Skip to main content

Full text of "Water Waves The Mathematical Theory With Applications"

See other formats


TEXT  FLY  WITHIN 
THE  BOOK  ONLY 


u<  OU 

^  GO 


DO 

160917>m 


STOKER 


WATER  WAVES 


OSMANIA  UNIVERSITY  LIBRARY 

Call  No.   ^VL-'Sj  S£7  lO     •  Accession  No. 
Author       S-Jfc  fcfcT,      T  - 

Title         CO<X^L/L 
This  book  should  be  returned  on  or  before  the  date  last  marked  below. 


WATER  WAVES 

The    Mathematical   Theory   with    Applications 


PURE  AND  APPLIED  MATHEMATICS 

A  Series  of  Texts  and  Monographs 

Edited  by 
R.  COURANT   •    L.  BERS   .    J.  J.  STOKER 


VOLUME  IV 


Waves  about  a  harbor 


WATER  WAVES 

The  Mathematical  Theory 
with  Applications 


J.  J.  STOKER 

INSTITUTE  OF  MATHEMATICAL  SCIENCES 
NEW  YORK  UNIVERSITY,  NEW  YORK 


19    f   fill     57 


INTERSCIENCE  PUBLISHERS,  INC.,  NEW  YORK 
INTERSCIENCE  PUBLISHERS  LTD.,  LONDON 


All  Rights  Reserved 

LIBRARY  OF  CONGRESS  CATALOG  CARD  NUMBER 
56-8228 


INTERSCIENCE  PUBLISHERS     INC. 
250  Fifth  Avenue,   New   York  1,   N.Y. 

For  Great  Britain  and  Northern  Ireland: 

INTERSCIENCE  PUBLISHERS  LTD. 
88/90  Chancery  Lane,  London  W.   C.  2 

PRINTED    IN    THE    NETHERLANDS 
BY   LATE   HOITSEMA    BROTHERS,    GRONINGEN 


To 

NANCY 


Introduction 


1.  Introduction 


The  purpose  of  this  book  is  to  present  a  connected  account  of  the 
mathematical  theory  of  wave  motion  in  liquids  with  a  free  surface 
and  subjected  to  gravitational  and  other  forces,  together  with  ap- 
plications to  a  wide  variety  of  concrete  physical  problems. 

Surface  wave  problems  have  interested  a  considerable  number  of 
mathematicians  beginning  apparently  with  Lagrange,  and  con- 
tinuing with  Cauchy  and  Poisson  in  France.*  Later  the  British  school 
of  mathematical  physicists  gave  the  problems  a  good  deal  of  atten- 
tion, and  notable  contributions  were  made  by  Airy,  Stokes,  Kelvin, 
Rayleigh,  and  Lamb,  to  mention  only  some  of  the  better  known.  In 
the  latter  part  of  the  nineteenth  century  the  French  once  more  took 
up  the  subject  vigorously,  and  the  work  done  by  St.  Venant  and 
Boussinesq  in  this  field  has  had  a  lasting  effect:  to  this  day  the 
French  have  remained  active  and  successful  in  the  field,  and  par- 
ticularly in  that  part  of  it  which  might  be  called  mathematical 
hydraulics.  Later,  Poincar^  made  outstanding  contributions  par- 
ticularly with  regard  to  figures  of  equilibrium  of  rotating  and  gravi- 
tating liquids  (a  subject  which  will  not  be  discussed  in  this  book); 
in  this  same  field  notable  contributions  were  made  even  earlier 
by  Liapounoff.  One  of  the  most  outstanding  accomplishments  in  the 
field  from  the  purely  mathematical  point  of  view  —  the  proof  of  the 
existence  of  progressing  waves  of  finite  amplitude  —  was  made  by 
Nckrassov  [N.I],  [N.lajf  in  1921  and  independently  by  a  different 
means  by  Levi-Civita  [L.7]  in  1925. 

The  literature  concerning  surface  waves  in  water  is  very  extensive. 
In  addition  to  a  host  of  memoirs  and  papers  in  the  scientific  journals, 
there  are  a  number  of  books  which  deal  with  the  subject  at  length. 
First  and  foremost,  of  course,  is  the  book  of  Lamb  [L.3],  almost 
a  third  of  which  is  concerned  with  gravity  wave  problems.  There 
are  books  by  Bouasse  [B.15],  Thorade  [T.4],  and  Sverdrup  [S.39] 

*  This  list  would  be  considerably  extended  (to  include  Euler,  the  Bernoullis, 

and  others)  if  hydrostatics  were  to  be  regarded  as  an  essential  part  of  our  subject. 

t  Numbers  in  square  brackets  refer  to  the  bibliography  at  the  end  of  the  book. 


X  INTRODUCTION 

devoted  exclusively  to  the  subject.  The  book  by  Thorade  consists 
almost  entirely  of  relatively  brief  reviews  of  the  literature  up  to 
1931  —  an  indication  of  the  extent  and  volume  of  the  literature 
on  the  subject.  The  book  by  Sverdrup  was  written  with  the  special 
needs  of  oceanographers  in  mind.  One  of  the  main  purposes  of  the 
present  book  is  to  treat  some  of  the  more  recent  additions  to  our 
knowledge  in  the  field  of  surface  wave  problems.  In  fact,  a  large  part 
of  the  book  deals  with  problems  the  solutions  of  which  have  been 
found  during  and  since  World  War  II;  this  material  is  not  available 
in  the  books  just  now  mentioned. 

The  subject  of  surface  gravity  waves  has  great  variety  whether 
regarded  from  the  point  of  view  of  the  types  of  physical  problems 
which  occur,  or  from  the  point  of  view  of  the  mathematical  ideas 
and  methods  needed  to  attack  them.  The  physical  problems  range 
from  discussion  of  wave  motion  over  sloping  beaches  to  flood  waves 
in  rivers,  the  motion  of  ships  in  a  sea-way,  free  oscillations  of  enclosed 
bodies  of  water  such  as  lakes  and  harbors,  and  the  propagation  of 
frontal  discontinuities  in  the  atmosphere,  to  mention  just  a  few. 
The  mathematical  tools  employed  comprise  just  about  the  whole  of 
the  tools  developed  in  the  classical  linear  mathematical  physics 
concerned  with  partial  differential  equations,  as  well  as  a  good  part 
of  what  has  been  learned  about  the  nonlinear  problems  of  mathe- 
matical physics.  Thus  potential  theory  and  the  theory  of  the  linear 
wave  equation,  together  with  such  tools  as  conformal  mapping  and 
complex  variable  methods  in  general,  the  Laplace  and  Fourier 
transform  techniques,  methods  employing  a  Green's  function,  integral 
equations,  etc.  are  used.  The  nonlinear  problems  arc  of  both  elliptic 
and  hyperbolic  type. 

In  spite  of  the  diversity  of  the  material,  the  book  is  not  a  collection 
of  disconnected  topics,  written  for  specialists,  and  lacking  unity  and 
coherence.  Instead,  considerable  pains  have  been  taken  to  supply 
the  fundamental  background  in  hydrodynamics  —  and  also  in  some 
of  the  mathematics  needed  —  and  to  plan  the  book  in  order  that  it 
should  be  as  much  as  possible  a  self-contained  and  readable  whole. 
Though  the  contents  of  the  book  are  outlined  in  detail  below,  it  has 
some  point  to  indicate  briefly  here  its  general  plan.  There  arc  four 
main  parts  of  the  book: 

Part  I,  comprising  Chapters  1  and  2,  presents  the  derivation  of 
the  basic  hydrodynamic  theory  for  non-viscous  incompressible  fluids, 
and  also  describes  the  two  principal  approximate  theories  which  form 


INTRODUCTION  XI 

the  basis  upon  which  most  of  the  remainder  of  the  book  is  built. 

Part  II,  made  up  of  Chapters  3  to  9  inclusive,  is  based  on  the  ap- 
proximate theory  which  results  when  the  amplitude  of  the  wave 
motions  considered  is  small.  The  result  is  a  linear  theory  which  from 
the  mathematical  point  of  view  is  a  highly  interesting  chapter  in 
potential  theory.  On  the  physical  side  the  problems  treated  include 
the  propagation  of  waves  from  storms  at  sea,  waves  on  sloping 
beaches,  diffraction  of  waves  around  a  breakwater,  waves  on  a 
running  stream,  the  motion  of  ships  as  floating  rigid  bodies  in  a  sea- 
way. Although  this  theory  was  known  to  Lagrange,  it  is  often  referred 
to  as  the  Cauchy-Poisson  theory,  perhaps  because  these  two  mathe- 
maticians were  the  first  to  solve  interesting  problems  by  using  it. 

Part  III,  made  up  of  Chapters  10  and  11,  is  concerned  with  problems 
involving  waves  in  shallow  water.  The  approximate  theory  which 
results  from  assuming  the  water  to  be  shallow  is  not.  a  linear  theory, 
and  wave  motions  with  amplitudes  which  are  not  necessarily  small 
can  be  studied  by  its  aid.  The  theory  is  often  attributed  to  Stokes 
and  Airy,  but  was  really  known  to  Lagrange.  If  linearized  by  making 
the  additional  assumption  that  the  wave  amplitudes  are  small,  the 
theory  becomes  the  same  as  that  employed  as  the  mathematical 
basis  for  the  theory  of  the  tides  in  the  oceans.  In  the  lowest  order 
of  approximation  the  nonlinear  shallow  water  theory  results  in  a 
system  of  hyperbolic  partial  differential  equations,  which  in  im- 
portant special  cases  can  be  treated  in  a  most  illuminating  way  with 
the  aid  of  the  method  of  characteristics.  The  mathematical  methods 
are  treated  in  detail  in  Chapter  10.  The  physical  problems  treated  in 
Chapter  10  are  quite  varied;  they  include  the  propagation  of  unsteady 
waves  due  to  local  disturbances  into  still  water,  the  breaking  of 
waves,  the  solitary  wave,  floating  breakwaters  in  shallow  water.  A 
lengthy  section  on  the  motions  of  frontal  discontinuities  in  the 
atmosphere  is  included  also  in  Chapter  10.  In  Chapter  11,  entitled 
Mathematical  Hydraulics,  the  shallow  water  theory  is  employed  to 
study  wave  motions  in  rivers  and  other  open  channels  which,  unlike 
the  problems  of  the  preceding  chapter,  are  largely  conditioned  by 
the  necessity  to  consider  resistances  to  the  flow  due  to  the  rough 
sides  and  bottom  of  the  channel.  Steady  flows,  and  steady  progressing 
waves,  including  the  problem  of  roll  waves  in  steep  channels,  are 
first  studied.  This  is  followed  by  a  treatment  of  numerical  methods 
of  solving  problems  concerning  flood-waves  in  rivers,  with  the  object 
of  making  flood  predictions  through  the  use  of  modern  high  speed 


XII  INTRODUCTION 

digital  computers.  That  such  methods  can  be  used  to  furnish  accurate 
predictions  has  been  verified  for  a  flood  in  a  400-mile  stretch  of  the 
Ohio  River,  and  for  a  flood  coming  down  the  Ohio  River  and  passing 
through  its  junction  with  the  Mississippi  River. 

Part  IV,  consisting  of  Chapter  12,  is  concerned  with  problems 
solved  in  terms  of  the  exact  theory,  in  particular,  with  the  use  of  the 
exact  nonlinear  free  surface  conditions.  A  proof  of  the  existence  of 
periodic  waves  of  finite  amplitude,  following  Levi-Civita  in  a  general 
way,  is  included. 

The  amount  of  mathematical  knowledge  needed  to  read  the  book 
varies  in  different  parts.  For  considerable  portions  of  Part  II  the 
elements  of  the  theory  of  functions  of  a  complex  variable  are  assumed 
known,  together  with  some  of  the  standard  facts  in  potential  theory. 
On  the  other  hand  Part  III  requires  much  less  in  the  way  of  specific 
knowledge,  and,  as  was  mentioned  above,  the  basic  theory  of  the 
hyperbolic  differential  equations  used  there  is  developed  in  all  detail 
in  the  hope  that  this  part  would  thus  be  made  accessible  to  engineers, 
for  example,  who  have  an  interest  in  the  mathematical  treatment  of 
problems  concerning  flows  and  wave  motions  in  open  channels. 

In  general,  the  author  has  made  considerable  efforts  to  try  to 
achieve  a  reasonable  balance  between  the  mathematics  and  the 
mechanics  of  the  problems  treated.  Usually  a  discussion  of  the  physical 
factors  and  of  the  reasons  for  making  simplified  assumptions  in  each 
new  type  of  concrete  problem  precedes  the  precise  formulation  of  the 
mathematical  problems.  On  the  other  hand,  it  is  hoped  that  a  clear 
distinction  between  physical  assumptions  and  mathematical  deduc- 
tions —  so  often  shadowy  and  vague  in  the  literature  concerned 
with  the  mechanics  of  continuous  media  —  has  always  been  main- 
tained. Efforts  also  have  been  made  to  present  important  portions 
of  the  book  in  such  a  way  that  they  can  be  read  to  a  large  extent 
independently  of  the  rest  of  the  book;  this  was  done  in  some  cases 
at  the  expense  of  a  certain  amount  of  repetition,  but  it  seemed  to 
the  author  more  reasonable  to  save  the  time  and  efforts  of  the  reader 
than  to  save  paper.  Thus  the  portion  of  Chapter  10  concerned  with 
the  dynamics  of  the  motion  of  fronts  in  meteorology  is  largely 
self-contained.  The  same  is  true  of  Chapter  11  on  mathematical 
hydraulics,  and  of  Chapter  9  on  the  motion  of  ships. 

Originally  this  book  had  been  planned  as  a  brief  general  introduc- 
tion to  the  subject,  but  in  the  course  of  writing  it  many  gaps  and 
inadequacies  in  the  literature  were  noticed  and  some  of  them  have 


INTRODUCTION  XIII 

been  filled  in;  thus  a  fair  share  of  the  material  presented  represents 
the  result  of  researches  carried  out  quite  recently.  A  few  topics  which 
are  even  rather  speculative  have  been  dealt  with  at  some  length 
(the  theory  of  the  motion  of  fronts  in  dynamic  meteorology,  given 
in  Chapter  10.12,  for  example);  others  (like  the  theory  of  waves  on 
sloping  beaches)  have  been  treated  at  some  length  as  much  because 
the  author  had  a  special  fondness  for  the  material  as  for  their  intrinsic 
mathematical  interest.  Thus  the  author  has  written  a  book  which  is 
rather  personal  in  character,  and  which  contains  a  selection  of 
material  chosen,  very  often,  simply  because  it  interested  him,  and 
he  has  allowed  his  predilections  and  tastes  free  rein.  In  addition, 
the  book  has  a  personal  flavor  from  still  another  point  of  view  since 
a  quite  large  proportion  of  the  material  presented  is  based  on  the  work 
of  individual  members  of  the  Institute  of  Mathematical  Sciences  of 
New  York  University,  and  on  theses  and  reports  written  by  students 
attending  the  Institute.  No  attempt  at  completeness  in  citing  the 
literature,  even  the  more  recent  literature,  was  made  by  the  author; 
on  the  other  hand,  a  glance  at  the  Bibliography  (which  includes 
only  works  actually  cited  in  the  book)  will  indicate  that  the  recent 
literature  has  not  by  any  means  been  neglected. 

In  early  youth  by  good  luck  the  author  came  upon  the  writings 
of  scientists  of  the  British  school  of  the  latter  half  of  the  nineteenth 
century.  The  works  of  Tyndall,  Huxley,  and  Darwin,  in  particular, 
made  a  lasting  impression  on  him.  This  could  happen,  of  course,  only 
because  the  books  were  written  in  an  understandable  way  and  also 
in  sucli  a  way  as  to  create  interest  and  enthusiasm:  —  but  this  was 
one  of  the  principal  objects  of  this  school  of  British  scientists. 
Naturally  it  is  easier  to  write  books  on  biological  subjects  for  non- 
specialists  than  it  is  to  write  them  on  subjects  concerned  with  the 
mathematical  sciences  —  just  because  the  time  and  effort  needed  to 
acquire  a  knowledge  of  modern  mathematical  tools  is  very  great. 
That  the  task  is  not  entirely  hopeless,  however,  is  indicated  by  John 
TyndaU's  book  on  sound,  which  should  be  regarded  as  a  great  classic 
of  scientific  exposition.  On  the  whole,  the  British  school  of  popularizers 
of  science  wrote  for  people  presumed  to  have  little  or  no  foreknow- 
ledge of  the  subjects  treated.  Now-a-days  there  exists  a  quite  large 
potential  audience  for  books  on  subjects  requiring  some  knowledge 
of  mathematics  and  physics,  since  a  large  number  of  specialists  of 
all  kinds  must  have  a  basic  training  in  these  disciplines.  The  author 
hopes  that  this  book,  which  deals  with  so  many  phenomena  of  every 


XIV  INTRODUCTION 

day  occurrence  in  nature,  might  perhaps  be  found  interesting,  and 
understandable  in  some  parts  at  least,  by  readers  who  have  some 
mathematical  training  but  lack  specific  knowledge  of  hydro- 
dynamics.* For  example,  the  introductory  discussion  of  waves  on 
sloping  beaches  in  Chapter  5,  the  purely  geometrical  discussion  of 
the  wave  patterns  created  by  moving  ships  in  Chapter  8,  great  parts 
of  Chapters  10  and  11  on  waves  in  shallow  water  and  flood  waves  in 
rivers,  as  well  as  the  general  discussion  in  Chapter  10  concerning 
the  motion  of  fronts  in  the  atmosphere,  are  in  this  category. 

2.  Outline  of  contents 

It  has  already  been  stated  that  this  book  is  planned  as  a  coherent 
and  unified  whole  in  spite  of  the  variety  and  diversity  of  its  contents 
on  both  the  mathematical  and  the  physical  sides.  The  possibility  of 
achieving  such  a  purpose  lies  in  the  fortunate  fact  that  the  material 
can  be  classified  rather  readily  in  terms  of  the  types  of  mathematical 
problems  which  occur,  and  this  classification  also  leads  to  a  reasonably 
consistent  ordering  of  the  material  with  respect  to  the  various  types 
of  physical  problems.  The  book  is  divided  into  four  main  parts. 

Part  I  begins  with  a  brief,  but  it  is  hoped  adequate,  development 
of  the  hydrodynamics  of  perfect  incompressible  fluids  in  irrotational 
flow  without  viscosity,  with  emphasis  on  those  aspects  of  the  subject 
relevant  to  flows  with  a  free  surface.  Unfortunately,  the  basic  general 
theory  is  unmanageable  for  the  most  part  as  a  basis  for  the  solution 
of  concrete  problems  because  the  nonlinear  free  surface  conditions 
make  for  insurmountable  difficulties  from  the  mathematical  point 
of  view.  It  is  therefore  necessary  to  make  restrictive  assumptions 
which  have  the  effect  of  yielding  more  tractable  mathematical 
formulations.  Fortunately  there  are  at  least  two  possibilities  in  this 
respect  which  are  not  so  restrictive  as  to  limit  too  drastically  the 
physical  interest,  while  at  the  same  time  they  are  such  as  to  lead  to 
mathematical  problems  about  which  a  great  deal  of  knowledge  is 
available. 

One  of  the  two  approximate  theories  results  from  the  assumption 
that  the  wave  amplitudes  are  small,  the  other  from  the  assumption 

*  The  book  by  Rachel  Carson  [C.I 6]  should  be  referred  to  here.  This  book  is 
entirely  nonmathematical,  but  it  is  highly  recommended  for  supplementary 
reading.  Parts  of  it  are  particularly  relevant  to  some  of  the  material  in 
Chapter  6  of  the  present  book. 


INTRODUCTION  XV 

that  it  is  the  depth  of  the  liquid  which  is  small  —  in  both  cases,  of 
course,  the  relevant  quantities  are  supposed  small  in  relation  to  some 
other  significant  length,  such  as  a  wave  length,  for  example.  Both  of 
these  approximate  theories  are  derived  as  the  lowest  order  terms 
of  formal  developments  with  respect  to  an  appropriate  small  dimen- 
sionless  parameter;  by  proceeding  in  this  way,  however,  it  can  be 
seen  how  the  approximations  could  be  carried  out  to  include  higher 
order  terms.  The  remainder  of  the  book  is  largely  devoted  to  the 
working  out  of  consequences  of  these  two  theories,  based  on  concrete 
physical  problems:  Part  II  is  based  on  the  small  amplitude  theory, 
and  Part  III  deals  with  applications  of  the  shallow  water  theory. 
In  addition,  there  is  a  final  chapter  (Chapter  12)  which  makes  up 
Part  IV,  in  which  a  few  problems  are  solved  in  terms  of  the  basic 
general  theory  and  the  nonlinear  boundary  conditions  are  satisfied 
exactly;  this  includes  a  proof  along  lines  due  to  Levi-Civita,  of  the 
existence,  from  the  rigorous  mathematical  point  of  view,  of  progressing 
waves  of  finite  amplitude. 

Part  II,  which  is  concerned  with  the  first  of  the  possibilities, 
might  be  called  the  linearized  exact  theory,  since  it  can  be  obtained 
from  the  basic  exact  theory  simply  by  linearizing  the  free  surface 
conditions  on  the  assumption  that  the  wave  motions  studied  con- 
stitute a  small  deviation  from  a  constant  flow  with  a  horizontal  free 
surface.  Since  we  deal  only  with  irrotational  flows,  the  result  is  a 
theory  based  on  the  determination  of  a  velocity  potential  in  the  space 
variables  (containing  the  time  as  a  parameter,  however)  as  a  solution 
of  the  Laplace  equation  satisfying  certain  linear  boundary  and  initial 
conditions.  This  linear  theory  thus  belongs,  generally  speaking,  to 
potential  theory. 

There  is  such  a  variety  of  material  to  be  treated  in  Part  II,  which 
comprises  Chapters  3  to  9,  that  a  further  division  of  it  into  sub- 
divisions is  useful,  as  follows:  1)  subdivision  A,  dealing  with  wave 
motions  that  arc  simple  harmonic  oscillations  in  the  time;  2)  sub- 
division B,  dealing  with  unsteady,  or  transient,  motions  that  arise 
from  initial  disturbances  starting  from  rest;  and  3)  subdivision  C, 
dealing  with  waves  created  in  various  ways  on  a  running  stream, 
in  contrast  with  subdivisions  A  and  B  in  which  all  motions  are 
assumed  to  be  small  oscillations  near  the  rest  position  of  equilibrium 
of  the  fluid. 

Subdivision  A  is  made  up  of  Chapters  3,  4,  and  5.  In  Chapter  3 
the  basically  important  standing  and  progressing  waves  in  liquids 


XVI  INTRODUCTION 

of  uniform  depth  and  infinite  lateral  extent  are  treated;  the  important 
fact  that  these  waves  are  subject  to  dispersion  comes  to  light,  and 
the  notion  of  group  velocity  thus  arises.  The  problem  of  the  uniqueness 
of  the  solutions  is  considered  —  in  fact,  uniqueness  questions  are 
intentionally  stressed  throughout  Part  II  because  they  are  interesting 
mathematically  and  because  they  have  been  neglected  for  the  most 
part  until  rather  recently.  It  might  seem  strange  that  there  could  be 
any  interesting  unresolved  uniqueness  questions  left  in  potential 
theory  at  this  late  date;  the  reason  for  it  is  that  the  boundary  con- 
dition at  a  free  surface  is  of  the  mixed  type,  i.e.  it  involves  a  linear 
combination  of  the  potential  function  and  its  normal  derivative,  and 
this  combination  is  such  as  to  lead  to  the  occurrence  of  non-trivial 
solutions  of  the  homogeneous  problems  in  cases  which  would  in  the 
more  conventional  problems  of  potential  theory  possess  only  iden- 
tically constant  solutions.  In  fact,  it  is  this  mixed  boundary  con- 
dition at  a  free  surface  which  makes  Part  II  a  highly  interesting 
chapter  in  potential  theory  —  quite  apart  from  the  interest  of  the 
problems  on  the  physical  side.  Chapter  4  goes  on  to  treat  certain 
simple  harmonic  forced  oscillations,  in  contrast  with  the  free  oscil- 
lations treated  in  Chapter  3.  Chapter  5  is  a  long  chapter  which  deals 
with  simple  harmonic  waves  in  cases  in  which  the  depth  of  the  water 
is  not  constant.  A  large  part  of  the  chapter  concerns  the  propagation 
of  progressing  waves  over  a  uniformly  sloping  beach;  various  methods 
of  treating  the  problem  are  explained  —  in  part  with  the  object  of 
illustrating  recently  developed  techniques  useful  for  solving  boundary 
problems  (both  for  harmonic  functions  and  functions  satisfying  the 
reduced  wave  equation)  in  which  mixed  boundary  conditions  occur. 
Another  problem  treated  (in  Chapter  5.5)  is  the  diffraction  of  waves 
around  a  vertical  wedge.  This  leads  to  a  problem  identical  with  the 
classical  diffraction  problem  first  solved  by  Sommerfeld  [S.I 2]  for 
the  special  case  of  a  rigid  half-plane  barrier.  Here  again  the  uniqueness 
question  comes  to  the  fore,  and,  as  in  many  of  the  problems  of  Part  II, 
it  involves  consideration  of  so-called  radiation  conditions  at  infinity.  A 
uniqueness  theorem  is  derived  and  also  a  new,  and  quite  simple  and 
elementary,  solution  for  Sommerfeld's  diffraction  problem  is  given. 
It  is  a  curious  fact  that  these  gravity  wave  problems,  the  solutions 
of  which  are  given  in  terms  of  functions  satisfying  the  Laplace 
equation,  nevertheless  require  for  the  uniqueness  of  the  solutions 
that  conditions  at  infinity  of  the  radiation  type,  just  as  in  the  more 
familiar  problems  based  on  the  linear  wave  equation,  be  imposed; 


I  INTRODUCTION  XVII 

ordinarily  in  potential  theory  it  is  sufficient  to  require  only  boundedness 
conditions  at  infinity  to  ensure  uniqueness. 

In  subdivision  B  of  Part  II,  comprised  of  Chapter  6,  a  variety  of 
problems  involving  transient  motions  is  treated.  Here  initial  con- 
ditions at  the  time  t  —  0  are  imposed.  The  technique  of  the  Fourier 
transform  is  explained  and  used  to  obtain  solutions  in  the  form  of 
integral  representations.  The  important  classical  cases  (treated  first 
by  Cauchy  and  Poisson)  of  the  circular  waves  due  to  disturbances  at 
a  point  of  the  free  surface  in  an  infinite  ocean  are  studied  in  detail. 
For  this  purpose  it  is  very  useful  to  discuss  the  integral  representations 
by  using  an  asymptotic  approximation  due  to  Kelvin  (and,  indeed, 
developed  by  him  for  the  purpose  of  discussing  the  solutions  of  just 
such  surface  wave  problems)  and  called  the  principle,  or  method,  of 
stationary  phase.  These  results  then  can  be  interpreted  in  a  striking 
way  in  terms  of  the  notion  of  group  velocity.  Recently  there  have 
been  important  applications  of  these  results  in  oceanography:  one 
of  them  concerns  the  type  of  waves  called  tsunamis,  which  are 
destructive  waves  in  the  ocean  caused  by  earthquakes,  another 
concerns  the  location  of  storms  at  sea  by  analyzing  wave  records 
on  shore  in  the  light  of  the  theory  at  present  under  discussion.  The 
question  of  uniqueness  of  the  transient  solutions  —  again  a  problem 
solved  only  recently  —  is  treated  in  the  final  section  of  Chapter  6. 
An  opportunity  is  also  afforded  for  a  discussion  of  radiation  con- 
ditions (for  simple  harmonic  waves)  as  limits  as  t  ->  oo  in  appropriate 
problems  concerning  transients,  in  which  boundedness  conditions  at 
infinity  suffice  to  ensure  uniqueness. 

The  final  subdivision  of  Part  II,  subdivision  C,  deals  with  small 
disturbances  created  in  a  stream  flowing  initially  with  uniform 
velocity  and  with  a  horizontal  free  surface.  Chapter  7  treats  waves  in 
streams  having  a  uniform  depth.  Again,  in  the  case  of  steady  motions, 
the  question  of  appropriate  conditions  of  the  radiation  type  arises; 
the  matter  is  made  especially  interesting  here  because  the  circum- 
stances with  respect  to  radiation  conditions  depend  radically  on  the 
parameter  U2/g/*,  with  U  and  h  the  velocity  and  depth  at  infinity,  res- 
pectively. Thus  if  U2/gh  >  1,  no  radiation  conditions  need  be  im- 
posed, if  U2/gh  <  1  they  are  needed,  while  if  U2jgh  =  1  something 
quite  exceptional  occurs.  These  matters  are  studied,  and  their  physical 
interpretations  are  discussed  in  Chapter  7.3  and  7.4.  In  Chapter  8 
Kelvin's  theory  of  ship  waves  for  the  idealized  case  of  a  ship  regarded 
as  a  point  disturbance  moving  over  the  surface  of  the  water  is  treated 


XVIII  INTRODUCTION 

in  considerable  detail.  The  principle  of  stationary  phase  leads  to  a 
beautiful  and  elegant  treatment  of  the  nature  of  ship  waves  that  is 
purely  geometrical  in  character.  The  cases  of  curved  as  well  as 
straight  courses  are  considered,  and  photographs  of  ship  waves  taken 
from  airplanes  are  reproduced  to  indicate  the  good  accord  with 
observations.  Finally,  in  Chapter  9  a  general  theory  (once  more  the 
result  of  quite  recent  investigations)  for  the  motion  of  ships,  regarded 
as  floating  rigid  bodies,  is  presented.  In  this  theory  no  restrictive 
assumptions  —  regarding,  for  example,  the  coupling  (or  lack  of 
coupling,  as  in  an  old  theory  due  to  Krylov  [K.20]  between  the 
motion  of  the  sea  and  the  motion  of  the  ship,  or  between  the  various 
degrees  of  freedom  of  the  ship  —  are  made  other  than  those  needed  to 
linearize  the  problem.  This  means  essentially  that  the  ship  must  be 
regarded  as  a  thin  disk  so  that  it  can  slice  its  way  through  the  water 
(or  glide  over  the  surface,  perhaps)  with  a  finite  velocity  and  still 
create  waves  which  do  not  have  large  amplitudes;  in  addition,  it 
is  necessary  to  suppose  that  the  motion  of  the  ship  is  a  small  oscil- 
lation relative  to  a  motion  of  translation  with  uniform  velocity.  The 
theory  is  obtained  by  making  a  formal  development  of  all  conditions 
of  the  complete  nonlinear  boundary  problem  with  respect  to  a  para- 
meter which  is  a  thickness-length  ratio  of  the  ship.  The  resulting 
theory  contains  the  classical  Michell-IIavelock  theory  for  the  wave 
resistance  of  a  ship  in  terms  of  the  shape  of  its  hull  as  the  simplest 
special  case. 

We  turn  next  to  Part  III,  which  deals  with  applications  of  the 
approximate  theory  which  results  from  the  assumption  that  it  is  the 
depth  of  the  liquid  which  is  small,  rather  than  the  amplitude  of  the 
surface  waves  as  in  Part  II.  The  theory,  called  here  the  shallow 
water  theory,  leads  to  a  system  of  nonlinear  partial  differential 
equations  which  are  analogous  to  the  differential  equations  for  the 
motion  of  compressible  gases  in  certain  cases.  We  proceed  to  outline 
the  contents  of  Part  III,  which  is  composed  of  two  long  chapters. 

In  Chapter  10  the  mathematical  methods  based  on  the  theory  of 
characteristics  are  developed  in  detail  since  they  furnish  the  basis 
for  the  discussion  of  practically  all  problems  in  Part  III;  it  is  hoped 
that  this  preparatory  discussion  of  the  mathematical  tools  will  make 
Part  III  of  the  book  accessible  to  engineers  and  others  who  have  not 
had  advanced  training  in  mathematical  analysis  and  in  the  methods 
of  mathematical  physics.  In  preparing  this  part  of  the  book  the 
author's  task  was  made  relatively  easy  because  of  the  existence  of  the 


INTRODUCTION  XIX 

book  by  Courant  and  Friedrichs  [C.9],  which  deals  with  gas  dynamics; 
the  presentation  of  the  basic  theory  given  here  is  largely  modeled 
on  the  presentation  given  in  that  book.  The  concrete  problems  dealt 
with  in  Chapter  10  are  quite  varied  in  character,  including  the 
propagation  of  disturbances  into  still  water,  conditions  for  the 
occurrence  of  a  bore  and  a  hydraulic  jump  (phenomena  analogous  to 
the  occurrence  of  shock  waves  in  gas  dynamics),  the  motion  resulting 
from  the  breaking  of  a  dam,  steady  two  dimensional  motions  at 
supercritical  velocity,  and  the  breaking  of  waves  in  shallow  water. 
The  famous  problem  of  the  solitary  wave  is  discussed  along  the  lines 
used  recently  by  Friedrichs  and  Hyers  [F.13]  to  prove  rigorously 
the  existence  of  the  solitary  wave  from  the  mathematical  point  of 
view;  this  problem  requires  carrying  the  perturbation  series  which 
formulate  the  shallow  water  theory  to  terms  of  higher  order.  The 
problem  of  the  motion  of  frontal  discontinuities  in  the  atmosphere, 
which  lead  to  the  development  of  cyclonic  disturbances  in  middle 
latitudes,  is  given  a  formulation  —  on  the  basis  of  hypotheses  which 
simplify  the  physical  situation  —  which  brings  it  within  the  scope 
of  a  more  general  "shallow  water  theory".  Admittedly  (as  has  already 
been  noted  earlier)  this  theory  is  somewhat  speculative,  but  it  is 
nevertheless  believed  to  have  potentialities  for  clarifying  some  of 
the  mysteries  concerning  the  dynamical  causes  for  the  development 
and  deepening  of  frontal  disturbances  in  the  atmosphere,  especially 
if  modern  high  speed  digital  computing  machines  are  used  as  an  aid 
in  solving  concrete  problems  numerically. 

Chapter  10  concludes  with  the  discussion  of  a  few  applications  of 
the  linearized  version  of  the  shallow  water  theory.  Such  a  linearization 
results  from  assuming  that  the  amplitude  of  the  waves  is  small.  The 
most  famous  application  of  this  theory  is  to  the  tides  in  the  oceans 
(and  also  in  the  atmosphere,  for  that  matter);  strange  though  it 
seems  at  first  sight,  the  oceans  can  be  treated  as  shallow  for  this 
phenomenon  since  the  wave  lengths  of  the  motions  are  very  long 
because  of  the  large  periods  of  the  disturbances  caused  by  the  moon 
and  the  sun.  This  theory,  as  applied  to  the  tides,  is  dealt  with  only 
very  summarily,  since  an  extended  treatment  is  given  by  Lamb 
[L.3].  Instead,  some  problems  connected  with  the  design  of  floating 
breakwaters  in  shallow  water  are  discussed,  together  with  brief 
treatments  of  the  oscillations  in  certain  lakes  (the  lake  at  Geneva 
in  Switzerland,  for  example)  called  seiches,  and  oscillations  in  harbors. 

Finally,  Part  III  concludes  with  Chapter  11  on  the  subject  of 


XX  INTRODUCTION 

mathematical  hydraulics,  which  is  to  be  understood  here  as  referring 
to  flows  and  wave  motions  in  rivers  and  other  open  channels  with 
rough  sides.  The  problems  of  this  chapter  are  not  essentially  different, 
as  far  as  mathematical  formulations  go,  from  the  problems  treated 
in  the  preceding  Chapter  10.  They  differ,  however,  on  the  physical 
side  because  of  the  inclusion  of  a  force  which  is  just  as  important  as 
gravity,  namely  a  force  of  resistance  caused  by  the  rough  sides  and 
bottom  of  the  channels.  This  force  is  dealt  with  empirically  by 
adding  a  term  to  the  equation  expressing  the  law  of  conservation  of 
momentum  that  is  proportional  to  the  square  of  the  velocity  and 
with  a  coefficient  depending  on  the  roughness  and  the  so-called 
hydraulic  radius  of  the  channel.  The  differential  equations  remain  of 
the  same  type  as  those  dealt  with  in  Chapter  10,  and  the  same  under- 
lying theory  based  on  the  notion  of  the  characteristics  applies. 

Steady  motions  in  inclined  channels  are  first  dealt  with.  In  par- 
ticular, a  method  of  solving  the  problem  of  the  occurrence  of  roll 
waves  in  steep  channels  is  given;  this  is  done  by  constructing  a 
progressing  wave  by  piecing  together  continuous  solutions  through 
bores  spaced  at  periodic  intervals,  This  is  followed  by  the  solution 
of  a  problem  of  steady  motion  which  is  typical  for  the  propagation 
of  a  flood  down  a  long  river;  in  fact,  data  were  chosen  in  such  a  way 
as  to  approximate  the  case  of  a  flood  in  the  Ohio  River.  A  treatment 
is  next  given  for  a  flood  problem  so  formulated  as  to  correspond 
approximately  to  the  case  of  a  flood  wave  moving  down  the  Ohio 
to  its  junction  with  the  Mississippi,  and  w'th  the  result  that  distur- 
bances are  propagated  both  upstream  and  downstream  in  the  Missis- 
sippi and  a  backwater  effect  is  noticeable  up  the  Ohio.  In  these 
problems  it  is  necessary  to  solve  the  differential  equations  numerically 
(in  contrast  with  most  of  the  problems  treated  in  Chapter  10,  in 
which  interesting  explicit  solutions  could  be  given),  and  methods  of 
doing  so  are  explained  in  detail.  In  fact,  a  part  of  the  elements  of 
numerical  analysis  as  applied  to  solving  hyperbolic  partial  differential 
equations  by  the  method  of  finite  differences  is  developed.  The  results 
of  a  numerical  prediction  of  a  flood  over  a  stretch  of  400  miles  in 
the  Ohio  River  as  it  actually  exists  are  given.  The  flood  in  question 
was  the  1945  flood  —  one  of  the  largest  on  record  —  and  the  predic- 
tions made  (starting  with  the  initial  state  of  the  river  and  using  the 
known  flows  into  it  from  tributaries  and  local  drainage)  by  numerical 
integration  on  a  high  speed  digital  computer  (the  Univac)  check 
quite  closely  with  the  actually  observed  flood.  Numerical  predictions 


INTRODUCTION  XXI 

were  also  made  for  the  case  of  a  flood  (the  1947  flood  in  this  case) 
coming  down  the  Ohio  and  passing  through  its  junction  with  the 
Mississippi;  the  accuracy  of  the  prediction  was  good.  This  is  a  case 
in  which  the  simplified  methods  of  the  civil  engineers  do  not  work 
well.  These  results,  of  course,  have  important  implications  for  the 
practical  applications. 

Finally  Part  IV,  made  up  of  Chapter  12,  closes  the  book- with  a 
few  solutions  based  on  the  exact  nonlinear  theory.  One  class  of  problems 
is  solved  by  assuming  a  solution  in  the  form  of  power  series  in  the 
time,  which  implies  that  initial  motions  and  motions  for  a  short  time 
only  can  be  determined  in  general.  Nevertheless,  some  interesting 
cases  can  be  dealt  with,  even  rather  easily,  by  using  the  so-called 
Lagrange  representation,  rather  than  the  Euler  representation  which 
is  used  otherwise  throughout  the  book.  The  problem  of  the  breaking 
of  a  dam,  and,  more  generally,  problems  of  the  collapse  of  columns 
of  a  liquid  resting  on  a  rigid  horizontal  plane  can  be  treated  in  this 
way.  The  book  ends  with  an  exposition  of  the  theory  due  to  Levi- 
Civita  concerning  the  problem  of  the  existence  of  progressing  waves  of 
finite  amplitude  in  water  of  infinite  depth  which  satisfy  exactly  the 
nonlinear  free  surface  conditions. 


Acknowledgments 


Without  the  support  of  the  Mathematics  Branch  and  the  Mechanics 
Branch  of  the  Office  of  Naval  Research  this  book  would  not  have  been 
written.  The  author  takes  pleasure  in  acknowledging  the  help  and 
encouragement  given  to  him  by  the  ONR  in  general,  and  by  Dr.  Joa- 
chim Wcyl,  Dr.  Arthur  Grad,  and  Dr.  Philip  Eiscnberg  in  particular. 
Although  she  is  no  longer  working  in  the  ONR,  it  is  neverthe- 
less appropriate  at  this  place  to  express  special  thanks  to  Dean 
Mina  llecs,  who  was  head  of  the  Mathematics  Branch  when  this 
book  was  begun. 

Among  those  who  collaborated  with  the  author  in  the  preparation 
of  the  manuscript,  Dr.  Andreas  Troesch  should  be  singled  out  for 
special  thanks.  His  careful  and  critical  reading  of  the  manuscript  re- 
sulted in  many  improvements  and  the  uncovering  and  correction  of 
errors  and  obscurities  of  all  kinds.  Another  colleague,  Professor  E. 
Isaacson,  gave  almost  as  freely  of  his  time  and  attention,  and  also 
aided  materially  in  revising  some  of  the  more  intricate  portions  of  the 
book.  To  these  fellow  workers  the  author  feels  deeply  indebted. 

Miss  Helen  Samoraj  typed  the  entire  manuscript  in  a  most  efficient 
(and  also  good-humored)  way,  and  uncovered  many  slips  and  in- 
consistencies in  the  process. 

The  drawings  for  the  book  were  made  by  Mrs.  Beulah  Marx  and 
Miss  Lark  in  Joyner.  The  index  was  prepared  by  Dr.  George  Booth  and 
Dr.  Walter  Littman  with  the  assistance  of  Mrs.  Halina  Montvila. 

A  considerable  part  of  the  material  in  the  present  book  is  the  result 
of  researches  carried  out  at  the  Institute  of  Mathematical  Sciences  of 
New  York  University  as  part  of  its  work  under  contracts  with  the 
Office  of  Naval  Research  of  the  U.S.  Department  of  Defense,  and  to  a 
lesser  extent  under  a  contract  with  the  Ohio  River  Division  of  the 
Corps  of  Engineers  of  the  U.S.  Army.  The  author  wishes  to  express  his 
thanks  generally  to  the  Institute;  the  cooperative  and  friendly  spirit 
of  its  members,  and  the  stimulating  atmosphere  it  has  provided  have 
resulted  in  the  carrying  out  of  quite  a  large  number  of  researches  in 
the  field  of  water  waves.  A  good  deal  of  these  researches  and  new 
results  have  come  about  through  the  efforts  of  Professors  K.  O.  Fried- 

xxiii 


XXIV  ACKNOWLEDGMENTS 

richs,  Fritz  John,  J.  B.  Keller,  H.  Lewy  (of  the  University  of  Cali- 
fornia), and  A.  S.  Peters,  together  with  their  students  or  with  visitors 
at  the  Institute. 

.1.  J.  STOKER 
New  York,  N.Y. 
January,  1957. 


