 

Spiros G. Lekoudis, Lockheed-Georgia Company, 
Marietta, USA 

John A. LeRoy, Australian Naval Attache Office, 
Washington, USA 

Wen-Chin Lin, David Taylor Naval Ship R & D 
Center, Bethesda, USA 

Robert R. Long, Johns Hopkins University, Baltimore, 
USA 

Hans J. Lugt, David Taylor Naval Ship R & D Center, 
Bethesda, USA 

Justin McCarthy, David Taylor Naval Ship R & D 
Center, Bethesda, USA 

John M. Macha, Texas A & M University, College 
Station, USA 

Leslie Mack, California Institute of Technology, 
Pasadena, USA 

Toshio Maeda, Mitsubishi Heavy Industries, Ltd., 
Kobe, Japan 

Allen H. Magnuson, Virginia Polytechnic Institute, 
Blacksburg, USA 

Robert W. Manning, Naval Sea Systems Command, 
Washington, USA 


Chiang C. Mei, Massachusetts Institute of Technology, 


Cambridge, USA 

Kenneth R. Meldahl, The Boeing Company, Seattle, 
USA 

John W. Miles, University of California, San Diego, 
USA 

Robert J. Mindak, Office of Naval Research, 
Washington, USA 

Erik Mollo-Christensen, Massachusetts Institute 
of Technology, Cambridge, USA 

Vincent Monacella, David Taylor Naval Ship R & D 
Center, Bethesda, USA 

Alan W. Moore, Admiralty Marine Technology 
Establishment, Teddington, England 

David D. Moran, David Taylor Naval Ship R & D 
Center, Bethesda, USA 

William B. Morgan, David Taylor Naval Ship R & D 
Center, Bethesda, USA 

Kazuhiro Mori, Hiroshima University, Hiroshima, 
Japan 

Parma Mungur, Lockheed-Georgia Company, Marietta, 
USA 

Walter H. Munk, University of California, San 
Diego, USA 

Hitoshi Murai, Tohoku University, Sendai, Japan 

Paul M. Naghdi, University of California, Berkeley, 
USA 

Ali H. Nayfeh, Virginia Polytechnic Institute, 
Blacksburg, USA 

J. Nicholas Newman, Massachusetts Institute of 
Technology, Cambridge, USA 

Francis Noblesse, Massachusetts Institute of 
Technology, Cambridge, USA 

David J. Norton, Texas A & M University, College 
Station, USA 

John A. Norton, Bird-Johnson Company, Walpole, USA 

John F. O'Dea, David Taylor Naval Ship R & D 
Center, Bethesda, USA 

Denis C. O'Neill, Ministry of Defence, Bath, 
England 

Marinus W. C. Oosterveld, Netherlands Ship Model 
Basin, Wageningen, Netherlands 

Blaine R. Parkin, Pennsylvania State University, 
State College, USA 

Virendra C. Patel, University of Iowa, Iowa City, 
USA 


631 


Mariano Perez, Canal de Experiencias Hidrodinamicas, 
Madrid, Spain 

Gonzalo Perez Gomez, Spanish Society of Naval 
Architects, Madrid, Spain 

Frank B. Peterson, David Taylor Naval Ship R & D 
Center, Bethesda, USA 

Owen M. Phillips, Johns Hopkins University, 
Baltimore, USA 

Ennio Piantini, Ministero Difesa Marina, Rome, 
Italy 

Pao C. Pien, David Taylor Naval Ship R & D Center, 
Bethesda, USA 

Leonard J. Pietrafesa, North Carolina State 
University, Raleigh, USA 

Gregory Platzer, David Taylor Naval Ship R & D 
Center, Bethesda, USA 

Allen Plotkin, University of Maryland, College 
Park, USA 

Alan Powell, David Taylor Naval Ship R & D Center, 
Bethesda, USA 

Jaakko V. Pylkkanen, Helsinki University of 
Technology, Helsinki, Finland 

Arthur M. Reed, David Taylor Ship R & D Center, 
Bethesda, USA 

Sidney R. Reed, Office of Naval Research, 
Washington, USA 

Bernd Remmers, Kempf & Remmers, Hamburg, Federal 
Republic of Germany 

Eli Reshotko, Case Western Reserve University, 
Cleveland, USA 

Wolfgang Reuter, Naval Ship Engineering Center, 
Washington, USA 

M. B. Ricketts, David Taylor Naval Ship R & D 
Center, Bethesda, USA 

Joel C. W. Rogers, Johns Hopkins University, 
Applied Physics Laboratory, Laurel, USA 