Contents 


PART  I 

CHAPTKR  PACK 

Introduction ix 

Acknowledgments xxiii 

1.  Basic  Hydrodynamics 3 

1.1  The  laws  of  conservation  of  momentum  and  mass 3 

1.2  Helmholty/s  theorem 7 

1.3  Potential  flow  and  Bernoulli's  law 9 

1.4  Boundary  conditions 10 

1.5  Singularities  of  the  velocity  potential 12 

1.0  Notions  concerning  energy  and  energy  flux 13 

1.7     Formulation  of  a  surface  wave  problem 15 

2.  The  Two  Basic  Approximate  Theories 19 

2.1  Theory  of  waves  of  small  amplitude 19 

2.2  Shallow  water  theory  to  lowest  order.  Tidal  theory     ....  22 

2.3  Gas  dynamics  analogy 25 

2.4  Systematic  derivation  of  the  shallow  water  theory 27 


PART  II 

Subdivision  A 
Waves  Simple  Harmonic  in  the  Time 

3.  Simple  Harmonic  Oscillations  in  Water  of  Constant  Depth    .    .        37 

3.1  Standing  waves 37 

3.2  Simple  harmonic  progressing  waves 45 

3.3  Energy  transmission  for  simple  harmonic  waves  of  small  ampli- 
tude    47 

3.4  Group  velocity.  Dispersion 51 

4.  Waves  Maintained  by  Simple  Harmonic  Surface  Pressure  in 

Water  of  Uniform  Depth.  Forced  Oscillations 55 

4.1  Introduction 55 

4.2  The  surface  pressure  is  periodic  for  all  values  of  a? 57 


XXVI  CONTENTS 

CHAPTER  PAGE 

4.8    The  variable  surface  pressure  is  confined  to  a  segment  of  the 

surface 58 

4.4    Periodic  progressing  waves  against  a  vertical  cliff 07 

5.  Waves  on  Sloping  Beaches  and  Past  Obstacles 69 

5.1  Introduction  and  summary 69 

5.2  Two-dimensional  waves  over  beaches  sloping  at  angles  co  =  n/2n  77 

5.3  Three-dimensional  waves  against  a  vertical  cliff 84 

5.4  Waves  on  sloping  beaches.  General  case 95 

5.5  Diffraction  of  waves  around  a  vertical  wedge.  Sommerfeld's 
diffraction  problem 109 

5.6  Brief  discussions  of  additional  applications  and  of  other  methods 

of  solution 133 


Subdivision  B 
Motions  Starting  from  Rest.  Transients 

6.  Unsteady  Motions 149 

6.1  General  formulation  of  the  problem  of  unsteady  motions  .    .  149 

6.2  Uniqueness  of  the  unsteady  motions  in  bounded  domains      .  150 

6.3  Outline  of  the  Fourier  transform  technique 153 

6.4  Motions  due  to  disturbances  originating  at  the  surface  ...  156 

6.5  Application  of  Kelvin's  method  of  stationary  phase     ....  163 

6.6  Discussion  of  the  motion  of  the  free  surface  due  to  disturbances 
initiated  when  the  water  is  at  rest 167 

6.7  Waves  due  to  a  periodic  impulse  applied  to  the  water  when 
initially  at  rest.  Derivation  of  the  radiation  condition  for  purely 
periodic  waves 174 

6.8  Justification  of  the  method  of  stationary  phase 181 

6.9  A  time-dependent  Green's  function.  Uniqueness  of  unsteady 
motions  in  unbounded  domains  when  obstacles  are  present  .  187 


Subdivision  C 
Waves  on  a  Running  Stream.  Ship  Waves 

7.  Two-dimensional  Waves  on  a  Running  Stream  in  Water  of 

Uniform  Depth 198 

7.1     Steady  motions  in  water  of  infinite  depth  with  p  =  0  on  the 

free  surface 199 


CONTENTS  XXVII 

CHAPTER  PAGE 

7.2  Steady  motions  in  water  of  infinite  depth  with  a  disturbing  pres- 
sure on  the  free  surface 201 

7.3  Steady  waves  in  water  of  constant  finite  depth 207 

7.4  Unsteady  waves  created  by  a  disturbance  on  the  surface  of  a 
running  stream 210 

8.  Waves  Caused  by  a  Moving  Pressure  Point.  Kelvin's  Theory  of 

the  Wave  Pattern  created  by  a  Moving  Ship 219 

8.1  An  idealized  version  of  the  ship  wave  problem.  Treatment  by 

the  method  of  stationary  phase 219 

8.2  The  classical  ship  wave  problem.  Details  of  the  solution  .    .  224 

9.  The  Motion  of  a  Ship,  as  a  Floating  Rigid  Body,  in  a  Seaway  245 

9.1  Introduction  and  summary 245 

9.2  General  formulation  of  the  problem 264 

9.3  Linearization  by  a  formal  perturbation  procedure 269 

9.4  Method  of  solution  of  the  problem  of  pitching  and  heaving  of  a 
ship  in  a  seaway  having  normal  incidence 278 


PART  III 

10.  Long  Waves  in  Shallow  Water 291 

10.1  Introductory  remarks  and  recapitulation  of  the  basic  equations  291 

10.2  Integration  of  the  differential  equations  by  the  method  of  char- 
acteristics       293 

10.3  The  notion  of  a  simple  wave 300 

10.4  Propagation  of  disturbances  into  still  water  of  constant  depth  305 

10.5  Propagation  of  depression  waves  into  still  water  of  constant 
depth 308 

10.6  Discontinuity,  or  shock,  conditions 314 

10.7  Constant  shocks:  bore,  hydraulic  jump,  reflection  from  a  rigid 
wall 326 

10.8  The  breaking  of  a  dam 333 

10.9  The  solitary  wave 342 

10.10  The  breaking  of  waves  in  shallow  water.  Development  of  bores  351 

10.11  Gravity  waves  in  the  atmosphere.  Simplified  version  of  the 
problem  of  the  motion  of  cold  and  warm  fronts 374 

10.12  Supercritical  steady  flows  in  two  dimensions.  Flow  around 
bends.  Aerodynamic  applications 405 

10.13  Linear  shallow  water  theory.  Tides.   Seiches.  Oscillations  in 
harbors.  Floating  breakwaters 414 


XXVIII  CONTENTS 

CHAPTER  PAGE 

11.  Mathematical  Hydraulics 451 

11.1  Differential  equations  of  flow  in  open  channels 452 

11.2  Steady  flows.  A  junction  problem 456 

11.3  Progressing  waves  of  fixed  shape.  Roll  waves 461 

11.4  Unsteady  flows  in  open  channels.  The  method  of  characteristics  409 

11.5  Numerical  methods  for  calculating  solutions  of  the  differential 
equations  for  flow  in  open  channels 474 

11.6  Flood  prediction  in  rivers.  Floods  in  models  of  the  Ohio  River 

and  its  junction  with  the  Mississippi  River 482 

1 1 .7  Numerical  prediction  of  an  actual  flood  in  the  Ohio,  and  at  its 
junction  with  the  Mississippi.  Comparison  of  the  predicted  with 

the  observed  floods 408 

Appendix  to  Chapter  11.    Expansion  in  the  neighborhood  of  the  first 

characteristic 505 


PART  IV 

12.  Problems  in  which  Free  Surface  Conditions  are  Satisfied  Exactly. 

The  Breaking  of  a  Dam.  Levi-Civita's  Theory 513 

12.1  Motion  of  water  due  to  breaking  of  a  dam,  and  related  problems  513 

12.2  The  existence  of  periodic  waves  of  finite  amplitude     ....  522 

12.2a  Formulation  of  the  problem 522 

12.2b  Outline  of  the  procedure  to  be  followed  in  proving  the  existence 

of  the  function  a)(%) 526 

12.2c  The  solution  of  a  class  of  linear  problems 529 

12.2d  The  solution  of  the  nonlinear  boundary  value  problem   .    .    .  537 

Bibliography 545 

Author  Index 561 

Subject  Index 563 


PART  I 


CHAPTER  1 


Basic  Hydrodynamics 

1.1.  The  laws  of  conservation  of  momentum  and  mass 

As  has  been  stated  in  the  introduction,  we  deal  exclusively  in  this 
book  with  flows  in  water  (and  air)  which  are  of  such  a  nature  as 
to  make  it  unnecessary  to  take  into  account  the  effects  of  viscosity 
and  compressibility.  As  a  consequence  of  the  neglect  of  internal 
friction,  or  in  other  words  of  neglect  of  shear  stresses,  it  is  well 
known  that  the  stress  system*  in  the  liquid  is  a  state  of  uniform 
compression  at  each  point.  The  intensity  of  the  compressive  stress 
is  called  the  pressure  p. 

The  equation  of  motion  of  a  fluid  particle  can  then  be  obtained  on 
the  basis  of  Newton's  law  of  conservation  of  momentum,  as  follows. 
A  small  rectangular  element  of  the  fluid  is  shown  in  Figure  1.1.1 


Fig.  1.1.1.  Pressure  on  a  fluid  element 


with  the  pressure  acting  on  the  faces  normal  to  the  o?-axis.  Newton's 
law  for  the  ^-direction  is  then 


[  —  (P 


—  Qa(*) 


*  We  assume  that  the  usual  concepts  of  the  general  mechanics  of  continuous 
media  are  known. 


WATER   WAVES 


in  which  X  is  the  external  or  body  force  component  per  unit  mass 
and  a(x)  is  the  acceleration  component,  both  in  the  ^-direction,  and 
Q  is  the  density.  The  quantities  p9  X,  and  a(x)  are  in  general  functions 
of  x9  y,  z,  and  t.  Here,  as  always,  we  shall  use  letter  subscripts  to 
denote  differentiation,  and  this  accounts  for  the  symbol  a(x}  to  denote 
the  component  of  a  vector  in  the  ^-direction.  Upon  passing  to  the 
limit  in  allowing  dx,  dy,  dz  to  approach  zero  we  obtain  the  equation 
of  motion  for  the  ^-direction  in  the  form  —  px  +  qX  =  Qa(x}9  and 
analogous  expressions  for  the  two  other  directions.  Thus  we  have  the 
equations  of  motion 


(1.1.1) 

or,  in  vector  form: 
(1.1.2) 


Px  +  %  = 

-Pv  +Y  = 


—  -P* 


—  grad  p  +  F  =  a, 
Q 


with  an  obvious  notation.  The  body  force  F  plays  a  very  important 
role  in  our  particular  branch  of  hydrodynamics  — in  fact  the  main 
results  of  the  theory  are  entirely  conditioned  by  the  presence  of  the 
gravitational  force  F  =  (0,  —  g,  0),  in  which  g  represents  the  acceler- 
ation of  gravity.  It  should  be  observed  that  we  consider  the  positive 
y-axis  to  be  vertically  upward,  and  the  x,  z-plane  therefore  to  be  horizontal 
(usually  it  will  be  taken  as  the  undisturbed  water  surface).  This  con- 
vention regarding  the  disposition  of  the  coordinate  axes  will  be  main- 
tained, for  the  most  part,  throughout  the  book. 

The  differential  equations  (1.1.1)  are  in  what  is  called  the  Lagrang- 
ian  form,  in  which  one  has  in  mind  a  direct  description  of  the  motion 
of  each  individual  fluid  particle  as  a  function  of  the  time.  It  is  more 
useful  for  most  purposes  to  work  with  the  equations  of  motion  in  the 
so-called  Eulerian  form.  In  this  form  of  the  equations  one  concen- 
trates attention  on  the  determination  of  the  velocity  distribution  in 
the  region  occupied  by  the  fluid  without  trying  to  follow  the  motion  of 
the  individual  fluid  particles,  but  rather  observing  the  velocity 
distribution  at  fixed  points  in  space  as  a  function  of  the  time.  In 


BASIC   HYDRODYNAMICS  5 

other  words,  the  velocity  field,  with  components  u9  v9  w9  is  to  be 
determined  as  a  function  of  the  space  variables  and  the  time.  After- 
wards, if  that  is  desired,  the  motion  of  the  individual  particles  can 
be  obtained  by  integrating  the  system  of  ordinary  differential  equa- 
tions x  =  u,  y  —  v,  z  —  w9  in  which  the  dot  over  the  quantities  x9  y, 
z  means  differentiation  with  respect  to  the  time  in  following  the 
motion  of  an  individual  particle.- 

In  order  to  restate  the  equations  of  motion  (1.1.1)  in  terms  of  the 
Euler  variables  u,  v,  w9  and  in  order  to  carry  out  other  important 
operations  as  well,  it  is  necessary  to  calculate  time  derivatives  of 
various  functions  associated  with  a  given  fluid  particle  in  following 
the  motion  of  the  particle.  For  example,  we  need  to  calculate  the 
time  derivative  of  the  velocity  of  a  particle  in  order  to  obtain  the 
acceleration  components  occurring  in  (1.1.1),  and  quite  a  few  other 
quantities  will  occur  later  on  for  which  such  particle  derivatives  will 
be  needed.  Suppose,  then,  that  F(x9  y,  z;  t)  is  a  function  associated 
with  a  particle  which  follows  the  path  given  by  the  vector 

x  =  (x(t),y(t)9z(t)); 
it  follows  that 

x  =  (x(t)9  y(t)9  z(t))  =  (u,  v9  w) 

is  the  velocity  vector  associated  with  the  particle.  For  this  particle 
the  arguments  x9  y9  z  of  the  function  F  are  of  course  the  functions 
of  t  which  characterize  the  motion  of  the  particle;  as  a  consequence 
we  have 

^  =  Fxx  +  Fyy  +  Fzz  +  Ft 


and  hence  the  operation  of  taking  the  particle  derivative  d/dt  is 
defined  as  follows: 

(1.1.3)          1(     )  =  u(     )m+v(     )y+w(     ).  +  (     ),. 
at 

The  distinction  between  dF/dt  and  dF/dt  =  Ft  should  be  carefully 
noted. 

Since  the  acceleration  a  of  a  particle  is  given  by  a  =  (du/dt9  dv/dt9 
dw/dt)9  in  which  (u9  v9  w)  are  the  components  of  the  velocity  v  of 


6  WATER   WAVES 

the  particle,  it  follows  from  (1.1.3)  that  the  component  a(x)  —  du/dt 
is  given  by 

du 

—  =  uux  +  vuy  +  wuz  +  ut, 

at 

with  similar  expressions  for  the  other  components.  The  equations  of 
motion  (1.1.1)  are  therefore  given  as  follows  in  terms  of  the  Euler 
variables: 


(1.1.4) 


1 

Ut  +  UUX  +  VUy  +  WUZ  =   —   --  px , 

6 
1 

Vt    +  ^^    +  TOV    +    WO*    =    —    -  Py   —  g, 

e 
i 

I0f  +  U^  +  U^y  +  WWZ  =  —  -  pz 

Q 


when  we  specify  the  external  or  body  force  to  consist  only  of  the 
force  of  gravity. 

Equations  (1.1.4)  form  a  set  of  three  nonlinear  partial  differential 
equations  for  the  five  quantities  u9  v,  w9  g,  and  p.  Since  the  fluid  is 
assumed  to  be  incompressible,  the  density  Q  can  be  taken  as  a  known 
constant.  At  the  same  time,  the  assumption  of  incompressibility  leads 
to  a  relatively  simple  differential  equation  expressing  the  law  of 
conservation  of  mass,  and  this  equation  constitutes  the  needed  fourth 
equation  for  the  determination  of  the  velocity  components  and  the 
pressure.  Perhaps  the  simplest  way  to  derive  the  mass  conservation 
law  is  to  start  from  the  relation 


which  states  that  the  mass  flux  outward  through  any  fixed  closed 
surface  enclosing  a  region  in  which  no  liquid  is  created  or  destroyed 
is  zero.  (By  vn  we  mean  the  velocity  component  taken  positive  in  the 
direction  of  the  outward  normal  to  the  surface.)  An  application  of 
Gauss's  divergence  theorem: 

(1.1.5)  JJei>n  dS  =  JJJdiv  (ev)  dr 

S  R 

to  the  above  integral  leads  to  the  relation 

fdiv  (QV)  dr  =  0 


R 


BASIC   HYDRODYNAMICS  7 

for  any  arbitrary  region  R.  It  follows  therefore  that  div  (gv)  =  0 
everywhere,  and  since  Q  =  constant,  we  have  finally 

(1.1.6)  divv  =  ux  +vv  +  wz  =  0 

as  the  expression  of  the  law  of  conservation  of  mass.  The  equation 
(1.1.6)  is  also  frequently  called  the  equation  of  continuity. 

Equations  (1.1.4)  and  (1.1.6)  are  sufficient,  once  appropriate 
initial  and  boundary  conditions  (to  be  discussed  shortly)  are  imposed, 
to  determine  the  velocity  components  u,  v,  w,  and  the  pressure  p 
uniquely. 


1.2.  Helmholtz's  theorem 

Before  discussing  boundary  conditions  it  is  preferable  to  for- 
mulate a  few  additional  conservation  laws  which  are  consequences 
of  the  assumptions  made  so  far—  in  particular  of  the  assumption  that 
internal  fluid  friction  can  be  neglected. 

The  first  of  these  laws  to  be  discussed  is  the  law  of  conservation 
of  circulation.  The  notion  of  circulation  is  defined  as  follows.  Consider 
a  closed  curve  C  which  moves  with  the  fluid  (that  is,  C  consists 
always  of  the  same  particles  of  the  fluid).  The  circulation  F  ==  F(t) 
around  C  is  defined  by  the  line  integral 


(1.2.1)  r(t)  =  j>udx+vdy+wdz 


=  <p  vs  ds 
c 

in  which  vs  is  the  velocity  component  of  the  fluid  tangent  to  C, 
and  ds  is  the  element  of  arc  length  of  C.  The  curve  C  is  considered 
as  given  by  the  vector  x(a,  t)  with  a  a  parameter  on  C  such  that 
0  ^  a  ^  1  and  x(0,  t)  =  x(l,  /).  We  are  thus  operating  in  terms  of 
the  Lagrange  system  of  variables  rather  than  in  terms  of  the  Euler 
system,  and  fixing  a  value  of  a  has  the  effect  of  picking  out  a  specific 

particle  on  C. 

i 

We  may  write  F(t)  =  \  v  •  xada  in  which  v  •  xa  is  a  scalar  product 

o 
and  xa,  as  usual,  refers  to  differentiation  with  respect  to  a.  For  the 

time  derivative  F  we  have  therefore 


8  WATER   WAVES 

1 

f(t)  =  J(v  •  xa  +  v  - 

0 

From  the  equation  of  motion  (1.1.2)  in  the  Lagrangian  form  with 
a  ==  v,  F  =  (0,  -g,  0)  =  —  grad  (gy),  and  from  xa  =  va,  the  last 
equation  yields 

i 

(1.2.2)  f(t)  =   n Xa  '  grad  p  -  gxa  •  grad  y  +  v  •  v J  da 

o 
i 

=  J     -  ~  Po  -  &Va  +  -  (v  •  v)a  Ida 

0 

=  0, 

since  the  values  of  p,  j/,  and  v  coincide  at  a  =  0  and  or  =  1,  and  p 
and  g  are  constants.  The  last  equation  evidently  states  that  in  a 
nonviscous  fluid  the  circulation  around  any  closed  curve  consisting  of 
the  same  fluid  particles  is  constant  in  time.  This  is  the  theorem  of  Helm- 
holtz.  The  assumption  of  zero  viscosity  entered  into  our  derivation 
through  the  use  of  (1.1.2)  as  equation  of  motion.* 

In  this  book  we  are  interested  in  the  special  case  in  which  the 
circulation  for  all  closed  curves  is  zero.  This  case  is  very  important 
in  the  applications  because  it  occurs  whenever  the  fluid  is  assumed 
to  have  been  at  rest  or  to  have  been  moving  with  a  constant  velocity 
at  some  particular  time,  so  that  v  =  const,  holds  at  that  time,  and 
hence  F  vanishes  for  all  time.  The  cases  in  which  the  fluid  motion 
begins  from  such  states  are  obviously  very  important. 

The  assumption  that  F  vanishes  for  all  closed  curves  has  a  number 
of  consequences  which  are  basic  for  all  that  follows  in  this  book. 
The  first  conclusion  from  JT  =  0  follows  almost  immediately  from 
Stokes's  theorem: 

(1.2.3)  r=j>vsds  =  JJ(curl  v)n  dA, 

c  s 

in  which  the  surface  integral  is  taken  over  any  surface  S  spanning  the 
curve  C.  If  71  =  0  for  all  curves  C,  as  we  assume,  it  follows  easily  by  a 
well-known  argument  that  the  vector  curl  v  vanishes  everywhere: 

*  It  should  be  added  that  the  law  of  conservation  of  circulation  holds  under 
much  wider  conditions  than  were  assumed  here  (cf.  [C.9],  p.  19). 


BASIC    HYDRODYNAMICS  9 

(1.2.4)  curl  v  =  (wy—vz,  uz—wx,  vx—uy)  =  0, 

and  the  flow  is  then  said  to  be  irrotational.  In  other  words,  a  motion 
in  a  nonviscous  fluid  which  is  irrotational  at  one  instant  always 
remains  irrotational.  Throughout  the  rest  of  this  book  we  shall  assume 
all  flows  to  be  irrotational. 

1.3.  Potential  flow  and  Bernoulli's  law 

The  assumption  of  irrotational  flow  results  in  a  number  of  sim- 
plifications in  our  theory  which  are  of  the  greatest  utility.  In  the 
first  place,  the  fact  that  curl  v  =  0  (cf.  (1.2.4))  ensures,  as  is  well 
known,  the  existence  of  a  single-valued  velocity  potential  0(x,  y,  2;  t) 
in  any  simply  connected  region,  from  which  the  velocity  field  can  be 
derived  by  taking  the  gradient: 

(1.3.1)  v  =  grad  0  =  (0X,  0y,  0Z), 
or,  in  terms  of  the  components  of  v: 

(1.3.2)  u  =  0X9  v  =  0y9  w  =  0Z. 

The  velocity  potential  is,  indeed,  given  by  the  line  integral 

<e,v,z 

0(x9  y,  2;  t)  =  I     u  dx  +  v  dy  +  w  dz. 

The  vanishing  of  curl  v  ensures  that  the  expression  to  be  integrated 
is  an  exact  differential.  Once  it  is  known  that  the  velocity  com- 
ponents are  determined  by  (1.3.2),  it  follows  from  the  continuity 
equation  (1.1.6),  i.e.  div  v  =  0,  that  the  velocity  potential  0  is  a 
solution  of  the  Laplace  equation 

(1.3.3)  V2<Z>  =  0XX  +  0VV  +  0ZZ  =  0, 

as  one  readily  sees,  and  0  is  thus  a  harmonic  function.  This  fact 
represents  a  great  simplification,  since  the  velocity  field  is  derivable 
from  a  single  function  satisfying  a  linear  differential  equation  which 
has  been  very  much  studied  and  about  which  a  great  deal  is  known. 
Still  another  important  consequence  of  the  irrotational  character 
of  a  flow  can  be  obtained  from  the  equations  of  motion  (1.1.4).  By 
making  use  of  (1.2.4),  it  is  readily  verified  that  the  equations  of 
motion  (1.1.4)  can  be  written  in  the  following  vector  form: 

1  p 

grad  0t  +  ~  grad  (u2  +  v2  +  w2)  =  —  grad  -    —  grad  (gy), 

2  Q 

use  having  been  made  of  the  fact  that  Q  =  constant.  Integration 


10  WATER  WAVES 

of  this  relation  leads  to  the  important  equation  expressing  what  is 
called  Bernoulli9  s  law: 

(1.3.4)  0t  +  I  (u*  +  v*  +  te;2)  +  ?  +  gy  =  C(t), 

2  Q 

in  which  C(t)  may  depend  on  t,  but  not  on  the  space  variables.  There 
are  two  other  forms  of  Bernoulli's  law  for  the  case  of  steady  flows, 
one  of  which  applies  along  stream  lines  even  though  the  flow  is 
not  irrotational,  but  since  we  make  no  use  of  these  laws  in  this  book 
we  refrain  from  formulating  them. 

The  potential  equation  (1.3.3)  together  with  Bernoulli's  law 
(1.8.4)  can  be  used  to  take  the  place  of  the  equations  of  motion 
(1.1.4)  and  the  continuity  equation  (1.1.6)  as  a  means  of  determining 
the  velocity  components  u,  u,  10,  and  the  pressure  p:  in  effect,  u,  v9 
and  w  are  determined  from  the  solution  0  of  (1.3.3),  after  which  the 
pressure  p  can  be  obtained  from  (1.3.4).  It  is  true  that  the  pressure 
appears  to  be  determined  only  within  a  function  which  is  the  same 
at  each  instant  throughout  the  fluid.  On  physical  grounds  it  is, 
however,  clear  that  a  function  of  t  alone  added  to  the  pressure  p 
has  no  effect  on  the  motion  of  the  fluid  since  no  pressure  gradients 
result  from  such  an  addition  to  the  pressure.  In  fact,  if  we  set 


=  0*  -f  fC(f  )  df  ,     then     0*     is     a     harmonic     function 


with 


grad  0  =  grad  0*  and  the  Bernoulli  law  with  reference  to  it  has  a 
vanishing  right  hand  side.  Thus  we  may  take  C(t)  =  0  in  (1.3.4) 
without  any  essential  loss  of  generality. 

While  it  is  true  that  the  Laplace  equation  is  a  linear  differential 
equation,  it  does  not  follow  that  we  shall  be  able  to  escape  all  of  the 
difficulties  arising  from  the  nonlinear  character  of  the  basic  differen- 
tial equations  of  motion  (1.1.4).  As  we  shall  sec,  the  problems  of 
interest  here  remain  essentially  nonlinear  because  the  Bernoulli  law 
(1.3.4),  and  another  condition  to  be  derived  in  the  next  section,  give 
rise  to  nonlinear  boundary  conditions  at  free  surfaces.  In  the  next 
section  we  take  up  the  important  question  of  the  boundary  con- 
ditions appropriate  to  various  physical  situations. 

1.4.  Boundary  conditions 

•  We  assume  the  fluid  under  consideration  to  have  a  boundary 
surface  S,  fixed  or  moving,  which  separates  it  from  some  other 
medium,  and  which  has  the  property  that  any  particle  which  is  once 


BASIC    HYDRODYNAMICS  11 

on  the  surface  remains  on  it.*  Examples  of  such  boundary  surfaces 
of  importance  for  us  are  those  in  which  S  is  the  surface  of  a  fixed 
rigid  body  in  contact  with  the  fluid— the  bottom  of  the  sea,  for 
example— or  the  free  surface  of  the  water  in  contact  with  the  air. 
If  such  a  surface  S  were  given,  for  example,  by  an  equation 
£(#,  j/,  z;  t)  —  0,  it  follows  from  (1.1.3)  that  the  condition 

(1.4.1)  ^  =  u£x  +  <,  +  «£.  +  C«  =  0 

dt 

would  hold  on  S.  From  (1.3.2)  arid  the  fact  that  the  vector  (£X9  £y,  £J 
is  a  normal  vector  to  S  it  follows  that  the  condition  (1.4.1)  can  be 
written  in  the  form 

(,.4.2),  <»<•=..     =„., 

a»        v#  +  ti+  c; 

in  which  d/dn  denotes  differentiation  in  the  direction  of  the  normal 
to  S  and  vn  means  the  common  velocity  of  fluid  and  boundary 
surface  in  the  direction  normal  to  the  surface. 

In  the  important  special  case  in  which  the  boundary  surface  S 
is  fixed,  i.e.  it  is  independent  of  the  time  t,  we  have  the  condition 

d0 

(1.4.3)  —  =  0  on  S. 

on 

This  is  the  appropriate  boundary  condition  at  the  bottom  of  the  sea, 
or  at  the  walls  of  a  tank  containing  water. 

Another  extremely  important  special  case  is  that  in  which  S  is  a 
free  surface  of  the  liquid,  i.e.  a  surface  on  which  the  pressure  p  is 
prescribed  but  the  form  of  the  surface  is  not  prescribed  a  priori. 
We  shall  in  general  assume  that  such  a  free  surface  is  given  by  the 
equation 

(1.4.4)  y  =  ri(x,z\t). 

On  such  a  surface  f  =  y  —  r)(a\  3;  t)  —  0  for  any  particle,  and  hence 
(1.4.1)  yields  the  condition 

(1.4.5)  0xr,x  -&y  +  0Z*)Z  +  17,  =  0  on  S. 

In  addition,  as  remarked  above,  we  assume  that  the  pressure  p  is 
given  on  S;  as  a  consequence  the  Bernoulli  law  (1.3.4)  yields  the 
condition: 

*  Actually,  this  property  is  a  consequence  of  the  basic  assumption  in  con- 
tinuum mechanics  that  the  motion  of  the  fluid  can  be  described  mathematically 
as  a  topological  deformation  which  depends  continuously  on  the  time  t. 


12  WATER   WAVES 

(1.4.6)  ff,  +  0t  +  I  (01  +  02y  +  01)  +  P-  =  0   OR   S. 

2  Q 

(As  remarked  earlier,  we  may  take  the  quantity  C(t)  =  0  in  (1.3.4).) 
Thus  the  potential  function  0  must  satisfy  the  two  nonlinear  boundary 
conditions  (1.4.5)  and  (1.4.6)  on  a  free  surface.  This  is  in  sharp  con- 
trast to  the  single  linear  boundary  condition  (1.4.3)  for  a  fixed 
boundary  surface,  but  it  is  not  strange  that  two  conditions  should 
be  prescribed  in  the  case  of  the  free  surface  since  an  additional 
unknown  function  rj(x,  z;  t)9  the  vertical  displacement  of  the  free 
surface,  is  involved  in  the  latter  case. 

Later  on  we  shall  also  be  concerned  with  problems  involving  rigid 
bodies  floating  in  the  water  and  S  will  be  the  portion  of  the  rigid 
body  in  contact  with  the  water.  In  such  cases  the  function  r)(x,  z\  t) 
will  be  determined  by  the  motion  of  the  rigid  body,  which  in  turn 
will  be  fixed  (through  the  dynamical  laws  of  rigid  body  mechanics) 
by  the  pressure  p  between  the  body  and  the  water  in  accordance 
with  (1.4.6).  The  detailed  conditions  for  such  cases  will  be  worked 
out  later  on  at  an  appropriate  place. 

1.5.  Singularities  of  the  velocity  potential 

In  our  discussion  up  to  now  it  has  been  tacitly  assumed  that  all 
quantities  such  as  the  pressure,  velocity  potential,  velocity  com- 
ponents, etc.  are  regular  functions  of  their  arguments.  It  is,  however, 
often  useful  to  permit  singularities  of  one  kind  or  another  to  occur 
as  an  idealization  of,  or  an  approximation  to,  certain  physical  situations. 
Perhaps  the  most  useful  such  singularity  is  the  point  source  or  sink 
which  is  given  by  the  harmonic  function 

(1.5.1)  0 


in  three  dimensions,  and  by 

(1.5.2)  0  =  —  log  r,       r2  =  x*  +  y* 

2n 

in  two  dimensions.  Both  of  these  functions  yield  flows  which  are 
radially  outward  from  the  origin,  and  for  which  the  flux  per  unit 
time  across  a  closed  surface  (for  (1.5.1))  or  a  closed  curve  (for  (1.5.2)) 
surrounding  the  origin  has  the  value  c,  as  one  readily  verifies  since 
d0/dn  =  d0fdr  for  r  =  constant.  That  these  functions  represent  at 


BASIC   HYDRODYNAMICS  13 

best  idealizations  of  the  physical  situations  implied  in  the  words 
source  and  sink  is  clear  from  the  fact  that  they  yield  infinite  velocities 
at  r  =  0.  Nevertheless,  it  is  very  useful  here— as  in  other  branches 
of  applied  mathematics  —to  accept  such  infinities  with  the  reservation 
that  the  results  obtained  are  not  to  be  taken  too  literally  in  the 
immediate  vicinity  of  the  singular  point. 

We  shall  have  occasion  to  deal  with  other  singularities  than  sources 
or  sinks,  such  as  dipoles  and  multipoles,  but  these  will  be  introduced 
when  needed. 


1.6.  Notions  concerning  energy  and  energy  flux 

In  dealing  with  surface  gravity  waves  in  water  it  is  important 
and  useful  to  analyze  in  some  detail  the  flow  of  energy  in  the  fluid 
past  a  given  surface  S.  Let  R  be  the  region  occupied  by  water  and 
bounded  by  a  "geometric"  surface  S  which  may,  or  may  not,  move 
independently  of  the  liquid.  The  energy  E  contained  in  R  consists 
of  the  kinetic  energy  of  the  water  particles  in  R  and  their  potential 
energy  due  to  gravity;  hence  E  is  given  by 

(1.6.1)       E  -  Q  R  (0|  +  01  +  01)  +  gy\  dx  dy  dz, 


or,  alternatively,  by 

(1.6.2)  E  =  -  JJJ(P  +  e^t)dx  dy  dz 

R 

upon  applying  Bernoulli's  law  (1.3.4)  with  C(t)  =  0. 

We  wish  to  calculate  dE/dt,  having  in  mind  that  the  region  R 
is  not  necessarily  fixed,  but  may  depend  on  the  time  t.  Quite  generally, 

if  E  =  f  f  f  f(x,  y,  z;  l)dx  dy  dz,  it  is  well  known  that 


R  S 

in  which  vn  denotes  the  normal  velocity  of  the  boundary  S  of  R 
taken  positive  in  the  direction  outward  from  R.  In  applying  the 
formula  for  dE/dt  we  make  use  of  the  definition  of  the  function  / 
implied  in  (1.6.1)  in  the  first  term,  but  take  /  from  (1.6.2)  for  the 
second  term.  The  result  is 


14  WATER    WAVES 


/ITT  P  P  P 

~dt   =  Q  J  J  J  ( 


s 
The  integrand  in  the  first  integral  can  be  expressed  in  the  form 

&x(®t)x  +  ^y(^t)y  +  &z(®t)z  =  grad  &  •  grad  0t 
and  hence  the  integral  can  be  written  as  the  following  surface  integral: 

\  ^  dS, 

s 

in  view  of  Green's  formula  and  the  fact  that  V20  =  0.  Thus  the 
expression  for  dE/dt,  the  rate  of  change  of  the  energy  in  R9  can  be 
put  into  the  following  form: 


J  Tjl  /*     /» 

-f-JJ 


(1.6.3)  .  =          [Q0t(0n  -  vn)  -  pvn]dS. 


s 

We  recall  that  vn  means  the  normal  velocity  component  of  S,  and 
0n  refers  to  the  velocity  component  of  the  fluid  taken  in  the  direction 
of  the  normal  to  S  which  points  outward  from  R. 

It  happens  frequently  that  the  boundary  surface  S  of  R  is  made 
up  of  a  number  of  different  pieces  which  have  different  properties 
or  for  which  various  different  conditions  are  prescribed.  Suppose 
first  that  a  portion  SP  of  S  is  a  "physical"  boundary  containing 
always  the  same  fluid  particles.  Then  0n  and  vn  are  identical  (cf. 
(1.4.2))  and 


(1.6.4.) 


dE 


dS. 


dt 

SP 

If,  in  addition,  the  surface  SP  is  fixed  in  space,  i.e.  vn  =  0,  the 
contribution  of  SP  to  dE/dt  evidently  vanishes,  as  it  should,  since  no 
energy  flows  through  a  fixed  boundary  containing  always  the  same 
fluid  particles.  Similarly,  the  contribution  to  the  energy  flux  also 
vanishes  in  the  important  special  case  in  which  SF  is  a  free  surface 
on  which  the  pressure  p  vanishes;  this  result  also  accords  with  what 
one  expects  on  physical  grounds. 

Suppose  now  that  SG  is  a  "geometric"  surface  fixed  in  space,  but 
not  necessarily  consisting  of  the  same  particles  of  water.  In  this 
case  we  have  vn  =  0  and  the  flow  of  energy  through  SG  is  given  by 


,,.6.5) 


BASIC    HYDRODYNAMICS  15 

dE 


-//< 


An  important  special  case  for  us  is  that  in  which  0  is  the  velocity 
potential  for  a  plane  progressing  wave  given,  for  example,  by 

(1.6.6)  0(x,  y,  z;  t)  =  <p(x-ct,  y,  z), 

which  represents  a  wave  moving  with  constant  velocity  c  in  the 
direction  of  the  #-axis.  The  flux  through  a  fixed  plane  surface  S 
orthogonal  to  the  #-axis  is  easily  seen  from  (1.6.5)  to  be  given  by 

(iv 

(1.6.7)  -^  = 

S 

The  negative  sign  results  since  our  stipulations  amount  to  saying 
that  the  region  R  occupied  by  the  fluid  lies  on  the  negative  side 
of  S  (i.e.  on  the  side  away  from  the  positive  normal,  the  #-axis); 
and  consequently  the  energy  flux  through  S  due  to  a  progressing 
wave  moving  in  the  positive  direction  of  the  normal  (so  that  c  is 
positive)  is  such  as  to  decrease  the  energy  in  jR,  as  one  would  expect. 
It  is  to  be  noted  that  there  is  always  a  flow  of  energy  through  a 
surface  S  orthogonal  to  the  direction  of  a  progressing  wave  if 
0n  E^=  0  — even  though  the  motion  of  the  individual  particles  of  the 
fluid  should  happen,  for  example,  to  be  such  that  the  particles  move 
in  a  direction  opposite  to  that  of  the  progressing  wave. 

1.7.  Formulation  of  a  surface  wave  problem 

It  is  perhaps  useful— although  somewhat  discouraging,  it  must  be 
admitted— to  sum  up  the  above  discussion  concerning  the  fun- 
damental mathematical  basis  for  our  later  developments  by  formula- 
ting a  rather  general,  but  typical,  problem  in  the  hydrodynamics  of 
surface  waves.  The  physical  situation  is  indicated  in  Figure  1.7.1; 
what  is  intended  is  a  situation  like  that  on  any  ocean  beach.  The 
water  is  assumed  to  be  initially  at  rest  and  to  fill  the  space  R  defined  by 

—  h((r,  2)^2/^0,   —  oo  <  z  <  oo, 

and  extending  to  +  oo  in  the  ^-direction.  At  the  time  t  =  0,  a  given 
disturbance  is  created  on  the  surface  of  the  water  over  a  region  D 
(by  the  wind,  perhaps),  and  one  wishes  to  determine  mathematically 
the  subsequent  motion  of  the  water;  in  particular,  the  form  of  the 


16 


WATER   WAVES 


Fig.  1.7.1.  A  very  general  surface  wave  problem 

free  surface  y  =  r\(x,  z;  t)  is  to  be  determined.  On  the  basis  of  these 
assumptions  the  following  conditions  should  be  satisfied:  First  of 
all,  the  differential  equation  to  be  satisfied  by  0  is,  of  course,  the 
Laplace  equation 


(I.T.I)     V*0  =  0XX  +  0y 


(xs(z9  t)  ^  x  <  oo 
^Q  for     -  h(x9  z)^y^  rj(^  z;  t) 

[ 


00   <  Z   <   00 


It  is  to  be  noted  that  xs(z;  t)9  the  abscissae  of  the  water  line  on  shore, 
and  rj(x9  z;  t),  the  free  surface  elevation,  are  not  known  in  advance 
but  are  rather  to  be  determined  as  an  integral  part  of  the  solution. 
As  boundary  condition  to  be  satisfied  at  the  bottom  of  the  sea  we 
have 


(1-7.2) 


— -  =  0  for  y  =  —  h(x9  z)9 
on 


while  the  free  surface  conditions  are  the  kinematic  condition  (cf. 
(1.4.5)) 

(1.T.3)        0xrjx  -0y  +  &zr]z  +  rjt  =  0  for  y  =  rj(x9  z;  t)9 

and  the  dynamic  condition 

(1.T.4)      gq  +  0t  +  \(0l  +01+  0%)  =  F(x9  z;  t)   on  y  -  rj(x9  z;  t), 

with  F(x9  z;  t)  =  0  everywhere  except  over  the  region  D  where  the 
disturbance  is  created.  At  oo,  i.e.  for  x  ->  oo  and  |  z  |  ->  oo,  we 
might  prescribe  that  0  and  y  remain  bounded,  or  perhaps  even  that 
they  and  certain  of  their  derivatives  tend  to  zero.  Next  we  have  the 
initial  conditions 

(1.T.5)  qfa  z;  t)  =  0  for  t  =  0, 

(1.T.6)  0X  =  0y  =  0Z  s=  0  for  t  =  0, 


BASIC    HYDRODYNAMICS  17 

appropriate  to  the  condition  of  rest  in  an  equilibrium  position. 
Finally  we  must  prescribe  conditions  fixing  the  disturbance;  this 
could  be  done,  for  example,  by  giving  the  pressure  p  over  the 
disturbed  region  D  of  the  surface,  in  other  words  by  prescribing  the 
function  F  in  (1.7.4)  appropriately  there. 

One  has  only  to  write  down  the  above  formulation  of  our  problem 
to  realize  how  difficult  it  is  to  solve  it.  In  the  first  place  the  problem 
is  nonlinear,  but  what  makes  for  perhaps  even  greater  difficulties  is 
the  fact  that  the  free  surface  is  not  known  a  priori  and  hence  the 
domain  in  which  the  velocity  potential  is  to  be  determined  is  not 
known  in  advance— aside  from  the  fact  that  its*  boundary  varies 
with  the  time. 

These  are,  however,  not  the  only  difficulties  in  the  above  problem. 
If  we  assume  that  the  function  0  is  regular  throughout  the  interior 
of  R  and  uniformly  bounded  (together  with  some  of  its  derivatives, 
perhaps)  in  R,  the  formulation  of  the  problem  given  above  would 
seem  to  be  reasonable  from  the  point  of  view  of  mechanics.  However, 
the  solution  would  probably  not  exist  for  all  t  >  0  for  the  following 
reason:  everyone  who  has  visited  an  ocean  beach  is  well  aware  that 
the  waves  do  not  come  in  smoothly  all  the  way  to  the  shore  (except 
possibly  in  very  calm  weather),  but,  rather,  they  steepen  in  front, 
curl  over,  and  eventually  break.  In  other  words,  any  mathematical 
formulation  of  the  problem  which  would  fit  the  commonly  observed 
facts  even  for  a  limited  time  would  necessitate  postulating  the 
existence  of  singularities  of  unknown  location  in  both  space  and  time. 

Because  of  the  difficulty  of  the  general  nonlinear  theory  very  little 
progress  has  been  made  in  solving  concrete  problems  which  employ 
it.  An  exception  is  the  problem  of  proving  the  existence  of  two- 
dimensional  periodic  progressing  waves  in  water  of  uniform  depth. 
This  was  done  first  by  Nckrassov  [N.I],  [N.I a]  and  by  Levi-Civita 
[L.7]  for  water  of  infinite  depth,  and  later  by  Struik  [S.29]  for  water  of 
constant  finite  depth.  In  Chapter  12  an  account  of  Levi-Civita's  theory 
is  given.  In  both  cases  the  authors  prove  rigorously  the  existence  of 
waves  having  amplitudes  near  to  zero  by  showing  that  perturbation 
series  in  the  amplitude  converge.  Another  exception  to  the  above 
statement  is  the  problem  of  the  solitary  wave,  the  existence  of  which, 
from  the  mathematical  point  of  view,  has  been  proved  recently  by 
Lavrentieff  [L.4]  and  by  Friedrichs  and  Hyers  [F.13];  an  account  of 
the  work  of  the  latter  two  authors  is  given  in  Chapter  10.9. 

It  seems  likely  that  solutions  of  problems  in  the  full  nonlinear 


18  WATER   WAVES 

version  of  the  theory  will,  for  a  long  time  to  come,  continue  to  be 
of  the  nature  of  existence  theorems  for  motions  of  a  rather  special 
nature. 

In  order  to  make  progress  with  the  theory  of  surface  waves  it  is 
in  general  necessary  to  simplify  the  theory  by  making  special  hypoth- 
eses of  one  kind  or  another  which  suggest  themselves  on  the  basis 
of  the  general  physical  circumstances  contemplated  in  a  given  class 
of  problems.  As  we  have  already  explained  in  the  introduction,  up 
to  now  attention  has  been  concentrated  almost  exclusively  upon  the 
two  approximate  theories  which  result  when  either  a)  the  amplitude 
of  the  surface  waves  is  considered  small  (with  respect  to  wave  length, 
for  example),  or  b)  the  depth  of  the  water  is  considered  small  (again 
with  respect,  say,  to  wave  length).  The  first  hypothesis  leads  to  a 
linear  theory  and  to  boundary  value  problems  more  or  less  of  classical 
type;  while  the  second  leads  to  a  nonlinear  theory  for  initial  value 
problems,  which  in  lowest  order  is  of  the  type  employed  in  wave 
propagation  in  compressible  gases.  If  both  hypotheses  are  made,  the 
result  is  a  linear  theory  involving  essentially  the  classical  linear 
wave  equation;  the  present  theory  of  the  tides  belongs  in  this  class 
of  problems. 

In  the  next  chapter  we  derive  the  approximate  theories  arising 
from  the  two  hypotheses  by  starting  from  the  general  theory  and 
then  developing  formally  with  respect  to  an  appropriate  parameter— 
essentially  the  surface  wave  amplitude  in  one  case  and  the  depth 
of  the  water  in  the  other— and  in  subsequent  chapters  we  continue 
by  treating  a  variety  of  special  problems  in  each  of  the  two  classes. 


CHAPTER  2 

The  Two  Basic  Approximate  Theories 

2.1.  Theory  of  waves  of  small  amplitude 

It  has  already  been  stated  that  the  theory  of  waves  of  small 
amplitude  can  be  derived  as  an  approximation  to  the  general  theory 
presented  in  Chapter  I  on  the  basis  of  the  assumption  that  the 
velocity  of  the  water  particles,  the  free  surface  elevation  y—r)(x,  z;  t), 
and  their  derivatives,  are  all  small  quantities.  We  assume,  in  fact, 
that  the  velocity  potential  0  and  the  surface  elevation  rj  possess  the 
following  power  series  expansions  with  respect  to  a  parameter  e: 

(2.1.1)  0  =  e0(l)  +  e20(2)  +  e30(3)  +  .  .  .,  and 

It  follows  first  of  all  that  each  of  the  functions  0<k)(x,  y,  z;  t)  is 
a  solution  of  the  Laplace  equation,  i.e. 

(2.1.3)  V20><*>  —  0. 

We  turn  next  to  the  discussion  of  the  boundary  conditions.  At  a 
fixed  physical  boundary  (cf.  section  1.4)  of  the  fluid  we  have  clearly 
the  conditions 


(2.1.4)  -a--  =  0, 

in   which  d/dn  represents  differentiation  along  the  normal   to  the 
boundary  surface. 

At  a  free  surface  S:  y  —  r](x>  z;  t)  on  which  the  pressure  is  zero  we 
have  two  boundary  conditions.  One  of  them  arises  from  the  Bernoulli 
law  and  has  the  form 


grj  +  0t  +  i(<Z>*  +  *I  +  01)  =  0  on  S. 

Upon  insertion  of  (2.1.1)  and  (2.1.2)  in  this  condition  and  developing 
&t>  &x>  etc'  systematically  in  powers  of  e  (due  regard  being  paid  to 
the  fact  that  the  functions  0^k)9  0(®,  etc.  are  to  be  evaluated  for 

19 


20  WATER  WAVES 

y  =  ri(x,  z\  t)  and  that  77  in  its  turn  is  given  in  terms  of  e   by 
it)  =  77<°>  -)-  erjM  -\-  .  .  .  )  one  finds  readily  the  conditions 

(2.1.5)  ?7«»  =  0, 

(2.1.6)  gqM  +  0]11  =  0. 

(2.1.7)  W«>  +  <P«  +  i[(0»>  )2  +  (0<1}  )2  +  (#»>  )•]  +  i7««*j«  =  0 


to  be  satisfied  for  y  =  T?(O),  and  since  ?y(0)  =  0  from  (2.1.5)  it  follows 
that  the  conditions  (2.1.6),  (2.1.7),  etc.  are  all  to  be  satisfied  on  the 
originally  undisturbed  surface  of  the  water  y  =  0.  The  other  boundary 
condition  on  S  arises  from  the  fact  that  the  water  particles  stay 
on  S  (cf.  section  1.4);  it  is  expressed  in  the  form 

#«ifc  +  #.%  +*]t  =  ®y  on  S. 

Insertion  of  the  power  series  for  0  and  77  in  this  expression  leads  to 
the  conditions 

(2.1.8)  fjW  -  0, 

(2.1.9) 

(2.1.10) 


which  are  also  to  be  satisfied  for  y  =  0. 

In  view  of  the  fact  that  iy(0)  =  0,  the  free  surface  conditions  can 
be  put  in  the  form 

(2.1.11)  gqW+Qp^O, 

(2.1.12)  OTC»  +  0W  =  _  J[(0W)«  +  (0^)2  +  ((PW  )•]  «  ,<i)0W, 

(2.1.13)  g^<w>  +  0|w)  =  jP^-D, 

in  which  the  symbol  pin-u  refers  to  a  certain  combination  of  the 
functions  rj(k)  and  0(k)  with  fc  j£  n  —  1,  and  all  conditions  are  to  be 
satisfied  for  y  =  0.  Similarly,  the  other  set  of  free  surface  conditions 
becomes 


TUP:   TWO    BASIC   APPROXIMATE   THEORIES  21 


(2.1.14)  fi 

(2.1.15)  ,»  =  *»  -  «>«,<«  -  *«,»>  + 


(2.1.16)  T?<W) 

in  which  G(n~l)  depends  only  upon  functions  ?y(fc)  and  0(k)  with 
k  ^  n  —  I9  and  once  more  all  conditions  are  to  be  satisfied  for 
y  _=  o.  This  theory  therefore  is  a  development  in  the  neighborhood 
of  the  rest  position  of  equilibrium  of  the  water. 

The  relations  (2.1.11)  to  (2.1.16)  thus,  in  principle,  furnish  a  means 
of  calculating  successively  the  coefficients  of  the  series  (2.1.1)  and 
(2.1.2),  assuming  that  such  series  exist:  The  conditions  (2.1.11)  and 
(2.1.14)  at  the  free  surface  together  with  appropriate  conditions  at 
other  boundaries,  and  initial  conditions  for  /  =  0,  would  in  conjunc- 
tion with  V20(1)  —  0  lead  to  unique  solutions  r/(l)  and  <2>(1).  Once 
rfl)  and  0(l)  are  determined,  they  can  be  inserted  in  the  conditions 
(2.1.12)  and  (2.1.15)  to  yield  two  conditions  for  rj(2)  and  0(2>  which 
with  the  subsidiary  boundary  and  other  conditions  on  0(2)  serve  to 
determine  them,  etc.  One  could  interpret  the  work  of  Levi-Civita  [L.7] 
and  Struik  [S.29]  referred  to  in  section  1.7  as  a  method  of  proving 
the  existence  of  progressing  waves  which  are  periodic  in  x  by  showing 
that  the  functions  0  and  rj  can  indeed  be  represented  as  convergent 
power  series  in  e  for  e  sufficiently  small. 

In  what  follows  in  Part  II  of  this  book  we  shall  content  ourselves 
in  the  main  with  the  degree  of  approximation  implied  in  breaking 
off  the  perturbation  series  after  the  terms  e0(l}  and  erj(l)  in  the  series 
(2.1.1)  and  (2.1.2),  i.e.  we  set  0  =  e0(l)  and  rj  =  er](l).  With  this 
stipulation  the  free  surface  conditions  (2.1.11)  and  2.1.14)  yield 

(2.1.17)  gl?+0t  =  o) 

I  for  y  =  0. 

(2.1.18)  rit    -0tf  =  oJ 

By  elimination  of  r\  between  these  two  relations  the  single  condition 
on  0: 

(2.1.19)  0tt  +  g0y  =  0         for  y  =  0 

is  obtained;  this  condition  is  the  one  which  will  be  used  mainly  in 
Part  II  in  order  to  determine  0  from  V20  =  0,  after  which  the  free 
surface  elevation  r\  can  be  determined  from  (2.1.17).  The  usual 
method  of  obtaining  the  last  three  conditions  is  to  reject  all  but 


22  WATER   WAVES 

the  linear  terms  in  77  and  0  and  their  derivatives  in  the  kinematic 
(cf.  (1.4.5))  and  dynamic  (cf.  (1.4.6))  free  surface  boundary  con- 
ditions. By  proceeding  in  this  way  we  can  obtain  a  first  approximation 
to  the  pressure  p  (which  was  not  considered  in  the  above  general 
perturbation  scheme)  in  the  form: 

(2.1.20)  -=  -gy-0t.  * 

Q 

We  can  now  see  the  great  simplifications  which  result  through 
the  linearization  of  the  free  surface  conditions:  not  only  does  the 
problem  become  linear,  but  also  the  domain  in  which  its  solution 
is  to  be  determined  becomes  fixed  a  priori  and  consequently  the 
surface  wave  problems  in  this  formulation  belong,  from  the  mathe- 
matical point  of  view,  to  the  classical  boundary  problems  of  potential 
theory. 

2.2.  Shallow  water  theory  to  lowest  order.  Tidal  theory 

A  different  kind  of  approximation  from  the  foregoing  linear  theory 
of  waves  of  small  amplitude  results  when  it  is  assumed  that  the 
depth  of  the  water  is  sufficiently  small  compared  with  some  other 
significant  length,  such  as,  for  example,  the  radius  of  curvature  of 
the  water  surface.  In  this  theory  it  is  not  necessary  to  assume  that 
the  displacement  and  slope  of  the  water  surface  are  small,  and  the 
resulting  theory  is  as  a  consequence  not  a  linear  theory.  There  are 
many  circumstances  in  nature  under  which  such  a  theory  leads  to 
a  good  approximation  to  the  actual  occurrences,  as  has  already  been 
mentioned  in  the  introduction.  Among  such  occurrences  are  the  tides 
in  the  oceans,  the  "solitary  wave"  in  sufficiently  shallow  water,  and 
the  breaking  of  waves  on  shallow  beaches.  In  addition,  many  pheno- 
mena met  with  in  hydraulics  concerning  flows  in  open  channels  such 
as  roll  waves,  flood  waves  in  rivers,  surges  in  channels  due  to  sudden 
influx  of  water,  and  other  kindred  phenomena,  belong  in  the  nonlinear 
shallow  water  theory.  Chapters  10  and  11  are  devoted  to  the  working 
out  of  consequences  of  the  shallow  water  theory. 

The  shallow  water  theory  is,  in  its  lowest  approximation,  the  basic 
theory  used  in  hydraulics  by  engineers  in  dealing  with  flows  in  open 

*  In  case  the  surface  pressure  pa(xt  z;  t)  is  not  zero  one  finds  readily  that  (2.1.17) 
•is  replaced  by 

(2.1.20)!  gri  +  &t  =  -  Po/Q, 

while  (2.1.18)  remains  unaltered. 


THE    TWO    BASIC    APPROXIMATE   THEORIES  23 

channels,  and  also  the  theory  commonly  referred  to  in  the  standard 
treatises  on  hydrodynamics  as  the  theory  of  long  waves.  We  begin  by 
giving  first  a  derivation  of  the  theory  for  two-dimensional  motion 
along  essentially  the  lines  followed  by  Lamb  [L.3],  p.  254.  As  usual, 
the  undisturbed  free  surface  of  the  water  is  taken  as  the  o?-axis  and 
the  t/-axis  is  taken  vertically  upwards.  The  bottom  is  given  by 
y  =  —  h(x),  so  that  h  represents  the  variable  depth  of  the  undisturbed 
water.  The  surface  displacement  is  given  by  y  =  r](x,  t).  The  velocity 
components  are  denoted  by  u(x,  y,  t)  and  v(x,  y,  t). 
The  equation  of  continuity  is 

(2.2.1)  Ux  +  Vy  =  0. 

The  conditions  to  be  satisfied  at  the  free  surface  are  the  kinematical 
condition: 

(2.2.2)  (r)t  +  ur,x  -  v)  \y=ri  =  0; 
and  the  dynamical  condition  on  the  pressure: 

(2.2.3)  p  !,_„  =  0. 
At  the  bottom  the  condition  is 

(2.2.4)  (uhx  +  v)  |V=_A  =  0. 
Integration  of  (2.2.1)  with  respect  to  y  yields 

(2.2.5)  f  (Ux)dy  +v\\  =  Q. 

J  —h 

Use  of  the  condition  (2.2.2)  at  y  =  77  and  (2.2.4)  at  y  —  —  h  yields 
the  relation 

(2.2.6)  r   uxdy  +  r)t  +  u  \    •  ^  +  u  \_h  •  hx  =  0. 
J  —h 

We  introduce  the  relation 

g    p(«)  rn 

(2.2.7)  ^-  u  dy  =  u  \       •  rjx  +  u  \y=_h  •  hx  +  \      ux  dy. 
v®  J  _fc(jB)  J  -h 

and  combine  it  with  (2.2.6)  to  obtain 
(2.2.8) 


Up  to  this  point  no  approximations  have  been  introduced. 

The  shallow  water  theory  is  an  approximate  theory  which  results 
from  the  assumption  that  the  y-  component  of  the  acceleration  of 


24  WATER   WAVES 

the  water  particles  has  a  negligible  effect  on  the  pressure  p,  or, 
what  amounts  to  the  same  thing,  that  the  pressure  p  is  given  as  in 
hydrostatics  *  by 

(2.2.9)  p  =  gety  -  y). 

The  quantity  Q  is  the  density  of  the  water.  A  number  of  consequences 
of  (2.2.9)  are  useful  for  our  purposes.  To  begin  with,  we  observe  that 

(2.2.10)  px  =  ggjfc  , 

so  that  px  is  independent  of  y.  It  follows  that  the  ^-component  of 
the  acceleration  of  the  water  particles  is  also  independent  of  y\ 
and  hence  u,  the  ^-component  of  the  velocity,  is  also  independent 
of  y  for  all  t  if  it  was  at  any  time,  say  at  t  =  0.  We  shall  assume 
this  to  be  true  in  all  cases —it  is  true  for  example  in  the  important 
special  case  in  which  the  water  was  at  rest  at  t  =  0— so  that  u= u(x,  t) 
depends  only  on  x  and  t  from  now  on.  As  equation  of  motion  in  the 
^-direction  we  may  write,  therefore,  in  view  of  (2.2.10): 

(2.2.11)  ut  +  uux  =  -gife. 

This  is  simply  the  usual  equation  of  motion  in  the  Eulerian  form, 
use  having  been  made  of  uy  =  0.  In  addition,  (2.2.8)  may  now  be 
written 

(2.2.12)  [u(ri  +  h)]x  =  -iji9 

prj  pr\ 

since        u  dy  =  u\     dy  on  account  of  the  fact  that  u  is  independent 

J  —  h  J  —h 

of  y.  The  two  first  order  differential  equations  (2.2.11)  and  (2.2.12) 
for  the  functions  u(x,  t)  and  r\(x,  t)  are  the  differential  equations  of 
the  nonlinear  shallow  water  theory.  Once  the  initial  state  of  the  fluid 
is  prescribed,  i.e.  once  the  values  of  u  and  77  at  the  time  t  =  0  are 
given,  the  equations  (2.2.11)  and  (2.2.12)  yield  the  subsequent 
motion. 

If  in  addition  to  the  basic  assumption  of  the  shallow  water  theory 
expressed  by  (2.2.9)  we  assume  that  u  and  77,  the  particle  velocity 
and  free  surface  elevation,  and  their  derivatives  are  small  quantities 
whose  squares  and  products  can  be  neglected  in  comparison  with 
linear  terms,  it  follows  at  once  that  equations  (2.2.11)  and  (2.2.12) 
simplify  to 

(2.2.13)  ut  =  -  gjfc  , 

(2.2.14)  (uh)x  =  -  ifc  , 

We  have  pv  =  —  gg  and  (2.2.9)  results  through  the  use  of  p  =  0  for  y  =  r). 


THE   TWO   BASIC   APPROXIMATE   THEORIES  25 

from  which  77  can  be  eliminated  to  yield  for  u  the  equation 

(2.2.15)  (uh)xx  ~  —  utt  =  0. 

S 

If,  in  addition,  the  depth  h  is  constant  it  follows  readily  that  u 
satisfies  the  linear  wave  equation 

(2.2.16)  uxx-—utt  =  0. 

In  this  case  77  satisfies  the  same  equation.  One  observes  therefore 
the  important  result  that  the  propagation  speed  of  a  disturbance  is 

given  by  Vgh.  In  principle,  this  linearized  version  of  the  shallow  water 
theory  is  the  one  which  has  always  been  used  as  the  basis  for  the 
theory  of  the  tides.  Of  course,  the  tidal  theory  for  the  oceans  requires 
for  its  complete  formulation  the  introduction  of  the  external  forces 
acting  on  the  water  due  to  the  gravitational  attraction  of  the  moon 
and  the  sun,  and  also  the  Coriolis  forces  due  to  the  rotation  of  the 
earth,  but  nevertheless  the  basic  fact  about  the  tidal  theory  from 
the  standpoint  of  mathematics  is  that  it  belongs  to  the  linear  shallow 
water  theory.  The  actual  oceans  do  not  from  most  points  of  view 
impress  one  as  being  shallow;  in  the  present  connection,  however, 
the  depth  is  actually  very  small  compared  with  the  curvature  of 
the  tidal  wave  surface  so  that  the  shallow  water  approximation  is  an 
excellent  one.  That  the  tidal  phenomena  should  be  linear  to  a  good 
approximation  would  also  seem  rather  obvious  on  account  of  the 
small  amplitudes  of  the  tides  compared  with  the  dimensions  of  the 
oceans.  A  few  additional  remarks  about  tidal  theory  and  some  other 
applications  of  the  linearized  version  of  the  shallow  water  theory  to 
concrete  problems  (seiches  in  lakes,  and  floating  breakwaters,  for 
example)  are  given  in  Chapter  10.13. 

2.3.  Gas  dynamics  analogy 

It  is  possible  to  introduce  a  different  set  of  dependent  variables 
in  such  a  way  that  the  equations  of  the  shallow  water  theory  become 
analogous  to  the  fundamental  differential  equations  of  gas  dynamics  for 
the  case  of  a  compressible  flow  involving  only  one  space  variable  x. 
(This  seems  to  have  been  noticed  first  by  Riabouchinsky  [R.8].) 
To  this  end  we  introduce  the  mass  per  unit  area  given  by 

(2.3.1)  Q  =  Q(rj  +h). 


26  WATER   WAVES 

Since  h  depends  only  on  x  we  have 

(2.3.2)  ft  ==e^. 

We  next  define  the  force  p  per  unit  width: 

(2.3.3)  p  =  \\pdy, 

J  —  n 

which,  in  view  of  (2.2.9)  and  (2.8.1),  leads  to 

(2.3.4)  P  =  ^(n+h}*=-j^Q*. 

The  relation  between  p  and  Q  is  thus  of  the  form  p  =  AQY  with 
y==2,  that  is,  the  "pressure"  p  and  the  "density"  Q  are  connected 
by  an  "adiabatic"  relation  with  the  fixed  exponent  2. 
Equation  (2.2.11)  may  now  be  written 

q(r)  +  h)(ut  +  uux)  =  -  gQfa  +  h)qx 

and  this,  in  turn,  may  be  expressed  through  use  of  (2.3.1)   and 

(2.3.4)  as  follows: 

(2.3.5)  Q(ut  +  uu9)  =  -  px  +  gQhx  , 

as  one  can  readily  verify. 

The  equation  (2.2.12)  may  be  written  as 

(2.3.6) 


in  view  of  (2.3.2)  as  well  as  (2.3.1).  The  differential  equations  (2.3.5) 
and  (2.3.6),  together  with  the  "adiabatic"  lawp  =  gg*/2g  given  by 
(2.3.4),  are  identical  in  form  with  the  equations  of  compressible  gas 
dynamics  for  a  one-dimensional  flow  except  for  the  term  gQhx  on 
the  right  hand  side  of  (2.3.5),  and  this  term  vanishes  if  the  original 
undisturbed  depth  h  of  the  water  is  constant.  The  "sound  speed"  c 
corresponding  to  our  equations  (2.3.5)  and  (2.3.6)  is,  in  analogy  with 
gas  dynamics,  given  by  c  =  VdpldQ,  and  this  from  (2.3.4)  and  (2.3.1  ) 
has  the  value 


(2.8.7)  C=—  =  Vg(rj  +  h). 

It  will  be  seen  later  that  c(x,  t)  represents  the  local  speed  at  which 
a  small  disturbance  advances  relative  to  the  water. 


THE   TWO   BASIC   APPROXIMATE   THEORIES  27 

2.4.  Systematic  derivation  of  the  shallow  water  theory 

It  is  of  course  a  matter  of  importance  to  know  under  what  cir- 
cumstances the  shallow  water  theory  can  be  expected  to  furnish 
sufficiently  accurate  results.  The  only  assumption  made  above  in 
addition  to  the  customary  assumptions  of  hydrodynamics  was  that 
the  pressure  is  given  as  in  hydrostatics  by  (2.2.9),  but  no  assumption 
was  made  regarding  the  magnitude  of  the  surface  elevation  or  the 
velocity  components.  Consequently  the  shallow  water  theory  may 
be  accurate  for  waves  whose  amplitude  is  not  necessarily  small, 
provided  that  the  hydrostatic  pressure  relation  is  not  invalidated. 
The  above  derivation  of  the  shallow  water  theory  is,  however,  open 
to  the  objection  that  the  role  played  by  the  undisturbed  depth  of  the 
water  in  determining  the  accuracy  of  the  approximation  is  not  put 
in  evidence.  In  fact,  since  we  shall  see  later  on  that  all  motions  die 
out  rather  rapidly  in  the  depth,  it  would  at  first  sight  seem  reasonable 
to  expect  that  the  hydrostatic  law  for  the  pressure  would  be,  on  the 
whole,  more  accurate  the  deeper  the  water.  That  this  is  not  the 
case  in  general  is  well  known,  since  the  solutions  for  steady  progressing 
waves  of  small  amplitude  (i.e.  for  solutions  obtained  by  the  linearized 
theory)  in  water  of  uniform  but  finite  depth  are  approximated 
accurately  by  the  solutions  of  the  shallow  water  theory  (when  it  also 
is  linearized)  only  when  the  depth  of  the  water  is  small  compared 
with  the  wave  length  (cf.  Lamb  [L.3],  p.  368).  It  is  possible  to  give 
a  quite  different  derivation  of  the  shallow  water  theory  in  which  the 
equations  (2.2.11)  and  (2.2.12)  result  from  the  exact  hydrodynamical 
equations  as  the  approximation  of  lowest  order  in  a  perturbation 
procedure  involving  a  formal  development  of  all  quantities  in  powers 
of  the  ratio  of  the  original  depth  of  the  water  to  some  other 
characteristic  length  associated  with  the  horizontal  direction.*  The 
relation  (2.2.9)  is  then  found  to  be  correct  within  quadratic  terms 
in  this  ratio.  In  this  section  we  give  such  a  systematic  derivation 
of  the  shallow  water  theory,  following  K.  O.  Friedrichs  (see  the 
appendix  to  [S.19]),  which,  unlike  the  derivation  given  in  section 

*  In  this  book  the  parameter  of  the  shallow  water  theory  is  defined  in  two 
different  ways:  in  dealing  with  the  breaking  of  waves  in  Chapter  10.10  it  is 
the  ratio  of  the  depth  to  a  significant  radius  of  curvature  of  the  free  surface; 
in  dealing  with  the  solitary  wave,  however,  it  is  essentially  the  ratio  of  the  depth 
to  the  quantity  Uzjg,  with  U  the  propagation  speed  of  the  wave,  and  in  this 
case  the  development  is  carried  out  for  Ua  /gh  near  to  one.  In  still  other 
problems  it  might  well  be  defined  differently  in  terms  of  parameters  that  are 
characteristic  for  such  problems. 


28 


WATER   WAVES 


2.2  above,  is  capable  of  yielding  higher  order  approximations. 
The  disposition  of  the  coordinate  axes  is  taken  in  the  usual  manner, 
with  the  x,  3-plane  the  undisturbed  water  surface  and  the  i/-axis 
positive  upward.  The  free  surface  elevation  is  given  by  y  —  rj(x9  z,  t) 
and  the  bottom  surface  by  y  =  —  h(x,  z).  We  recapitulate  for  the 
sake  of  convenience  the  differential  equations  and  boundary  con- 
ditions in  terms  of  the  Euler  variables,  that  is,  the  equations  of 
continuity  and  motion,  the  vanishing  of  the  rot  at  ion,  and  the  boundary 
conditions: 


(2.4.1) 


(2.4.2) 


wz  =  0, 


wt  +  uwx 


vu 


vv 


vw 


px 


wvz  =  ---  py 


wwz  —   ----  pz 


(2.4.3) 
(2.4.4) 
(2.4.5) 
(2.4.6) 


=  wx,  vx  =  u 


nt  +  urix  +  wr]z  =  v  at  y  =  77, 
p  =  0  at  y  =  77, 
uhx  -\-v-\~  whz  =  0  at  y  =  —  /i. 

We  now  introduce  dimensionless  variables  through  the  use  of  two 
lengths  d  and  k,  with  d  intended  to  represent  a  typical  depth  and  k 
a  typical  length  in  the  horizontal  direction— it  is  characteristic  of 
the  procedure  followed  here  that  the  horizontal  and  vertical  direc- 
tions are  not  treated  in  the  same  way.  The  new  independent  variables 
are  as  follows: 

(2.4.7)  x  =  x/k,  y  =  y/d,  z  =  z/k,  r  =  t  Vgd/k, 
while  the  new  dimensionless  dependent  variables  are 

u  =  ( Vgd)~lu,     v  =  (kVgd/d)-1  v,     w  =  ( Vgd)~l  w 

(2.4.8)  p  =  —  p, 

^  =  7^/d,     A  =  h/d. 
In  addition,  we  introduce  the  important  parameter 

(2.4.9)  a  = 


THE    TWO    BASIC    APPROXIMATE    THEORIES  29 

in  terms  of  which  all  quantities  will  be  developed;  when  this  parameter 
is  small  the  water  is  considered  to  be  shallow.  This  means,  of  course, 
that  d  is  small  compared  with  fc,  and  hence  that  the  x  and  z  coor- 
dinates (cf.  (2.4.7))  are  stretched  differently  from  the  y  coordinate 
and  in  a  fashion  which  depends  upon  the  development  parameter. 
Since  it  is  the  horizontal  coordinate  which  is  strongly  stretched 
relative  to  the  depth  coordinate,  it  seems  reasonable  to  refer  to  the 
resulting  theory  as  a  shallow  water  theory.  The  stretching  process 
combined  with  a  development  with  respect  to  a  is  the  characteristic 
feature  of  what  we  call  the  shallow  water  theory  throughout  this 
book.  The  dimensionless  development  parameter  a  has  a  physical 
significance,  of  course,  but  its  interpretation  will  vary  depending 
on  the  circumstances  in  individual  cases,  as  has  already  been  noted 
above.  For  example,  consider  a  problem  in  which  the  motion  is  to  be 
predicted  starting  from  rest  with  initial  elevation  y  —  770(#,  j/,  z) 
prescribed;  from  (2.4.8)  we  have 

y  ^  r)0  =--  dfjQ(x9  y,  z) 

i-  ix  y 

^'/I?0U'  "<*' 
from  which  we  obtain 


It  is  natural  to  assume  that  the  dimensionless  second  derivative 
tfficx  will  be  at  least  bounded  and  consequently  one  sees  that  the 
assumption  that  a  is  small  might  be  interpreted  in  this  case  as 
meaning  that  the  product  of  the  curvature  of  the  free  surface  of  the 
water  and  a  typical  depth  is  a  small  quantity. 

The  object  now  is  to  consider  a  sequence  of  problems  depending 
on  the  small  parameter  a  and  then  develop  in  powers  of  or.  Introduc- 
tion of  the  new  variables  in  the  equations  (2.4.1)  to  (2.4.6)  yields 

(2.4.1)'  aux  +  vy  +  awz  =  0, 

a[ut  +  uux  +  wuz  +  px]  +  vuy  =  0, 

(2.4.2)'  •  a[vt  +  uvx  +  wvz  +  py  +  1]  +  TO,  =  0, 

.  a[wt  +  uwx  +  wwz  +  pg]  +  vwy  =  0, 

(2.4.3)'  wy  =  VZ9  uz  =  wx,  vx  =  uy, 

(2.4.4)'  a[ijt  +  ur)x  +  wr]z]  =  v  at  y  =  77, 


80 


WATER   WAVES 


(2.4.5)' 
(2.4.6)' 


a[uhx 


p  =  0  at  y  = 
whz]  -j-t;  =  Oa 


=—  h, 


when  bars  over  all  quantities  are  dropped  and  r  is  replaced  by  t. 
The  next  step  is  to  assume  power  series  developments  for  u,  v, 
w,  rj,  and  p: 


(2.4.10) 


v  = 

w   = 


=  p(0) 


-f  .  .  ., 


4-  .  .  ., 


and  insert  them  in  the  equations  (2.4.1)'  to  (2.4.6)'  to  obtain,  by 
equating  coefficients  of  like  powers  of  a,  equations  for  the  successive 
coefficients  in  the  series,  which  are  of  course  functions  of  x,  y,  z, 
and  t.  The  terms  of  zero  order  yield  the  equations 


(2.4.1  )„ 

(2.4.2); 


T.(0)  n 

y  —      ' 

/  a«%<°>  =  o, 

7,(0)7,(°)  _  o 

\vvv  —  u» 


(2.4.3); 
(2-4.4); 
(2-4.5); 
(2.4.6); 


M(0)    _  ^(0)         (0)    _       (0) 

Uy  Wy.          ,  VX  Uy          , 


«v  == 

u(0>  =  0  at  y  =  T?(O), 
p(o)  =  o  at  y  =  ^(0), 
rj(o)  =  o  at  y  =  —  A. 
These  equations  yield  the  following: 

(2.4.11)  i><°>  =0, 

(2.4.12)  w«»  =  w<°>(ff,  z,  <)» 

(2.4.13)  u<°)  =  u(0)(o;,  2,  i), 

(2.4.14)  p<°>(0,  i^W),  a,*)  =  0, 

which  contain  the  important  results  that  the  vertical  velocity  com- 
ponent is  zero  and  the  horizontal  velocity  components  are  independent 
of  the  vertical  coordinate  y  in  lowest  order. 

The  first  order  terms  arising  from  (2.4.1)'  to  (2.4.6)'  in  their  turn 


THE    TWO   BASIC   APPROXIMATE   THEORIES  81 

yield  the  equations 

(2.4.1);  u«+»«=-*W, 

t  tt<0)  +  w(<%f  +  w(0)tt<0)  +  pf  =  0, 
(2.4.2);  p<,°>  +1=0, 

.  K>!0)  +  u^w(x°}  +  a;«»a><0)  +  p<°>  -  0, 
(2.4.4);          rjf  +  u^rif  +  w™r)W  =  »<«  at  y  =  »?«», 
(2.4.6)i  u^hx  +  ojtWA,  +  »<«  =  0  at  y  =  -  h, 

upon  making  use  of  (2.4.11),  (2.4.12),  and  (2.4.13).  Equation  (2.4.1  )J 
can  be  integrated  at  once  since  uw  and  w(0)  are  independent  of  y 
to  yield 

(2.4.15)  yd)  -  -  (««>  +  o,<°> )  z,  +  F(<c,  z,  t), 

with  F  an  arbitrary  function  which  can  be  determined  by  using 
(2.4.6);;  the  result  for  u(1>  is  then 

(2.4.16)  »<«  =  -  (M<°>  +  u,<0) )  y  -  [(«<»>/*),  +  (»«»)*).],._.  . 

To  second  order  the  vertical  component  of  the  velocity  is  thus  linear 
in  the  depth  coordinate.  In  similar  fashion  the  second  of  the  equations 
(2.4.2)j  can  be  integrated  and  the  additive  arbitrary  function  of 
x,  z,  t  determined  from  (2.4.14);  the  result  is 

(2.4.17)  p«»(0f  y,  z,  t)  -  ijW(x9  z,t)-y 

which  is  obviously  the  hydrostatic  pressure  relation  (in  dimension- 
less  form). 

In  the  derivation  of  the  shallow  water  theory  given  in  the  preceding 
section  this  relation  was  taken  as  the  starting  point;  here,  it  is 
derived  as  the  lowest  order  approximation  in  a  formal  perturbation 
scheme.  However,  it  is  of  course  not  true  that  we  have  proved  that 
(2.4.17)  is  in  some  sense  an  appropriate  assumption:  instead,  it 
should  be  admitted  frankly  that  our  dimensionless  variables  were 
introduced  in  just  such  a  way  that  (2.4.17)  would  result.  If  it  could 
be  shown  that  our  perturbation  procedure  really  does  yield  a  correct 
asymptotic  development  (that  the  development  converges  seems 
unlikely  since  the  equations  (2.4.1)'  to  (2.4.6)'  degenerate  in  order 
so  greatly  for  a  =  0)  then  the  hydrostatic  pressure  assumption  could 
be  considered  as  having  been  justified  mathematically.  A  proof  that 
this  is  the  case  would  be  of  great  interest,  since  it  would  give  a 
mathematical  justification  for  the  shallow  water  theory;  to  do  so  in 


82  WATER  WAVES 

a  general  way  would  seem  to  be  a  very  difficult  task,  but  Friedrichs 
and  Hyers  [F.18]  have  shown  that  the  development  does  yield 
the  existence  of  the  solution  in  the  important  special  case  of  the 
solitary  wave  (cf.  Chapter  10.9).  (Keller  [K.6]  had  shown  earlier 
that  the  formal  procedure  yields  the  solitary  wave.)  The  problem 
is  of  considerable  mathematical  interest  also  because  of  the  following 
intriguing  circumstance:  the  approximation  of  lowest  order  to  the 
solution  of  a  problem  in  potential  theory  is  sought  in  the  form  of  a 
solution  of  a  nonlinear  wave  equation,  and  this  means  that  the 
solution  of  a  problem  of  elliptic  type  is  approximated  (at  least  in  the 
lowest  order)  by  the  solution  of  a  problem  of  hyperbolic  type. 

The  values  of  v(l)  and  p(0)  given  by  (2.4.16)  and  (2.4.17)  are  now 
inserted  in  the  first  and  third  equations  of  (2.4.2)J  and  in  (2.4.4)!' 
to  yield  finally 

(2.4.18)  uf  +  tt<°>t*W  +  w^uf  +  rif  =  0, 

(2.4.19)  w><0)  +  i*<°>  H>W  +  w^wf  +  i?<°>  -  0, 

(2.4.20)  T?<°>  +  [ttWfoW  +  h)]x  +  [w<°>fo<o)  +  h)],  =  0, 

as  definitive  equations  for  t^0),  w(Q\  and  iy(0)— all  of  which,  we  repeat, 
depend  only  upon  #,  2,  and  /.  If  the  superscript  is  dropped,  w{Q)  is 
taken  to  be  zero,  and  it  is  assumed  that  all  quantities  are  independent 
of  z,  one  finds  readily  that  these  equations  become  identical  with 
equations  (2.2.11)  and  (2.2.12)  except  for  the  factor  g  in  (2.2.11) 
which  is  missing  here  because  of  our  introduction  of  a  dimensionless 
pressure. 

It  is  clear  that  the  above  process  can  be  continued  to  obtain  the 
higher  order  approximations.  An  example  of  such  a  calculation  will 
be  given  later  in  Chapter  10.9,  where  we  shall  see  that  the  first  non- 
trivial  term  in  the  development  which  yields  the  solitary  wave  is  of 
second  order. 


PART  II 


Summary 


In  Part  II  we  treat  a  variety  of  problems  in  terms  of  the  theory 
which  arises  through  linearization  of  the  free  surface  condition  (cf. 
the  preceding  chapter);  thus  the  problems  refer  to  waves  of  small 
amplitude.  To  this  theory  the  names  of  Cauchy  and  Poisson  are 
usually  attached.  The  material  falls  into  three  different  types,  or 
classes,  of  problems,  as  follows:  A)  Waves  that  are  simple  harmonic 
in  the  time.  These  problems  are  treated  in  Chapters  3,  4,  and  5  and 
they  include  a  study  of  the  classical  standing  and  progressing  wave 
solutions  in  water  of  uniform  depth,  and  waves  over  sloping  beaches 
and  past  obstacles  of  one  kind  or  another.  The  mathematical  tools 
employed  here  comprise,  aside  from  classical  methods  in  potential 
theory,  a  thorough-going  use  of  integrals  in  the  complex  domain. 
B)  Waves  created  by  disturbances  initiated  at  an  instant  when  the  water 
is  at  rest.  These  problems,  which  are  treated  in  Chapter  6,  comprise 
a  variety  of  unsteady  motions,  including  the  propagation  of  waves 
from  a  point  impulse  and  from  an  oscillatory  source.  Uniqueness 
theorems  for  the  unsteady  motions  are  derived.  The  principle  mathe- 
matical tool  used  in  solving  these  problems  is  the  Fourier  transform. 
The  method  of  stationary  phase  is  justified  and  used.  C)  Waves 
arising  from  obstacles  immersed  in  a  running  stream.  This  category 
of  problems  differs  from  the  first  two  in  that  the  motion  to  be  in- 
vestigated is  a  small  oscillation  in  the  neighborhood  of  a  uniform 
flow,  while  the  former  cases  concern  small  oscillations  near  the  state 
of  rest.  This  difference  is  in  one  respect  rather  significant  since  the 
problems  of  the  first  two  types  require  no  restriction  on  the  shape  of 
immersed  bodies,  or  obstacles,  while  the  third  type  of  problem 
requires  that  the  immersed  bodies  should  be  in  the  form  of  thin  disks, 
since  otherwise  the  flow  velocity  would  be  changed  by  a  finite  amount, 
and  a  linearization  of  the  free  surface  condition  would  not  then  be 
justified.  In  other  words,  the  problems  of  this  third  type  require 

35 


36  WATER   WAVES 

a  linearization  based  on  assuming  a  small  thickness  for  any  immersed 
bodies,  as  well  as  a  linearization  with  respect  to  the  amplitude  of  the 
surface  waves.  These  problems  are  treated  in  Chapters  7,  8,  and  9. 
The  classical  case  of  the  waves  created  by  a  small  obstacle  in  a  running 
stream  of  uniform  depth  is  first  treated.  This  includes  the  classical 
shipwave  problem,  discussed  in  Chapter  8,  in  which  the  "ship"  is 
treated  as  though  it  could  be  replaced  by  a  point  singularity.  A 
treatment  is  given  in  Chapter  9  of  the  problem  of  the  waves  created 
by  a  ship  moving  through  a  sea  of  arbitrary  waves,  assuming  the 
ship  to  be  a  floating  rigid  body  with  six  degrees  of  freedom  and  with 
its  motion  determined  by  the  propeller  thrust  and  the  pressure  of  the 
water  on  its  hull. 

Finally,  in  an  Appendix  to  Part  II  a  brief  summary  of  some  of 
the  more  recent  literature  concerned  with  the  above  types  of  problems 
is  given,  since  the  cases  selected  for  detailed  treatment  here  do  not 
by  any  means  exhaust  the  interesting  problems  which  have  been 
solved. 


SUBDIVISION  A 
WAVES  SIMPLE  HARMONIC  IN  THE  TIME 

CHAPTER  3 

Simple  Harmonic  Oscillations  in  Water 
of  Constant  Depth 


3.1.  Standing  waves 

In  Chapter  2  we  have  derived  the  basic  theory  of  irrotational 
waves  of  small  amplitude  with  the  following  results  (in  the  lowest 
order,  that  is).  Assuming  the  #,  2-plane  to  coincide  with  the  free 
surface  in  its  undisturbed  position,  with  the  t/-axis  positive  upward, 
the  velocity  potential  0(x,  y,  z;  t)  satisfies  the  following  conditions: 

(8.1.1  )  V2<Z>  =  0XX  +  0 


vy 


in  the  region  bounded  above  by  the  plane  y  =  0  and  elsewhere  by 
any  other  given  boundary  surfaces.  The  free  surface  condition  under 
the  assumption  of  zero  pressure  there  is 

(8.1.2)  0tt  +  g0y  =  0  for  y  =  0. 

The  condition  at  fixed  boundary  surfaces  is  that  d0/dn  =  0;  for 
water  of  uniform  depth  h  =  const,  we  have  therefore  the  condition 

(8.1.3)  0y  =  0  for  y  =  -  h. 

Once  the  velocity  potential  0  has  been  determined  the  elevation 
Y)(x,  z;t)  of  the  free  surface  is  given  by 

(3.1.4)  7?  =  --#,(*,  0,  *;  f). 

% 

Conditions  at  oo  as  well  as  appropriate  initial  conditions  at  t  =  0 
must  also  be  prescribed. 

In  this  section  we  are  interested  in  those  special  types  of  standing 
waves  which  are  simple  harmonic  in  the  time;  we  therefore  write 

37 


88  WATER    WAVES 

(3.1.5)  0(x,  y,  z;  t)  =  eM<p(x,  y,  z)  * 

with  <p  a  real  function,  and  with  the  understanding  that  either  the 
real  or  the  imaginary  part  of  the  right  hand  side  is  to  be  taken. 
The  problems  to  be  treated  here  thus  belong  to  the  theory  of  small 
oscillations  of  dynamical  systems  in  the  neighborhood  of  an  equilib- 
rium position. 

The  conditions  on  0  given  above  translate  into  the  following 
conditions  on  <p: 

(3.1.6)  V2<p  =  0,   —  h  <y  <Q9   —  oo  <  #,  *  <  oo, 

(3.1.7)  <p    -—<p  =  Q,  y  =  Q, 

S 

(3.1.8)  <py  =  0,         y=-h. 

As  conditions  at  oo  we  assume  that  y  and  (py  are  uniformly  bounded.** 
Arbitrary  initial  conditions  cannot  now  be  prescribed,  of  course, 
since  we  have  assumed  the  behavior  of  our  system  to  be  simple 
harmonic  in  the  time.  The  free  surface  elevation  is  given  by 

tcs 

(3.1.9)  ri  =  --  eM  .  <p(x,  0,  z). 

S 

We  look  first  for  standing  wave  motions  which  are  two-dimensional, 
so  that  (jp  depends  only  upon  x  and  y:  <p  —  y(x,  y),  and  also  consider 
first  the  case  of  water  of  infinite  depth,  i.e.  h  =-  oo.  One  verifies 
readily  that  the  functions 


(3.1.10) 

\  (p  =  emv  sin  mx 

are  harmonic  functions  which  satisfy  the  free  surface  condition 
(3.1.7)  provided  that  the  constant  m  satisfies  the  relation 

(3.1.11)  m  =  o*/g. 

In  addition,  the  conditions  at  oo  are  satisfied.  In  particular,  it  is  of 
interest  to  observe  that  the  oscillations  die  out  exponentially  in  the 
depth.  The  free  surface  elevation  is  then  given  by 

*  The  most  general  standing  wave  would  be  given  by  0  =  f(t)(p(x,  y,  z).  This 
means,  of  course,  that  the  shape  of  the  wave  in  space  is  fixed  within  a  multiplying 
factor  depending  only  on  the  time.  Thus  nodes,  maxima  and  minima,  etc.  occur 
at  the  same  points  independent  of  the  time. 

**  This  means  that  the  vertical  components  of  the  displacement  and  velocity 
are  bounded  at  oo.  One  could  prescribe  more  general  conditions  at  oo  without 
impairing  the  uniqueness  of  the  solutions  of  our  boundary  value  problems,  but 
it  does  not  seem  worth  while  to  do  so  in  this  case. 


SIMPLE    HARMONIC    OSCILLATIONS  39 


(3.1.12)  ^_^ 

g  \  sin  mx 

It  should  be  pointed  out  specifically  that  our  boundary  problem, 
though  it  is  linear  and  homogeneous,  has  in  addition  to  the  solution 
<p  ==  0  a  two-parameter  set  of  "non  trivial"  solutions  obtained  by 
taking  linear  combinations  of  the  two  solutions  given  in  (3.1.10). 
The  surface  waves  given  by  (3.1.10)  are  thus  simple  harmonic  in 
x  as  well  as  in  t.  The  relation  (3.1.11)  furthermore  states  the  very 
important  fact  that  the  wave  length  A  given  by 

(3.1.13)  A  ==  2n/m  =  <2ng/o2 

is  not  independent  of  the  frequency  of  the  oscillation,  but  varies 
inversely  as  its  square. 

The  above  discussion  yields  standing  wave  solutions  of  physically 
reasonable  type,  but  one  nevertheless  wonders  whether  there  might 
not  be  others—  for  example,  standing  waves  which  are  not  simple 
sine  or  cosine  functions  of  x,  but  rather  waves  with  amplitudes  which, 
for  example,  die  out  as  x  tends  to  infinity.  Such  waves  do  not  occur 
in  two  dimensions,*  however,  in  the  sense  that  all  solutions  for  water 
of  infinite  depth,  except  cp  =  0,  of  the  homogeneous  boundary  problem 
formulated  in  (3.1.6)  and  (3.1.7)  together  with  the  condition  that  <p 
and  (py  are  uniformly  bounded  at  oo  are  given  by  (3.1.10)  with  m 
satisfying  (3.1.11).  This  is  a  point  worth  pausing  to  prove,  especially 
since  the  method  of  proof  foreshadows  a  mode  of  attack  on  our 
problems  which  will  be  used  in  a  more  essential  way  later  on.  The 
first  step  in  the  uniqueness  proof  is  to  introduce  the  function  y(x,  y) 
defined  by 

(3.1.14)  \p  =  (jpy  —  my,         m  >  0. 

Since  q>  is  a  harmonic  function,  obviously  yj  is  also  a  harmonic  func- 
tion. In  addition,  y)  vanishes  for  y  =  0  on  account  of  its  definition 
and  (3.1.7).  Hence  \p  can  be  continued  by  reflection  over  the  iT-axis 
into  a  potential  function  which  is  regular  and  defined  as  a  single- 
valued  function  in  the  entire  x,  j/-plane.  Since  <p  and  q>y  were  assumed 
to  be  uniformly  bounded  in  the  entire  lower  half  plane  it  follows  that 
\p  is  bounded  in  the  entire  x,  i/-plane  since  reflection  in  the  #-axis 
does  not  destroy  boundedness  properties.  Thus  ^  is  a  potential  func- 
tion which  is  regular  and  bounded  in  the  entire  x,  t/-plane.  By  Liou- 

*  This  statement  is  not  valid  in  three  dimensions  as  we  shall  see  later  on  in 
this  section. 


40  WATER   WAVES 

ville's  theorem  it  is  therefore  a  constant,  and  since  y  =  0  for  y  =  0, 
the  constant  must  be  zero.  Hence  y  vanishes  identically.  From  (3.1.14) 
it  therefore  follows  that  any  solutions  (p(x,  y)  of  our  boundary 
value  problem  are  also  solutions  of  the  differential  equation 

(3.1.15)  <pv  —  nup  =  0,         —  oo  <  y  <  0. 

The  most  general  solution  of  this  differential  equation  is  given  by 

(3.1.16)  <p  =  c(x)emv 

with  c(x)  an  arbitrary  function  of  x  alone.  However,  <p(x,  y)  is  a 
harmonic  function  and  hence  c(x)  is  a  solution  of 


(8.1.17)  —   +  m*c  =  0 


which,  in  turn,  has  as  its  general  solution  the  linear  combinations 
of  sin  mx  and  cos  mx.  It  follows,  therefore,  that  the  standing  wave 
solutions  of  our  problem  are  indeed  all  of  the  form  Aemv  cos  (w#+oc),* 
with  a  and  A  arbitrary  constants  fixing  the  "phase"  and  the  amplitude 
of  the  wave,  while  m  is  a  fixed  constant  which  determines  the  wave 
length  A  in  terms  of  the  given  frequency  a  through  (3.1.13). 

In  water  of  uniform  finite  depth  h  it  is  also  quite  easy  to  obtain 
two-dimensional  standing  wave  solutions  of  our  boundary  value 
problem.  One  has,  corresponding  to  the  solutions  (3.1.10)  for  water  of 
infinite  depth,  the  harmonic  functions 

f  (p  =  cosh  m(y  +  h)  cos  mx, 

(d.l.lS)  4  . 

[<p  =  cosh  m(y  +  h)  sm  mx, 

as  solutions  which  satisfy  the  boundary  condition  at  the  bottom, 
while  the  free  surface  condition  is  satisfied  provided  that  the  con- 
stant m  satisfies  the  relation 

(3.1.19)  a2  =  gm  tanh  mh 

instead  of  the  relation  (3.1.11),  as  one  readily  sees.  Since  tanh  mh-+I 
as  h  ->  oo  it  is  clear  that  the  relation  (3.1.19)  yields  (3.1.11)  as  limit 
relation  for  water  of  infinite  depth.  The  uniqueness  of  the  solutions 
(3.1.18)  for  the  two-dimensional  case  under  the  condition  of  boun- 
dedness  at  oo  was  first  proved  by  A.  Weinstein  [W.7]  by  a  method 

*  It  can  now  be  seen  that  the  negative  sign  in  the  free  surface  condition  (3.1.7) 
is  decisive  for  our  results:  if  this  sign  were  reversed  one  would  find  that  the 
solution  g?  analogous  to  (8.1.16)  would  be  bounded  at  oo  only  for  c(x)  =  0, 
because  (3.1.16)  would  now  be  replaced  by  c(x)e~mv,  with  m  >  0. 