Richard R. Rojas, Naval Research Laboratory, 
Washington, USA 

Olle G. A. Rutgersson, Swedish State Shipbuilding 
Experimental Tank, Goteborg, Sweden 

Manley St. Denis, U. S. Naval Academy, Annapolis, 
USA 

Nils Salvesen, David Taylor Naval Ship R & D 
Center, Bethesda, USA 

Geert H. Schmidt. University of Technology, Delft, 
Netherlands 

Michael Schmiechen, VWS Berlin Model Basin, Berlin, 
Federal Republic of Germany 

Joanna W. Schot, David Taylor Naval Ship R & D 
Center, USA 

Paul Sclavounous, Massachusetts Institute of 
Technology, Cambridge, USA 

Carl A. Scragg, Science Applications Inc., San 
Diego, USA 

Som D. Sharma, Massachusetts Institute of Tech- 
nology, Cambridge, USA 

Young T. Shen, David Taylor Naval Ship R & D 
Center, Bethesda, USA 

Vincent G. Sigillito, Johns Hopkins University, 
Applied Physics Laboratory, Laurel, USA 

Leslie Sinclair, Stone Manganese Marine Ltd., 
London, England 

Olav H. Slaattelid, Ship Research Institute of 
Norway, Trondheim, Norway 

Neill S. Smith, Naval Coastal Systems Center, 
Panama City, USA 

J. A. Sparenberg, University of Groningen, 
Groningen, Netherlands 

Nicholas R. Stark, Beltsville, USA 

Frank X. Stora, U. S. Army, Fort Belvoir, USA 


632 


Albert M. Sturrman, Royal Netherlands Navy, The 
Hague, Netherlands 

Ming-Yang Su, U. S. Navy, NORDA, Bay St. Louis, 
USA 

Hiraku Tanaka, Ship Research Institution, Tokyo, 
Japan 

Stephen A. Thorpe, Institute of Oceanographic 
Sciences, Surrey, England 

Yoshiaki Toba, Tohoku University, Sendai, Japan 

Ernest O. Tuck, University of Adelaide, Adelaide, 
Australia 

Marshall P. Tulin, Hydronautics, Inc., Laurel, 
USA 

Ka-Kit Tung, Dynatech, Torrance, USA 

J. Stewart Turner, Australian National University, 
Canberra, Australia 

Willem van Berlekom, Swedish State Shipbuilding 
Experimental Tank, Goteborg, Sweden 

Jan D. van Manen, Netherlands Ship Model Basin, 
Wageningen, Netherlands 

Pieter van Oossanen, Netherlands Ship Model Basin, 
Wageningen, Netherlands 

Jan van der Meulen, Netherlands Ship Model Basin, 
Wageningen, Netherlands 

Christian von Kerczek, David Taylor Naval Ship 
R & D Center, Bethesda, USA 


Alice Vucinic, Rijeka University, Rijeka, Yugoslavia 


Nicholas Vytlacil, Westinghouse Electric COED, 
Annapolis, USA 

David A. Walden, U. S. Coast Guard, Washington, 
USA 


Lawrence W. Ward, Webb Institute of Naval Archi- 
tecture, Glen Cove, USA 

Richard M. Wargelin, U. S. Navy, Suitland, USA 

John V. Wehausen, University of California, 
Berkeley, USA 

Michael A. Weissman, Flow Industries, Inc., Kent, 
USA 

Ernst-August Weitendorf, University of Hamburg, 
Hamburg, Federal Republic of Germany 

John R. Weske, University of Maryland, College 
Park, USA 

Robert E. Whitehead, Office of Naval Research, 
Washington, USA 

Sheila Widnall, Massachusetts Institute of 
Technology, Cambridge, USA 

Karl Wieghardt, University of Hamburg, Hamburg, 
Federal Republic of Germany 

Colin B. Wills, Admiralty Marine Technology 
Establishment, Haslar, England 

Theodore Y. Wu, California Institute of Technology, 
Pasadena, USA 

Chia-Shun Yih, University of Michigan, Ann Arbor, 
USA 

Bohyun Yim, David Taylor Naval Ship R & D Center, 
Bethesda, USA 

Hajime Yuasa, Mitsui Engineering & Shipbuilding 
Co., Ltd., Tokyo, Japan 

P. Richard Zarda, David Taylor Naval Ship R & D 
Center, Bethesda, USA 


PREVIOUS BOOKS IN THE NAVAL HYDRODYNAMICS SERIES 


"First Symposium on Naval Hydrodynamics." National 
Academy of Science-Nation Research Council, 
Publication 515, 1957. Washington, D. C.; 
PB133732. 