SIMPLE   HARMONIC   OSCILLATIONS  41 

different  from  the  method  used  above  for  water  of  infinite  depth 
which  can  not  be  employed  in  this  case  (see  [B.  12]). 

It  is  of  interest  to  calculate  the  motion  of  the  individual  water 
particles.  To  this  end  let  6x  and  dy  represent  the  displacements  from 
the  mean  position  (x,  y)  of  a  given  particle.  Our  basic  assumptions 
mean  that  dx9  dy  and  their  derivatives  are  small  quantities;  it  follows 
therefore  that  we  may  write 


—  =  u(xy  y)  =  0X  —  —  m  A  cos  at  cosh  m(y  +  h)  sin  mx 
dt 

ddy 

—-  =  v(x,  y)  =  0y  =  mA  cos  erf  sinh  m(y  +  A)  cos  mx 

within  the  accuracy  of  our  basic  approximation.  The  constant  A  is 
an  arbitrary  factor  fixing  the  amplitude  of  the  wave.  Hence  we 
have  upon  integration 

dx  =  —  sin  at  cosh  m(y  +  h)  sin  mx, 

(3.1.20) 

7??  /-i 

dy  = sin  at  sinh  m(y  +  /i)  cos  w#. 

a 

The  motion  of  each  particle  takes  place  in  a  straight  line  the  direction 
of  which  varies  from  vertical  under  the  wave  crests  (cos  mx  =  1)  to 
horizontal  under  the  nodes  (cos  mx  =  0).  The  motion  also  naturally 
becomes  purely  horizontal  on  approaching  the  bottom  y  =  —  h. 
These  consequences  of  the  theory  are  verified  in  practice,  as  indicated 
in  Fig.  3.1.1,  taken  from  a  paper  by  Ruellan  ami  Wallet  (cf.  [R.12]). 
The  photograph  at  the  bottom  makes  the  particle  trajectories  visible  in 
a  standing  wave;  this  is  the  final  specimen  in  a  series  of  photographs  of 
particle  trajectories  for  a  range  of  cases  beginning  with  a  pure  pro- 
gressing wave  (ef.  see.  3.2),  and  continuing  with  superpositions  of  pro- 
gressing waves  traveling  in  opposite  directions  and  having  the  same 
wave  length  but  not  the  same  amplitudes,  finally  ending  with  a 
standing  wave  when  the  wave  amplitudes  of  the  two  trains  are  equal. 
We  proceed  next  to  study  the  special  class  of  three-dimensional 
standing  waves  that  are  simple  harmonic  in  the  time,  arid  which 
depend  only  on  the  distance  r  from  the  t/-axis.  In  other  words,  we 
seek  standing  waves  having  cylindrical  symmetry.  Again  we  seek 
solutions  of  (3.1.6)  which  satisfy  (3.1.7).  Only  the  case  of  water  of 
infinite  depth  will  be  treated  here,  and  hence  (3.1.8)  is  replaced  by 


42 


WATER    WAVES 


Fig.  3.1.1.  Particle  trajectories  in  progressing  and  standing  waves 

the  condition  that  the  solutions  be  bounded  at  oo  in  the  negative 
//-direction  as  well  as  in  the  x-  and  ^-directions.  It  is  once  more  of 
interest  to  derive  all  possible  standing  wave  solutions  which  are 
everywhere  regular  and  bounded  at  oo  because  of  the  fact  that  the 
solutions  in  the  present  case  behave  quite  differently  from  those 
obtained  above  for  motions  that  are  independent  of  the  ^-coordinate. 


SIMPLE    HARMONIC    OSCILLATIONS  43 

In  particular,  we  shall  see  that  all  bounded  standing  waves  with 
cylindrical  symmetry  die  out  at  oo  like  the  inverse  square  root  of 
the  distance,  while  in  two  dimensions  we  have  seen  that  the  assump- 
tion that  the  wave  amplitude  dies  out  at  oo  leads  to  waves  of  zero 
amplitude  everywhere. 

It  is  natural  to  make  use  of  cylindrical  coordinates  in  deriving 
our  uniqueness  theorem.  Thus  we  write  (3.1.6)  in  the  form 

(3.1.21)  *  _  (r  -?^  +  -?  =  0,  0  ^  j/  >  -  oo,  0  ^  r  <  oo 
r  or  \    dr  J       dy2 

with  r  the  distance  from  the  j/-axis.  The  assumption  that  q>  depends 
only  upon  r  and  y  and  not  on  the  angle  6  has  already  been  used. 
For  our  purposes  it  is  useful  to  introduce  a  new  independent  variable 
Q  replacing  r  by  means  of  the  relation 

(3.1.22)  g  =  logr, 
in  terms  of  which  (3.1.21)  becomes 


3V        2a> 

(3.1.23)  <?-*  -Z  +  -Z  =  0,  y  <  0,   -  oo  <  0  <  oo. 

OQ*       dy2 

This  equation  holds,  we  observe,  in  the  half-plane  y  <  0  of  the 
j/,  g-plane.  The  boundary  condition  to  be  satisfied  at  y  =  0  is  (cf. 
(3.1.7)): 

(3.1.24)  cpy  —  mq)  =  0,  m  =  a2/g. 

We  wish  to  find  all  regular  solutions  of  (3.1.23)  satisfying  (3.1.24) 
for  which  99  and  <py  are  bounded  at  oo.  To  this  end  we  proceed  along 
much  the  same  lines  as  above  (cf.  (3.1.14)  and  the  reasoning  imme- 
diately following  it)  for  the  case  of  two  dimensions,  and  introduce 
the  function  y(g,  y)  by  the  identity 

(3.1.25)  y  =  <py  —  m<p,     2/<0,         —  oo  <  p  <  oo. 

Since  y;  involves  only  a  derivative  of  <p  with  respect  to  y  and  not 
with  respect  to  Q  it  follows  at  once  that  y  is  a  solution  of  (3.1.23). 
Since  if)  vanishes  at  y  =  0  from  (3.1.24)  it  follows  easily  that  it  can 
be  continued  analytically  into  the  upper  half-plane  y  >  0  by  setting 
y(£>  y)  =  —  y(£>  —y)  an(l  that  the  resulting  function  will  be  a 
solution  of  (3.1.23)  in  the  entire  p,  t/-plane.  The  function  \p  thus 
obtained  will  be  bounded  in  the  entire  plane,  since  it  was  bounded 
in  the  lower  half  plane  by  virtue  of  the  boundedness  assump- 
tions with  respect  to  <p.  A  theorem  of  S.  Bernstein  now  yields 


44  WATER   WAVES 

the  result  that  \p  is  everywhere  constant*  if  it  is  a  uniformly  bounded 
solution  of  (3.1.23)  in  the  entire  Q,  t/-plane.  Since  ip  vanishes  on  the 
t/-axis  it  follows  that  y  vanishes  identically.  Consequently  we  con- 
clude from  (3.1.25)  that  9?  satisfies  the  relation 

(3.1.26)  (py  —  m<p  =  0,         y  <  0. 

The  most  general  function  <p(g,  y)  satisfying  this  equation  is 

(3.1.27)  (p  =  emvf($)  =  emvf(log  r)  =  em^g(r) 

with  g(r)  an  arbitrary  function.  But  (p(r,  y)  is  also  a  solution  of 
(3.1.21)  and  hence  g(r)  is  a  solution  of  the  ordinary  differential 
equation 


(3.1.28)  -:?    I',! 

r  dr  \   dr 

or,  in  other  words,  g(r)  is  a  Bessel  function  of  ofrder  zero: 

(3.1.29)  g(r)  =  AJ0(mr)  +  BY,(mr). 

Since  we  restricted  ourselves  to  bounded  solutions  only  it  follows 
that  all  solutions  <p(r,  y)  of  our  problem  are  given  by 

(3.1.30)  g(r,  y)  =  Ae^J^(mr),         m  =  cr2/g, 

with  A  an  arbitrary  constant.  Upon  reintroduction  of  the  time 
factor  we  have,  therefore,  as  the  only  bounded  velocity  potentials 
the  functions 

(3.1.31)  0(r,  y;  t)  =  AeiatemvJQ(mr). 

As  is  well  known,  these  functions  behave  for  large  values  of  r  as 
follows: 

(3.1.32)  0(r,  y;  t)  ~  Aeiaie™*  •  I/—-  cos  [mr  —  -\ 

'  nmr         \  4/ 

and  thus  they  die  out  like  l/Vr>  as  stated  above. 

In  two  dimensions  we  were  able  to  find  bounded  standing  waves 
of  arbitrary  phase  (in  the  space  variable)  at  oo.  In  the  present  case 
of  circular  waves  we  have  found  bounded  waves  with  only  one  phase 
at  oo.  However,  if  we  were  to  permit  a  logarithmic  singularity  at  the 

*  What  is  needed  is  evidently  a  generalization  of  Liouville's  theorem  to  the 
elliptic  equation  (3.1.23)  which  has  a  variable  coefficient.  The  theorem  of  Bern- 
stein referred  to  is  much  more  general  than  is  required  for  this  special  purpose, 
but  it  is  also  not  entirely  easy  to  prove  (cf.  E.  Hopf  [H.17]  for  a  proof  of  it). 


SIMPLE    HARMONIC    OSCILLATIONS  45 

axis  r  =  0  and  thus  admit  the  singular  Bessel  function  YQ(mr)  as 
a  solution  of  (3.1.28),  we  would  have  as  possible  velocity  potentials 
the  functions 

(3.1.33)  0(r,  y;  t)  =  Beiatem^Y0(mr) 
which  behave  for  large  r  as  follows: 

(3.1.34)  0(r,  y;  t)  ~  Beiate™v  ]/—  sin  (mr  -  -\. 

*  nmr         \  4/ 

Admitting  solutions  with  a  logarithmic  singularity  on  the  i/-axis 
thus  leads  to  standing  waves  which  behave  at  oo  in  the  same  way 
as  those  which  are  everywhere  bounded,  except  that  they  differ  by 
90°  in  phase  at  oo.  Thus  waves  having  an  arbitrary  phase  at  oo  can 
be  constructed,  but  not  without  allowing  a  singularity.  It  has,  however 
not  been  shown  that  (3.1.31)  and  (3.1.33)  yield  all  solutions  with  this 
property. 

3.2.  Simple  harmonic  progressing  waves 

Since  our  boundary  problem  is  linear  and  homogeneous  we  can 
reintroduce  the  time  factors  cos  at  and  sin  at  and  take  appropriate 
linear  combinations  of  the  standing  waves  (3.1.5)  to  obtain  simple 
harmonic  progressing  wave  solutions  in  water  of  uniform  depth  of 
the  form 

(3.2.1)  0  =  A  cosh  m(y  +  h)  cos  (mx  ±  at  +  a) 
with  m  and  a  satisfying 

(3.2.2)  a2  =  gm  tanh  mh, 
as  before. 

The  wave,  or  phase  9  speed  c  is  of  course  given  by 

(3.2.3)  c  =  a/m, 

or,  in  terms  of  the  wave  length  A  —  2n/m  by 


(3.2.3),  c  =  tanh  *!*. 

'    %7l  / 

It  is  useful  to  write  the  relation  (3.2.2)  in  terms  of  the  wave  length 
X  =  2n/m  and  then  expand  the  function  tanh  mh  in  a  power  series 
to  obtain 


46  WATER   WAVES 

We  see  therefore  that 

(3.2.5)  a2  ->  (y)2^  =  m*&h  as  \  -*°* 
and  hence  that 

(3.2.6)  c^=L  Vgh  if  A/A  is  small. 

This  last  relation  embodies  the  important  fact  that  the  wave  speed 
becomes  independent  of  the  wave  length  when  the  depth  is  small  compared 
with  the  wave  length,  but  varies  as  the  square  root  of  the  depth.  This 
fact  is  in  accord  with  what  resulted  in  Chapter  2  upon  linearizing 
the  shallow  water  theory  (cf.  (2.2.16))  and  the  sentence  immediately 

following),  which  led  to  the  linear  wave  equation  and  to  c  =  Vgh 
as  the  propagation  speed  for  disturbances.  We  can  gain  at  least  a 
rough  idea  of  the  limits  of  accuracy  of  the  linear  shallow  water 
theory  by  comparing  the  values  of  c  given  by  c2  =  gh  with  those 
given  by  the  exact  formula 


(3.2.7)  c*  =        tanh 

2n  A 

for  various  values  of  the  ratio  A/A.  One  finds  that  c  as  given  by 
Vgh  is  in  error  by  about  6  %  if  the  wave  length  is  ten  times  the 
depth  and  by  less  than  2  %  if  the  wave  length  is  twenty  times  the 
depth.  The  error  of  course  increases  or  decreases  with  increase  or 
decrease  in  A/A. 

In  water  of  infinite  depth,  on  the  other  hand,  we  have  already 
observed  (cf.   (3.2.2))  that 

(3.2.8)  c2  =  gA/2jr. 

Actually,  the  error  in  c  as  computed  by  the  formula  c2  =  gA/2rc  is 
already  less  than  1/2  %  if  A/A  >  |.  One  might  therefore  feel  justified 
in  concluding  that  variations  in  the  bottom  elevation  will  have  but 
slight  effect  on  a  progressing  wave  provided  that  they  do  not  result 
in  depths  which  are  less  than  half  of  the  wave  length,  and  observations 
seem  to  bear  this  out.  In  other  words,  the  wave  would  not  "feel"  the 
bottom  until  the  depth  becomes  less  than  about  half  a  wave  length. 
It  is  of  interest  to  determine  the  paths  of  the  individual  water 
particles  as  the  result  of  the  passage  of  a  progressing  wave.  As  in  the 
preceding  section  we  take  dx  and  dy  to  represent  the  displacements 
of  a  particle  from  its  average  position,  and  determine  those  dis- 
placements from  the  equations 


SIMPLE    HARMONIC    OSCILLATIONS  47 

—  =  0X  =  —  Am  cosh  m(y  +  h)  sin  (mx  -\-at-\-  a), 
dt 

ddy 

—  -.-.-  0y  =  Am  sinh  m(y  +  ^)  cos  (mx  +  at  +  a), 
(zr 

since  0  is  given  by  (3.2.1)  in  the  present  case.  Integration  of  these 
equations  yields 


(3.2.9) 


dx  =  -  —  cosh  m(y  +  /i)  cos  (rna:  +  at  +  a), 


dy  —  ---  sinh  m(y  +  ^)  sin  (m#  +  (rf  +  a), 
a 


so  that  the  path  of  a  particle  at  depth  y  is  an  ellipse 

dx2    ,fy2_l 
a?        ~¥  ~ 

with  semi-axes  a  and  6  given  by 


. 

a  =  -  cosh  m(y  +  h) 
a 

b  =  —  sinh  m(y  +  h). 
a 

On  the  bottom,  y  —  —  h,  the  ellipse  degenerates  into  a  horizontal 
straight  line,  as  one  would  expect.  Both  axes  of  the  ellipse  shorten 
with  increase  in  the  depth.  For  experimental  verification  of  these 
results,  the  discussion  with  reference  to  Fig.  3.1.1  should  be  con- 
sulted. In  water  of  infinite  depth  the  particle  paths  would  be  circles, 
as  one  can  readily  verify.  The  fact  that  the  displacement  of  the 
particles  dies  out  exponentially  in  the  depth  explains  why  a  submarine 
need  only  submerge  a  slight  distance  below  the  surface—  a  half  wave 
length,  say  —  in  order  to  remain  practically  unaffected  even  by  severe 
storms. 


3.3.    Energy    transmission    for    simple    harmonic    waves   of  small 
amplitude 

In  Chapter  1  the  general  formulas  for  the  energy  E  stored  in  a 
fluid  and  its  flux  or  rate  of  transfer  F  across  given  surfaces  were 
derived  for  the  most  general  types  of  motion.  In  this  section 


48  WATER   WAVES 

we  apply  these  formulas  to  the  special  motions  considered  in  the 
present  chapter,  that  is,  under  the  assumption  that  the  free  surface 
conditions  are  linearized.  The  formula  for  the  energy  E  stored  in  a 
region  R  is  (cf.  (1.6.1)): 


(3.3.1)  E  =  e          [J(<^2  +  0;  +  <*>*2)  +  gy]dxdydz; 

R 

while  the  flux  of  energy  F  in  a  time  T  across  a  surface  SG  fixed  in 
space  is  given  by  (cf.  (1.6.5)): 

(8.3.2)  F- 

We  suppose  first  that  the  motion  considered  is  the  superposition 
of  two  standing  waves  which  are  simple  harmonic  in  the  time,  as 
follows: 

(3.3.3)  <P(x,  y,  z;  t)  =  9^(0?,  y,  z)  cos  at  +  99,  (#,  j/,  z)  sin  at. 

Insertion  of  this  in  (3.3.2)  with  T  =  2jc/a9  i.e.  for  a  time  interval 
equal  to  the  period  of  the  oscillation,  leads  at  once  to  the  following 
expression  for  the  energy  flux  F  through  SG: 


One  observes  that  the  energy  flux  over  a  period  is  zero  if  either  (pl 
or  9?2  vanishes,  i.e.  if  the  motion  is  a  standing  wave:  a  fact  which 
is  not  surprising  since  one  expects  an  actual  transport  of  energy 
only  if  the  motion  has  the  character  of  a  progressing  wave.  Still 
another  fact  can  be  verified  from  (3.3.4)  in  our  present  cases,  in  which 
(pl  and  (p2  are,  as  we  know,  harmonic  functions:  if  SG  is  a  fixed  closed 
surface  in  the  fluid  enclosing  a  region  R  Green's  formula  states  that 


provided  that  9^  and  <p2  have  no  singularities—  sources  or  sinks  for 
example—  in  R.  In  this  case  the  energy  flux  F  clearly  vanishes  since 
9?!  and  <p2  are  harmonic.  Also  one  sees  by  a  similar  reasoning  that  the 
flux  F  over  a  period  remains  constant  if  SG  is  deformed  without 
passing  over  singularities.  In  particular,  the  energy  flux  through  a 
vertical  plane  passing  from  the  bottom  to  the  free  surface  of  the  water 


SIMPLE    HARMONIC   OSCILLATIONS  49 

in  a  two-dimensional  motion  would  be  the  same  (per  unit  width  of 
the  plane)  for  all  positions  of  the  plane  provided  that  no  singularities 
are  passed  over.  This  fact  makes  it  possible,  if  one  wishes,  to  con- 
sider the  energy  in  the  fluid  as  though  the  energy  itself  were  an 
incompressible  fluid,  and  to  speak  of  its  rate  of  flow. 

In  the  literature  dealing  with  waves  in  all  sorts  of  media,  but 
particularly  in  dispersive  media,  it  is  indeed  commonly  the  custom 
to  introduce  the  notion  of  the  velocity  of  the  flow  of  energy  ac- 
companying a  progressing  wave,  and  to  bring  this  velocity  in  relation 
to  the  kinematic  notion  of  the  group  velocity  (to  be  discussed  in 
the  next  section).  The  author  has  found  it  difficult  to  reconcile  him- 
self to  these  discussions,  and  feels  that  it  would  be  better  to  discard 
the  difficult  concept  of  the  velocity  of  transmission  of  energy,  since 
this  notion  is  not  of  primary  importance,  and  nothing  can  be  ac- 
complished by  its  use  which  cannot  be  done  just  as  well  by  using 
the  well-founded  and  clear-cut  concept  of  the  flux  of  energy  through 
a  given  surface.  On  the  other  hand,  the  notion  is  used  in  the  literature 
(and  probably  will  continue  to  be  used)  and  consequently  a  dis- 
cussion of  it  is  included  here,  following  pretty  much  the  derivation 
given  by  Rayleigh  in  an  appendix  to  the  first  volume  of  his  Sound 
[R.4],  In  the  next  section,  where  the  notion  of  group  velocity  is 
introduced,  some  further  comments  about  the  concept  of  the  velocity 
of  transmission  of  energy  will  be  made. 

We  consider  the  energy  flux  per  unit  breadth  across  a  vertical 
plane  in  the  case  of  a  simple  harmonic  progressing  wave  in  water 
of  uniform  depth  (or,  in  view  of  the  above  remarks,  across  any  surface 
of  unit  breadth  extending  from  the  bottom  to  the  free  surface).  The 
velocity  potential  0  is  given  by  (cf.  (3.2.1)) 

(3.3.5)  0  =  A  cosh  m(y  +  h)  cos  (mx  +  at  +  a) 
and  (3.3.2)  yields 

(3.3.6)  F  =  A*Q<jm  J*+2*/<J  ft*    Cosh2  m(y  +  h)dy\  sin2  (mx  +  at)  dt 

for  the  flux  across  a  strip  of  unit  breadth  in  the  time  T  =  2jt/a9 
the  period  of  the  oscillation.  Hence  the  average  flux  per  unit  time 
is  given  by 

(*  q  7^  w          F 

(3.3.7)  tav  =  - 


since  the  average  of  sin2  6  over  a  period  is  1/2.  We  have  also  taken 


50  WATER   WAVES 

r)  =  0  in  the  upper  limit  of  the  integral  in  (3.3.6)  and  thus  neglected 
a  term  of  higher  order  in  the  amplitude.  It  is  useful  to  rewrite  the 
formula  (3.8.7)  in  the  following  form  through  use  of  the  relations 
or*  =  gm  tanh  mh  and  c  =  a/m: 


(8.8.8)  Fav  =  —si-  cosh*  mh  •  U, 

O 

with  U  a  quantity  having  the  dimensions  of  a  velocity  and  given 
by  the  relation 

(3.3.9)  U 

Next  we  calculate  the  average  energy  stored  in  the  water  as  a  result 
of  the  wave  motion  with  respect  to  the  length  in  the  direction  of 
propagation  of  the  wave.  This  is  obtained  from  (3.3.1)  by  calculating 
first  the  energy  JEj  over  a  wave  length  A  =  2n/m  at  any  arbitrary 
fixed  time.  In  the  present  case  we  have 

EK  —  E0  =  m2Q  r  JA  [$A2  sinh2  m(y  +  h)  cos2  (mx  +  at  +  a) 

(3.3.10)  +  £  A2  cosh2  m(y  +  h)  sin2  (mx  +  at  +  a)]  dxdy 

+ 


in  which  the  constant  EQ  refers  to  the  potential  energy  of  the  water 
of  depth  h  when  at  rest.  On  evaluating  the  integrals,  and  ignoring 
certain  terms  of  higher  order,  we  obtain  for  the  energy  between  two 
planes  a  wave  length  apart  arising  from  the  passage  of  the  wave  the 
expression 


(3.3.11  )  EI  -  E0  =       -  A  cosh2  nth, 

2g 

as  one  finds  without  difficulty.  Thus  the  average  energy  Eav  in  the 
fluid  per  unit  length  in  the  ^-direction  which  results  from  the  motion 
is  given  by 


(3.3.12)  Eav  =    -       cosh2  mh. 

2g 

Upon  comparison  with  equation  (3.3.8)  we  observe  that  Eav  is 
exactly  the  coefficient  of  U  in  the  formula  (3.3.8)  for  the  average 
energy  flux  per  unit  time  across  a  vertical  plane.  It  therefore  follows, 
assuming  that  no  energy  is  created  or  destroyed  within  the  fluid 


SIMPLE   HARMONIC   OSCILLATIONS  51 

itself,  that  the  energy  is  transmitted  in  the  direction  of  propagation  of 
the  wave  on  the  average  with  the  velocity  U.  As  we  see  from  (3.3.9) 
the  velocity  U  is  not  the  same  as  the  phase  or  propagation  velocity  c; 
in  fact,  U  is  always  less  than  c:  for  water  of  infinite  depth  it  has  the 
value  c/2  and  it  increases  with  decrease  in  depth,  approaching  the 
phase  velocity  c  as  the  depth  approaches  zero. 

3.4.  Group  velocity.  Dispersion 

In  any  body  of  water  the  motion  of  the  water  in  general  consists 
of  a  superposition  of  waves  of  various  amplitudes  and  wave  lengths. 
For  example,  the  motion  of  the  water  due  to  a  disturbance  over  a 
restricted  area  of  the  surface  can  be  analyzed  in  terms  of  the  super- 
position of  infinitely  many  simple  harmonic  wave  trains  of  varying 
amplitude  and  wave  length;  such  an  analysis  will  in  fact  be  carried 
out  in  Chapter  6.  However,  we  know  from  our  previous  discussion 
(cf.  (3.2.7))  that  the  propagation  speed  of  a  train  of  waves  is  an 
increasing  function  of  the  wave  length—  in  other  words,  the  wave 
phenomena  with  which  we  are  concerned  arc  subject  to  dispersion— 
and  thus  one  might  expect  that  the  waves  would  be  sorted  out  as 
time  goes  on  into  various  groups  of  waves  such  that  each  group 
would  consist  of  waves  having  about  the  same  wave  length.  We  wish 
to  study  the  properties  of  such  groups  of  waves  having  approximately 
the  same  wave  length. 

Suppose,  for  example,  that  the  motion  can  be  described  by  the 
superposition  of  two  progressing  waves  given  by 


(341) 

=  A  sin  ([m  +  dm]x  —  [a  +  da]t) 

with  dm  and  da  considered  to  be  small  quantities.  The  superposition 
of  the  two  wave  trains  yields 

0  =  2A  cos  -  (xdm  —  tda)  sin  I    m  +  —  \x  —  \a  -\  --  \t\ 
(3.4.2)  2V  \L  2  J  L          2j    / 

=  B  sin  (m'x  —  o't) 

with  m'  =  m  +  dm/2,  a'  =  a  +  da/2.  Since  dm  and  da  are  small  it 
follows  that  the  function  B  varies  slowly  in  both  x  and  t  so  that  <P 
is  an  amplitude-modulated  sine  curve  at  each  instant  of  time,  as 
indicated  schematically  in  Figure  3.4.1.  In  addition,  the  "groups" 
of  waves  thus  defined—  in  other  words  the  configuration  represented 


52 


WATER   WAVES 


by  the  dashed  curves  of  Figure  8.4.1. —advance  with  the  velocity 
da/dm  in  the  0-direction.  In  our  problem  a  will  in  general  be  a  function 


Fig.  3.4.1.  Wave  groups 

of  m  so  that  the  velocity  U  of  the  group  is  given  approximately 
by  da/dm9  or,  in  terms  of  the  wave  length  A  =  27t/m  and  wave  velocity 
c  =  a/m,  by 

d(mc)  .  dc 

dm  dk 


(3.4.3) 


U  = 


The  matter  can  also  be  approached  in  the  following  way  (cf. 
Sommerfeld  [S.13]),  which  comes  closer  to  the  more  usual  cir- 
cumstances. Instead  of  considering  the  superposition  of  only  two 
progressing  waves,  consider  rather  the  superposition,  by  means  of 
an  integral,  of  infinitely  many  waves  with  amplitudes  and  wave 
lengths  which  vary  over  a  small  range: 

(3.4.4)  0  =  r*+*A(m)  exp  (i(mx  -  at)}  dm. 

J  WQ— e 

The  quantity  mx  —  at  can  be  written  in  the  form 

(3.4.5)  mx  —  at  =  m#c  —  aQt  +  (m  —  m0)x  —  (a  —  aQ)t. 
From  (3.4.4)  one  then  finds 

(3.4.6)  0  =  C  exp  {i(m<p  —  a0t)}> 

in  which  the  amplitude  factor  C  is  given  by 

(8.4.7)  C  =  r*+*A(m)  exp{i[(m  -  mQ)x  -  (a  -  a0)*]}  dm. 

J  WQ— « 

We  are  interested  here  in  seeking  out  those  places  and  times  (if  any) 
where  the  function  C  represents  a  wave  progressing  with  little  change 
in  form,  since  (3.4.6)  will  then  furnish  what  we  call  a  group  of  waves. 
Since  x  and  t  occur  only  in  the  exponential  term  in  (8.4.7),  it  follows 
'that  the  values  of  interest  are  those  for  which  this  term  must  be 
nearly  constant,  i.e.  those  for  which  (m  —  w0)#  —  (a  —  a0)t~  const. 


SIMPLE   HARMONIC   OSCILLATIONS  58 

It  follows  that  the  propagation  speed  of  such  a  group  is  given  by 
(a  —  (70)/(ra  —  ra0),  and  if  (m  —  w0)  is  small  enough  we  obtain  again 
the  formula  (3.4.3). 

Evidently,  it  is  important  for  this  discussion  of  the  notion  of  group 
velocity  that  the  motion  considered  should  consist  of  a  superposition 
of  waves  differing  only  slightly  in  frequency  and  amplitude.  In 
practice,  the  motions  obtained  in  most  cases  —through  use  of  the 
Fourier  integral  technique,  for  example,—  are  the  result  of  super- 
position of  waves  whose  frequencies  vary  from  zero  to  infinity  and 
whose  amplitudes  also  vary  widely.  However,  as  we  shall  see  in 
Chapter  6,  it  happens  very  frequently  that  the  motion  at  certain 
places  and  times  is  approximated  with  good  accuracy  by  integrals 
of  the  type  given  in  (3.4.4)  with  e  arbitrarily  small.  (This  is,  indeed, 
the  sense  of  the  principle  of  stationary  phase,  to  be  treated  in  Chap- 
ter 6.)  In  such  cases,  then,  groups  of  waves  do  exist  and  the  dis- 
cussion above  is  pertinent. 

In  our  problems  the  relation  between  wave  speed  and  wave  length 
is  given  by  (3.2.2)  and  consequently  the  velocity  U  of  a  group  is 
readily  found,  from  (3.4.3),  to  be 


2 

We  observe  that  the  group  velocity  has  the  same  value  as  was  given 
in  the  preceding  section  for  the  average  rate  of  propagation  of  energy 
in  a  uniform  train  of  waves  having  the  same  wave  length  as  those 
of  the  group.  In  other  words,  the  rate  at  which  energy  is  propagated 
is  given  by  the  group  velocity  and  not  the  phase  velocity.  This  is 
often  considered  as  the  salient  fact  with  respect  to  the  notion  of 
group  velocity.  As  indicated  already  in  the  preceding  section,  the 
author  does  not  share  this  view,  but  feels  rather  that  the  kinematic 
concept  of  group  velocity  is  of  primary  significance,  while  the  notion 
of  velocity  of  propagation  of  energy  might  better  be  discarded.  It 
is  true  that  the  two  velocities,  in  spite  of  the  fact  that  one  is  derived 
from  dynamics  while  the  other  is  of  purely  kinematic  origin,  turn 
out  to  be  the  same—  not  only  in  this  case,  but  in  many  others  as 
well*—  but  it  is  also  true  that  they  are  not  always  the  same—  for 
example,  the  two  velocities  are  not  the  same  if  there  is  dissipation 
of  energy  in  the  medium.  In  addition,  we  have  seen  in  the  preceding 
section  that  the  notion  of  velocity  of  energy  can  be  derived  when  no 

A  general  analysis  of  the  reason  for  this  has  been  given  by  Broer   [B.I  8]. 


54  WATER   WAVES 

wave  group  exists  at  all —we  in  fact  derived  this  velocity  for  the  case 
of  a  wave  having  but  one  harmonic  component. 

In  Chapter  6  we  shall  have  occasion  to  see  how  illuminating  the 
kinematic  concept  of  a  group  and  its  velocity  can  be  in  interpreting 
and  understanding  the  complicated  unsteady  wave  motions  which 
arise  when  local  disturbances  propagate  into  still  water. 


CHAPTER  4 


Waves  Maintained  by  Simple  Harmonic  Surface  Pressure 
in  Water  of  Uniform  Depth*  Foreed  Oscillations 


4.1.   Introduction 

In  our  previous  discussions  we  have  considered  always  that  the 
pressure  at  the  free  surface  was  constant  (usually  zero)  in  both  space 
and  time.  In  other  words,  only  the  free  oscillations  were  treated  and 
the  problems  were,  correspondingly,  linear  and  homogeneous  boun- 
ary  value  problems.  Here  we  wish  to  consider  two  problems  in  which 
the  surface  pressure  p0  is  simple  harmonic  in  the  time  and  the  resulting 
motions  are  thus  forced  oscillations;  the  problems  then  also  have  a 
nonhomogeneous  boundary  condition.  In  the  first  such  problem  we 
assume  that  the  motion  is  two-dimensional  and  that  the  surface  pres- 
sure is  a  periodic  function  of  the  space  coordinate  x  over  the  entire 
#-axis;  in  the  second  problem  the  surface  pressure  is  assumed  to  be 
zero  except  over  a  segment  of  finite  length  of  the  #-axis.  In  these 
problems  the  depth  of  the  water  is  assumed  to  be  everywhere  infinite, 
but  the  corresponding  problems  in  water  of  constant  finite  depth 
can  be,  and  have  been,  solved  by  much  the  same  methods. 

The  formulation  of  the  first  two  problems  is  as  follows.  A  velocity 
potential  0(x,  y;t)  is  to  be  determined  which  is  simple  harmonic  in 
the  time  t  and  satisfies 

(4.1.1)  V20  =  0         for  y  <  0. 
The  surface  pressure  p(x;  t)  is  given  by 

(4.1.2)  p(x;  t)  =  p(x)  sin  at, 

and  the  boundary  conditions  at  the  free  surface  are  the  dynamical 
condition  (cf.  (2.1.20)!) 

(4.1.3)  r]=  -—&t 

& 

55 


50  WATER   WAVES 

and  the  kinematic  condition 

(4.1.4)  rit  =  ®y. 

The  last  condition  means  that  no  kinematic  constraint  is  imposed 
on  the  surface— it  can  deform  freely  subject  to  the  given  pressure 
distribution.  In  addition,  we  require  that  <Pt  and  0y  should  be 
uniformly  bounded  at  oo.  This  means  effectively  that  the  vertical 
displacement  and  vertical  velocity  components  are  bounded.  In 
section  4.3,  the  amplitude  p(x)  of  the  surface  pressure  p  will  have 
discontinuities  at  two  points  and  we  shall  impose  appropriate  con- 
ditions on  0  at  these  points  when  we  consider  this  case. 

We  seek  the  most  general  simple  harmonic  solutions  of  our  problem; 
they  have  the  form 

(4.1.5)  0  =  <p(x,  y)  cos  at  +  y(x,  y)  sin  at. 

The  functions  <p  and  y;  are  of  course  harmonic  in  the  lower  half 
plane.  The  conditions  (4.1.2),  (4.1.3),  and  (4.1.4)  arc  easily  seen  to 
yield  for  the  function  <p  the  boundary  condition 

a  __ 

(4.1.6)  (py  —  m<p  =  —  —  p(x)         for  y  —  0 

Ci  O 

with  the  constant  m  defined  by 

(4.1.7)  m  =  a2/g; 
while  for  y  they  yield  the  condition 

(4.1.8)  \py  —  my  =  0         for  y  =  0. 

The  phase  sin  at  assumed  for  p  in  (4.1.2)  has  the  effect  that  ^  satisfies 
the  homogeneous  free  surface  condition,  as  one  sees. 

We  know  from  the  first  section  of  the  preceding  chapter  that  the 
only  bounded  and  regular  harmonic  functions  y  which  satisfy  the 
condition  (4.1.8)  are  given  by 

f  cos  mx 

(4.1.9)  V(%,y) 


\« 

1- 


sin  mx 


In  the  next  two  sections  we  shall  determine  the  function  (p(x,  y), 
.  i.e.    that   part   of  0   which   has   the   phase   cos  at,    in   accordance 
with  two  different  choices  for  the  amplitude  p(x)  of  the  surface 
pressure  p. 


SIMPLE    HARMONIC    SURFACE    PRESSURE  57 

4.2.  The  surface  pressure  is  periodic  for  all  values  of  x 

We  consider  now  the  case  in  which  the  surface  pressure  is  periodic 
in  x  such  that  p(x)  in  (4.1.2)  and  (4.1.6)  is  given  by 

(4.2.1)  p(x)  =  PsinAtf,          —  oo  <x  <  oo. 
One  verifies  at  once  that  the  following  function  (p(x,  y): 

aP       e*v 

(4.2.2)  (p(x,  y)  = sin  fa 

Qg     m  —  A 

is  a  harmonic  function  which  satisfies  the  free  surface  boundary 
condition  (4.1.0)  imposed  in  the  present  case.  Since  the  difference  # 
of  two  solutions  q>l9  q>2  both  satisfying  all  of  our  conditions  would 
satisfy  the  homogeneous  boundary  condition  %v  —  m%  —  0,  it  follows 
that  all  solutions  <p  of  our  boundary  value  problem  can  be  obtained  by 
adding  to  the  special  solution  given  by  (4.2.2)  any  solution  of  the 
homogeneous  problem,  and  these  latter  solutions  are  the  functions 
y  given  by  (4.1.9)  since  jj  satisfies  the  same  conditions  as  y.  Therefore 
the  most  general  simple  harmonic  solutions  of  the  type  (4.1.5)  are 
given  in  the  present  case  by 

VaP     eXv  fcos0M?]l 

(4.2.3)  0(<r,  ?/;  t)  - sin  fa  +  Ae™*  \    .  \\  cos  at 

[_Qg  m  —  A  [sin  mx  jj 

{cos  mx ] 
\  sin  at, 
sin  mx  J 

with  A  and  B  constants  which  are  at  our  disposal.  In  other  words, 
the  resulting  motions  are,  as  usual  in  linear  vibrating  systems,  a 
linear  combination  of  the  forced  oscillation  and  the  free  oscillations. 
These  solutions— without  the  uniqueness  proof— seem  to  have  been 
given  first  by  Lamb  [L.2]. 

We  observe  that  the  case  A  —  m  must  be  excluded,  and  that  if  A 
is  near  to  m  large  amplitudes  of  the  surface  waves  arc  to  be  expected. 
This  means  physically,  as  one  sees  immediately,  that  waves  of  large 
amplitude  are  created  if  the  periodic  surface  pressure  distribution 
has  nearly  the  wave  length  which  belongs  to  a  surface  wave  of  the 
same  frequency  for  pressure  zero  at  the  surface— that  is,  the  wave 
length  of  the  corresponding  free  oscillation. 

If  instead  of  (4.1.2)  we  take  the  surface  pressure  as  a  progressing 
wave  of  the  form 

(4.2.4)  p(x\  t)  =  H  sin  (at  —  fa) 


58  WATER   WAVES 

it  is  readily  found  that  progressing  surface  waves  result  which  are 
given  by 

Ha     e^ 

(4.2.5)  0(x,  y;t)  =  --  -  cos  (at  -  fa). 

Qg  m  —  /. 

To  this  one  may,  of  course,  add  any  of  the  wave  solutions  which 
occur  under  zero  surface  pressure.  Again  one  observes  an  odd  kind 
of  "resonance"  phenomenon:  large  amplitudes  are  conditioned  by 
the  wave  length  in  space  of  the  applied  pressure  once  the  frequency 
has  been  fixed. 

4.3.  The  variable  surface  pressure  is  confined  to  a  segment  of  the 
surface 

In  this  section  we  consider  the  case  in  which  the  surface  pressure  p 

I  x  \  <  a 


{P  sin  at. 


with  P  a  constant.  Some  of  the  motions  which  can  arise  under  such 
circumstances  are  discussed  by  Lamb  [L.2]  in  the  paper  quoted  above. 
However,  here  as  elsewhere,  Lamb  assumes  fictitious  damping 
forces*  in  order  to  be  rid  of  the  free  oscillations  and  thus  achieve  a 
unique  solution,  and  he  also  makes  use  of  the  Fourier  integral  tech- 
nique which  we  prefer  to  replace  by  a  different  procedure.  In  fact, 
the  present  problem  is  a  key  problem  for  this  Part  II  and  a  peg  upon 
which  a  variety  of  observations  important  for  other  discussions  in  later 
chapters  will  be  hung.  As  we  shall  see,  the  present  problem  is  also 
decidedly  interesting  for  its  own  sake,  although  Lamb  strangely 
enough  made  no  attempt  in  his  paper  to  point  out  the  really  striking 
results. 

In  addition  to  prescribing  the  pressure  p  through  (4.3.1)  it  is 
necessary  to  add  to  the  conditions  imposed  in  section  4.1  appropriate 
conditions  at  the  points  (it  #,  0)  where  p  has  discontinuities.  In 
view  of  (4.1.3)  it  is  clear  that  a  finite  discontinuity  in  0t  or  r\  or 
both  must  be  admitted  and  it  seems  also  likely  that  the  derivatives 
0X  and  0y  of  0  would  be  unbounded  near  these  points.  We  shall  make 

*  Lamb  assumes  resistances  which  are  proportional  to  the  velocity.  In  this 
way  the  irrotational  character  of  the  flow  is  preserved,  but  it  is  difficult  to  see 
how  such  resistances  can  be  justified  mechanically.  It  would  seem  preferable 
to  secure  the  uniqueness  of  the  solution  in  unbounded  domains  by  imposing 
physically  reasonable  conditions  on  the  behavior  of  the  waves  at  infinity. 


SIMPLE    HARMONIC    SURFACE    PRESSURE  59 

the  following  requirements 

(4.3.2)  0t  bounded;  0y  =  O(Q-I^£),         e  >  0 


in  a  neighborhood  of  the  points  (^  a,  0)  with  Q  the  distance  from 
these  points.  This  means,  in  particular,  that  the  surface  elevation  is 
bounded  near  these  points  and  that  the  singularity  admitted  is  not 
as  strong  as  that  of  a  source  or  sink.  We  recall  that  0t  and  0y  were 
required  to  be  uniformly  bounded  at  oo. 

The  stipulations  made  so  far  do  not  ensure  the  uniqueness  of  the 
solution  0  of  our  problem  any  more  than  similar  conditions  ensured 
uniqueness  of  the  solution  of  the  problem  treated  in  the  preceding 
section.  However,  we  have  in  mind  now  a  physical  situation  in  which 
we  expect  the  solution  to  be  unique:  We  imagine  the  motion  resulting 
from  the  applied  surface  pressure  p  given  by  (4.3.1)  to  be  the  limit 
approached  after  a  long  time  subsequent  to  the  application  of  p  to 
the  water  when  initially  at  rest.  Under  these  circumstances  one  feels 
instinctively  that  the  motion  of  the  water  far  away  from  the  source 
of  the  disturbance  should  have  the  character  of  a  progressing  wave 
moving  away  from  the  source  of  the  disturbance,  since  at  no  time 
is  there  any  reason  why  waves  should  initiate  at  infinity.  (We  shall 
show  (of.  (6.7))  that  the  motion  of  the  water  arising  from  such  initial 
conditions  actually  does  approach,  as  the  time  increases  without 
limit,  the  motion  to  be  obtained  here.)  Consequently  we  add  to  our 
conditions  on  0  the  condition— often  called  the  Sommerfeld  condition 
in  problems  concerning  electromagnetic  wave  propagation— that  the 
zvares  should  behave  at  oo  like  progressing  waves  moving  away  from 
the  source  of  the  disturbance.  As  we  shall  see,  this  qualitative  condition 
leads  to  a  unique  solution  of  our  problem. 

In  solving  our  problem  there  are  some  advantages  to  be  gained  by 
not  stipulating  at  the  outset  that  the  Sommerfeld  condition  should 
be  satisfied,  but  to  obtain  first  all  possible  solutions  of  the  form 
(4.1.5),  and  only  afterwards  impose  the  condition.  We  have  therefore 
to  find  the  harmonic  functions  (p  which  satisfy  the  condition  (cf. 
(4.1.6)  and  (4.3.1)) 

(  c,        |  x  \  ^  a 

(4.3.3)  <p    -  my  =  \  ,        y  =  0 
rv                   (0,        |  x  |  >  a 

with 

Pa 

(4.3.4)  ra  =  a2/g,         c  =  —  — 

68 


60  WATER   WAVES 

on  the  free  surface,  and  the  boundedness  conditions  which  follow 
from  those  imposed  on  0: 

{cp  and  cpv  bounded  at  oo, 
i  I  V  7 

<p  bounded  and  <py  =  0(e~"1+e),      e  >  0,     at  x  =  ±  a. 

The  functions  y  in  (4.1.5),  i.e.  those  which  yield  the  waves  of  phase 
sin  at  in  <Z>,  satisfy  the  same  conditions  as  in  section  4.1  and  are 
therefore  given  by  (4.1.9).  We  have  therefore  only  to  determine  the 
functions  <p.  To  this  end  it  is  convenient  to  introduce  new  dimen- 
sionless  quantities 

(4.3.6)  xl  =  mx,         yl  =  my,         a±  =  ma 

together  with  c^  =  c/m  so  that  the  free  surface  condition  (4.3.3) 
takes  the  form 

(  cl  ,     |  xl  \  ^  «! 

(4.3.7)  V     _  v  =  ,       ft  -  0. 


In  what  follows  we  use  the  condition  in  this  form  but  drop  the  sub- 
scripts for  the  sake  of  convenience. 

In  most  of  the  two-dimensional  problems  treated  in  the  remainder 
of  Part  II  we  make  use  of  the  fact  that  any  harmonic  function 
(p(x,  y)  can  be  taken  as  the  real  part  of  an  analytic  function  f(z)  of 
the  complex  variable  z  =  x  -\-  iy  and  write 

(4.3.8)  f(z)  =  <p(x,  y)  +  iy(x,  y)  =  f(x  +  iy). 

In  our  present  problem  f(z)  is  defined  and  analytic  in  the  lower  half 
plane.  To  express  the  surface  condition  (4.3.7)  in  terms  of  f(z)  we 
write 


id-~%  -  /)  =  die  (if  -  f), 


in  which  the  symbol  8&e  means  that  the  real  part  of  what  follows  is 
to  be  taken.  Consequently  the  free  surface  condition  has  the  form: 

(4.3.9)       V.-V 


SIMPLE    HARMONIC   SURFACE    PRESSURE  61 

We  now  introduce  a  new  analytic  function  F(z)  by  the  equation* 
(4.3.10)  F(z)  =  if'(z)  -  /(«) 

and  seek  to  determine  F(z)  uniquely  through  the  conditions  imposed 
on  9?  —  Ste  J(z).  We  observe  to  begin  with  that  F(z)  satisfies  the 
condition 

f  r          \  T  I   <.  n 
c9        |  x  i  ^  a 


in  view  of  (4.3.9).  We  show  now  that  F(z)  is  uniquely  determined 
within  an  additive  pure  imaginary  constant,  as  follows:  Suppose 
that  G(z)  -  Fi(z)  —  F2(z)  is  the  difference  of  two  functions  F(z) 
satisfying  the  conditions  resulting  from  (4.3.10)  through  those  on 
f(z).  Then  die  G(z)  would  vanish  on  the  entire  real  axis,  except 
possibly  at  x  --  i  a,  as  one  sees  from  (4.3.11).  Hence  3&eG(z)  is  a 
potential  function  which  can  be  continued  analytically  by  reflection 
over  the  real  axis  into  the  entire  upper  half  plane;  it  will  then  be 
defined  and  single-valued  in  the  whole  plane  except  for  the  points 
(i  fl,  0).  At  oo,  8&e  G(z)  is  bounded  in  the  lower  half  plane,  while 
£Jle  G(z)  =  0(g~~1+e),  e  >  0,  at  x  =  ±  a  in  view  of  the  regularity 
conditions  and  the  definition  of  G(z).  These  boundedness  properties 
are  evidently  preserved  in  the  analytic  continuation  into  the  upper 
half  plane.  Consequently  Sfce  G(z)  has  a  removable  singularity  at  the 
points  x  —  i  a  on  the  real  axis  since  the  singularity  is  weaker  than 
a  pole  of  first  order  and  the  function  is  single-valued  in  the  neigh- 
borhood of  these  points.  Thus  3te  G(z)  is  a  potential  function  which 
is  regular  and  bounded  in  the  entire  plane,  and  is  zero  on  the  real 
axis;  by  Liouville's  theorem  it  is  therefore  zero  everywhere.  Con- 
sequently the  analytic  function  G(z)  is  a  pure  imaginary  constant, 
and  the  result  we  want  is  obtained.  On  the  other  hand  it  is  rather 
easy  to  find  a  function  F(z)  which  has  the  prescribed  properties— for 
example  by  first  finding  its  real  part  from  (4.3.11)  through  use  of 
the  Poisson  integral  formula.  We  simply  give  it: 

ic        z  —  a 
(4.3.12)  F(z)  =  _logr:_; 

one  verifies  readily  that  it  has  all  of  the  required  properties.  We 
take  that  branch  of  the  logarithm  which  is  real  for  (z  —-  a)/(z  +  a) 

*  This  device  has  been  used  by  Kotschin  [K.I 4],  and  it  was  exploited  by  Lewy 
[L.8]  and  the  author  [S.18]  in  studying  waves  on  sloping  beaches. 


62 


WATER   WAVES 


real  and  positive. 

Ojice  F(z)  has  been  uniquely  determined,  the  complex  velocity 
potential  f(z)  is  restricted  to  the  solutions  of  the  first  order  ordinary 
differential  equation  (4.3.10),  which  means  that  the  solutions  depend 
only  on  the  arbitrary  constant  which  multiplies  the  non-vanishing 
solution  e~iz  of  the  homogeneous  equation  if'(z)  —  f  =  0.  But 
9te  (A  +  iB)e~iz  =  ev(A  cos  x  +  B  sin  x)  and  these  are  the  standing 
wave  solutions  for  the  case  of  surface  pressure  p  =  0.  The  most 
general  solution  of  (4.3.10),  with  F(z)  given  by  (4.3.12),  can  be 
written,  as  one  can  readily  verify,  in  the  form 


(4.3.13) 


c         Cz  t  —  a 

/(*)=-*-«      e«log— —  dt, 
n        J  zn  t  +  a 


with  the  initial  point  *0  and  the  path  of  integration  any  arbitrary 
path  in  the  slit  plane.  Changing  #0  obviously  would  have  the  effect 
of  changing  the  additive  solution  of  the  homogeneous  equation.  It 
is  convenient  to  replace  (4.3.13)  by  the  following  expression,  obtained 
through  an  integration  by  parts: 

z  -  a  C*  /I  1 


(4.3.14)  /(*) 
v  y/v; 


ci  f 
=  —    - 

rc  L 


log 

fo 


/I 
(  -- 
\t-a 


t+a 


and  at  the  same  time  to  fix  the  path  of  integration  as  indicated  in 


t-  plane 


(a)  (b) 

Fig.  4.3.1a,b.    Path  of  integration  in  f-plane 

Figure  4.3.1.  This  path  comes  from  oo  along  the  positive  imaginary 
axis  and  encircles  the  origin,  leaving  it  and  the  point  (-—a,  0)  to 


SIMPLE    HARMONIC    SURFACE    PRESSURE 


63 


the  left.  Use  has  been  made  of  the  fact  that  log  (z— -a)/(*+a)  ->0 
when  z  ->  oo;  we  observe  also  that  the  integrals  converge  on  account 
of  the  exponential  factor. 

That  95(3?,  y)  =  &te  f(z)  as  given  through  (4.3.14)  satisfies  the 
boundary  conditions  imposed  at  the  free  surface  and  the  regularity 
condition  at  the  points  (i  a,  0)  is  easy  to  verify.  We  proceed  to 
discuss  the  behavior  of  f(z)  at  oo  (always  for  z  in  the  lower  half  plane). 
For  this  purpose  it  suffices  to  discuss  the  integrals 

Pii 


tOO 


t±  a 


dt  since  the  function  log  I  — 


behaves  like  l/z  at  oo  (as  one  readily  sees).  To  this  end  we  integrate 
once  by  parts  to  obtain 

i  Cz      ei(t~z} 

J(fy\      *        I fit 

2(z)  —  —  "r—  —  *  ,.   ,    f,\2atm 

£»  i  fl          J  £00  \*  i"  #) 

We  suppose  that  the  curved  part  of  the  path  of  integration  in  Figure 
4. 3. la  is  an  arc  of  a  circle.  It  follows  at  once  that  the  complex  number 
t  —  z  has  a  positive  imaginary  part  on  the  path  of  integration  as 
long  as  the  real  part  of  z  is  negative,  and  hence  we  have 


1 

±  # 


f 

J  foo 


dt 


\t±a\ 


f 

J    0 


dt 


\t±a\ 


Consequently  I(z)  behaves  like  l/z  at  infinity  when  the  real  part  of 
z  is  negative,  and  f(z)  likewise.  The  situation  is  different,  however, 
if  the  real  part  of  z  is  positive.  To  study  this  case,  we  add  and  subtract 
circular  arcs,  as  indicated  in  Figure  4.3.1b,  in  order  to  have  an 
integral  over  the  entire  circle  enclosing  the  singularities  at  ±  a 
as  well  as  over  a  path  symmetrical  to  the  path  in  Figure  4.3. la. 
By  the  same  argument  as  above,  the  contribution  of  the  integral 
over  the  latter  path  behaves  like  1/2  at  oo,  and  hence  the  non- 
vanishing  contribution  arises  as  a  sum  of  residues  at  the  points  ±  #• 
These  contributions  to  I(z)  are  at  once  seen  to  have  the  values 
2nie^ize^ia.  Thus  we  may  describe  the  behavior  of  f(z)  as  given 
by  (4.3.14)  at  oo  as  follows: 


(4.3.15)      /(*)  = 


'(7) 


for  Ste  z  <  0 


—  4ci  (sin  a)  e~iz  +  O  I  —  I  for  die  z  >  0. 


WATER   WAVES 


From  (4.3.10)  and  (4.3.12)  one  sees  that  f(z)  has  the  same  behavior 
at  oo  as  f(z)9  except  for  a  factor  —  i.  Hence  f(z)9  and  with  it 
(jp(x,  y)  =  3te  /(*),  has  the  postulated  behavior  at  oo.  It  is  convenient 
to  write  down  explicitly  the  behavior  of  <p(x,  y)  at  oo: 


for  x  <  0, 

(4.3.16)  <p(x9  y)  =  9te  f(z)  =  }      V  r  '  . 

4c  sin  a  ev  sin  x  +  O  I  —  1   for  x  >  0. 
\rj 

It  follows  that  all  simple  harmonic  solutions  of  our  problem  arc 
given  by  linear  combinations  of 

(4.3.17)  0(x9  y;  t)  =  (die  f(z)  +  Ae*  sin  x  +  Be*  cos  x)  cos  at 
and 

(4.3.18)  <P(x9  y\  t)  =  (Cey  sin  x  +  Dev  cos  x)  sin  at 

in  which  A,  B,  C,  and  1)  are  arbitrary  constants,  and  f(z)  is  given 
by  (4.3.14).  In  other  words,  the  standing  waves  <p(x9  y)  cos  at  just 
found  above,  together  with  the  standing  waves  which  exist  for 
vanishing  free  surface  pressure,  constitute  all  possible  standing  waves. 
We  now  impose  the  condition  that  the  wave  0(x9  y\  t)  we  seek 
behaves  like  an  outgoing  progressing  wave  at  oo,  i.e.  that  it  behaves 
like 

S_:     ev(H  sin  (x  +  at)  +  K  cos  (x  +  erf))   at  x  =  —  oo 
and  like 

S+:     ev(L  sin  (x  —  at)  +  M  cos  (x  —  at))   at  x  =  +  oo. 

In  view  of  the  behavior  of  (p(x,  y)  =  <%ef(z)  at  x  =  —  oo  (of.  (4.3.16)), 
i.e.  the  fact  that  it  dies  out  there,  it  is  clear  that  we  may  combine 
the  standing  wave  solutions  (4.3.17)  and  (4.3.18)  in  such  a  way  as 
to  obtain  a  progressing  wave  solution 

(4.3.19)  0(x9  y;  t)  =  e*(H  sin  (x  +  at)  +  K  cos  (x  +  at)) 

+  <p(x9  y)  cos  at 

valid  everywhere  and  which  satisfies  the  condition  £_,  with  the  two 
constants  H  and  K  still  arbitrary.  At  x  =  +  oo  this  wave  has  the 
behavior 

(4.3.20)  &(x,  y;  t)  =  e*[(H  sin  (x  +  at)  +  K  cos  (x  +  at)) 

—  4c  sin  a  sin  x  cos  at]  +  O  1  —  1 


SIMPLE   HARMONIC   SURFACE    PRESSURE  65 

in  view  of  (4.3.16).  In  order  that  S+  should  hold  for  this  solution 
for  all  t  one  sees  readily  that  the  constants  //  and  K  must  satisfy 
the  linear  equations 

L  —  H  —  4e  sin  a 
(4.3.21  )  L  =  -  H 

\M  =  K,  M  =  -  K, 
from  which  we  conclude  that 

f  L  —  —  -  2c  sin  a,      //  =  2c  sin  a 

<«•»>  {*-*_.. 

Thus  the  solution  is  now  uniquely  determined  through  imposition 
of  the  Sommerfcld  condition,  and  can  be  expressed  as  follows: 


(4.3.23)     0(x,  y;  t)  ~  ----  sin  maemv  sin  (mx  ~ot)  +  O\—\,    x  >  0 


/  1  \ 
O\—\,    x 

\  r  / 


upon  rcintroduction  of  the  original  variables  and  parameters  (cf. 
(4.3.6)),  with  0(l/r)  representing  a  function  which  dies  out  at 
infinity  like  1/r.  The  function  0  of  course  yields  a  wave  with  sym- 
metrical properties  with  respect  to  the  i/-axis.  We  observe  that 
the  wave  length  A  =  'Infm  of  these  waves  at  oo  is  the  same  as  that  of 
free  oscillations  of  the  same  frequency,  as  one  would  expect. 

The  most  striking  thing  about  the  solution  is  the  fact  that  for 
certain  frequencies  and  certain  lengths  of  the  segment  over  which 
the  periodic  pressure  differs  from  zero,  the  amplitude  of  the  progressing 
wave  is  zero  at  oo;  this  occurs  obviously  for  sin  ma  =  0,  i.e.  for 
ma  —  kn,  k  —  -  1,  2,  3,  .  .  ..  Since  m  —  2n/h  with  A  the  wave  length 
of  a  free  oscillation  of  frequency  a,  it  follows  that  the  amplitude  of 
the  progressing  wave  at  oo  vanishes  when 

(4.3.24)  2a  -  frA,         k  =  1,  2  ____  , 

i.e.  when  the  length  of  the  segment  on  which  the  pressure  is  applied 
is  an  integral  multiple  of  the  wave  length  of  the  free  oscillation  having 
the  same  frequency  as  the  periodic  pressure.  This  does  not  of  course 
mean  that  the  entire  disturbance  vanishes,  but  only  that  the  motion 
in  this  case  is  a  standing  wave  given  by 

(4.3.25)  &(x,  y;  t)  =  y(x,  y)  cos  at, 

since  the  quantities  //  and  K  in  (4.3.19)  are  now  both  zero.  Since 
<p  now  behaves  like  1/r  at  both  infinities,  the  amplitude  of  the  standing 


66  WATER  WAVES 

wave  tends  to  zero  at  infinity.  A  wave  generating  device  based  on  the 
physical  situation  considered  here  would  thus  be  ineffective  at  certain 
frequencies.  It  is  clear  that  no  energy  is  carried  off  to  infinity  in 
this  case,  and  hence  that  the  surface  pressure  p  on  the  segment 
—  a  5*  x  ^  +  a  can  do  no  net  work  on  the  water  on  the  average. 
Since  r/t  =  0V  it  follows  that  the  rate  at  which  work  is  done  by  the 
pressure  p  (per  unit  width  at  right  angles  to  the  #,  t/-plane)  is 

rpq>y  cos  atdx,  and  since  p  has  the  phase  sin  at  it  is  indeed  clear 
•a 

that  the  average  rate  of  doing  work  is  zero  in  this  case. 

There  is  a  limit  case  of  the  present  problem  which  has  considerable 
interest  for  us.  It  is  the  limit  case  in  which  the  length  of  the  segment 
over  which  p  is  applied  shrinks  to  zero  while  the  amplitude  P  of  p 
increases  without  limit  in  such  a  way  that  the  product  2aP  approaches 
a  finite  limit.  In  this  way  we  obtain  the  solution  for  an  oscillating 
pressure  point.  One  sees  easily  that  the  function  f(z)  given  by  (4.3.13), 
which  yields  the  forced  oscillation  in  our  problem,  takes  the  following 
form  in  the  limit: 

(4.3.26)  i(z)  =  ^e-«^dt, 

with  C  the  real  constant  2aPa/gg.  At  oo  this  function  behaves  as 
follows 

0  (— )  for  die  z  <  0, 

(4.3.27)  /(«)  =  '      V*7 

2Ci  e~iz  +  0  ( —  I  for  S&e  z  >  0. 

In  this  limit  case  of  an  oscillating  pressure  point  we  see  that  there  are 
no  exceptional  frequencies:  application  of  the  external  force  always 
leads  to  transmission  of  energy  through  progressing  waves  at  oo. 
The  singularity  of  f(z)  at  the  origin  is  clearly  a  logarithmic  singularity 
since  f(z)  behaves  near  the  origin  like 

C          C*dt 

(4.3.28)  /(*)  =  -*-"      _+.... 

n         J     t 

We  see  that  a  logarithmic  singularity  is  appropriate  at  a  source  or 
sink  of  energy  when  the  motion  is  periodic  in  the  time. 


SIMPLE    HARMONIC    SURFACE    PRESSURE  67 

4.4.  Periodic  progressing  waves  against  a  vertical  clift 

With  the  aid  of  the  complex  velocity  potential  defined  by  (4.3.13) 
we  can  discuss  a  problem  which  is  different  from  the  one  treated  in 
the  preceding  section.  The  problem  in  question  is  that  of  the  deter- 
mination of  two-dimensional  progressing  waves  moving  toward  a 
vertical  cliff,  as  indicated  in  Figure  4.4.1.  The  cliff  is  the  vertical 


Fig.  4.4.1.  Waves  against  a  vertical  cliff 

plane  containing  the  //-axis.  As  in  the  preceding  section,  we  assume 
also  that  a  periodic  pressure  (cf.  (4.3.1))  is  applied  over  the  segment 
0  <j|  x  ^  a  at  the  free  surface.  To  solve  the  problem  we  need  only 
combine  the  standing  waves  given  by  (4.3.17)  and  (4.3.18)  in  such 
a  way  as  to  obtain  progressing  waves  which  move  inward  from  the 
two  infinities,  and  this  can  be  done  in  the  same  way  as  in  section  4.3. 
The  result  will  be  again  a  wave  symmetrical  with  respect  to  the 
t/-axis,  and  hence  one  for  which  0X  =  0  along  the  j/-axis;  thus  such 
a  wave  satisfies  the  boundary  condition  appropriate  to  the  vertical 
cliff.  We  would  find  for  the  velocity  potential  0  the  expression,  valid 
for  x  >  0: 

2Pa  / 1  \ 

(4.4.1 )      0(x,  ij;  t)  = sin  maem^[sm  (nix  +  at)]  +  0  I  — I 

to  »      ' 

with  0(l/r)  a  function  behaving  like  1/r  at  oo  but  with  a  singularity 
at  (a,  0).  In  order  to  obtain  a  system  of  waves  which  are  not  reflected 
back  to  oo  by  the  vertical  cliff  it  was  necessary  to  employ  a  mechanism 
—the  oscillating  pressure  over  the  segment  0  ^  x  ^  a  on  the  free 


68  WATER   WAVES 

surface— which  absorbs  the  energy  brought  toward  shore  by  the  in 
coming  wave.  However,  the  particular  mechanism  chosen  here,  i.e. 
one  involving  an  oscillatory  pressure  having  the  same  frequency  as 
the  wave,  will  not  always  serve  the  purpose  since  the  amplitude  A 
of  the  surface  elevation  of  the  progressing  wave  at  oo  is  given,  from 

(4.4.1)  and  (4.1.4),  by 

2P 

(4.4.2)  A  =  —  sin  ma. 

68 

Thus  the  ratio  of  the  pressure  amplitude  P  applied  on  the  water 
surface  near  shore  to  the  amplitude  of  the  wave  at  oo  would  obviously 
become  oo  when  sin  ma  =  0.  In  other  words,  such  a  mechanism  would 
achieve  its  purpose  for  waves  whose  wave  length  A  at  oo  satisfies 
the  relation  a  —  k  A/2,  with  k  any  integer,  only  if  infinite  pressure 
fluctuations  at  the  shore  occur.  Presumably  this  should  be  interpreted 
as  meaning  that  for  these  wave  lengths  the  mechanism  at  shore  is 
not  capable  of  absorbing  all  of  the  incoming  energy,  or  in  other  words, 
some  reflection  back  to  oo  would  occur.  This  remark  has  a  certain 
practical  aspect:  a  device  to  obtain  power  from  waves  coming  toward 
a  shore  based  on  the  mechanism  considered  here  would  function 
differently  at  different  wave  lengths. 

It  is  of  interest  in  the  present  connection  to  consider  the  same  limit 
case  as  was  discussed  at  the  end  of  the  preceding  section,  in  which 
the  segment  of  length  a  shrinks  to  zero  while  Pa  remains  finite.  In 
this  case  no  exceptional  wave  lengths  or  frequencies  occur.  However, 
the  limit  complex  potential  now  has  a  logarithmic  singularity  at  the 
shore  line,  as  we  noticed  in  the  preceding  section,  and  the  amplitude 
of  the  surface  would  therefore  also  be  infinite  at  the  shore  line.  What 
would  really  happen,  of  course,  is  that  the  waves  would  break  along 
the  shore  line  if  no  reflection  of  wave  energy  back  to  oo  occurred, 
and  the  infinite  amplitude  obtained  with  our  theory  represents  the 
best  approximation  to  such  an  essentially  nonlinear  phenomenon 
that  the  linear  theory  can  furnish. 

This  limit  case  represents  the  simplest  special  case  of  the  problem 
of  progressing  waves  over  uniformly  sloping  beaches  which  will  be 
treated  more  generally  in  the  next  chapter.  However,  the  present 
case  has  furnished  one  important  insight:  a  singularity  of  the  complex 
velocity  potential  is  to  be  expected  at  the  shore  line  if  the  condition 
at  oo  forbids  reflection  of  the  waves  back  to  oo,  and  the  singularity 
should  be  at  least  logarithmic  in  character. 


CHAPTER  5 


Waves  on  Sloping  Beaches  and  Past  Obstacles 

5.1.  Introduction  and  summary 

Perhaps  the  most  striking— and  perhaps  also  the  most  fascinating- 
single  occurrence  among  all  water  wave  phenomena  encountered  in 
nature  is  the  breaking  of  ocean  waves  on  a  gently  sloping  beach. 
The  purpose  of  the  present  chapter  is  to  analyze  mathematically  the 
behavior  of  progressing  waves  over  a  uniformly  sloping  beach  insofar 
as  that  is  possible  within  the  accuracy  of  the  linearized  theory  for 
waves  of  small  amplitude;  that  is,  within  the  accuracy  of  the  theory 
with  which  we  are  concerned  in  the  present  Part  II.  Later,  in  Chapter 
10.10,  we  shall  discuss  the  breaking  of  waves  from  the  point  of  view 
of  the  nonlinear  shallow  water  theory. 

To  begin  with,  it  is  well  to  recall  the  main  features  of  what  is  often 
observed  on  almost  any  ocean  beach  in  not  too  stormy  weather. 
Some  distance  out  from  the  shore  line  a  train  of  nearly  uniform 
progressing  waves  exists  having  wave  lengths  of  the  order  of  say 
fifty  to  several  hundred  feet.  These  waves  can  be  considered  as  simple 
sine  or  cosine  waves  of  small  amplitude.  As  the  waves  move  toward 
shore,  the  line  of  the  wave  crests  and  troughs  becomes  more  and 
more  nearly  parallel  to  the  shore  line  (no  matter  whether  this  was 
the  case  in  deep  water  or  not),  and  the  distance  between  successive 
wave  crests  shortens  slightly.  At  the  same  time  the  height  of  the 
waves  increases  somewhat  and  their  shape  deviates  more  and  more 
from  that  given  by  a  sine  or  cosine— in  fact  the  water  in  the  vicinity 
of  the  crests  tends  to  steepen  and  in  the  troughs  to  flatten  out  until 
finally  the  front  of  the  wave  becomes  nearly  vertical  and  eventually 
the  water  curls  over  at  the  crest  and  the  wave  breaks.  These  ob- 
servations are  all  clearly  borne  out  in  Figures  5.1.1,  5.1.2,  which 
are  photographs,  given  to  the  author  by  Walter  Munk  of  the  Scripps 
Institution  of  Oceanography,  of  waves  on  actual  beaches.  At  the 
same  time,  it  should  be  stated  here  that  the  breaking  of  waves  also 
occurs  in  a  manner  different  from  this— a  fact  which  will  be  discussed 

69 


5.1,1.  Waves  on,  a 


Fig.  5.1.2.  Breaking  and  diffraction  of  waves  at  an  inlet 


WAVES    ON    SLOPING    BEACHKS    AND    PAST   OBSTACLES  71 

in  Chapter  10.10  on  the  basis  of  other  photographs  of  actual  waves 
and  a  nonlinear  treatment  of  the  problem. 

It  is  clear  that  the  linear  theory  we  apply  here  can  not  in  principle 
yield  large  departures  from  the  sine  or  cosine  form  of  the  waves  in 
deep  water,  and  still  less  can  it  yield  the  actual  breaking  phenomena: 
obviously  these  are  nonlinear  in  character.  On  the  other  hand  the 
linear  theory  is  to  be  applied  and  should  yield  a  good  approximation 
for  deep  water  and  for  the  intermediate  zone  between  deep  water  and 
the  actual  surf  region.  However,  the  fact  that  breakers  do  in  general 
occur  in  nature  cannot  by  any  means  be  neglected  even  in  formulating 
the  problems  in  terms  of  the  linear  theory,  for  the  following  reasons. 
Suppose  we  consider  a  train  of  progressing  waves  coming  from  deep 
water  in  toward  shore.  As  we  know  from  Chapter  3,  such  a  train  of 
waves  is  accompanied  by  a  flow  of  energy  in  the  direction  toward  the 
shore.  If  we  assume  that  there  is  little  or  no  reflection  of  the  waves 
from  the  shore— which  observations  show  to  be  largely  the  case  for  a 
gently  sloping  beach*  —it  follows  that  there  must  exist  some  mecha- 
nism which  absorbs  the  incoming  energy;  and  that  mechanism  is  of 
course  the  breaking  of  the  waves  which  converts  the  incoming  wave 
energy  partially  into  heat  through  turbulence  and  partially  into  the 
energy  of  a  different  type  of  flow,  i.e.  the  undertow.  In  terms  of  the 
linear  theory  about  the  only  expedient  which  we  have  at  our  disposal 
to  take  account  of  such  an  effect  in  a  rough  general  way  is  to  permit 
that  the  wave  amplitude  may  become  very  large  at  the  shore  line,  or, 
in  mathematical  terms,  that  the  velocity  potential  should  be  per- 
mitted to  have  an  appropriate  singularity  at  the  shore  line.  As  we 
have  already  hinted  at  the  end  of  the  preceding  chapter,  the  ap- 
propriate singularity  for  a  two-dimensional  motion  seems  to  be 
logarithmic,  and  hence  the  wave  amplitude  would  be  logarithmically 
infinite  at  the  shore  line.  Indeed,  it  turns  out  that  no  progressing 
wave  solutions  without  reflection  from  the  shore  line  exist  at  all 
within  the  framework  of  the  linear  theory  unless  a  singularity  at 
least  as  strong  as  a  logarithmic  singularity  is  admitted  at  the  shore 
line. 

The  actual  procedure  works  out  as  follows:  Once  the  frequency 
of  the  wave  motion  has  been  fixed,  two  different  types  of  standing 

*  This  fact  is  also  used  in  laboratory  experiments  with  water  waves:  the 
experimental  tanks  are  often  equipped  with  a  sloping  ''beach"  at  one  or  more 
of  the  ends  in  order  to  absorb  the  energy  of  the  incoming  waves  through  breaking, 
and  thus  prevent  reflection  from  the  ends  of  the  tank.  This  makes  it  possible  to 
perform  successive  experiments  without  long  waits  for  the  motions  to  subside. 


72  WATER   WAVES 

waves  are  obtained,  one  of  which  has  finite  amplitude,  the  other 
infinite  amplitude,  at  the  shore  line.  These  two  different  types  of 
standing  waves  behave  at  oo  like  the  simple  standing  wave  solu- 
tions for  water  of  infinite  depth  obtained  in  Chapter  3;  i.e.  one  of 
them  behaves  like  emv  sin  (mx  +  a)  while  the  other  behaves  like 
emy  cos  {mx  -}-  a);  hence  the  two  may  be  combined  with  appropriate 
time  factors  to  yield  arbitrary  simple  harmonic  progressing  waves 
at  oo.  If  the  amplitude  at  oo  is  prescribed,  and  also  the  condition 
(cf.  the  last  two  sections  of  the  preceding  chapter)  requiring  that  the 
wave  at  oo  be  a  progressing  wave  moving  toward  shore,  then  the 
solution  is  uniquely  determined;  in  particular,  the  strength  of  the 
logarithmic  singularity  at  the  shore  line  is  uniquely  fixed  once  the 
amplitude  of  the  incoming  wave  is  prescribed  at  oo. 

The  fact  that  progressing  waves  over  uniformly  sloping  beaches 
can  be  uniquely  characterized  in  the  simple  way  just  stated  is  not 
a  thing  which  has  been  known  for  a  long  time,  but  represents  rather 
an  insight  gained  in  relatively  recent  years  (cf.  the  author's  paper 
[S.18]  of  1947  and  the  other  references  given  there).  The  method 
employed  in  the  author's  paper  makes  essential  use  of  an  idea  due 
to  H.  Lewy  to  obtain  the  actual  solutions  for  the  case  of  two-dimen- 
sional waves  over  beaches  sloping  at  the  angles  Tt/Vn,  with  n  an 
integer;  H.  Lewy  [L.8]  extended  his  method  also  to  solve  the  problem 
for  slope  angles  (p/2n)n9  with  p  an  odd  integer  and  n  any  integer 
such  that  p  <  2n.  For  the  special  slope  angles  7t/2n  the  progressing 
wave  solutions  were  obtained  first  by  Miche  [M.8]  (unknown  to  the 
author  at  the  time  because  of  lack  of  communications  during  World 
War  II),  and  somewhat  later  by  Bondi  [B.14],  but  without  uniqueness 
statements.  Actually,  the  special  standing  wave  solutions  for  these 
same  slope  angles  which  are  finite  at  the  shore  line  had  already 
been  obtained  by  Hanson  [H.3]. 

All  of  these  solutions  for  the  slope  angles  eo  ==  jr/2n,  become  more 
complicated  and  cumbersome  as  n  becomes  larger,  that  is,  as  the 
beach  slope  becomes  smaller.  In  fact,  the  solutions  consist  of  finite 
sums  of  complex  exponentials  and  exponential  integrals,  and  the 
number  of  the  terms  in  these  sums  increases  with  n.  Actual  ocean 
beaches  usually  slope  rather  gently,  so  that  many  of  the  interesting 
cases  are  just  those  in  which  the  slope  angle  is  small— of  the  order 
of  a  few  degrees,  say.  It  is  therefore  important  to  give  at  least  an 
approximate  representation  of  the  solution  of  the  problem  valid  for 
small  angles  eo  independent  of  the  integer  n.  Such  a  representation 


WAVES    ON    SLOPING    BEACHES    AND    PAST   OBSTACLES  73 

has  been  given  by  Friedrichs  [F.I 4],  To  derive  it  the  exact  solution 
is  first  obtained  for  integer  n  in  the  form  of  a  single  complex  integral, 
which  can  in  turn  be  treated  by  the  saddle  point  method  to  yield 
asymptotic  solutions  valid  for  large  n,  that  is,  for  beaches  with  small 
slopes.  The  resulting  asymptotic  representation  turns  out  to  be  very 
accurate.  A  comparison  with  the  exact  numerical  solution  for  co  =  6° 
shows  the  asymptotic  solution  to  be  practically  identical  with  the 
exact  solution  all  the  way  from  infinity  to  within  a  distance  of  less 
than  a  wave  length  from  the  shore  line.  Eckart  [E.2,  3]  has  devised 
an  approximate  theory  which  gives  good  results  in  both  deep  and 
shallow  water. 

For  slope  angles  which  are  rational  multiples  of  a  right  angle  of 
the  special  form  co  =  pji/2n  with  p  any  odd  integer  smaller  than  2n, 
the  problem  of  progressing  waves  has  been  treated  by  Lewy,  as  was 
mentioned  above.  Thus  the  theory  is  available  for  cases  in  which  co 
is  greater  than  jr/2,  so  that  the  ''beach"  becomes  an  overhanging 
cliff.  The  solution  for  a  special  case  of  this  kind,  i.e.  for  co  =  135° 
or  p  =  3,  n  =  2,  has  been  carried  out  numerically  by  E.  Isaacson 
[1.2].  It  turns  out  that  there  is  at  least  one  interesting  contrast  with 
the  solutions  for  waves  over  beaches  in  which  co  <  n/2.  In  the  latter 
case  it  has  been  found  that  as  a  progressing  wave  moves  in  toward 
shore  the  amplitude  first  decreases  to  a  value  below  the  value  at  oo, 
before  it  increases  and  becomes  very  large  at  the  shore  line.  (This 
fact  has  also  often  been  verified  experimentally  in  wave  tanks). 
The  same  thing  holds  for  standing  waves:  at  a  certain  distance  from 
shore  there  exists  always  a  crest  which  is  lower  than  the  crests  at  oo. 
In  the  case  of  the  overhanging  cliff  with  co  =  135°,  however,  the 
reverse  is  found  to  be  true:  the  first  maximum  going  outward  from 
the  shore  line  is  about  1  %  higher  than  the  height  of  the  crests  at  oo. 
Still  another  fact  regarding  the  behavior  of  the  solutions  near  the 
shore  line  is  interesting.  In  all  cases  there  exists  just  one  standing 
wave  solution  which  has  a  finite  amplitude  at  the  shore  line;  Lewy 
[L.8]  has  shown  that  the  ratio  of  the  amplitude  there  to  the  am- 
plitude at  oo  is  given  in  terms  of  the  angle  co  by  the  formula  (n/2co)11*. 
Thus  for  angles  co  less  than  n/2  the  amplitude  of  the  standing  wave 
with  finite  amplitude  is  greater  on  shore  than  it  is  at  infinity  (becoming 
very  large  as  co  becomes  small)  while  for  angles  co  greater  than  jr/2 
the  amplitude  on  shore  is  less  than  it  is  at  oo.  Since  the  observations 
indicate  that  the  standing  wave  of  finite  amplitude  is  likely  to  be  the 
wave  which  actually  occurs  in  nature  for  angles  co  greater  than 


74  WATER   WAVES 

about  40°,  the  above  results  can  be  used  to  give  a  rational  explanation 
for  what  might  be  called  the  "wine  glass"  effect:  wine  is  much  more 
apt  to  spill  over  the  edge  of  a  glass  with  an  edge  which  is  flared  out- 
ward than  from  a  glass  with  an  edge  turned  over  slightly  toward  the 
inside  of  the  glass. 

A  limit  case  of  the  problem  of  the  overhanging  cliff  has  a  special 
interest,  namely  the  case  in  which  co  approaches  the  value  n  and  the 
problem  becomes  what  might  be  called  the  "dock  problem":  the 
water  surface  is  free  up  to  a  certain  point  but  from  there  on  it  is 
covered  by  a  rigid  horizontal  plane.  The  solutions  given  by  Lewy 
are  so  complicated  as  p  and  n  become  large  that  it  seems  hopeless 
to  consider  the  limit  of  his  solutions  as  co  ->  n.  Friedrichs  and  Lewy 
[F.I 2]  have,  however,  attacked  and  solved  the  dock  problem  directly 
for  two-dimensional  waves.  For  three-dimensional  waves  in  water  of 
constant  finite  depth  the  problem  has  been  solved  by  Hcins  [H.I 3] 
(also  see  [H.12]). 

It  would  be  somewhat  unsatisfying  to  have  solutions  for  the  sloping 
beach  problem  only  for  slope  angles  which  are  rational  multiples 
of  n:  it  is  clear  that  this  limitation  is  imposed  by  the  methods  used 
to  solve  the  problem  and  not  by  any  inherent  characteristics  of  the 
problem  itself.  The  two-dimensional  problem  has,  in  fact,  been  solved 
for  all  slope  angles  by  Isaacson  [I.I].  Isaacson  obtained  an  integral 
representation  of  Lewy's  solutions  for  the  angles  pn/2n  analogous 
to  the  representation  obtained  by  Friedrichs  for  the  angles  n/2n, 
and  then  observed  that  his  representation  depended  only  upon  the 
ratio  of  p  to  n  and  not  on  these  quantities  separately.  Thus  the 
solutions  for  all  angles  are  given  by  this  representation.  Peters  [P.5] 
has  solved  the  same  problem  by  an  entirely  different  method,  which 
makes  no  use  of  solutions  for  the  special  slope  angles  pyi/2n. 

The  problem  of  two-dimensional  progressing  waves  over  sloping 
beaches  thus  has  been  completely  solved  as  far  as  the  theory  of 
waves  of  small  amplitude  is  concerned.  Only  one  solution  for  three- 
dimensional  motion  has  been  mentioned  so  far,  i.e.  the  solution  by 
Heins  for  three-dimensional  motion  in  the  case  of  the  dock  problem. 
For  certain  slope  angles  co  =  n/2n  the  method  used  by  the  author 
[S.18]  can  be  extended  in  such  a  way  as  to  solve  the  problem  of 
three-dimensional  waves  on  sloping  beaches;  in  the  paper  cited  the 
solution  is  carried  out  for  the  case  co  =  n/29  i.e.  for  the  case  of  waves 
approaching  at  an  angle  and  breaking  on  a  vertical  cliff.  Roseau 
[R.9]  has  used  the  same  method  for  the  case  co  =  jr/4.  Subsequently 


WAVES    ON    SLOPING    BEACHES    AND    PAST    OBSTACLES  75 

the  problem  of  three-dimensional  waves  on  sloping  beaches  has  been 
solved  by  Peters  [P.6]  and  Roseau  [R.9],  who  make  use  of  a  certain 
functional  equation  derived  from  a  representation  of  the  solution 
by  a  Laplace  integral.  In  section  5.4.  we  shall  give  an  account  of 
this  method  of  attack.  Roseau  [R.9]  has  solved  the  problem  of  waves 
in  an  ocean  having  different  constant  depths  at  the  two  different 
infinities  in  the  .r-direction  which  are  connected  by  a  bottom  of 
variable  depth. 

Before  outlining  the  actual  contents  of  the  present  chapter,  it 
may  be  well  to  summarize  the  conclusions  which  have  been  obtained 
from  studying  numerical  solutions  of  the  problems  being  considered 
here,  which  have  been  carried  out  (cf.  [S.18])  for  two-dimensional 
waves  for  slope  angles  co  =  185°,  90°,  45°,  and  6°,  and  for  three- 
dimensional  waves  for  the  case  o>  =  90°.  The  results  for  the  case  of  an 
overhanging  cliff  with  a)  --  135°  have  already  been  discussed  earlier. 
In  the  other  three  cases  the  most  striking  and  important  result  is 
the  following:  The  wave  lengths  and  amplitudes  change  very  little 
from  their  values  at  oo  until  points  about  a  wave  length  from  shore 
have  been  reached.  Closer  inshore  the  amplitude  becomes  large,  as 
it  must  in  accord  with  our  theory.  It  is  a  curious  fact  (already  men- 
tioned earlier)  that  the  amplitude  of  a  progressing  wave  becomes 
less  (for  r/j  -  6°  about  10  %  less)  at  a  point  near  shore  than  its  value 
at  oo,  although  it  becomes  infinite  as  the  shore  is  approached.  This 
effect  has  often  been  observed  experimentally.  This  statement  holds 
for  the  three-dimensional  waves  against  a  vertical  cliff  (with  an 
amplitude  decrease  of  about  2  %),  as  well  as  for  the  two-dimensional 
cases. 

The  exact  numerical  solution  for  the  case  of  a  beach  sloping  at 
6°  is  useful  for  the  purpose  of  a  comparison  with  the  results  obtained 
from  the  linear  shallow  water  theory  (treated  in  Chapter  (10.13)  of 
Part  III)  and  from  the  asymptotic  approximation  to  the  exact  theory 
obtained  by  Friedrichs  [F.I 4],  The  linear  shallow  water  theory,  as 
its  name  indicates,  can  in  principle  not  furnish  a  good  approximation 
to  the  waves  on  sloping  beaches  in  the  deep  water  portion  since  it 
yields  waves  whose  amplitude  tends  to  zero  at  oo.  For  a  beach  sloping 
at  6°,  for  example,  it  is  found  that  the  shallow  water  theory  furnishes 
a  good  approximation  to  the  exact  solution  for  a  distance  of  two  or 
three  wave  lengths  outward  from  the  shore  line  if  the  wave  length 
is,  say,  about  eight  times  the  maximum  depth  of  the  water  in  this 
range;  but  the  amplitudes  furnished  by  the  shallow  water  theory 


76  WATER    WAVES 

would  be  50  to  60  percent  too  small  at  about  15  wave  lengths  away 
from  the  shore  line.  One  of  the  asymptotic  approximations  to  the 
exact  theory  given  by  Friedrichs  yields  a  good  approximation  over 
practically  the  whole  range  from  the  shore  line  to  infinity  (it  is  in- 
accurate only  very  close  to  shore);  this  approximation,  which  even 
yields  the  decrease  in  amplitude  under  the  value  at  oo  mentioned  above, 
is  almost  identical  with  one  obtained  by  Rankine  (cf.  Miche  [M.8, 
p.  287])  which  is  based  upon  an  argument  using  energy  flux  con- 
siderations in  connection  with  the  assumption  that  the  speed  of  the 
energy  flux  can  be  computed  at  each  point  in  water  of  slowly  varying 
depth  by  using  the  formula  (cf.  (3.3.9))  which  is  appropriate  in  water 
having  everywhere  the  depth  at  the  point  in  question.  Friedrichs  thus 
gives  a  mathematical  justification  for  such  a  procedure  on  beaches 
of  small  slope. 

It  has  already  been  made  clear  that  the  discussion  in  this  chapter 
cannot  yield  information  about  the  breaking  of  waves,  which  is  an 
essentially  nonlinear  phenomenon.  However,  it  is  possible  to  analyze 
the  breaking  phenomena  in  certain  cases  and  within  certain  limitations 
by  making  use  of  the  nonlinear  shallow  water  theory,  as  we  shall  see 
in  Part  III.  For  this  purpose,  one  needs  to  know  in  advance  the 
motion  at  some  point  in  shallow  water,  and  this  presumably  could 
be  done  by  using  the  methods  of  the  present  chapter,  combined 
possibly  with  the  methods  provided  by  the  linear  shallow  water 
theory. 

The  material  in  the  subsequent  sections  of  this  chapter  is  ordered 
as  follows.  In  section  5.2.  the  problem  of  two-dimensional  progressing 
waves  over  beaches  sloping  at  the  angles  7t/2n,  n  an  integer,  is  discussed 
following  the  method  of  Lewy  [L.8]  and  the  author  [S.18].  In  section 
5.3  the  problem  of  three-dimensional  waves  against  a  vertical  cliff 
is  treated,  also  using  the  author's  method.  The  reasons  for  including 
these  treatments  in  spite  of  the  fact  that  they  yield  results  that  are 
included  in  the  more  general  treatments  of  Peters  [P.6]  and  Roseau 
[R.9]  is  that  they  are  interesting  in  themselves  as  an  example  of 
method,  and  also  they  can  be  applied  to  other  problems,  such  as  the 
problem  of  plane  barriers  inclined  at  the  angles  n/2n  (cf.  F.  John 
[J.4]),  which  have  not  been  treated  by  other  methods.  In  section  5.4, 
the  general  problem  of  three-dimensional  waves  on  beaches  sloping 
at  any  angle  is  treated  following  essentially  the  ideas  of  Peters. 

In  section  5.5  the  problem  of  diffraction  of  waves  around  a  rigid 
vertical  wedge  is  treated;  in  case  the  wedge  reduces  to  a  plane  the 


WAVES    ON   SLOPING   BEACHES   AND    PAST   OBSTACLES 


77 


problem  becomes  the  classical  diffraction  problem  of  Sommerfeld 
[S.I 2]  for  the  case  of  diffraction  of  plane  waves  in  two  dimensions 
around  a  half-plane  barrier.  A  new  uniqueness  theorem  and  a  new  and 
elementary  solution  for  the  problem  are  given.  Methods  of  analyzing 
the  solution  are  also  discussed;  photographs  of  the  waves  in  such  cases 
and  comparisons  of  theory  and  experiment  are  made. 

Finally,  in  section  5.6  a  brief  survey  of  a  variety  of  solved  and 
unsolved  problems  which  might  have  been  included  in  this  chapter, 
with  references  to  the  literature,  is  given.  Included  are  brief  references 
to  researches  in  oceanography,  seismology,  and  to  a  selection  of  papers 
dealing  with  simple  harmonic  waves  by  using  mathematical  methods 
different  from  those  employed  otherwise  in  this  chapter.  In  parti- 
cular, a  number  of  papers  employing  integral  equations  as  a  basic 
mathematical  tool  are  mentioned  and  occasion  is  taken  to  explain 
the  Wiener-Hopf  technique  of  solving  certain  singular  integral 
equations. 