"Second Symposium on Naval Hydrodynamics: Hydro- 
dynamic Noise and Cavity Flow," Office of Naval 


Research, Department of the Navy, ACR-38, 1958; 
PB157668. 
"Third Symposium on Naval Hydrodynamics: High 


Performance Ships," Office of Naval Research, 
Department of the Navy, ACR-65, 1960; AD430729. 

"Fourth Symposium on Naval Hydrodynamics: Propul- 
sion and Hydroelasticity," Office of Naval 
Research, Department of the Navy, ACR-92, 
AD447732. 

"The Collected Papers of Sir Thomas Havelock on 
Hydrodynamics," Office of Naval Research, 
Department of the Navy, ACR-103, 1963; AD623589. 

"Fifth Symposium on Naval Hydrodynamics: Ship 
Motions and Drag Reduction," Office of Naval 
Research, Department of the Navy, ACR-112, 1964; 
AD640539. 

"Sixth Symposium on Naval Hydrodynamics: Physics 
of Fluids, Maneuverability and Ocean Platforms, 
Ocean Waves, and Ship-Generated Waves and Wave 
Resistance," Office of Naval Research, Depart- 
ment of the Navy, ACR-136, 1966; AD676079. 

"Seventh Symposium on Naval Hydrodynamics: Unsteady 
Propeller Forces, Fundamental Hydrodynamics, 


1962; 


633 


Unconventional Propulsion," Office of Naval 
Research, Department of the Navy, DR-148, 1968; 
AD721180. 

"Eighth Symposium on Naval Hydrodynamics: Hydro- 
dynamics in the Ocean Environment," Office of 
Naval Research, Department of the Navy, ACR-179, 
1970; AD748721. 

"Ninth Symposium on Naval Hydrodynamics: Unconven- 
tional Ships, Ocean Engineering, Frontier Problems," 
Office of Naval Research, Department of the 
Navy, ACR-203, 1972; Two Volumes; Vol. 1, ADA- 
010505; Vol 2, ADAO10506. 

"Tenth Symposium on Naval Hydrodynamics: Hydrody- 
namics for Safety, Fundamental Hydrodynamics," 
Office of Naval Research, Department of the Navy, 
ACR-204, 1974; ADA0O22379. 

"Eleventh Symposium on Naval Hydrodynamics: Unsteady 
Hydrodynamics of Marine Vehicles," Office of 
Naval Research, Department of the Navy. Also 
available from Mechanical Engineering Publications 
Limited, London and New York. 


The above books are avilable on microfilm 
and microfiche from the National Technical 
Information Service, U. S. Department of 
Commerce, Springfield, Virginia 22151. 

Some early issues are also available in paper 
copies. The catalog numbers, as of the 

date of this issue, are shown for each book. 


NOTE: 


i 
i mal 
f 
i 
te a 
a “ 
é Ty 
P \ 
* Fi i 
« 3 
= 
ie e 
{ 
y 
3 
Mee = 
= ‘ os 
ii wi Ste | 
iT > y 
Sill ri 
SMa ff 
> wish 
aaa 
i nt 
Ag D ie ce ies 
' re 
1 


Panipat 
ae 
Pat tle 


i 
i 

Ma 
on i 


Gye) 
(ae As 


} iy 

hee Wit 
iy) 14 
we rei 


1uly 
i 


oe 
ptt 


4 


ie 


l— 


Ailis 


inl 
‘ f 


i “ 
1 atl 


rey 


lu 
Bia 


With 
nn, r i vt With r 
PORN Nae Ne 


} 
ie. 
Fi 


\ 


Ve ANF 
AMR). Mere na ; 


lity 4 x 
yee 


an | 
PA Pits of 7 
P Nai) id 


th " t 
ivr un 
th 


Mites 
i |