5.2.  Two-dimensional  waves  over  beaches  sloping  at  angles  a)=n/2n 

We  consider  first  the  problem  of  two-dimensional  progressing 
waves  over  a  beach  sloping  at  the  angle  co  =  nf2n  with  n  an  integer 


Fig.  5.2.1.  Sloping  beach  problem 

(cf.  Figure  5.2.1),  in  spite  of  the  fact  that  the  problem  can  be  solved, 
as  was  mentioned  in  the  preceding  section,  by  a  method  which  is  not 


78  WATER   WAVES 

restricted  to  special  angles  (cf.  Peters  [P.6],  and  Roseau  [R.9]). 
The  problem  is  solved  here  by  a  method  which  makes  essential  use  of 
the  fact  that  the  slope  angle  has  the  special  values  indicated  because 
the  method  has  some  interest  in  itself,  and  it  yields  representations 
which  have  been  evaluated  numerically  in  certain  cases.  In  addition, 
the  relevant  uniqueness  theorems  are  obtained  in  a  very  natural  way. 
We  assume  that  the  velocity  potential  0(x9  y;  t)  is  taken  in  the 
form  0  =  eiat<p(x,  y).  Hence  <p(x*  y)  is  a  harmonic  function  in  the 
sector  of  angle  co  —  n/2n.  The  free  surface  boundary  condition  then 
takes  the  form 

a2 

(5.2.1)  <py  —  —  y  =  0,         for  y  =  0,         x  >  0, 

S 
as  we  have  often  seen  (cf.  (3.1.7)),  while  the  condition  at  the  bottom  is 

(5.2.2)  ??  =  0. 

on 

It  is  useful  to  introduce  the  same  dimensionlcss  independent  variables 
as  were  used  in  the  preceding  chapter: 

(5.2.3)  xl  =  mx,         y±  =  my,         m  —  a*/g. 

The  function  q>(x,  y)  obviously  remains  harmonic  in  these  variables, 
and  conditions  (5.2.1)  and  (5.2.2)  become 

(5.2.1)'  <py  -  <p  =  0,         y  -  0,         x  >  0, 

(5.2.2)'  cpn  =  0, 

after  dropping  subscripts. 

The  simple  harmonic  standing  waves  in  water  of  infinite  depth 
everywhere  are  given  by 

cos  (x  +  a) 
sin  (x  +  a)' 

we  write  these  down  because  we  expect  that  they  will  represent  the 
behavior  of  the  standing  waves  in  our  case  at  large  distances  from 
the  origin,  that  is,  far  away  from  the  shore  line. 

The  solution  of  the  problem  is  obtained  in  terms  of  the  complex 
potential  f(z)  defined  by 

(5.2.5)  /(*)  =  f(x  +  iy)  =  <p(x,  y)  +  iX(x,  y). 

The  function  f(z)  should,  like  q>9  be  regular  and  analytic  in  the 


{oo 
SH 


WAVES    ON    SLOPING    BEACHES    AND    PAST    OBSTACLES  79 

entire   sector    (including   the   boundaries,*   except  for  the  origin). 
The  boundary  conditions  (5.2.1)'  and  (5.2.2)'  are  given  in  terms  of 


(5.2.6)    <py  -  <p  =  &  Ij-  - 


=  31  e  (if    -  /)  =  0  for  z  real  and  positive, 


(5.2.7)  Vn  =  3le     -  (f(z))  =  0fc  (-  iexp  (-  in/2n)  /') 

on 

=  0  for  z  —  r  exp  {—  w/2n}9     r  >  0. 
The  second  condition  results  from 

•**  ~  (/(*))  -  #*  (-  -  ~  (/(*)))  =  #*  {-'>*/'(*)}. 

3fl  1     r  dO  ) 

We  introduce  the  two  following  linear  differential  operators: 

(5.2.8)  Li(D)  =  -  i  exp  {-  inftn)  D, 

(5.2.9)  L2n(D)  =  i/)  -  1 

with  Z>  meaning  d/dz.  The  basic  idea  of  the  method  invented  by 
H.  Lcwy  is  to  find  additional  linear  operators,  L2,  L3,  .  .  .,  £2n-i 
such  that  the  operation  L^  •  L2  •  .  .  .  •  L2n  applied  on  /(s)  yields  a 
function  F(z)  whose  real  part  vanishes  on  both  boundaries  of  our  sector. 
Once  this  has  been  done,  the  function  F(z)  can  be  continued  analyti- 
cally over  the  boundaries  of  the  sector  by  successive  reflections  to 
yield  a  single-valued  function  defined  in  the  entire  complex  plane 
except  possibly  the  origin.  It  can  then  be  shown  (see  [S.18]),  essen- 
tially by  using  Liouville's  theorem,  that  the  function  F(z)  is  uniquely 
determined  within  a  constant  multiplying  factor  by  boundedness 
conditions  on  the  complex  potential  f(z)  at  oo  together  with  the  order 
of  the  singularity  admitted  at  the  origin.  After  F(z)  has  been  thus 
determined,  the  complex  potential  f(z)  is  obtained  as  a  solution  of 
the  ordinary  differential  equation  L^L2  .  .  .  L2nf(z)  =  F(z).  Of  course, 
it  is  necessary  in  the  end  to  determine  the  arbitrary  constants  in  the 
general  solution  of  this  differential  equation  in  such  a  way  as  to 
satisfy  all  conditions  of  the  problem,  and  this  can  in  fact  be  done 
explicitly.  It  turns  out  that  the  resulting  solution  behaves  at  oo  like 

*  Far  less  stringent  conditions  at  the  boundaries  could  be  prescribed,  since 
analytic  continuations  over  the  boundaries  can  easily  be  obtained  explicitly  in 
the  present  ease. 


80  WATER    WAVES 

the  known  solutions  for  waves  in  water  having  infinite  depth  every- 
where and  that  it  is  uniquely  determined  by  prescribing  the  amplitude 
of  the  wave  at  oo  together  with  the  assumption  that  it  should  be, 
say,  an  incoming  wave. 

We  proceed  to  carry  out  this  program,  without  however  giving 
all  of  the  details  (which  can  be  found  in  the  author's  paper  [S.18]). 
To  begin  with,  the  ordinary  differential  equation  for  f(z)  and  the 
operators  Lt  are  given  by 

(5.2.10)       L(D)f  =  L!  •  L2  •  L8  •  .  .  .  -  L2nf 

nD  -  I)/ 


with  the  complex  constants  v.k  defined  by 

(5.2.11  )  a*  =  e~in  (L  +  1}  ,         fc  =  1,  2,  .  .  .,  2n. 

One  observes  that  L±(D)  and  L2n(D)  coincide  with  the  definitions 
given  in  (5.2.8)  and  (5.2.9).  It  is,  in  fact,  not  difficult  to  verify  that 

(5.2.12)  &eF(z)  =  0 

on  both  boundaries  of  the  sector,  by  making  use  of  the  properties 
of  the  numbers  a*  and  of  the  fact  that  Ste  L^D)  and  Ste  L2n(D)f 
vanish  on  the  bottom  and  the  free  surface,  respectively,  by  virtue 
of  the  boundary  conditions  (5.2.7)  and  (5.2.6). 

So  far  we  have  not  prescribed  conditions  on  f(z)  at  oo  and  at  the 
origin,  and  we  now  proceed  to  do  so.  At  the  origin  we  assume,  in 
accordance  with  the  remarks  made  in  section  5.1  and  the  discussion 
in  the  last  section  of  the  preceding  chapter,  that  f(z)  has  at  most  a 
logarithmic  singularity;  we  interpret  this  to  mean  that  |  dkf(z)/dzk  \  < 
Mk/\  z  \k  in  a  neighborhood  of  the  origin  for  k  =  1,  2,  .  .  .,  2/i,  with 
Mk  certain  constants.  At  oo  we  require  that  <p  =  ffle  f(z)  together 
with  |  dkf(z)/dzk  |  for  k  =  1,  2,  .  .  .,  2n  be  uniformly  bounded  when 
z  ->  oo  in  the  sector.  (These  conditions  could  be  weakened  con- 
siderably, but  they  are  convenient  and  are  satisfied  by  the  solutions 
we  obtain.  )  In  other  words,  although  we  expect  the  solutions  of  our 
problem  to  behave  at  oo  in  accordance  with  (5.2.4)  it  is  not  necessary 
to  prescribe  the  behavior  at  oo  so  precisely  since  the  boundedness 
conditions  yield  solutions  having  this  property  automatically.  Once 
these  conditions  on  f(z)  have  been  prescribed  we  see  that  the  function 
F(z)  defined  by  (5.2.10)  has  the  following  properties:  1)  |  F(z)  \  is 


WAVES   ON   SLOPING   BEACHES   AND   PAST   OBSTACLES  81 

uniformly  bounded  in  the  sector,  and  2)  |  F(z)  \  =  0(1  jz2n)  in  the 
neighborhood  of  the  origin. 

We  have  already  observed  that  3te  F(z)  =  0  on  both  boundaries 
of  the  sector  and  that  F(z)  can  therefore  be  continued  as  a  single- 
valued  function  into  the  whole  plane,  except  the  origin,  by  the 
reflection  process.  Here  we  make  decisive  use  of  the  assumption 
that  co,  the  angle  of  the  sector,  is  n/2n  with  n  an  integer.  Since  the 
boundedness  properties  of  F(z)  at  oo  and  the  origin  are  preserved 
in  the  reflection  process,  it  is  clear  from  well-known  results  concerning 
analytic  functions  that  F(z)  is  an  analytic  function  over  the  whole 
plane  having  a  pole  of  order  at  most  2n  at  the  origin.  Since  in  ad- 
dition the  real  part  of  F(z)  vanishes  on  all  rays  z  =  r  exp{ikn/2n}, 
k  —  1,  2,  .  .  .,  4tt,  it  follows  that  F(z)  is  given  uniquely  by 

(5.2.13)  F(z)  = 


Z2n 


with  A2n  an  arbitrary  real  constant  which  may  in  particular  have 
the  value  zero.  Thus  the  complex  potential  f(z)  we  seek  satisfies  the 
differential  equation 

(5.2.14)     faDfoJ)  -  1) ...  (a2n.1D)(a2nZ)  -  I)/  -  ^. 


Our  problem  is  reduced  to  finding  a  solution  f(z)  of  this  differential 
equation  which  satisfies  all  of  the  conditions  imposed  on  f(z).  From 
the  discussion  of  section  5.1  we  expect  to  find  two  solutions  f^z) 
and  f2(z)  of  our  problem  which  behave  differently  at  the  origin  and 
at  oo;  at  the  origin,  in  particular,  we  expect  to  find  one  solution, 
say  fi(z),  to  be  bounded  and  the  other,  /2(s),  to  have  a  logarithmic 
singularity. 

The  regular  solution  f^z)  is  the  solution  of  (5.2.14)  which  one 
obtains  by  taking  for  the  real  constant  A2n  the  value  zero,  while 
f2(z)  results  for  A2n  ^  0.  In  other  words  the  solution  of  the  non- 
homogeneous  equation  contains  the  desired  singularity  at  the  origin. 
One  finds  for  f^z)  the  solution 


in  which  the  constants  ck  and  ftk  are  the  following  complex  numbers: 


82  WATER   WAVES 


(5.2.16) 


.     n  +  l     k\\        n        2n  (k-l 

=  exp  {ml  ---  --  1}  cot  —  cot  —  .  .  .cot- 

*\     \     4          2/J        2n        2n  2n 


-l)n 
--  , 
2n 

A:  =  2,  3,  ...,n 

=  cn. 

The  constants  ck  are  obtained  by  adjusting  the  arbitrary  constants 
in  the  solution  of  (5.2.14)  so  that  the  boundary  conditions  on  f(z) 
at  the  free  surface  and  the  bottom  are  satisfied;  that  such  a  result 
can  be  achieved  by  choosing  a  finite  number  of  constants  is  at  first 
sight  rather  startling,  but  it  must  be  possible  if  it  is  true  that  a  func- 
tion f(z)  having  the  postulated  properties  exists  since  such  a  function 
must  satisfy  the  differential  equation  (5.2.14).  The  calculation  of  the 
constants  ck  is  straightforward,  but  not  entirely  trivial.  The  function 
f^z)  is  uniquely  given  by  (5.2.15)  within  a  real  multiplying  factor. 
As  |  z  |  ->  oo  in  the  sector,  all  terms  clearly  die  out  exponentially 
except  the  term  for  k  =  n,  which  is  cn  exp  {—  iz}9  since  all  (iks 
except  f}n  have  negative  real  parts.  Even  the  term  for  k  —  n  dies 
out  exponentially  except  along  lines  parallel  to  the  real  axis.  (The 
value  of  cn,  by  the  way,  is  exp  {—  -  in(n  —  l)/4}  since  the  cotangents 
in  (5.2.16)  cancel  each  other  for  k  =  n.)  This  term  thus  yields  the 
asymptotic  behavior  of  fi(z): 


The  solution  /2(*)  of  the  nonhomogeneous  equation  (5.2.14)  which 
satisfies  the  boundary  conditions  is  as  follows: 

(5.2.18)  f2(z) 


n     r      r  **Pk  **'  n 

=  £  a*  ^k  —dt-  me**** 

*-l  L  J«00  t  J 


for  the  case  in  which  the  real  constant  A2n  is  set  equal  to  one.  The 
constants  {ik  are  defined  in  (5.2.16);  and  the  constants  ak  are  defined  by 

(5.2.19)  ak  =  ck/{(n  -  I)\Vn}> 

that  is,  they  are  a  fixed  multiple  (for  given  n)  of  the  constants  ck 
defined  in  (5.2.16).  The  constants  ak9  like  the  ck,  are  uniquely  deter- 
mined within  a  real  multiplying  factor.  The  path  of  integration  for 
all  integrals  in  (5.2.18)  is  indicated  in  Figure  5.2.2.  That  the  points 
•  izftk  lie  in  the  lower  half  of  the  complex  plane  (as  indicated  in  the 
figure)  can  be  seen  from  our  definition  of  the  constants  (ik  and  the 
fact  that  z  is  restricted  to  the  sector  —  jr/2n  ^  arg  2^0. 


WAVES    ON    SLOPING    BEACHES    AND    PAST    OBSTACLES 


83 


Fig.  5.2.2.  Path  of  integration  in  f-plane 

The  behavior  of  f2(z)  at  oo  of  course  depends  on  the  behavior 
of  the  functions  in  (5.2.18).  It  is  not  hard  to  show— for  example, 
by  the  procedure  used  in  arriving  at  the  result  given  by  (4.3.15) 
in  the  preceding  chapter— that  these  functions  behave  asymptotically 
as  follows: 


(5.2.20) 


fi*(*keit  (()(    }  9 

>*i*k\        —dt~\    VI/ 

I  J  I  .  /1\ 

%/too     *  [2m — o[~] , 


0, 


0,    Jm  (iz0k)  ^  0. 


Once  this  fact  is  established  it  is  clear  from  (5.2.19)  and  (5.2.18) 
that  f2(z)  behaves  asymptotically  as  follows: 


(5.2.21) 


n 


, 

(n  — 


since  the  term  for  k  =  n  dominates  all  others  (cf.  (5.2.20))  and 
Ste(izpk)  >  0  in  this  case.  Comparison  of  (5.2.21)  with  (5.2.17)  shows 
that  the  real  parts  of  f^z)  and  f2(z)  would  be  90°  out  of  phase  at  oo. 
That  the  derivatives  of  /2(z)  behave  asymptotically  in  the  same 
fashion  as  f2(z)  itself  is  easily  seen,  since  the  only  terms  in  the  deriva- 
tives of  (5.2.18)  of  a  type  different  from  those  in  (5.2.18)  itself  are 
of  the  form  bk/zk9  k  an  integer  ^  1.  Finally,  it  is  clear  that  f2(z) 
has  a  logarithmic  singularity  at  the  origin.  Hence  f^z)  and  f2(z) 
satisfy  all  requirements.  Just  as  in  the  90°  case  (cf.  the  last  section 


84  WATER   WAVES 

of  the  preceding  chapter)  it  is  now  clear  that  f(z)  =  b^^z)  +  b%f2(z), 
with  bi  and  62  any  real  constants,  yields  all  standing  wave  solutions 
of  our  problem. 

The  relations  (5.2.17)  and  (5.2.21  )  yield  for  the  asymptotic  behavior 
of  the  real  potential  functions  (p^  and  q>2  the  relations: 


(5.2.22)     p^x,  y)  =  9keji~  ~t  -  ~\T~r  e"  cos  \x  +  —7—  n\ 

\  4         / 


t 
(n  — 

7T  /  *W         _    1         \ 

(5.2.23)  <p2(x,  y)  =  3tef2~  -  -  -.7-7-  **  sin  (a?  +  —  -  —  n\ 

(n  —  -  IJiyn  \  4         / 

when  it  is  observed  that  cn  =  exp  {—  CT(/&  —  l)/4).  It  is  now  possible 
to  construct  either  standing  wave  or  progressing  wave  solutions  which 
behave  at  oo  like  the  known  solutions  for  steady  progressing  waves 
in  water  which  is  everywhere  infinite  in  depth.  In  particular  we 
observe  that  it  makes  sense  to  speak  of  the  wave  length  at  oo  in  our 
cases  and  that  the  relation  between  wave  length  and  frequency 
satisfies  asymptotically  the  relation  which  holds  everywhere  in  water 
of  infinite  depth.  For  this,  it  is  only  necessary  to  reintroduce  the 
original  space  variables  by  replacing  x  and  y  by  mx  and  my,  with 
m  =  o2lg  (cf.  (5.2.3)),  and  to  take  note  of  (5.2.22)  and  (5.2.23). 

Finally,  we  write  down  a  solution  0(x,  y\  t)  which  behaves  at  oo 
like  ey  cos  (x  +  t  +  a),  i.e.  like  a  steady  progressing  wave  moving 
toward  shore: 

(5.2.24)  0(x,  y;  t)  =  A[cpi(x,  y)  cos  (t  +  a)  —  <p2(x,  y)  sin  (t  +  a)]. 

As  our  discussion  shows,  this  solution  is  uniquely  determined  as  soon 
as  the  amplitude  is  prescribed  at  oo  (i.e.  as  soon  as  A  is  fixed)  since 
<pi(x,  y)  and  (p2(x,  y)  yield  the  only  standing  wave  solutions  of  our 
problem  and  they  are  determined  also  within  a  real  factor.  As  we 
have  already  stated  in  the  preceding  section,  the  progressing  wave 
solutions  (5.2.24)  have  been  determined  numerically  (cf.  [S.18])  for 
slope  angles  to  =  90°,  45°,  and  6°,  with  results  whose  general  features 
were  already  discussed  in  that  section. 

5.3.  Three-dimensional  waves  against  a  vertical  cliff 

It  is  possible  to  treat  some  three-dimensional  problems  of  waves 
pver  sloping  beaches  by  a  method  similar  to  the  method  used  in  the 
preceding  section  for  two-dimensional  waves,  in  spite  of  the  fact 
that  it  is  now  no  longer  possible  to  make  use  of  the  theory  of  analytic 


WAVES    ON    SLOPING   BEACHES   AND    PAST   OBSTACLES  85 

functions  of  a  complex  variable.  In  this  section  we  illustrate  the 
method  by  treating  the  problem  of  progressing  waves  in  an  infinite 
ocean  bounded  on  one  side  by  a  vertical  cliff  when  the  wave  crests 
at  oo  may  make  any  angle  with  the  shore  line  (cf.  [S.18]). 

We  seek  solutions  0(x,  y,  z;  t)  of  V2(x>1/^)0  =  0  in  the  region 
x  *^Q,  y  ^  0,  —  oo  <  2  <  oo  with  the  j/-axis  taken  normal  to  the 
undisturbed  free  surface  of  the  water  and  the  z-axis*  taken  along 
the  "shore",  i.e.  at  the  water  line  on  the  vertical  cliffs  =  0.  Progressing 
waves  moving  toward  shore  are  to  be  found  such  that  the  wave 
crests  (or  other  curves  of  constant  phase)  at  large  distances  from 
shore  tend  to  a  straight  line  which  makes  an  arbitrary  angle  with 
the  shore  line.  For  this  purpose  we  seek  solutions  of  the  form 

(5.3.1)  0(x,  y,  z;  t)  =  exp  {i(at  +  kz  +  ^)}(p(x9  y) 

that  is,  solutions  in  which  periodic  factors  in  both  z  and  t  are  split  off. 
As  in  the  preceding  section,  we  introduce  new  variables  and  para- 
meters through  the  relations  xl  —  MX,  yl  ~-  my,  zl  =  mz,  ^  =-  k/m, 
m  =  o2jg  and  cbtain  for  <p  the  differential  equation 

(5.3.2)  Vf,^  -  Arty  =-  0 
and  the  free  surface  condition 

(5.3.3)  (py  —  (p  =  0         for  y  =  0, 

after  dropping  the  subscript  1  on  all  quantities.  The  condition  at 
the  cliff  is,  of  course, 

(5.3.4)  --^    =0         for  x  =  0. 

ox 

At  the  origin  x  ~~  0,  y  —  0  (i.e.  at  the  shore  line  on  the  cliff)  we 
require,  as  in  former  cases,  that  (p  should  be  of  the  form 

(5.3.5)  (p  —  Ip  log  r  +  <p,         r  <C  1, 

for  sufficiently  small  values  of  r  =  (x2  +  f/2)1/2f.  with  <p  and  ^  certain 
bounded  functions  with  bounded  first  and  second  derivatives  in  a 
neighborhood  of  the  origin.  The  functions  q>  and  !p  should  be  considered 
at  present  as  certain  given  functions;  later  on,  they  will  be  chosen 
specifically. 

For  large  values  of  r  we  wish  to  have  0(x,  y,  z;  t)  behave  like 

*  It  has  already  been  pointed  out  that  functions  of  a  complex  variable  are 
not  used  in  this  section,  so  that  the  reintroduction  of  the  letter  2  to  represent  a 
space  coordinate  should  cause  no  confusion  with  the  use  of  the  letter  z  as  a  complex 
variable  in  earlier  sections. 


86  WATER    WAVES 

ev  exp  {i(at  +  kz  +  cue  +  /?)}  with  &2  +  oc2  =  1  but  k  and  a  otherwise 
arbitrary  constants,  so  that  progressing  waves  tending  to  an  arbitrary 
plane  wave  at  oo  can  be  obtained.  This  requires  that  <p(x,  y)  should 
behave  at  oo  like  ev  exp  {i(onx  +^2)}  because  of  (5.3.1).  However, 
it  is  no  more  necessary  here  than  it  was  in  our  former  cases  to  require 
that  (p  should  behave  in  this  specific  way  at  oo;  it  suffices  in  fact 
to  require  that 

(5.8.6)  \<P\  +\<Px\  +  \<Pxy\  <M        for  r>  R0, 

i.e.  that  (p  and  the  two  derivatives  of  <p  occurring  in  (5.3.6)  should 
be  uniformly  bounded  at  oo.  As  we  shall  see,  this  requirement  leads 
to  solutions  of  the  desired  type. 

We  proceed  to  solve  the  boundary  value  problem  formulated  in 
equations  (5.3.2)  to  (5.3.6).  The  procedure  we  follow  is  analogous 
to  that  used  in  the  two-dimensional  cases  in  every  respect.  To  begin 
with,  we  observe  that 

/}    /  /)          \ 

(5.3.7)  —  ( —  _  i\<p  =  o         for  both  x  =  0  and  y  =  0, 
dx  \dy        J 

because  of  the  special  form  of  the  linear  operator  on  the  left  hand 
side  together  with  the  fact  that  (5.3.3)  and  (5.3.4)  are  to  be  satisfied. 
A  function  yj(x,  y)  is  introduced  by  the  relation 

(5-8-8)  "  -  T*  (TV  - 

The  essential  point  of  our  method  is  that  the  function  \p  is  determined 
uniquely  within  an  arbitrary  factor  if  our  function  9?,  having  the 
properties  postulated,  exists.  Furthermore,  y>  can  then  be  given  explic- 
itly without  difficulty.  The  properties  of  \p  are  as  follows. 

1.  \p  satisfies  the  same  differential  equation  as  y>,  i.e.  equation 
(5.3.2),  as  one  sees  from  the  definition  (5.3.8)  of  \p. 

2.  y)  is  regular  in  the  quadrant  x  >  0,  y  <  0  and  vanishes,  in  view 
of  (5.3.7),  on  x  =  0,  y  <  0   and   y  =  0,  x  >  0.   Hence  y  can  be 
continued  over  the  boundaries  by  the  reflection  process  to  yield  a 
continuous   and   single-valued  function  having  continuous   second 
derivatives  yxx  and  y>yv  (as  one  can  readily  see  since  V2y>  —  k2\p  =  0, 
and  \p  =  0  on  the  boundaries)  in  the  entire  x,  t/-planc  with  the  ex- 
ception of  the  origin.  (Here  we  use  the  fact  that  our  domain  is  a  sector 
of  angle  rc/2.) 


WAVES    ON    SLOPING   BEACHES   AND    PAST   OBSTACLES  87 

3.  At  the  origin,  \p  has  a  possible  singularity  which  is  of  the  form 
(jp(x,  t/)//*2,  with  99  regular,  as  one  can  see  from  (5.3.5)  and  (5.3.8). 
This  statement  clearly  holds  for  the  function  y  when  it  has  been 
extended  by  reflection  to  a  full  neighborhood  of  the  origin. 

4.  The  condition  (5.3.6)  on  q>  clearly  yields  for  \p  the  condition 
that  y  is  uniformly  bounded  at  oo  after  \p  has  been  extended  to  the 
whole  plane. 

Thus  \p  is  a  solution  of  V2y>  —  k\  =  0  in  the  entire  plane  which 
is  uniformly  bounded  at  oo.  At  the  origin  ip  —  q>/r2  +  <p  with  9?  and 
^  certain  regular  functions  (q>  —  0  not  excluded).  In  addition,  \p  =  0 
on  the  entire  x  and  y  axes.  We  shall  show,  following  Weinstein 
[W.5],*  that  the  function 


(5.3.9)     \p(x,  y)  =  AiH(£  (ikr)  sin  20,  r  =  Vx2  +  y2,  0  ^  k  ^  1 

is  the  unique  solution  for  \p  in  polar  coordinates  (r,  0)  with  A  an 
arbitrary  real  constant,  and  H^  the  Hankel  function  of  order  two 
which  tends  to  zero  as  r  ->  oo.  The  function  ip  has  real  values  for  r 
real.  (The  notation  given  in  Jahnke-Emde,  Tables  of  Functions,  is 
used.) 

The  solution  y  is  obtained  by  Weinstein  in  the  following  way. 
In  polar  coordinates  (r,  6)  the  differential  equation  for  ^  is 


_ 

3r2  ^  r  dr       r*  * 


For  any  fixed  value  of  r  the  function  \p  can  be  developed  in  the 
following  sine  series: 


n=l 

since  y  vanishes  for  0  =  0,  rc/2,  n,  3n/2;  and  the  coefficients  cn(r) 
are  given  by 

cn(r)  =  Cn  r/2y(r,  O)  sin  2nO  dd,         n  =  1,  2,  .  .  ., 
Jo 

with  Cn  a  normalizing  factor.  From  this  formula  one  finds  by  differen- 
tiations with  respect  to  r  and  use  of  the  differential  equation  for  ^ 
that  cn(r)  satisfies  the  equation 


*  In  the  author's  paper  the  solution  \p  was  obtained,  but  with  a  less  general 
uniqueness  statement. 


88  WATER   WAVES 

The  right  hand  side  of  this  equation  vanishes,  as  can  be  seen  by 
integrating  the  first  term  twice  by  parts  and  making  use  of  the 
boundary  conditions  \p  =  0  for  0  =  0  and  6  =  n/2.  Thus  the  functions 
cn(r)  are  Bessel  functions,  as  follows: 

cn(r)  =  Awi*^H<£(ikr)  +  BZaI2n(kr), 

with  A2n  and  B2n  arbitrary  real  constants.  The  functions  I2n  are 
unbounded  at  oo;  the  Hankel  functions  H^  behave  like  r~2n  for 
r  ->  0  and  tend  to  zero  exponentially  at  oo.  It  follows  therefore  that 
the  Fourier  series  for  ip  in  our  case  reduces  to  the  single  term  given 
by  (5.3.9)  because  of  the  boundedness  assumptions  on  \p. 

For  our  purposes  it  is  of  advantage  to  write  the  solution  \p  in  the 
following  form: 


32 


(5.3.10)  ip  =  Ai  --  H  (1)  (ikr),         r  =  Vx2  +  y2, 

oxoy 

in  which  A  is  any  real  constant  and  H(l)  is  the  Hankel  function  of 
order  zero  which  is  bounded  as  r  ->  oo.  It  is  readily  verified  that  this 
solution  differs  from  that  given  by  (5.3.9)  only  by  a  constant  multi- 
plier: for  example,  by  using  the  well-known  identities  involving  the 
derivatives  of  Bessel  functions  of  different  orders. 

Once  y>  is  determined  we  may  write  (5.3.8)  in  the  form 

(5.3.11)  —  (  --  I\w  =  Ai  -  HM  (ikr),     A  arbitrary. 
ox  \oy         /  oxoy 

This  means  that  our  function  <p,  if  it  exists,  must  satisfy  (5.3.11)  as 
well  as  (5.3.2).  By  integration  of  (5.3.11)  it  turns  out  that  we  are 
able  to  determine  q>  explicitly  without  great  difficulty  on  account 
of  the  simple  form  of  the  left  hand  side  of  (5.3.11).  This  we  proceed 
to  do. 

Integration  of  both  sides  of  (5.3.11)  with  respect  to  x  leads  to 


(5.3.12)  --iv  =  Ai-  H(»  (ikr)  +  g(y), 

in  which  g(y)  is  an  arbitrary  function.  But  g(y)  must  satisfy  (5.3.2), 
since  all  other  terms  in  (5.3.12)  satisfy  it.  Hence  d2g/dy2  —  k2g  =  0. 
In  addition  g(0)  =  0,  since  the  other  terms  in  (5.3.12)  vanish  for 
,y=0  because  of  (5.3.3)  and  the  fact  that  dH™  l9y=(ik)'^(ylr)dH^  /dr. 
Finally,  g(y)  is  bounded  as  y  ->  —  oo  because  of  condition  (5.3.6) 
and  the  fact  that  dH(V  /By  tends  to  zero  as  r  -+  oo.  The  function 


WAVES    ON    SLOPING   BEACHES   AND    PAST   OBSTACLES  89 

g(y)  is  therefore  readily  seen  to  be  identically  zero.  By  integration 
of  (5.3.12)  we  obtain  (after  setting  g(y)  =  0): 

(5.8.13)     <p  =  Aiey  T  e~*  ~  [H™  (ikVx2  +  t2)]dt  +  B(x)c*. 

J+oo          ut 

The  function  B(x)  and  the  real  constant  A  are  arbitrary.  The  integral 
converges,  since  d(H(V  )/dt  dies  out  exponentially  as  t  -*•  oo. 

We  shall  see  that  two  solutions  ^  (x,  y)  and  <p2(x,  y)  satisfying  all 
conditions  of  our  problem  can  be  obtained  from  (5.3.13)  by  taking 
A  =  0  in  one  case  and  A  ^  0  in  the  other  case,  and  that  these 
solutions  will  be  90°  "out  of  phase"  at  oo.  (This  is  exactly  analogous 
to  the  behavior  of  the  solutions  in  our  previous  two-dimensional  cases.  ) 
Consider  first  the  case  A  =  0.  The  function  9?  given  by  (5.3.13) 
satisfies  (5.3.2)  only  if 


(5.3.14)  -  +  (1  -  k2)B(x)  =  0. 


It  is  important  to  recall  that  k2  <  1.  The  boundary  condition 
(px  =  o  for  x  =  0  requires  that  Bx(0)  =  0.  The  condition  q>y  —  <p  =  0 
for  y  =  0  is  automatically  satisfied  because  of  (5.3.12)  and  g(y)  =  0. 
Hence  B(x)  =  Al  cos  Vl  —  k2x,  with  Al  arbitrary,  and  the  solution 


(5.3.15)  ^(cr,  y)  =  A^v  cos  Vl  —  k*x. 

This  leads  to  solutions  0l  in  the  form  of  standing  waves,*  as  follows: 

(5.3.15)'     0^,  j/,  2;  t)  =  Af^e*  cos  Vl~^~k*x  •  (C°S 

(sin 

for  k2  <  1.  If  k  =  1,  the  solution  0X  given  by  (5.3.15)'  continues  to 
be  valid. 

As  we  have   already   stated,   we   obtain  solutions  <p%(x,y)  from 

(5.3.13)  for  A  ^  0  which  behave  for  large  x  like  sin  Vl  —  k2  x  rather 
than  like  cos  Vl  —  k2  x9  and  with  these  two  types  of  solutions 
progressing  waves  approaching  an  arbitrary  plane  wave  at  oo  can 
be  constructed  by  superposition. 

We  begin  by  showing  that  (5.3.2)  is  satisfied  for  all  x  >  0,  y  <  0 
by  (p  as  given  in  (5.3.13)  with  A  ^  0,  provided  only  that  B(x) 

*  The  standing  wave  solutions  of  this  type  (but  not  of  the  type  with  a  singu- 
larity) for  beaches  sloping  at  angles  n/2n  were  obtained  by  Hanson  [H.3]  by  a 
quite  different  method. 


90  WATER   WAVES 

satisfies  (5.3.14).  Since  x  >  0,  it  is  permissible  to  differentiate  under 
the  integral  sign  in  (5.3.13),  even  though  t  takes  on  the  value  zero 
(since  the  upper  limit  y  is  negative).  By  differentiating  we  obtain 

(5.3.16)     V2  <p-  top  =  Ai  {«*  f  V«  I  fj^  +  (1  - 

(v  \ 
- 


Since  ffW  is  a  solution  of  (5.3.2)  the  operator  (92/9a?2  —  A;2)  oc- 
curring under  the  integral  sign  can  be  replaced  by  —  92/9j/2  and  hence 
the  integral  can  be  written  in  the  form 


We  introduce  the  following  notation 


and  obtain  through  two  integrations  by  parts  the  result 

e* 


in  which  we  have  made  use  of  the  fact  that  the  boundary  terms  arc 
zero  at  the  lower  limit  +  oo,  since  all  derivatives  of  H^  (ikr)  tend 
to  zero  as  r  ->  +  oo.  The  integral  of  interest  to  us  is  given  obviously 
by  II  —  /3  and  this  in  turn  is  given  by 

U) 

dt 


_ 

3  a*/2        a?/ 

i 

by  use  of  the  above  relations  for  Im.  Hence  the  quantity  in  the  first 
bracket  in  (5.3.16)  is  identically  zero—in  other  words  the  term 
containing  the  integral  on  the  right  hand  side  of  (5.3.13)  is  a  solution 
of  (5.3.2).  Hence  y  is  a  solution  of  (5.3.2)  in  the  case  A  ^  0  if 
B(x)  satisfies  (5.3.14).  Since  (5.3.12)  holds  and  g(y)  =  0  it  follows 
that  the  free  surface  condition  (5.3.3)  is  satisfied  by  <p  in  view  of 
the  fact  that  dH(V  (ikr)/dy  =  0  for  y  =  0. 

We  have  still  to  show  that  a  solution  B(x)  of  (5.3.14)  can  be  chosen 


WAVES   ON   SLOPING   BEACHES   AND   PAST   OBSTACLES  91 

so  that  (px  —  0  for  x  —  0,  and  that  g?  has  the  desired  behavior  for 
large  values  of  r.  Actually,  these  two  things  go  hand  in  hand.  An 
integration  by  parts  in  (5.3.13)  yields  the  following  for  q>: 


5.3.17) 


(p=Aie*  f  V«#£l)  (ikVx*+t 

J  QC 


provided  that  x  >  0.  It  should  be  recalled  that  the  upper  limit  y 
of  the  integral  is  negative;  thus  the  integrand  has  a  singularity  for 
x  =  0  since  t  =  0  is  included  in  the  interval  of  integration  and 
11^  (ikr)  is  singular  for  r  —  0.  We  shall  show  that  lim  dyjdx  =  0 

x->0 

provided  that  #,(0)  =  —  2 A  ^  0.  We  have,  for  x  >  0  and  y  <  0: 

(jW 

—  --  Aiev 


f  V*  |-  [HM  (ikVx*  +  t*)]dt 

J  00          OX 


dx 

The  second  term  on  the  right  hand  side  is  readily  seen  to  approach 
zero  as  x  ->  0  since  this  term  can  be  written  as  the  product  of  x  and 
a  factor  which  is  bounded  for  y  <  0.  For  the  same  reason  it  is  clear 
that  the  only  contribution  furnished  by  the  integral  in  the  limit 
as  x  -  >  0  arises  from  a  neighborhood  of  t  —  0  since  the  factor  x  may 
be  taken  outside  of  the  integral  sign.  We  therefore  consider  the  limit 


lim  f    V<  A  [i//£ 

a->oJe  OX 


lim  <        [i//>  (ikVx*  +72)]d*,        e  >  0. 


The  function  ill^  (ikr)  has  the  following  development  valid  near 
r  =  0: 


HI^  (ikr)  =  —  -  [JQ(ikr)  log  r  +  p(r)~\ 

o    v      J  n  L  ov      /     g     -r  FV  ;j 

in  which  p(r)  represents  a  convergent  power  series  containing  only 
even  powers  of  r,  and  J0  is  the  regular  Bessel  function  with  the 
following  development 

J0(ikr)  =  i  +.-L-  +.... 
It  follows  that 

—  [iHM  (ikr)]  =  -  -  |-  JQ(ikr)  +J'o(ikr)  -  logr  +xg(r)  1 
ox  n\r2  r  J 

1 
0  2 


92  WATER   WAVES 

in  which  g(r)  =  (l/r)dp/dr  is  bounded  as  o?->0  since  j/<0.  The  con- 
tribution of  our  integral  in  the  limit  is  therefore  easily  seen  to  be 
given  by 

2  f~*     i      *  2  f~fi      a? 

lim  —  -         £-'—  -  -  d*  =  lim  —  -         —  ---  -  dt. 


By  introducing  u  =  </#  as  new  integration  variable  and  passing  to 
the  limit  we  may  write 

2  f-e      x         ,  2  f-00     dtt 

lim  -  -         -  0  dt  =  -  -         -  -  2. 


It  therefore  follows  that  lim  d(p/dx  =  0  provided  that 


(5.3.18)  ^(0)  =  -  2A. 

The  function  B(o?)  which  satisfies  this  condition  and  the  differential 
equation  (5.3.14)  is 

o  A  __ 

(5.3.19)  B(x)  -  --  --  sin  Vl  - 

Vl  -k2 


Since  H^  (ikr)  dies  out  exponentially  as  r  ->  oo  it  follows  that  the 
solution  <p  given  by  (5.3.17)  with  B(x)  defined  by  (5.3.19)  behaves 
at  oo  like  e*  sin  [(1  -  k*)lf*x]. 

A  solution  9?2  of  our  problem  which  is  out  of  phase  with  q)l  (cf. 
(5.3.15))  is  therefore  given  by 


(5.3.20)    <pi(x,  y)  =  A2    ie*  [* 

n^y  ___      __       "I 

(ikVx*  +  i/2)  -  —.^.•=  sin  Vl  -  k2  x   , 
Vl-  k2  J 

with  ^f2  an  arbitrary  real  constant.  Standing  wave  solutions  02  are 
then  given  by 


(5.3.20)'  <Z>2 


fcos  kz] 

( V*    ii\  •  )  \. 

\x>  y)      1    .       ,     f  • 

[sin  kz\ 


By  taking  appropriate  values  of  k  progressing  waves  tending  at  oo 
to  any  arbitrary  plane  wave  solution  for  water  of  infinite  depth  can 
be  obtained  by  forming  proper  linear  combinations  of  solutions  of  the 
type  (5.3.15)'  and  (5.3.20)'.  For  a  progressing  wave  traveling  toward 
shore,  for  example,  we  may  write 


WAVES    ON    SLOPING   BEACHES   AND    PAST   OBSTACLES 


93 


(5.3.21  )  0(x,  y,  z;  t)  =  A 


[' 

-A\p 


A/1  —  k2  "1 

x,  y)  cos  kz  -\  --  q>2(x,  y)  sin  kz   cos  at 


Jfcs  "1 

992(#,  y)cos  kz  sin  at 

in  which  Al  and  A2  in  (5.3.15)  and  (5.3.20)  are  both  taken  equal 

to  ^.  The  solution  (5.3.21)  behaves  at  oo  like  Ae*  cos  (Vl  —k2x+kz+at) 
as  one  can  readily  verify  by  making  use  of  the  asymptotic  behavior 
of  q>i(x,  y)  and  q>2(x,  y)*  and  it  is  the  only  such  solution  since  <p^ 
and  9?2  are  uniquely  determined. 

The  special  case  k  =  I  has  a  certain  interest.  It  corresponds  to 
waves  which  at  oo  have  their  crests  at  right  angles  to  the  shore. 
One  readily  sees  from  (5.3.15)  and  (5.3.20)  that  as  k  ->  1  the  pro- 
gressing wave  solution  (5.3.21)  tends  to 

(5.3.22)  0(y,  z;  t)  =  Aev  cos  (z  +  at) 

that  is,  the  progressing  wave  solution  for  this  case  is  independent 
of  a?,  is  free  of  a  singularity  at  the  origin,  and  the  curves  of  constant 
phase  are  straight  lines  at  right  angles  to  the  shore  line—  all  properties 
that  are  to  be  expected. 

The  progressing  wave  solution  (5.3.21)  was  studied  numerically 


-2 


KO 


K2.)J 


Fig.  5.8.1.  Standing  wave  solution  for  a  vertical  cliff  (with  crests  at  an  angle 

of  30°  to  shore) 


*  We  remark  once  more  that  the  original  space  and  time  variables  can  be 
reintroduced  simply  by  replacing  «,  y,  z  by  ma?,  my,  mz  and  k  by  kjm. 


94 


WATER  WAVES 


for  k  =  1/2,  i.e.  for  the  case  in  which  the  wave  crests  tend  at  oo 
to  a  straight  line  inclined  at  30°  to  the  shore  line.  The  function 
<p%(%9  0)  is  plotted  in  Figure  5.3.1.  With  the  aid  of  these  values  the 
contours  for  0  were  calculated  and  are  given  in  Figure  5.3.2.  These 
are  also  essentially  contour  lines  for  the  free  surface  elevation  77, 

in  accordance  with  the  formula  rj  = <Pt  \  v.0 .  The  water  surface 

g 

is  shown  between  a  pair  of  successive  "nodes"  of  <Z>,  that  is,  curves 
for  which  0  =  0.  These  curves  go  into  the  2-axis  (the  shore  line) 
under  zero  angle,  as  do  all  other  contour  lines.  This  is  seen  at  once 
from  their  equation  (cf.  (5.3.21)  with  at  =  n/2) 


(5.3.23)     (pi(x,  0)  cos  kz  + 


#,  0)  sin  kz  =  const. 


Since  <p2  ->  oo  as  #  ->  0  while  9^  remains  bounded,  it  is  clear  that 
sin  kz  must  approach  zero  as  x  ->  0  on  any  such  curve.  That  the 
contours  are  all  tangent  to  the  s-axis  at  the  points  z  =  27W,  n  an 
integer,  is  also  readily  seen.  It  is  interesting  to  observe  that  the 


12 

10 

8 

6 

4 

2 

0 

•2 

-4 


^ 


-091 


.-09 

l' 

m  \ 

-  0.0 


0  I  2  3  4          x 

Fig.  5.3.2.  Level  lines  for  a  wave  approaching  a  vertical  cliff  at  an  angle 

height  of  the  wave  crest  is  lower  at  some  points  near  to  the  cliff  than 
it  is  at  oo.  It  may  be  that  the  wave  crest  is  a  ridge  with  a  number 
of  saddle  points. 


WAVES   ON   SLOPING   BEACHES   AND   PAST   OBSTACLES 


95 


It  should  be  pointed  out  that  we  are  no  more  able  to  decide  in  the 
present  case  than  we  were  in  the  two-dimensional  cases  whether 
the  waves  are  reflected  back  to  infinity  from  the  shore,  and  if  so  to 
what  extent.  Our  numerical  solution  was  obtained  on  the  assumption 
that  no  reflection  takes  place,  which  is  probably  not  well  justified 
for  the  case  of  a  vertical  cliff,  but  would  be  for  a  beach  of  small  slope. 


5.4.  Waves  on  sloping  beaches.  General  case 

We  discuss  here  the  most  general  case  of  periodic  waves  on  sloping 
beaches  which  behave  at  oo  like  an  arbitrary  progressing  wave— in 
particular,  a  wave  with  crests  at  an  arbitrary  angle  to  the  shore  line— 
and  for  a  beach  sloping  at  any  angle.  As  has  been  mentioned  earlier, 
this  problem  was  first  solved  by  Peters  and  Roseau  (cf.  the  remarks 
in  section  5.1). 

We  seek  a  harmonic  function  0(x,  y,  z;  t)  of  the  form  exp  {i(at+kz)} 
'  9>(#9  y)  in  the  region  indicated  in  cross  section  in  Figure  5.4.1.  At 


Fig.  5.4.1.  Sloping  beach  of  arbitrary  angle 

oo  the  function  0  should  behave  like  exp  {i(at+kz+vix)}  •  exp  {a2y/g} 
with  k  and  a  arbitrary.  The  function  <p(x,  y)  is  not  a  harmonic  func- 
tion, but  satisfies,  as  one  readily  sees,  the  differential  equation 

(5.4.1)  9^*  +  p™  -  *V  =  °> 

the  free  surface  condition 

a2 

(5.4.2)  cpy  —  m<p  =  0,         y  =  0,         m  =  — , 


96  WATER  WAVES 

and  the  condition  at  the  bottom* 

(5.4.3)  <pn  =  0,         y  =  —  x  tan  co. 

By  introducing  (as  we  have  done  before)  the  new  dimensionless 
quantities  xl  =  mx,  yl  =  my,  ax  =  a/w,  A?!  =  k/m  the  conditions  of 
the  problem  for  <p(x,  y)  can  be  put  in  the  form 

(5.4.1)!  <p9X  +  <pyy  —  k*<p  =  0,         0  ^  fc  ^  1, 

(5.4.2)!  <pv  -  9?  =  0,         y  =  0, 

(5.4.3)!  y>n  =  0,         y  =  —  x  tan  o> 

after  dropping  subscripts.  Since  we  require  99(0?,  t/)  to  behave  like 
ei&x  emy  —  exp  {IK^XI  +  t/J  at  oo,  it  follows  from  (5.4.1)  that 
—  a2  +  m2  —  &2  =  0  and  hence  that  a?  +  A;*  =  1.  Thus  fc  in  (5.4.1)! 
(really  it  is  A^)  is,  as  indicated,  restricted  to  the  range  0  ^  k  ^  1, 
and  this  fact  is  of  importance  in  what  follows.**  Finally,  we  know 
from  past  experience  that  a  singularity  must  be  permitted  at  the 
origin.  (In  the  problems  treated  earlier  in  this  chapter  we  have 
prescribed  only  boundedness  conditions  at  oo  in  a  way  which  led  to 
a  statement  concerning  the  uniqueness  of  the  solution.  In  the  present 
case  we  do  not  obtain  a  similar  uniqueness  theorem— in  fact,  as  has 
been  pointed  out  by  Ursell  [U.7,  8],  Stokes  showed  that  there  exist 
motions  different  from  the  state  of  rest  and  which  die  out  at  oo. 
For  these  motions,  however,  the  quantity  k  is  larger  than  unity). 
We  seek  functions  (p(x,  y)  satisfying  the  above  conditions  as  the 
real  or  the  imaginary  part  of  a  complex  function  f(z,  z)  which  is 
analytic  in  each  of  the  variables  z  =  x  +  iy  and  its  conjugate 
z  =  x  —  iy.  In  the  two-dimensional  cases,  it  was  sufficient  to  consider 
analytic  functions  f(z)  of  one  complex  variable,  but  in  the  present 
case  it  is  necessary  to  take  more  general  functions  since  q>(x9  y)  is 
not  a  harmonic  function.  Note  that  we  now  use  the  variable  z  in  a 
different  sense  than  above,  where  it  is  one  of  the  space  variables; 
no  confusion  should  result  since  the  space  variable  z  hardly  occurs 
again  in  the  discussion  to  follow.  It  is  useful  to  calculate  some  of  the 
derivatives  of  such  functions  with  respect  to  x  and  y;  we  have, 
clearly: 

*  Peters  [P.6]  solves  the  problem  when  the  condition  (5.4.3)  is  replaced  by 
the  more  general  mixed  boundary  condition  <pn  +  a<p  —  0,  a  =  const. 

**  Involved  in  this  remark  is  the  assumption  that  the  derivatives  of  the  solution 
btehave  asymptotically  the  same  as  the  derivatives  of  its  asymptotic  development; 
but  this  is  indeed  the  case,  as  we  could  verify  on  the  basis  of  our  final  represen- 
tation of  the  solution. 


WAVES    ON    SLOPING    BEACHES    AND    PAST    OBSTACLES  97 

fx  =/**x  +/!««  =/,  +/z, 
fy  =  i(h  -  /i), 
ABB  +  fyy  =  4/«Z  • 

Consequently  our  differential  equation  (5.4.1)!  can  be  replaced  by 
the  differential  equation 


since  the  real  or  the  imaginary  part  of  any  solution  of  it  is  clearly 
a  solution  of  (5.4.1)!. 

Among  the  solutions  of  the  last  equation  are  the  following  simple 
special  solutions  (obtained,  for  example,  by  separating  the  variables 
in  writing  /  =  f^z)  •  /,(«)): 

kz  2 

f(z,  z)  =  C&  +  T  c  ,         f  =  const., 
which,  when  £  —  —  i  for  example,  is  of  the  form 

Kkz\     \  [         (         k2\     \ 

I  +  —  I  y  !  •  exp  |  —  i  1  1  —  —  )  x  >  ,  and  this  is  a  solution  of 
4/     ]  {         \          4/     j 

(5.4.1  )l  which  has  the  proper  behavior  at  oo,  at  least.  (Actually, 
when  combined  with  the  factor  eikz9  with  z  once  more  the  space 
variable,  the  result  is  a  harmonic  function  yielding  a  plane  wave  in 
water  of  infinite  depth  and  satisfying  the  free  surface  condition). 
One  can  obtain  a  great  many  more  solutions  by  multiplying  the 
above  special  solution  by  an  analytic  function  g(£)  and  integrating 
along  a  path  P  in  the  complex  £-plane: 

(5.4.4)  /(*,  5)  -  ±-.  {  ez^  +  T  f  •  g(C)d£. 

'2m  J  P 

By  appropriate  choices  of  the  analytic  function  g(£)  and  the  path  P, 
we  might  hope  to  satisfy  the  boundary  conditions  and  the  condition 
at  oo.  This  does,  indeed,  turn  out  to  be  the  case. 

Still  another  way  to  motivate  taking  (5.4.4)  as  the  starting  point 
of  our  investigation  is  the  following.  It  would  seem  reasonable  to 
look  for  solutions  of  (5.4.1)  in  the  form  of  the  exponential  functions 
<p  =  exp  {mx  +  ly}.  However,  since  we  wish  to  work  with  analytic 
functions  of  complex  variables  it  would  also  seem  reasonable  to  express 
x  and  y  in  terms  of  z  =  x  +  iy  and  z  =  x  —  iy,  and  this  would  lead  to 

(/%  -4-  z\  /Z  _  Z\  \ 

m  j     ___*!  I  —  U  [  -  1  1  .    In    order    that    this   function 
\     2     /  \     2     /) 


98 


WATER   WAVES 


(which  is  clearly  analytic  in  z  and  z  separately)  be  a  solution  of 
(5.4.1)!  we  must  require  that  m*  -f-  Z2  —  k2  =  0,  and  this  leads  at 

{k2  z\ 
£z  + 1 ,  with  £  an  ar- 
4  £j 

bitrary  parameter,  as  one  can  readily  verify.  The  method  used  by 
Peters  [P.6]  to  arrive  at  a  representation  of  the  form  (5.4.4)  is  better 
motivated  though  perhaps  more  complicated,  since  he  operates  with 
(5.4.1)  in  polar  coordinates,  applies  the  Laplace  transform  with 
respect  to  the  radius  vector,  transforms  the  resulting  equation  to 
the  Laplace  equation,  and  eventually  arrives  at  (5.4.4). 

One  of  the  paths  of  integration  used  later  on  is  indicated  in  Figure 
5.4.2.  The  essential  properties  of  this  parth  are:  it  is  symmetrical 
with  respect  to  the  real  axis,  goes  to  infinity  in  the  negative  direction 


£-  plane 


Fig.  5.4.2.  The  path  P  in  the  f -plane 

of  the  real  axis,  enters  the  origin  tangentially  to  the  real  axis  and 
from  the  left,  and  contains  in  the  region  lying  to  the  left  of  it  a 
number  of  poles  of  g(£).  (The  path  is  assumed  to  enter  the  origin 
in  the  manner  indicated  so  that  the  term  z/£  in  the  exponential 
factor  will  not  make  the  integral  diverge).  Our  discussion  will  take 
the  following  course:  We  shall  assume  g(£)  to  be  defined  in  the  £-plane 
slit  along  the  negative  real  axis  (and  also  on  occasion  on  a  Riemann 
surface  obtained  by  continuing  analytically  over  the  slit).  The  choice 
of  the  symmetrical  path  P  leads  to  a  functional  equation  for  g(£) 
through  use  of  the  boundary  conditions  (5.4.2)!  and  (5.4.3)!,  and  vice 
versa  a  solution  g(£)  of  the  functional  equation  leads  to  a  function 


WAVES    ON    SLOPING   BEACHES    AND    PAST   OBSTACLES  99 


9>(#>  y)  =  ^te  /(*»*)  satisfying  the  boundary  conditions.  (By  the 
symbols  Jm  and  Ste  we  mean,  of  course,  that  the  imaginary,  or  real, 
part  of  what  follows  is  to  be  taken.  )  We  seek  a  solution  of  the  functional 
equation  which  is  defined  and  regular  in  the  slit  £-plane,  with  at 
most  poles  in  the  left  half-plane  (including  certain  first  order  poles 
on  the  negative  imaginary  axis),  and  dying  out  at  oo  like  1/f.  Once 
such  a  function  has  been  found,  the  prescribed  conditions  at  oo  will 
be  seen  to  follow  by  deforming  the  path  P  over  the  poles  into  a  path 
on  the  two  edges  of  the  slit  along  the  negative  real  axis:  the  residues 
at  the  poles  on  the  negative  imaginary  axis  clearly  would  yield  con- 
tributions of  the  type 

{k2      z      \ 
—  irz  H  ----  ~»     }  ,  r  >  0,  which  are  easily  seen  to  be 
4  (-tr)J 

of  the  desired  type  at  oo,  while  the  remaining  poles  and  the  integral 
over  the  deformed  path  will  be  found  to  yield  contributions  that  tend 
to  zero  when  3te  z  ->  +  oo. 

We  begin  this  program  by  expressing  the  boundary  conditions 
(5.4.2  )j  and  (5.4.3  )t  in  terms  of  the  function  f(z,  z).  The  first  of 
these  conditions  will  be  satisfied  if  the  following  condition  holds: 

(5.4.2)i  ^M(fz  —  /5  +  if)  =  0,         z  real,  positive, 

as  one  readily  sees.  The  condition  (5.4.3)1  will  be  satisfied  if 
n  •  grad  <p  =  0,  with  n  the  unit  normal  at  the  bottom  surface,  i.e. 
if  Ste  {n  •  grad  /}  =  0,  and  the  latter  is  given  by 

&t  {(/«  +  /•)  sin  <*>  +  {(fz  -  fz)  cos  «>}  =  0, 
or  finally,  in  the  form 

(5.4.3);       Sm  {fze-i(0  -  /-<?ift>}  -  0,         z  =  re~ta>9         r  >  0. 
Upon  making  use  of  (5.4.4)  in  (5.4.2)(  the  result  is 

(5.4.5)  Sm  ~  f  c*t  +  *  •  [C  -  £  +  tl  g(C)C  -  0, 

2m  Jp  I         4£        J 

z  real,  positive, 
while  (5.4.3  )[  yields 

(5.4.6)  Jm  —.  f  e*  +  ?  •  [V**  ~  ^  e^  g(C)dC  =  0, 

2m  Jp  I  4C      J 

z  =  re-*",        r  >  0. 

To  satisfy  the  boundary  condition  (5.4.5)  it  is  sufficient  to  require 
that  g(f  )  satisfies  the  condition 


100  WATER   WAVES 

r      k2      "i 

(5.4.7)  «//»£__+!    g(£)  =  0,         f  real,  positive. 

The  proof  is  as  follows:  If  (5.4.7)  holds,  then  the  integrand  G(z,  z,  £) 
in  (5.4.5)  is  real  for  real  z  and  real  positive  f  .  Hence  G  takes  on  values 
G,  G  at  conjugate  points  £,  £  which  are  themselves  conjugate,  by  the 
Schwarz  reflection  principle.  Since  the  path  P  is  symmetrical,  as 
shown  in  Figure  5.4.2,  it  follows  that  d£  takes  on  values  at  £,  £  that 

are  negative  conjugates.  Thus  the  integral  (1  /2m)  \   G  d£  is  real  when 

z  is  real  and  (5.4.7)  holds.  In  considering  next  (5.4.6)  we  first  introduce 
a  new  variable  s  =  £e~ia)  to  obtain  for  z  =  re~i<0  the  condition. 
replacing  (5.4.6): 

ir       _L  —  r      k2~\ 

(5.4.8)  Jm  —        e     ^  **  •  \*  --    g(seia>)  eiu>  ds  =  0,          r  real. 

2ni  J  p/  L         4*J 

Here  P'  is  the  path  obtained  by  rotating  P  (and  the  slit  in  the 
£-plane  as  well,  of  course)  clockwise  about  the  origin  through  the  angle 
a).  If  g  behaves  properly  at  oo,  and  if  the  rotation  of  P'  can  be 
accomplished  without  passing  over  any  poles  of  the  integrand,  we 
may  deform  P'  back  to  P  and  obtain 

(5.4.8)'      Jm—{  *r*  +  ~T  •  \s  -  —1  g(seim)  ei(0  ds  =  0,         r  real. 
p  L         4*J 


By  the  same  argument  as  before  we  now  see  that  the  condition 
(5.4.6)  will  be  satisfied  provided  that  g(£)  satisfies  the  condition 

(5.4.9)  Jmg^e™}?*  =  0,         f  real,  positive. 

Thus  if  the  function  g(£)  satisfies  the  conditions  (5.4.7)  and 
(5.4.9),  the  function  f(z9  z)  constructed  by  its  aid  will  satisfy  the 
boundary  conditions.  As  we  have  already  remarked,  g(£)  must 
satisfy  still  other  conditions  —at  oo,  for  example.  In  addition,  we 
know  from  earlier  discussions  in  this  and  the  preceding  chapter  that 
it  is  necessary  to  find  two  solutions  (p(x,  y)  and  <pi(x9  y)of  our  problem 
which  are  "out  of  phase  at  oo",  in  order  that  a  linear  combination 
of  them  with  appropriate  time  factors  will  lead  to  a  solution  having 
the  form  of  an  arbitrary  progressing  wave  at  oo.  In  this  connection 
we  observe  that  if  the  path  Pl  of  integration  (as  shown  in  Figure 
5.4.3)  is  taken  instead  of  the  path  P  (it  differs  from  P  only  in  reversal 
of  direction  of  the  portion  in  the  upper  half-plane),  and  if  we  define 
<pi(x,  y)  as  the  imaginary  part  of  /1(^,  z)  instead  of  its  real  part: 


WAVES    ON    SLOPING    BEACHES    AND    PAST   OBSTACLES 


101 


x9  y)  =  Jm  G(z9  2,  £)d£  = 


(5.4.10) 


with  G  the  same  integrand  as  before,  then  <pi(x,  y),  by  the  same 
argument  as  above,  will  satisfy  the  boundary  conditions  provided 
that  the  function  g(£)  also  in  this  case  satisfies  the  conditions  (5.4.7) 


-  plane 


Fig.  5.4.3.  The  path  Pl  in  the  f  -plane 

and  (5.4.9).  It  seems  reasonable  to  expect  that  the  integral  over  Pl 
will  behave  the  same  as  the  integral  over  P  when  @te  z  is  large  and 
positive  (since  the  poles  in  the  lower  half-plane  alone  determine  this 
behavior  and  the  paths  P  and  Pl  differ  only  in  the  upper  half- 
plane)  except  that  a  factor  i  will  appear,  and  hence  that  y>  and  <pl 
will  differ  in  phase  at  +  oo  (in  the  variable  #,  that  is)  by  90°.  This 
docs  indeed  turn  out  to  be  the  case. 

Thus  to  satisfy  the  boundary  conditions  for  both  types  of  standing 
wave  solutions  we  have  only  to  find  a  function  g(£)  satisfying  the  con- 
ditions (5.4.7)  and  (5.4.9)  which  behaves  properly  at  oo—  the  last 
condition  being  needed  in  order  that  the  path  of  integration  can  be 
rotated  in  the  manner  specified  in  deriving  (5.4.8)'.  To  this  end  we 
derive  a  functional  equation  for  g(£)  by  making  use  of  these  con- 
ditions. From  (5.4.7)  we  have,  clearly: 


(5.4.11)     U  ~  ~  +  i\  g(0  = 
while  from  (5.4.9)  we  have 


~  ^  -  *')  «(?)• 


real>  Positive, 


102  WATER   WAVES 


(5.4.12)          g(C)*-ia)  =  g(C*2ia>''a>,         £  ^al,  positive, 


both  by  virtue  of  the  reflection  principle.  Eliminating  g(£)  from  the 
two  equations  we  obtain 

(5.4.13) 

This  functional  equation  was  derived  for  £  real  and  positive,  but 
since  g(£)  is  analytic  it  is  clear  that  it  holds  throughout  the  domain 
of  regularity  of  g(£);  it  ig  the  basic  functional  equation  for  g(£),  a 
solution  of  which  will  yield  the  solution  of  our  problem.  Of  course, 
this  equation  is  only  a  necessary  condition  that  must  be  fulfilled  if 
the  boundary  conditions  are  satisfied;  later  on  we  shall  show  that  the 
solution  of  it  we  choose  also  satisfies  the  condition  (5.4.11  ),  and  hence 
the  condition  (5.4.12)  will  also  be  satisfied  since  (5.4.13)  holds, 

We  proceed  now  to  find  a  solution  g(£)  of  (5.4.13)  which  has  all  of 
the  desired  properties  needed  to  identify  (5.4.4)  and  (5.4.10)  as 
functions  furnishing  the  solution  of  our  problem,  as  has  been  done  by 
Peters  in  the  paper  cited  above. 

We  therefore  proceed  to  treat  the  functional  equation  (5.4.13), 
which  is  easily  put  in  the  form: 

<,teZi<»n  £2      I     ,/£   __  *! 

(5.4.14)  g(  *<"  4 

{       ) 


(C  -  «>i)(£  -  »> 


with  r1>2  -  --  -  -  . 

The  numbers  rlt2  are  real  since  we  know  that  k  lies  between  0  and  1. 
It  is  convenient  to  set 

(5.4.15)  ,(f,  -  *g> 

(C  +  Wi)(t  +  ^^•2) 

in  which  h(£),  like  g(£),  is  defined  in  the  £-plane  slit  along  the  negative 
real  axis.  The  function  &(£)  will  have  poles  in  the  left  half-plane,  but 
only  the  poles  at  f  =  —  irl  and  f  =  —  ir2  of  g(C)  will  be  found  to 
contribute  a  non-  vanishing  residue  of  f(z)  for  die  z  ->  +  oo,  and 
"this  in  turn  would  guarantee  that  f(z)  behaves  at  GO  on  the  free 
surface  like  Ae~**.  For  A(£)  we  have  from  (5.4.14)  and  (5.4.15)  the 
equation 


WAVES    ON    SLOPING    BEACHES    AND    PAST   OBSTACLES  103 


,.41ftv  2  ,n 

(5.4.16)  -  =  -  =  m(L). 

MO  (f-friXf-fr,)  k  ' 

This  equation  is  solved  by  introducing  the  function  /(£)  by 

(5.4.17)  log  A(C)  =  l(£), 

and  one  finds  at  once  that  /(£)  satisfies  the  difference  equation 

(5.4.18)  (      l(e**»t)  -  /(C)  -  log  m(C)  =  w(f). 

In  solving  this  equation  we  shall  begin  by  producing  a  solution 
free  of  singularities  in  the  sector  —  co  ^  arg  £  ^  co,  after  which  the 
function  /&(£)  —  which  is  (cf.  (5.4.17))  then  also  regular  in  the  same 
sector—  can  be  continued  analytically  into  the  whole  £-plane  slit 
along  the  negative  real  axis  (or,  if  desired,  into  a  Riemann  surface 
having  the  origin  as  its  only  branch  point)  by  using  (5.4.16).  As  an 
aid  in  solving  equation  (5.4.18)  we  set 

\co  =  OCTT,         0  <  a  ^  1, 


(5.4.19) 

V  1C  =  Ta,  *(*")  -  L(r),         Mr")  =  W(r)9 


and  operate  now  in  a  r-plane.  One  observes  that  the  sector  —  co 
<  arg  f  <  co  in  the  C-plane  corresponds  to  the  r-plane  slit  along 
its  negative  real  axis.  For  L(r)  one  then  finds  at  once  from  (5.4.18) 
the  equation 

(5.4.20)  L(re2ni)  -  L(r)  =  W(r). 

Our  object  in  putting  the  functional  equation  into  this  form 
(following  Peters)  is  that  a  solution  is  now  readily  found  by  making 
use  of  the  Cauchy  integral  formula.  Let  us  assume  for  this  purpose 
that  L(r)  is  an  analytic  function  in  the  closed  r-plane  slit  along  its 
negative  real  axis*  (which  would  imply  that  l(£)  is  regular  in 
the  sector  —  o>  ^  arg  f  ^  co,  as  we  see  from  (5.4.19));  in  such  a 
case  L(r)  can  be  represented  by  the  Cauchy  integral  formula: 

(5.4.21)  L(r)=  -L 


with  C  the  path  in  the  {-plane  indicated  by  Figure  5.4.4.  If  we 
suppose  in  addition  that  L(£)  dies  out  at  least  as  rapidly  as,  say, 
l/{  at  oo,  it  is  clear  that  we  can  let  the  radius  R  of  the  circular 
part  of  C  tend  to  infinity,  draw  the  path  of  integration  into  the  two 
edges  of  the  slit  and,  in  the  limit,  find  for  L(r)  the  representation 

We  shall  actually  produce  such  a  regular  solution  shortly. 


104  WATER   WAVES 


(-plane 


Fig.  5.4.4.  Path  C  in  the  {-plane 


(5-4.22)  L(r)  = 

2m  J       2m 

—  > 

in  readily  understandable  notation.  On  making  use  of  (5.4.20),  and 
drawing  the  two  integrals  together,  it  is  readily  seen  that  L(r)  is 
given  by 


(5.4.23)  L(r)  =  —  .          --        d£. 

2jwJ_aof  —  r 

The  path  of  integration  is  the  negative  real  {-axis,  and  W(£)  is  to 
be  evaluated  for  arg{  —  —  n.  Since  W(£)  has  no  singularities  (of. 
(5.4.18)),  it  follows  that  L(r)  as  given  by  (5.4.23)  is  indeed  regular 
in  the  slit  r-plane.  L(r)  also  has  no  singularity  on  the  slit  except 
at  the  origin,  where  it  has  a  logarithmic  singularity.  Since  the 
numerator  in  the  integrand  behaves  like  l/{a,  a  >  0,  at  oo  (cf. 
(5.4.19),  (5.4.18),  (5.4.16)),  it  is  clear  that  the  function  L(r)  dies 
out  like  I  /r  at  oo  in  the  r-plane.  This  function  therefore  has  all  of 
the  properties  postulated  in  deriving  (5.4.23)  from  (5.4.21),  and 
hence  is  a  solution  of  the  difference  equation  (5.4.20)  in  the  slit 
plane  including  the  lower  edge  of  the  slit. 

A  solution  of  (5.4.18)  can  now  be  written  down  through  use  of 
(5.4.19);  the  result  is: 

with  w(£a)  to  be  evaluated  for  arg  {  =  —  n.  This  solution  is  valid 
so  far  only  for  f  in  the  sector  —  a>  ^  arg  £  ^  co,  where  it  is  regular, 


WAVES    ON    SLOPING   BEACHES    AND    PAST   OBSTACLES  105 

as  we  know  from  the  discussion  above.  However,  it  is  necessary  to 
define  the  function  A(£)  =  el(^  (cf.  (5.4.17))  in  the  entire  slit  £- 
plane,  and  this  can  be  done  by  analytic  continuation  with  the  aid 
of  the  functional  equation  (5.4.16).  In  the  process  of  analytic  con- 
tinuation, starting  with  the  original  sector  in  which  /(£),  and  hence 
A(£),  is  free  of  singularities,  one  sees  that  the  only  singularities  which 
could  occur  in  continuing  into  the  upper  half-plane,  say,  would  arise 
from  the  function  on  the  right  hand  side  of  the  equation  (5.4.16). 
The  only  singularities  of  this  function  occur  obviously  at  £  =  irl  2. 
Consequently  no  singularity  of  A(£)  appears  in  the  analytic  con- 
tinuation into  the  upper  half-plane,  through  widening  of  the  sector 
in  which  &(£)  is  defined,  until  the  points  £  =  irl  and  £  =  ir2  have 
been  covered,  and  one  sees  readily  from  (5.4.16)  that  the  first  such 
singularities  of  /i(£)  —poles  of  first  order— -appear  at  the  points 
ri,2  exP  {*(2co  +  7T/2)},  the  next  at  r±  2  exp  (i(4co  +  ?r/2)},  etc., 
though  some,  or  all,  of  these  poles  may  not  appear  on  the  first  sheet 
of  the  slit  f -plane,  depending  on  the  value  of  the  angle  co.  The  con- 
tinuation into  the  lower  half-plane  is  accomplished  by  writing 
(5.4.16)  in  the  equivalent  form 

(5.4.16)' 


Again  we  see  that  poles  will  occur  in  the  lower  half-plane  in  the 
course  of  the  analytic  continuation,  this  time  at  rl%2  exp  {-—i(2co  +JT/2} 
ri,2  CXP  {—i(4a)  +yt/2)}9  etc.  The  situation  is  indicated  in  Figure 
5.4.5;  /i(£)  lacks  the  singularities  of  g(£)  at  the  points  —  ?>!  and 
—  ?V2  (cf.  (5.4.15)).  Thus  the  function  A(£)  is  defined  in  the  slit  £- 
plane.  (It  can  also  be  continued  analytically  over  the  slit  which 
permits  a  rotation  of  the  path  of  integration.)  We  see  that  /«(£) 
may  have  poles  in  the  open  left*  half-plane,  on  two  circles  of  radii 
r1  and  r2,  but  the  poles  closest  to  the  imaginary  axis  are  at  the 
angular  distance  2co  from  it.  There  is  also  a  simple  pole  of  A(£)  at 
the  origin,  but  g(£)  (cf.  (5.4.15))  is  regular  there. 

The  behavior  of  /i(£)  at  oo  in  the  slit  plane  is  now  easily  discussed: 
In  the  original  sector  we  know  from  (5.4.24)  that  /(£)  dies  out  at  oo 
like  l/£1/a.  Hence  A(£)  =  el(^  is  bounded  in  the  sector,  and  since  the 
right  hand  side  of  (5.4.16)  is  clearly  bounded  at  oo  it  follows  that 
A(£)  is  bounded  at  oo  in  the  £-pUme. 

The  function  g(£)  =  £*(£)/(£  +  fViHf  +  trt)  (cf.  (5.4.15))  can  now 
be  seen  to  have  all  of  the  properties  needed  to  identify  the  functions 


106 


WATER   WAVES 


{-plane 


Fig.  5.4.5.  The  singularities  of  /i(£)  and 


/(*)  in  (5.4.4)  and  f^z)  in  (5.4.10)  as  functions  whose  real  part  and 
imaginary  part,  respectively,  yield  the  desired  standing  wave  solutions 
of  our  problem.  To  this  end  we  write  down  the  integrals 


(5.4.25) 


-i*L' 


§) . 

(C 


over  the  paths  indicated  in  Figure  5.4.6,  where  the  direction  is  in- 
dicated only  on  the  part  of  the  path  in  the  lower  half-plane,  since 
the  paths  P9  Pl  differ  only  in  the  direction  in  which  the  remainder 
of  the  path  is  traversed. 

Since  /&(£)  is  bounded  at  oo,  and  0te  z  >  0,  the  integrals  clearly 
converge.  One  sees  also  that  the  paths  of  integration  can  be  rotated 
through  the  angle  co  about  the  origin  without  passing  over  singularities 
of  the  integrand,  and  also  without  changing  the  value  of  the  in- 
tegrals. (This  was  needed  in  deriving  (5.4.8)'.)  We  prove  next  that 
g(£)  satisfies  the  boundary  condition  (5.4.11).  To  begin  with,  we  shall 
show  that  J(£)  as  defined  by  (5.4.24)  is  real  when  £  is  real  and  positive. 
Once  this  is  admitted  to  be  true,  then  h(£)  as  given  by  (5.4.17)  would 
have  the  same  property,  and  the  function  g(£)  defined  by  (5.4.15) 
would  easily  be  seen  to  satisfy  the  condition  (5.4.11).  We  have,  then, 
only  to  show  that  /(£)  is  real  for  real  £,  and  this  can  be  seen  as  follows: 


WAVES    ON    SLOPING    BEACHES    AND    PAST   OBSTACLES  107 


-  plone 


Fig.  5.4.0.  The  paths  P,  P^  in  the  C-plane 


111  (5.4.21)  log  w(£a)  is  to  be  evaluated  for  arg  £  =  —  jr.  But  in  this 
case  one  sees  easily  from  the  equation  (5.4.16)  defining  w(fa)  (with 
a  =  ro/jr,  ef.  (5.4.19))  that  m(^a)  has  its  values  on  the  unit  eircle 
when  arg  £  —  —  rr,  and  hence  its  logarithm  is  pure  imaginary  on 
the  path  of  integration;  it  follows  at  once  from  (5.4.24)  that  /(£) 
is  real  for  £  real  and  positive.  Since  g(C)  was  constructed  in  such  a 
way  as  to  satisfy  (5.4.13)  we  know  that  (5.4.12)  is  satisfied  auto- 
matically. Thus  our  standing  wave  solutions  satisfy  the  boundary 
conditions. 

Finally,  we  observe  that  the  behavior  of  /  and  fl  for  &e  z  ->  oo 
is  what  was  prescribed.  To  this  end  we  deform  the  path  of  integration 
into  a  path  running  along  the  two  banks  of  the  slitted  negative 
real  axis.  The  residues  at  £  =  —  ir^  2  contribute  terms  already 
discussed  above  which  furnish  the  desired  behavior  for  3te  z  ->  +  oo. 
We  have,  then,  only  to  make  sure  that  the  residues  at  the  remaining 
poles  and  the  integrals  along  the  slit  make  contributions  which  die 
out  as  3te  z  ->  +  oo.  As  for  the  residues  at  the  poles  at  the  points 


108  WATER    WAVES 

£w  =  r1>2  exp  {±  i(2nco  +  rc/2)},  n  =  1,  2,  .  .  .,  we  observe  that  these 
contributions  are  of  the  form  Aez^n9  but  since  —  co  ^  arg  z  ^  0  it 
is  clear  that  these  contributions  die  out  exponentially  when  z  tends 
to  infinity  in  the  sector  —  CD  ^  arg  2  fg  0.  As  for  the  integrals  along 
the  slit,  they  are  known  to  die  out  like  1/2,  as  we  have  seen  in  similar 
cases  before,  or  as  one  can  verify  by  integration  by  parts.  Thus  all 
of  the  conditions  imposed  on  f(z)  and  f^z)  are  seen  to  be  satisfied. 
We  observe,  however,  that  the  integrals  in  (5.4.25)  over  the  paths 
P  and  Pl  converge  only  if  9&e  z  ^  0,  and  hence  this  representation 
of  our  solution  is  valid  only  if  the  bottom  slopes  down  at  an  angle 
^  jr/2.  For  an  overhanging  cliff,  when  co  >  n/29  the  solution  can  be 
obtained  by  first  swinging  the  path  of  integration  clockwise  through 
90°  (and  swinging  the  slit  also,  of  course);  the  resulting  integrals 
would  then  be  valid  for  all  z  such  that  ^m  z  ^  0  and  the  solutions 
would  hold  for  0  <  co  ^  n. 

It  is  perhaps  of  interest  to  bring  the  final  formulas  together  for 
the  simplest  special  case,  i.e.  the  dock  problem  for  two-dimensional 
motion  (first  solved  by  Friedrichs  and  Lewy  [F.12]),  in  which  the 

y 


Fig.  5.4.7.  The  dock  problem 

angle  co  has  the  value  n,  as  indicated  in  Figure  5.4.7.*  In  this  case 
the  function  /(£)  is  given  by 

(5.4.26) 

and  the  integral  defines  it  at  once  in  the  entire  slit  £-plane.  The 
standing  wave  solutions  <p(x,  y)  —  3le  f(z)  and  y1  (#,  y)  =Sm  f^z) 
are  determined  through 

*  As  was  mentioned  in  section  5.1,  the  dock  problem  in  water  of  uniform 
finite  depth  and  for  the  three-dimensional  case  was  first  solved  by  Heins  [H.I 3] 
with  the  aid  of  the  Wiener-Hopf  technique. 


WAVES   ON    SLOPING   BEACHES   AND    PAST   OBSTACLES 


109 


(5.4.27)  f(z)  =  ~  |    e^ 

i     r     . 

(5.4.28)  /t(z) 

with  A(C)  defined  by 
(5.4.29) 

As  was  remarked  above,  the  integrals  in  (5.4.27)  and  (5.4.28)  con- 
verge only  if  &e  z  ^  0.  However,  the  analytic  continuation  into  the 
entire  lower  half-plane  is  achieved  simply  by  swinging  the  paths 
P  and  P!  into  the  positive  imaginary  axis  (which  can  be  done  since 
A(C)  is  bounded  at  oo),  while  staying  on  the  Riemann  surface  of  A(£), 
and  these  integrals  are  then  valid  for  all  z  in  the  lower  half-plane. 
Finally,  it  is  also  of  interest  to  remark  that  the  functions  f(z) 
and  f^z)  do  not  behave  in  the  same  way  at  the  origin:  the  first  is 
bounded  there,  and  the  second  is  not,  and  this  behavior  holds  not 
only  for  the  spcciaj  case  of  the  dock  problem,  but  also  in  all  cases 
under  consideration  here. 


5.5.  Diffraction  of  waves  around  a  vertical  wedge. 
Sommerfeld's  diffraction  problem 

In  this  section  we  are  primarily  concerned  with  the  problem  of 
determining  the  effect  of  a  barrier  in  the  form  of  a  vertical  rigid 
wedge,  as  indicated  in  Fig.  5.5.1,  on  a  plane  simple  harmonic  wave 


y 


i 


Fig.  5.5.1.  Diffraction  of  a  plane  wave  by  a  vertical  wedge 

coming  from  infinity.  In  this  case  it  is  convenient  to  make  use  of 
cylindrical  coordinates  (r,  0,  y).  We  seek  a  harmonic  function 


110  WATER   WAVES 

&(r>  0>  y\  t)  in  the  region  0  <  0  <v,  —  h  <y  <Q,  i.e.  in  the  region 
exterior  to  the  wedge  of  angle  2n  —  v  and  in  water  of  finite  depth  h 
when  at  rest.  The  problem  is  reduced  to  one  in  the  two  independent 
variables  (r,  6)  by  setting 

(5.5.1)  <2>(r,  0,  yi  t)  =  /(r,  0)  cosh  m(y  +  h)eiot. 

The  boundary  conditions  0e  =  0  for  0  =  0,  6  =  v  corresponding 
to  the  rigid  walls  of  the  wedge  yield  for  /(r,  0)  the  boundary  con- 
ditions 

(5.5.2)  f0  =  0,         0-0,         Q=v. 

The  free  surface  condition  g0y  +  0tt  =  0  at  y  =-  0  yields  the  con- 
dition 

(5.5.3)  m  tanh  mh  —  o2/g, 

while  the  condition  0y  =  0  at  the  bottom  t/  =  —  A  is  satisfied 
automatically.  Once  any  real  value  for  the  frequency  a  is  prescribed, 
equation  (5.5.3)  is  used  to  determine  the  real  constant  m  —  which 
will  turn  out  to  be  the  wave  number  of  the  waves  at  oo  — ,  and  we 
note  that  (5.5.3)  has  exactly  one  real  solution  of  in  except  for  sign; 
if  the  water  is  infinitely  deep  we  have  in  —  <72/#,  and  the  function 
cosh  m(y  +  h)  in  (5.5.1)  is  replaced  by  emv. 

Thus  the  function  /(r,  0)  is  to  be  determined  as  a  solution  of  the 
reduced  wave  equation 

(5.5.4)  V^0)/  +  w2/  =  0,         0  <  r  <  oo,         0  <  0  <  v, 

subject  to  the  boundary  conditions  (5.5.2).  Actually,  we  shall  in  the 
end  carry  out  the  solution  in  detail  only  for  the  case  of  a  reflecting 
rigid  plane  strip  (i.e.  for  the  special  case  v  —  2jr),  but  it  will  be  seen 
that  the  same  method  would  furnish  the  result  for  any  wedge.  It  is 
convenient  to  introduce  a  new  independent  variable  p,  replacing  r, 
by  the  equation  r  =  Q/m;  in  this  variable  equation  (5.5.4)  has  the 
form 

(5.5.5)  V^0)/+/  =  0,         0<e<oo,         0<0<*, 

and  we  assume  this  equation  as  the  basis  for  the  discussion  to  follow. 
So  far  we  have  not  formulated  conditions  at  oo,  except  for  the 
vague  statement  that  we  want  to  consider  the  effect  of  our  wedge- 
Shaped  barrier  on  an  incoming  plane  wave  from  infinity.  Of  course, 
we  then  expect  a  reflected  wave  from  the  barrier  and  also  diffraction 
effects  from  the  sharp  corner  at  the  origin.  In  conformity  with  our 


WAVES    ON    SLOPING   BEACHES   AND    PAST   OBSTACLES  111 

general  practice  we  wish  to  formulate  these  conditions  at  oo  in  such 
a  way  that  the  solution  of  the  problem  will  be  uniquely  determined. 
It  has  some  point  to  consider  the  question  of  reasonable  conditions 
at  oo  which  determine  unique  solutions  of  the  reduced  wave  equation 
under  more  general  circumstances  than  those  considered  in  the 
physical  problem  formulated  above.  For  general  domains  it  is  not 
known  how  to  formulate  these  conditions  at  oo,  and,  in  fact,  it  would 
seem  to  be  a  very  difficult  task  to  do  so  since  such  a  formulation 
would  almost  certainly  require  consideration  of  many  special  cases. 
In  one  special  case,  however,  the  appropriate  condition  to  be  im- 
posed at  infinity  has  been  known  for  a  long  time.  This  is  the  case 
in  which  any  reflecting  or  refracting  obstacles  lie  in  a  bounded  domain 
of  the  plane,  or,  stated  otherwise,  it  is  the  case  in  which  a  full  neigh- 
borhood of  the  point  at  infinity  is  made  up  entirely  of  the  homogeneous 
medium  in  which  the  waves  propagate.  In  this  case,  the  condition 
at  oo  which  determines  the  "secondary"  waves  uniquely  is  Sommer- 
feld's  radiation  condition,  which  states,  roughly  speaking,  that  these 
waves  behave  like  a  cylindrical  outgoing  progressing  wave  at  oo. 
However,  if  the  reflecting  or  refracting  obstacles  extend  to  infinity, 
the  Soinmorfeld  condition  may  not  be  appropriate  at  all.  Consider, 
for  example,  the  case  in  which  the  entire  tT-axis  is  a  reflecting  barrier 
(i.e.  the  case  v  —  JT),  and  the  primary  wave  is  an  incoming  plane  wave 
from  infinity.  It  is  clear  on  physical  grounds  that  the  secondary  wave 
will  be  the  reflected  plane  wave,  which  certainly  does  not  behave 
at  oo  like  a  cylindrical  wave  since,  for  example,  its  amplitude  does 
not  even  tend  to  zero  at  oo.  Another  case  is  that  of  Sommerfeld's 
classical  di (Traction  problem  in  which  an  incoming  plane  wave  is 
reflected  from  a  barrier  consisting  of  the  positive  half  of  the  ^-axis. 
In  this  case,  the  secondary  wave  has  both  a  reflected  component 
which  has  a  non-zero  amplitude  at  oo,  and  a  diffracted  part  which 
dies  out  at  oo.  A  uniqueness  theorem  has  been  derived  by  Peters  and 
Stoker  [P.10]  which  includes  these  special  cases;  we  proceed  to  give 
this  proof  both  for  its  own  sake  and  also  because  it  points  the  way 
to  a  straightforward  and  elementary  solution  of  the  special  problem 
formulated  above.  In  Chapters  6  and  7  a  different  way  of  looking 
at  the  problem  of  determining  appropriate  radiation  conditions  is 
proposed;  it  involves  considering  simple  harmonic  waves  (Chapter  6), 
or  steady  waves  (Chapter  7)  as  limits  when  t  ->  oo  in  appropriately 
formulated  initial  value  problems  which  correspond  to  unsteady 
motions. 


112 


WATER   WAVES 


The  uniqueness  theorem,  which  is  general  enough  to  include  the 
problem  above,  is  formulated  in  the  following  way:  We  assume  that 
/(#,  y)  is  a  complex- valued  solution*  of  the  equation 

(5.5.6)  V2/  +  /  =  0 

in  a  domain  D  with  boundary  jT,  part  of  which  may  extend  to  in- 
finity. It  is  supposed  that  any  circle  C  in  the  x,  j/-plane  cuts  out  of 
D  a  domain  in  which  the  application  of  Green's  formula  is  legitimate, 
and,  in  addition,  that  the  boundary  curve  F  outside  a  sufficiently 


Fig.  5.5.2.  The  domain  D 

large  circle  consists  of  a  single  half-ray  R  going  to  oo  (cf.  Figure 
5.5.2).**  On  the  boundary  F  the  condition 

(5.5.7)  fn  =  0 

is  imposed,  i.e.  the  normal  derivative  of  /  vanishes,  corresponding 
to  a  reflecting  barrier.  (We  could  also  replace  this  condition  on  part, 
or  all,  of  F  by  the  condition  /  =  0. )  We  now  write  the  solutions  of 
f  in  D  which  satisfy  (5.5.6)  in  the  form 

(5.5.8)  f  =  g+h, 

in  order  to  formulate  the  conditions  at  oo  in  a  convenient  way. 
What  we  have  in  mind  is  to  separate  the  solution  into  a  part  h 
which  satisfies  a  radiation  condition  and  a  part  g  which  contains, 

*  It  is  natural  to  consider  such  complex  solutions,  since,  for  example,  a  plane 
'wave  is  obtained  by  taking  f(q,  6)  =  exp  {  IQ  cos  (0  -f-  a)}. 
**  Our  theorem  also  holds  if  D  is  the  more  general  domain  in  which  the  ray  K 
is  replaced  at  oo  by  a  sector,  and  the  uniqueness  proof  given  below  holds  with 
insignificant  modifications  for  this  case  also. 


WAVES   ON   SLOPING   BEACHES   AND    PAST   OBSTACLES  118 

roughly  speaking,  the  prescribed  incoming  wave  together  with  any 
secondary  reflected  or  refracted  waves  which  also  do  not  satisfy  a 
radiation  condition.  More  precisely,  we  require  h  to  satisfy  the 
following  radiation  condition: 


(5.5.9)  lim 


dh  ,  •,. 

—  +^h 


=  0. 


Here  C  is  taken  to  be  a  circle,  with  its  center  O  (cf.  Figure  5.5.2) 
on  the  ray  R  going  to  infinity,  and  with  radius  Q  so  large  that  all 
obstacle  curves  except  a  part  of  R  lie  in  its  interior.  This  condition 
clearly  follows  from  the  well-known  Sommerfeld  radiation  condition, 
which  requires  that 

(5.5.9),  lim  p*  (V  +  ih\  ->  0 

^oo        \OQ  J 

uniformly  in  0,  and,  incidentally,  this  is  a  condition  independent  of 
the  particular  point  from  which  Q  is  measured;  we  observe  that  if  h 
behaves  at  oo  like  e~~iQ/\/Q,  i.e.  like  an  outgoing  cylindrical  wave, 
then  condition  (5.5.9)!  is  satisfied.  We  shall  make  use  of  the  radiation 
condition  in  the  form  (5.5.9)  in  much  the  same  way  as  F.  John  [J.5] 
who  used  it  to  obtain  uniqueness  theorems  for  (5.5.6)  in  cases  other 
than  those  treated  here;  his  methods  were  in  turn  modeled  on  those 
of  Rcllich  [R.7]. 

The  behavior  of  the  function  g  at  infinity  is  prescribed  as  follows: 

(5.5.10)  g~g!  +ga  at   oo, 

with  gl  a  function  that  is  once  for  all  prescribed,*  while  g2  is  a  function 
satisfying  the  same  radiation  condition  as  /i,  i.e.  the  condition 
(5.5.9).  (That  the  behavior  of  g  at  oo  is  fixed  only  within  an  additive 
function  satisfying  the  radiation  condition  is  natural  and  inevitable. ) 
Finally,  we  prescribe  regularity  conditions  at  re-entrant  points 
(such  as  A,  B,  C  in  Figure  5.5.2)  of  the  boundary  of  D;  these  con- 
ditions are  that 

(5.5.11)  f(Q,0)~cv         fQ(Q*0)~^>         k<l> 

Qk 

with  (g,  6)  polar  coordinates  centered  at  the  particular  singular  point, 
arid  cl  and  c2  constants.  (These  conditions  on  /  mean  physically  that 
the  radial  velocity  component  may  be  infinite  at  a  corner,  but  not 

How  the  function  gl  should  be  chosen  is  a  matter  for  later  discussion. 


114  WATER   WAVES 

as  strongly  as  it  would  be  for  a  source  or  sink.)  At  other  boundary 
points  we  require  continuity  of  /  and  its  normal  derivative. 
We  can  now  state  our  theorem  as  follows: 

Uniqueness  theorem:  A  solution  /  of  (5.5.6)  in  D  is  uniquely  determined 
if  it  1)  satisfies  the  boundary  condition  (5.5.9);  2)  admits  of  a  decom- 
position of  the  form  (5.5.8)  with  h  a  function  satisfying  (5.5.9),  g  a 
function  behaving  as  prescribed  by  (5.5.10)  at  oo;  and  3)  satisfies  the 
regularity  conditions  at  the  boundary  of  D. 

The  proof  of  this  theorem  will  be  given  shortly,  but  we  proceed 
to  discuss  its  implications  here.  The  theorem  is  at  first  sight  somewhat 
unsatisfactory  since  it  involves  the  assumption  that  every  solution 
considered  can  be  decomposed  according  to  (5.5.8),  with  g(p,  6)  a 
certain  function  the  behavior  of  which  at  oo,  in  so  far  as  the  leading 
term  gl  (cf.  (5.5.10))  in  its  asymptotic  development  is  concerned,  is 
not  given  a  priori.  However,  it  is  not  difficult  in  some  instances  at 
least  to  guess,  on  the  basis  of  physical  arguments,  how  the  function 
g1(p,  6)  should  be  defined.  For  example,  suppose  the  domain  D 
consisted  of  the  exterior  of  bounded  obstacles  only.  In  such  a  case  it 
seems  clear  that  gi(g,  6)  should  be  defined  as  the  function  describing 
the  incoming  wave— either  as  a  plane  wave  from  infinity,  say,  or  a 
wave  originating  from  an  oscillatory  source -—since  bounded  obstacles 
give  rise  only  to  reflected  and  diffracted  components  which  die  out 
at  oo  and  which  could  be  expected  to  satisfy  the  radiation  condition. 
Even  if  there  is  a  ray  in  the  boundary  that  goes  to  oo  (as  was  postulated 
above),  it  still  would  seem  appropriate  to  take  gt(p,  0)  as  the  function 
describing  the  incoming  wave,  provided  that  it  arises  from  an  oscil- 
latory point  source,*  since  such  a  source  would  hardly  lead  to  reflcted 
or  refracted  secondary  waves  that  would  violate  the  radiation  con- 
dition. However,  if  the  incoming  wave  is  a  plane  wave  and  an 
obstacle  extends  to  oo,  one  expects  an  outgoing  reflected  wave  to 
occur  which  would  in  general  not  satisfy  the  radiation  condition; 
in  this  case  the  function  g1(g,  0)  should  be  taken  as  the  sum  of  the 
incoming  plane  wave  and  an  outgoing  reflected  wave.  For  example, 
one  might  consider  the  case  in  which  the  entire  #-axis  is  a  reflecting 
barrier,  as  in  Figure  5.5.3.  In  this  case  one  would  in  an  altogether 
natural  way  define  gi((),  6)  as  the  sum  of  the  incoming  and  of  the 
reflected  wave  as  follows: 

*  The  same  statement  would  doubtlessly  hold  if  the  disturbance  originated 
in  a  bounded  region,  since  this  case  could  be  treated  by  making  use  of  a  distribu- 
tion of  oscillatory  point  sources. 


WAVES    ON    SLOPING    BEACHES   AND    PAST    OBSTACLES  115 


(5.5.12)  g^g,  0)  ==  eiQ  cos  (*-a)  +  eiQ  " 

with  a  the  angle  of  incidence  of  the  incoming  plane  wave.  If  we  were 
then  to  set  /  =  gl  +  h  (i.e.  we  set  g  =  gl  everywhere)  and  prescribe 
that  h  should  satisfy  the  radiation  condition,  it  is  clear  that  we  would 


Fig.  5.5.3.  Infinite  straight  line  barrier 

have  a  unique  solution  by  taking  h  =  0.  Our  uniqueness  theorem 
does  not  apply  directly  here  since  there  are  two  infinite  reflecting 
rays  going  to  oo,  but  it  could  be  easily  modified  so  that  it  would 
apply  to  this  case.  Thus  we  have  — for  the  first  time,  it  seems— a 


/ 

/ 
/ 


Fig.  5.5.4.  Sommerf eld's  diffraction  problem 

uniqueness  theorem  for  this  particularly  simple  problem  of  the 
reflection  of  a  plane  wave  by  a  rigid  plane.  A  less  trivial  example  is 
the  classical  Sommerfeld  diffraction  problem  — in  effect,  a  special 


116  WATER   WAVES 

ease  of  the  problem  with  which  our  present  discussion  began  — in 
which  a  plane  wave  coming  from  infinity  at  angle  a  to  the  #-axis  is 
reflected  and  diffracted  by  a  rigid  half-plane  barrier  along  the 
positive  ir-axis,  as  indicated  in  Figure  5.5.4.  In  this  case  it  seems 
plausible  to  define  the  function  g1(g,  6)  as  follows: 

!eie  cos  (0-oe)    +  eig  cos  (0+<^          0   <  0    <  JT  —  OC 
**cos(e-a),  rc-a<0<7r+a 

0,  n  +  a  <  0  <  2n. 

This  function  is,  of  course,  discontinuous,  corresponding  to  the 
division  of  the  plane  into  the  regions  in  which  a)  the  incoming  wave 
and  its  reflection  from  the  barrier  coexist,  b)  the  region  in  which 
only  the  wave  transmitted  past  the  edge  of  the  barrier  exists,  and 
c)  the  region  in  the  shadow  created  by  the  barrier.  Again  we  would 
be  inclined  to  takeg  =  gl  (cf.  (5.5.8)  and  (5.5.10))  and  set/  =  gl  +  h, 
with  h  satisfying  the  radiation  condition.  Of  course,  the  function 
h(Q9  6)  in  (5.5.8)  representing  the  diffracted  wave  would  then  also 
be  discontinuous  in  that  case  since  the  sum  gl  +  h  is  everywhere 
continuous.  It  will  be  seen  that  the  well-known  solution  given  by 
Sommerfeld  can  be  decomposed  in  this  way  and  that  h  then  satisfies 
the  radiation  condition.  Our  uniqueness  theorem  will  thus  be  shown 
to  be  applicable  in  at  least  the  important  special  case  of  particular 
interest  in  this  section. 

One  might  hazard  a  guess  regarding  the  right  way  to  determine 
the  function  g  in  all  cases  involving  unbounded  domains:  it  seems 
highly  plausible  that  it  would  always  be  correctly  given  by  the 
methods  of  geometrical  optics.  By  this  we  mean,  from  the  mathemati- 
cal point  of  view,  that  g  would  be  the  lowest  order  term  in  an  asymp- 
totic expansion  of  the  solution  /  with  respect  to  the  frequency  of  the 
motion  that  is  valid  for  large  frequencies;  the  methods  of  geometrical 
optics  would  thus  be  available  for  determining  g.  However,  to  prove 
a  theorem  of  such  generality  would  seem  to  be  a  very  difficult  task 
since  it  would  probably  require  some  sort  of  representation  for  the 
solution  of  wave  propagation  problems  when  more  or  less  arbitrary 
domains  and  boundary  data  are  prescribed. 

Once  having  proved  that  the  solution  of  Sommerfeld's  diffraction 

problem  could  be  decomposed  in  the  way  indicated  above  into  the 

.sum   of  two   discontinuous   functions,    one   of  which   satisfies   the 

radiation  condition,  it  was  observed  that  the  latter  fact  opens  the 

way  to  a  new  solution  of  the  diffraction  problem  which  is  entirely 


WAVES   ON   SLOPING   BEACHES   AND   PAST   OBSTACLES  117 

elementary,  straightforward,  and  which  can  be  written  down  in  a  few 
lines.  In  other  words,  once  the  reluctance  to  work  with  discontinuous 
functions  is  overcome,  the  solution  of  the  problem  is  reduced  to 
something  quite  elementary  by  comparison  with  other  methods  of 
solution.  The  problem  was  solved  long  ago  by  Sommerfeld  [S.12], 
and  afterwards  by  many  others,  including  Macdonald  [M.I],  Bateman 
[B.5],  Copson  [C.4],  Schwinger  [S.5],  and  Karp  [K.3]. 

We  shall  first  prove  the  uniqueness  theorem.  Afterwards,  the  simple 
solution  of  Sommerf eld's  problem  just  referred  to  will  be  derived; 
this  solution  is  in  the  form  of  a  Fourier  series.  The  Fourier  series 
solution  is  next  transformed  to  furnish  a  variety  of  solutions  given 
by  integral  representations,  including  the  familiar  representation 
given  by  Sommerfeld.  The  new  representations  are  particularly 
convenient  for  the  purpose  of  discussing  a  number  of  properties  of 
the  solution.  In  particular,  two  such  representations  can  be  used  to 
show  that  the  function  h  in  the  decomposition  /  =  gl  +  h  (cf. 
(5.5.13))  satisfies  the  radiation  condition,  and  that  our  solution  / 
satisfies  the  regularity  conditions  at  the  origin;  thus  the  solution  is 
shown,  by  virtue  of  our  uniqueness  theorem,  to  be  the  only  one  which 
behaves  at  oo  like  gl  plus  a  function  satisfying  the  radiation  condition. 
The  Stokes'  phenomenon  encountered  in  crossing  the  lines  of  discon- 
tinuity of  the  functions  gl  and  h  is  also  discussed. 

The  uniqueness  theorem  formulated  above  is  proved  in  the  following 
way.  Suppose  there  were  two  solutions  /  and  /*  (cf.  (5.5.8))  with 
/*  given  by 

(5.5.U)  He,  °)  =  «*(e»  0)  +  **(e»  *)• 

We  introduce  the  difference  ^(^,  0)  of  these  solutions: 

(5.5.15)         *(e,0)=/(M)-/*(M) 

=  8(Q,  0)  -  g*(Q,  0)  +  h(e,  0)  -  A*(e,  0) 

and  observe  that  %(Q,  0)  satisfies  the  radiation  condition  (5.5.9),  by 
virtue  of  the  Schwarz  inequality,  since  h,  A*,  and  the  difference 
g  __  g*  ail  satisfy  it  by  hypothesis;  thus  we  have 

2 


(5.5.16)  lim 


im 

->oo  Jc 


,  • 
+«* 


=  0. 


The  complex-valued  function  #  is  decomposed  into  its  real  and 

imaginary  parts: 

(5.5.17)  x  =  Xi  +  *X» 

and  Green's  formula 


118 


WATER  WAVES 


(5.5.18) 


is  applied  to  #x  and  #2  in  the  domain  D*  indicated  in  Figure  5.5.5. 
The  domain  JD*  is  bounded  by  a  circle  C  so  large  as  to  include  all  of 
the  obstacles  in  its  interior  except  R,  by  curves  which  exclude  the 


Fig.  5.5.5.  The  domain  D* 

prolongation  of  R  into  the  interior  of  C,  and  by  curves  excluding 
the  other  bounded  obstacles.  By  (5.5.15),  #  is  a  solution  of  (5.5.6) 
which  also  clearly  satisfies  the  boundary  condition  (5.5.7).  Since 
V2Xi  =  —  Xi  and  V2#2  =  —  #2>  it  follows  that  the  integrand  of  the 
left-hand  side  of  (5.5.18)  vanishes.  Because  of  the  regularity  conditions 
at  boundary  points  we  are  permitted  to  deform  the  boundary  curve 
J1*  into  the  obstacle  curves,  and  it  then  follows  from  the  boundary 
condition  (5.5.7)  that 

(5.5.19) 

since  the  contributions  at  the  obstacles  all  vanish.  We  now  make 
use  of  the  easily  verified  identity 


WAVES   ON   SLOPING   BEACHES   AND    PAST   OBSTACLES  119 


(5.5.20) 

to  deduce  from  (5.5.19)  the  condition 

(5.5.21)  Jc(l*nl2  +  \X\2)d*  ~  jc\Xn  +  ix\*ds  =  0, 

from  which  we  obtain,  in  view  of  (5.5.16)  and  %n  =  dx/dg  on  C: 

(5.5.22)  lim   f    \%\*ds  =  lim   f 
->oo  *  c  ->oo  ^ 


From  the  boundary  condition  (5.5.7),  as  applied  on  R,  we  see  that 
%(Q,  0)  can  be  continued  as  a  periodic  function  of  period  4jr  in  6 
on  C;  hence  #  can  be  represented  for  all  sufficiently  large  values  of 
Q  by  the  Fourier  series 

00  nO 

(5.5.23)  *  =  I^n/2(0)cos-, 

0  * 

with  Ani2(())9  the  Fourier  coefficient,  a  certain  linear  combination 
of  the  Bessel  functions  Jw/2(g)  and  Fn/2(g),  since  %  is  a  solution  of 
(5.5.6).  The  Fourier  coefficients  are  given  by 

1   f2*  nO 

(5.5.24)  '<„/,(<?)  =  -        X(e,0)Cos-?.dO, 

n  Jo  2 

and  consequently  we  have 

(5.5.25)  le*^ 


dO 


It  follows  at  once  from  (5.5.22)  that  the  Fourier  coefficients  behave 
for  large  Q  as  follows: 

(5.5.26)  lim  e*  Anf2(e)  =  0. 

0— >00 

Since  the  Bessel  functions  «/n/2(o)  and  Ynj2((>)  all  behave  at  oo  like 
1/V(?»  i*  follows  that  all  of  the  coefficients  ^4n/2(g)  must  vanish. 
Consequently  %  vanishes  identically  outside  a  sufficiently  large  circle, 
hence  it  vanishes*  throughout  its  domain  of  definition,  and  the 

*  This  could  be  proved  in  standard  fashion  since  #  is  now  seen  to  satisfy 
homogeneous  boundary  conditions  in  the  domain  D*  of  Fig.  5.5.5. 


120  WATER   WAVES 

uniqueness  theorem  is  proved.  As  was  stated  above,  this  uniqueness 
proof  is  much  like  that  of  Rellich  [R.7]. 

The  above  proof  can  be  modified  easily  in  such  a  way  as  to  apply 
to  a  region  with  a  sector,  rather  than  a  ray,  cut  out  at  oo.  The  only 
difference  is  that  the  Fourier  series  for  #(g,  0)  would  then  not  have 
the  period  4jr  and  that  the  Bessel  functions  involved  would  not  be 
of  index  n/2. 

Once  it  has  become  clear  that  the  decomposition  of  the  solution 
into  the  sum  of  the  two  discontinuous  functions  g(p,  6)  and  A(g,  6) 
defined  earlier  is  a  procedure  that  is  really  natural  and  suitable  for 
this  problem,  one  is  then  led  to  the  idea  that  such  a  decomposition 
might  be  explicitly  used  in  such  a  way  as  to  determine  the  solution 
of  the  original  problem  ((cf.  (5.5.8))  in  a  direct  and  straightforward 
way.  Our  next  purpose  is  to  carry  out  such  a  procedure. 

We  set  (cf.  Figure  (5.5.4)  and  equation  (5.5.13)): 


(5.5.27)  /(0, 

with  g(p,  0)  defined  by 

eiQ  cos  (0-oo  _|_  £*<?  cos  (0+oc)^     o  <  0  <  tt  —  a 
(5.5.28) 


-*)9  7t-QL<0 

0,  n  +  oc  <  0  <  2n. 


In  addition  we  have 

(5.5.29)  f0  =  0  for  0  =  0,         0  ==  2rc 
and  we  also  require 

(5.5.30)  lim  <\/Q  (^-  +  <*)  =  0  uniformly  in  0, 
0-^00  we  / 

since  the  validity  of  the  radiation  condition  in  this  strong  form  can 
be  verified  in  the  end. 

The  desired  solution  will  be  found  by  developing  /(p,  0)  into  a 
Fourier  series  in  0  for  fixed  p,  and  determining  the  coefficients  of  the 
series  through  use  of  the  radiation  condition  in  the  strong  form 
(5.5.30);  afterwards,  the  series  can  easily  be  summed  to  yield  a 
convenient  integral  representation  of  the  solution.  That  such  a 
*  process  will  be  successful  can  be  seen  very  easily:  The  Fourier  series 
for  /(g,  0)  will,  on  account  of  the  boundary  condition  (5.5.29)  and  the 
fact  that  /  is  a  solution  of  the  reduced  wave  equation,  be  of  the  form 


WAVES    ON   SLOPING   BEACHES   AND    PAST   OBSTACLES  121 

]£cnJn/2(0)  cos  n0/2;   the  Fourier  coefficients  for  g(g,  6)  as  defined 
by  (5.5.28)  are  given  in  terms  of  integrals  of  the  form 

•2*1  „& 


f2 

ln=\ 

Jo 


since  this  function  also  satisfies  the  condition  ge  =  0  for  0  =  0,  2n. 
Since  «/n/2((?)>  and  its  derivatives  as  well,  behave  like  l/\/(?  f°r  large 
values  of  Q  and  the  integrals  In  —  by  a  straightforward  application 
of  the  method  of  stationary  phase,  for  example,  —  also  behave  in 
this  way,  it  is  clear  that  the  limit  relation  (5.5.30)  when  used  in 
connection  with  (5.5.27)  will  serve  to  determine  the  coefficients  cn. 
We  proceed  to  carry  out  this  program.  The  finite  Fourier  transform 
/  of  /  is  introduced  by  the  formula 

—  /*  2*i  Aj/5 

(5.5.31)  J(Q,  n)  =         f(o,  Q)  cos  —  dO.  . 

Jo  2 

Since  /9  —  0  for  6  —  0,  2jt  we  find  for  fm  the  transform 

(5.5.32)  J^-^-^J, 

4 

by  using  two  integrations  by  parts.  Since  /  is  a  solution  of 

(5.5.33)  Q*fee  +  efQ  +  foe  +  Q2f  =  0, 
it  follows  that  /  is  a  solution  of 

(5.5.34)  &QQ  +  QfQ  +    g2  ~          1  =  0, 


and  solutions  of  this  equation  are  given  by 

(5.5.35)  J(e.n)  =  anJnl2(Q). 

(The  Bcsscl  functions  Ynj2(())  of  the  second  kind  are  not  introduced 
because  they  are  singular  at  the  origin;  the  solution  we  want  is  in 
any  case  obtained  without  their  use.) 

The  transform  of  g(p,  0)  is,  of  course,  given  by 

/*  2*i  A|/5 

(5.5.36)  g(0,  n)  =        g(e,  6)  cos  —  d0, 

Jo  2 

and  we  have,  in  view  of  (5.5.8),  the  relation: 

C2n  nQ  f2*  n6 

(5.5.37)  h(Q,  6)  cos  —  dO  =  anJnl2(Q)  -        g(g9  6)  cos  —  dO 

Jo  2  Jo  2 

or,  also: 


122  WATER   WAVES 

(5.5.88)  h(e,  n)  =  anJnl2(g)  -  g(e,  n). 

We  must  next  apply  the  operation  yp  (9/^e  +0  to  both  sides 
of  (5.5.87)  and  then  make  the  passage  to  the  limit,  with  the  result* 

* 


(5.5.89)     0  =  lim  ^/Q  (|-  +  i)  \anJnl2(e)  -  (  *g(e,  6)  cos  ^  d6\. 

*-»«>  \OQ  /    L  JO  2          J 

Since  the  functions  t/n/2(g)  behave  asymptotically  as  follows: 

2 


and  since  these  asymptotic  expansions  can  be  differentiated,  we  have 
(5.5.40)  IL  +  i)  Jn/2(Q)  ~  I/A 


as  an  easy  calculation  shows.  The  behavior  of  the  integral  over  g 
can  be  found  easily  by  the  well-known  method  of  stationary  phase, 
which  (cf.  Ch.  6.8)  states  that 


f 

Ja 


Q   <P 

in  which  a  is  a  simple  zero  of  the  derivative  <p'(Q)  in  the  range 
a  <  6  <  ft,  and  the  ambiguous  sign  in  the  exponential  is  to  be  taken 
the  same  as  the  sign  of  ^"(a).  In  the  present  case,  in  which  g(p,  0) 
is  defined  by  (5.5.28)  one  sees  at  once  that  there  are  three  points 
of  stationary  phase,  i.e.  at  6  =  a,  0  =  n  —  a,  and  0  —  n  +  a.  Of 
the  three  contributions  only  the  first,  i.e.  the  contribution  at  6  =  a,** 
furnishes  a  non-vanishing  contribution  for  Q  ->  oo  when  the  operator 
Vp(9/9p  +  i)  is  applied  to  it;  one  finds,  in  fact: 

(5.5.41)  (I  +  A  rg(e,  6)  cos  ^  dQ  ~  2  V^  cos  ™  X«3. 

we      /Jo  2  f  e       2 

Use  of  (5.5.40)  and  (5.5.41)  in  (5.5.39)  furnishes,  finally,  the  coef- 
ficients an: 

(5.5.42)  an  =  2n  cos  —  e?*T. 

2 

The  Fourier  series  for  /(g,  0)  is 

*  It  should  be  noted  that  the  argument  goes  through  if  the  radiation  condition 
is  used  in  the  weak  form. 

**  This  has  physical  significance,  since  it  says  that  only  the  incoming  wave  is 
effective  in  determining  the  Fourier  coefficients  of  the  solution. 


WAVES    ON   SLOPING   BEACHES    AND    PAST   OBSTACLES 


123 


/(<?>  0)  =     /(g,  o)  +        /(e,  n)  cos 
aw  TT  ~i  2 

or,  from  (5.5.35)  and  (5.5.42), 


cos    -. 


It  is  not  difficult  to  sum  the  series  for  /(p,  0).  If  we  use  the  represen- 
tation (for  a  derivation,  see  Courant-Hilbert  [C.10,  p.  413]) 


(5.5.43) 


,(e)  =  J!  f  *-lH)  c—  i  dc, 

2^i  Jp 


where  P  is  the  path  in  the  complex  £-plane  shown  in  Figure  5.5.6, 
we  find  that  /(p,  0)  can  be  expressed  as  the  integral  of  the  sum  of 


£-plane 


Fig.  5.5.6.  The  path  P  in  the  f-plane 

a  constant  plus  four  geometric  series.  The  summation  of  the  geometric 
series  and  a  little  algebra  yields,  finally,  a  solution  in  the  form 


(5.5.44) 

1     f  e~Wt) 

ftp*  0)  = - --  ' 

Sjti  J  p       £ 


3w 


"^ 


37i 


We  proceed  to  analyze  the  solution  (5.5.44)  of  our  problem  with 
respect  to  its  behavior  at  oo  and  the  origin,  and  we  will  show  that 


124  WATER  WAVES 

the  conditions  needed  for  the  validity  of  the  uniqueness  theorem 
proved  above  are  satisfied.  We  will  also  transform  it  into  the  solution 
given  by  Sommerfeld  (cf.  equation  (5.5.47)).  Not  all  of  the  details 
of  these  calculations  will  be  given:  they  can  be  found  in  the  paper 
by  Peters  and  Stoker  [P.19]. 
If  we  set 


(5.5.45)  /(g,  0)  -  J(e,  0+ot)  +  /(e,  0-<x) 

and  define  /(g,  «)  by 


(5.5.46)    /(g,x)  =  —     e  2 


L 


-  1          i       3;i 


we  see  on  comparison  with  (5.5.44)  that  (5.5.45)  defines  /(p,  0) 
correctly  as  the  solution  we  wish  to  investigate. 

Let  us  first  obtain  the  solution  in  the  form  given  by  Sommerfeld. 
To  this  end,  the  denominators  of  the  fractions  in  square  brackets 
in  (5.5.46)  are  rationalized,  and  the  fractions  combined  to  yield 

Qf      l\  r  3     3*  1     .3n  -, 

1     re~2\C~C/    C2  +  2£2*14  cos-f +2t£S/4  cos-y  +  1  LJ> 

J(e>*)  =  7-;    — i —  Fi~r^ ; — 

4m  Jp      C       L  C2  +  2zC  cos  x  —  1  J 

One  can  then  verify  readily  that  /  satisfies  the  differential  equation 
dl  If  --(c--)r         --  t'~      x       .  -?  i-      *        21 

^008-    .         ....          .  3 


2  f    -^:--^       - 

—      ^  2V  c/  (f  2  +  {, 

2m     Jp 


If  we  use  (5.5.43)  and  the  well-known  trigonometric  formulas  for 
^i/2(?)>  J-HZ(Q)  we  see  ^^a*  tta  last  equation  is  equivalent  to 

dl  I/IT  ~  M 

—  2—  +  (2i  cos  K\!  =  —  ^  —  ^4  ^-^  cos  -  . 

^     v         ;         r^e  2 

A  solution  /#  of  the  non-homogeneous  equation  which  in  general 
vanishes  as  Q  ->  oo  is  readily  found: 

in 


»oo/,-<A(l4-cosx) 

=       — 

Vtoi 


v  /»oo/, 

coijf   f 
2Je 


WAVES   ON   SLOPING   BEACHES   AND    PAST   OBSTACLES  125 

Thus  for  /  the  appropriate  solution  of  the  differential  equation 
must  be 

1=6  +  •*#• 

Introduction  of  a  new  variable  of  integration  z  in  the  expression  for 
IN  through  the  relation  2A  cos2  x/2  =  z2,  and  use  of  the  formula 

_— 
e~iz2  dz  =  \/7t  e   4 

/— -00 

leads  with  no  difficulty  to  the  expression 

in 

—  /•     / —         x 

6*  f  V  2c  cos  -  a 

(5.5.47)  I(Q,  x)  =  el* cos  x  2  e~iz  dz 

and  this  leads,  in  conjunction  with  (5.5.45),  to  Sommerfeld's  solution. 

To  derive  the  asymptotic  behavior  of  7(p,  x)  as  Q  ->  oo  we  proceed 

a  little  differently.  The  fractions  in  the  square  brackets  in  (5.5.46) 

are  combined,  and  some  algebraic  manipulation  is  applied,  to  yield 


-ur 
(Qt  '  ~  4^   i 

4^  JP_1_  (C,/2 


A  new  integration  variable  A  is  now  introduced  by  the  equation 


•V/2  2v/2 

with  the  result 

-_ 


A  —  V2  ^4  cos  - 

2 

The  path  P  (cf.  Fig.  5.5.6)  is  transformed  into  the  path  L  shown  in 
Fig.  5.5.7,  as  one  readily  can  see.  The  path  L  leaves  the  circle  of 
radius  \/2  centered  at  the  origin  on  its  left.  This  representation  of 
the  function  7(g,  K)  is  obviously  a  good  deal  simpler  than  that 
furnished  by  (5.5.46),  and  it  is  quite  advantageous  in  studying  the 
properties  of  the  solution:  for  one  thing,  the  plane  waves  at  oo 
can  be  obtained  as  the  residues  at  the  poles 

3*<        Q  i  a 
(5.5.49)  A±  =  V2  c  *   cos  -=-• 


126 


WATER  WAVES 


In  fact,  if  there  is  a  pole  in  the  upper  half  of  the  A-plane  (and  there 
may  or  may  not  be,  depending  on  the  values  of  both  6  and  a)  one 


X-plane 


Fig.  5.5.7.  The  path  L  in  the  A-plane 

has,  after  deformation  of  the  path  L  over  it  and  into  the  real  axis, 
for  I(Q,  K)  the  result: 

.,  (*\      e"iQ 

*w-  — 


(5.5.50)     I(o,  9c)  =  e 


—  V2  e  4   cos  ~ 
v  2 


piq  cos  x  


"*  A  —  A/2  e  4  cos  - 

2 

If  x  =  7i  —  the  only  case  in  which  there  is  a  singularity  on  the  real 


z-plane 


Fig.  5.5.8.  The  path  C  in  the  z-plane 

axis,  i.e.  a  pole  at  A  =  0  —  we  assume  that  the  path  of  integration 
is  deformed  near  the  origin  into  the  upper  half-plane.  It  is  convenient 


WAVES    ON   SLOPING   BEACHES   AND   PAST   OBSTACLES  127 

to  introduce  the  variable  z  =  gA2  in  the  integral,  with  the  result 
(5.5.51)      /($,*)=    '-~-          "          '    C~*dZ 


with  Ax  =  \/2  e'3*'4  cos  x/2,  and  C  the  path  of  integration  shown 
in  Fig.  5.5.8  For  large  values  of  g,  and  assuming  Ax  ^  0,  the  square 
bracket  in  the  integrand  can  be  developed  in  powers  of  (Z/Q  )1/2,  and 
we  may  write 

e-*dz  I   C  e~zdz 


It  is  clear  that  we  may  allow  e  ->  0  (see  Fig.  5.5.8)  and  hence  the 
path  C  can  be  deformed  into  the  two  banks  of  the  slit  along  the 
real  axis;  each  of  the  terms  in  the  square  brackets  then  can  be 
evaluated  in  terms  of  the  /^-function  (cf.,  for  example,  MacRobert 
[M.2],  p.  143).  It  is  thus  clear  that  for  Ax  ^  0,  the  leading  term  in 
the  asymptotic  expansion  of  the  integral  in  (5.5.51)  behaves  like 
l/\/P»  *n  fact,  we  have  for  /(g,  «): 

(5.5.52) 


Since  /'(^)  =  V^  and  ^x  —  V^  ^  4   cos  ~   wc  have 


A 

(5.5.53)  7(p,  x)  ~  e**  cosx  -  --  _—-  4--  . 

zVzno  cos  - 
^         2 

Of  course,  this  holds  only  if  H  lies  in  the  range  0  ^  ^  <  n  since  a  pole 
occurs  in  the  upper  half  of  the  A-plane  only  when  cos  */2  is  positive 
(cf.  (5.5.50)).  We  must  also  exclude  the  value  K  —  n,  corresponding 
to  Ax  =  0.  Since  K  =  0  T  a,  we  see  that  the  values  0  =  n  ±  a  cor- 
respond to  the  exceptional  value  K  —  n,  and  these  values  of  0,  in 
turn,  are  those  which  yield  the  lines  in  the  physical  plane  across 
which  our  solution  /  behaves  discontinuously  at  oo.  (Cf.  Fig.  5.5.4). 


128  WATER  WAVES 

The  discussion  of  the  last  paragraph  yields  the  result,  in  conjunction 
with  equation  (5.5.45)  which  defines  our  solution  in  terms  of /(p,  H): 

(5.5.54 )     /(0,  6 )  ~  e* cos 


2V2no  cos  -  -       2V2no  cos  -~ 
*  2  *  2 

for  large  Q  and  for  angles  6  such  that  0  <  0  <  n  —  a,  and  a  in  the 
range  0  <  a  <  n\  only  in  this  case  are  there  poles  of  both  of  the 
integrals  in  (5.5.45)  in  the  upper  halfplane. 

The  discussion  of  the  behavior  of  the  solution  in  other  sectors  of 
the  physical  plane  and  along  the  exceptional  lines  can  be  carried  out 
in  the  same  way  as  above.  For  example,  if  A+  =  \/2  £i371/4  cos  (  0  +oc/2  )  =  0, 
and  hence  0  =  n  —  a,  it  follows  that  there  is  only  one  pole  in  the 
upper  halfplane  and  our  solution  /(g,  n—  a)  is  given  by  (cf.  5.5.48), 
(5.5.49)): 


e-iQ    f  e-Q& 

?_  M  L_ 
27riJL   A 


+ 

or  also  (cf.  (5.5.51)  and  Fig.  5.5.8)  by: 

e~zdz 


f(g9  n-*)  = 


e-iQ     /•    e-, 

—  dz. 

Anijc  z 


The  asymptotic  behavior  of  /  can  now  be  determined  in  the  same 
way  as  above;  the  result  is 

_•  _if! 
(5.5.55)      f(Q,  n  -  a) 


2  /  -        Jt  —  2oc 

2  V2jrp  cos  - 
*  2 

the  second  term  resulting  from  the  pole  at  the  origin. 

In  this  fashion  the  behavior  of  /(g,  6)  for  large  values  of  Q  is  deter- 
mined, and  leads  to 


WAVES   ON  SLOPING  BEACHES  AND   PAST  OBSTACLES  129 


(5.5.56) 


/(ft 


eiQ  cos  (<>-a) 


tg  cos  (0-a) 


,     ft  —  a  < 0  <ft  +  a 

0  ,     ft  +  a  <  0  <  2ft. 

This  is,  of  course,  a  verification  of  one  of  the  conditions  imposed 
at  oo.  In  addition,  the  next  terms  in  the  asymptotic  expansion,  of 
order  1/V0»  are  a'so  determined,  as  follows: 


(5.5.56)' 


in 


in 


0-1a      2V^?cos9-±-a 
2  ^2 


,  0<0<ft— a 


cos 


-_    _    O« 


in 


2  V/2ftg  i 


__  a 


,  0— ft — oc 


~ ,  ft-a<0<ft+a 

0  +  a 


2V2ftg 


t  ft  +  2a 

i 

2 


/ —        0-a  / —        0+a 

2v 2ft/)  cos 2V2ft0  cos 

2  2 

We  observe  that  these  expansions  do  not  hold  uniformly  in  0  because 
of  zeros  in  the  denominators  for  0  =  n  i  a,  i.e.  at  the  lines  of 
discontinuity  of  the  function  g(p,  0). 

With  the  aid  of  the  function  g(p,  0)  defined  in  (5.5.56)  we  define 
a  function  />(p,  0)  by  the  equation 
(5.5.57)  /({),  0)  =  g({>,  0)  +  h(Q,  0). 

Thus  h  is  of  necessity  a  discontinuous  function  since  /  is  continuous 
while  g  has  jump  discontinuities  along  the  lines  0  =  ft  ±  a.  The 
function  h  is  given  by  (cf.  (5.5.50)): 


130  WATER  WAVES 


(5.5.58) 


with  the  proviso  that  the  integrals  should  be  deformed  into  the 
upper  half-plane  in  the  vicinity  of  the  origin  in  case  either  A_  or  k+ 
vanishes:  i.e.,  in  case  0  has  one  of  its  two  critical  values  n±  a. 
That  the  sum  g  +  h  really  is  our  solution  /  is  rather  clear  in  the 
light  of  our  discussion  above;  and  that  it  has  jump  discontinuities 
which  just  compensate  those  of  g  in  order  to  make  /  continuous  can 
also  be  easily  verified.  We  shall  not  carry  out  the  calculation  here. 
The  function  A(g,  0),  in  view  of  (5.5.56)  and  (5.5.57)  thus  yields  what 
might  be  called  the  "scattered"  part  of  the  wave. 

In  order  to  show  that  our  solution  /  satisfies  the  conditions  of  the 
uniqueness  theorem  proved  above,  we  proceed  to  show  that  h  as 
defined  by  (5.5.58)  satisfies  the  radiation  condition  (5.5.9);  afterwards 
we  will  prove  that  /  behaves  at  the  origin  as  prescribed  by  (5.5.11). 
Our  solution  /  will  thus  be  proved  to  be  unique. 

That  the  function  A(p,  6)  defined  by  (5.5.58)  satisfies  the  radiation 
condition  is  not  at  all  obvious:  one  sees,  for  example  (cf.  (5.5.50)'), 
that  its  behavior  at  oo  is  far  from  being  uniform  in  the  angle  0. 
In  fact,  the  transformation  of  h  to  be  introduced  below  is  motivated 
by  the  desire  to  obtain  an  estimate  for  the  quantity  |  dh/dQ  +  ih  l> 
which  figures  in  the  radiation  condition,  that  is  independent  of  0; 
and  this  in  turn  means  an  estimate  independent  of  the  quantities 
A-  and  A+  defined  by  (5.5.49).  The  function  A(g,  0)  can  first  of  all 
be  put  in  the  form 


fc/  m   -*~*fj  r  e~Q*d*  L3  r 

Me*  0  =  —  —  M-  •     -„—  -.j-  +  V 

m     (       Jo  A2  -  A2_  Jo 

as  one  readily  verifies.  We  proceed  as  follows:  First  we  write 

,  6)  =  -  ~  h  •  ("V*1  fV^'J'dfatt  +  A,  •  fV^2  f  Y<^>' 

M  L       Jo          Jo  Jo          Jo 

then  carry  out  the  integrations  with  respect  to  A  to  obtain 


*~i(?  r,   r  **-**  ,  5   r00  ^** 

-7--U--       ----  i+V       --    i 

2Vm  L       Jo        +0*  Jo  ((?  +02 


(p  +0*  Jo  ((? 

From  this  representation  of  h  we  obtain 

3h  tr*    r,      f00  /-«dt     ,   .      f00  /'A  I 

+  tA  =  __    A_-       ----  3+^'       ------  ,  • 

5p  4V7ri  L      Jo       +t)i  Jo       +«)iJ 


WAVES    ON   SLOPING    BEACHES   AND    PAST   OBSTACLES 


131 


It  is  important  to  observe  that  A2_  and  Al  have  pure  imaginary  values, 
as  we  know  from  (5.5.49).  We  also  observe  that  the  exceptional 
lines  0  —  n  ±  a,  which  correspond  to  A^  =0,  simply  have  the  effect 
that  one  of  the  two  terms  in  the  brackets  in  the  last  equation  vanishes. 
From  the  Schwarz  inequality  we  have 

\dh 


\d~Q 


+  ih 


/*  GO          A     t     IA  2  I    1        I  2    '     /* ' 

f     e^-ldi  *        X,  r    f 

I I    i__±_!_     I 

Jo   (o  +t)l  &n     Jo 


Consider  the  first  term  on  the  right;  we  find: 


2 !  r°°  X-erf<  i2        2    f00 

Jo    (g  +  <)z  Jo    ( 


(It 


e« 


<«dt  '2 


Since  the  same  estimate  holds  for  the  second  term,  it  follows  that 

dh       . 


and  this  estimate  holds  for  all  values  of  0,  since  it  holds  for  the  two 
exceptional  values  6  =  n  ±  a  as  well  as  for  all  other  values  in  the 
range  0  fg  0  rgj  2n.  We  have  thus  verified  that  the  radiation  condition 
holds  —  in  fact,  we  have  shown  that  it  holds  in  the  strong  form. 
We  proceed  to  show  that  /(p,  6)  behaves  properly  at  the  origin. 
To  this  end,  we  start  with  the  solution  in  the  form  (cf.  (5.5.48)  and 
(5.5.45)): 


with  L  the  path  of  Fig.  (5.5.7).  The  transformation  A  =  \/'z  is  then 
made,  so  that  the  new  path  of  integration  D  is  like  the  path  C  in 
Fig.  5.5.8  except  that  the  circular  part  now  has  a  radius  large  enough 
to  include  the  singularities  of  the  integrands  in  its  interior.  We  may 
take  the  radius  of  the  circular  part  of  D  to  have  the  value  1/p,  since 
we  care  only  for  small  values  of  Q  in  the  present  consideration.The 
transformation  QZ  =  u  then  leads  to  the  following  formula  for  /(g,  Q): 


132 


WATER    WAVES 


with  Dj  a  path  of  the  same  type  as  D  except  that  the  circular  part 
of  Dl  is  now  the  circle  of  unit  radius.  For  small  values  of  Q  the  integrals 
in  the  last  expression  can  be  expressed  in  the  form 


u 


.  .  .  \du 


/*      ft — W 

+  e*^±     —  ^w 

From  this  expansion  we  see  clearly  that 


6—  du  +  .... 


as 


0. 


This  completes  the  verification  of  the  conditions  needed  for  the 
application  of  the  uniqueness  theorem  to  our  solution  /. 

It  has  been  shown  by  Putnam  and  Arthur  [P.18]  (see  also  Carr 
and  Stelzriede  [C.I])  that  the  theory  of  diffraction  of  water  waves 


Fig.  5.5.9.  Waves  behind  a  breakwater 

around  a  vertical  barrier  is  in  good  accord  with  the  physical  facts, 
the  accuracy  being  particularly  high  in  the  shadow  created  by  the 


WAVES    ON    SLOPING    BEACHES   AND   PAST   OBSTACLES  133 

breakwater.  Figure  5.5.9  is  a  photograph  (given  to  the  author  by 
J.  H.  Carr  of  the  Hydrodynamics  Laboratory  at  the  California 
Institute  of  Technology)  of  a  model  of  a  breakwater  which  gives  some 
indication  of  the  wave  pattern  which  results. 

5.6.  Brief  discussions  of  additional  applications  and  of  other  methods 
of  solution 

The  object  of  the  present  section  is  to  point  out  a  few  further 
problems  and  methods  of  dealing  with  problems  concerned,  for  the 
most  part,  with  simple  harmonic  waves  of  small  amplitude. 

The  first  group  of  problems  to  be  mentioned  belongs,  generally 
speaking,  to  the  field  of  oceanography.  For  general  treatments  of 
this  subject  the  book  of  Sverdrup,  Johnson,  and  Fleming  [S.32] 
should  be  consulted.  One  type  of  problem  of  this  category  which 
was  investigated  vigorously  during  World  War  II  is  the  problem  of 
wave  refraction  along  a  coast,  or,  in  other  terms,  the  problem  of 
the  modification  in  the  shape  of  the  wave  crests  and  in  the  amplitude 
of  ocean  waves  as  they  move  from  deep  water  into  shallow  water. 
We  have  seen  in  the  preceding  sections  that  it  is  not  entirely  easy 
to  give  exact  solutions  in  terms  of  the  theory  of  waves  of  small 
amplitude  even  in  relatively  simple  cases,  such,  for  example,  as  the 
case  of  a  uniformly  sloping  bottom.  As  a  consequence,  approximate 
methods  modeled  after  those  of  geometrical  optics  were  devised, 
beginning  with  the  work  of  Sverdrup  and  Munk  [S.35].  Basically, 
these  methods  boil  down  to  the  assumption  that  the  local  propagation 
speed  of  a  wave  of  given  length  is  known  at  any  point  from  the  for- 
mulas derived  in  Chapter  4  for  water  of  constant  depth  once  the 
depth  of  the  water  at  that  point  is  known;  and  that  Iluygens'  principle, 
or  variants  of  it,  can  be  used  to  locate  wave  fronts  or  to  construct 
the  rays  orthogonal  to  them.  The  errors  resulting  from  such  an  as- 
sumption should  not  be  very  great  in  practice  since  the  depth  varia- 
tions are  usually  rather  gradual.  Various  schemes  of  a  graphical 
character  have  been  devised  to  exploit  this  idea,  for  example  by 
Johnson,  O'Brien,  and  Isaacs  [J.7],  Arthur  [A.3],  Munk  and  Traylor 
[M.16],  Suquet  [S.30],  and  Pierson  [P.8].  Figure  5.6.1  is  a  refraction 
diagram  for  waves  passing  over  a  shoal  in  an  otherwise  level  bottom 
in  the  form  of  a  flat  circular  hump,  and  Fig.  5.6.2  is  a  picture  of  the 
actual  waves.  Both  figures  were  taken  from  a  paper  by  Pierson  [P.8], 
and  they  refer  to  waves  in  an  experimental  tank.  As  one  sees,  there 


134 


WATER   WAVES 


Fig.  5.6.1.  Theoretical  wave  crest-orthogonal  pattern  for  waves  passing  over  a 

clock  glass.  No  phase  shift 

is  fair  general  agreement  in  the  wave  patterns— even  good  agreement 
in  detail  over  a  good  part  of  the  area.  However,  near  the  center  of 
the  figures  there  are  considerable  discrepancies,  since  the  theoretical 
diagram  shows,  for  instance,  a  sharp  point  in  one  of  the  wave  crests 


WAVES    ON    SLOPING    BEACHES    AND    PAST    OBSTACLES 


135 


which  is  lacking  in  the  photograph.  The  fact  is  that  there  is  a  caustic 
in  the  rays  constructed  by  geometrical  optics  (i.e.  the  orthogonals 
to  the  wave  crests  have  an  envelope),  and  in  the  vicinity  of  such  a 
region  the  approximation  by  geometrical  optics  is  not  good.  One  of 


Fig.  5.6.2.  Shadowgraph  for  waves  of  moderate  length  passing  over  a  clock  glass 

the  interesting  features  of  Fig.  5.6.2  is  that  the  shoal  in  the  bottom 
results  in  wave  crests  which  cross  each  other  on  the  lee  side  of  the 
shoal,  although  the  oncoming  waves  form  a  single  train  of  plane 
waves.  Figure  5.6.3  is  an  aerial  photograph  (again  taken  from  the 


130 


WATER    WAVES 


paper  by  Pierson)  showing  the  same  effect  in  the  ocean  at  a  point 
off  the  coast  of  New  Jersey;  the  arrow  points  to  a  region  where  there 
would  appear  to  be  three  wave  trains  intersecting,  but  all  of  them 
appear  to  arise  from  a  single  train  coming  in  from  deep  water. 


Fig.  5.6.3.  Aerial  photograph  at  Great  Egg  Inlet,  New  Jersey 

In  the  case  of  sufficiently  shallow  water  Lowell  [L.I 6]  has  studied 
the  conditions  under  which  the  approximation  by  geometrical  optics 
is  valid;  his  starting  point  is  the  linear  shallow  water  theory  (for 
which  see  Ch.  10.13)  in  which  the  propagation  speed  of  waves  is 
Vgh,  with  h  the  depth  of  the  water,  and  it  is  thus  independent  of 
the  wave  length.  Eckart  [E.2,  3]  has  devised  an  approximate  theory 


WAVES    ON    SLOPING   BEACHES    AND    PAST   OBSTACLES  187 

which  makes  it  possible  to  deal  with  waves  in  both  deep  and  shallow 
water,  as  well  as  in  the  transition  region  between  the  two. 

There  is  an  interesting  application  of  the  theory  of  water  waves  to 
a  problem  in  seismology  which  will  be  explained  here  even  though  it 
is  necessary  to  go  somewhat  beyond  the  linear  theory  on  which  this 
part  of  the  book  is  based.  We  have  seen  in  Chapter  3  above  that  the 
displacements,  velocities,  and  pressure  variations  in  a  simple  harmonic 
standing  wave  die  out  exponentially  in  water  of  infinite  depth. 
However,  it  was  pointed  out  by  Miche  [M.  8]  that  this  is  true  only  of 
the  first  order  terms  in  the  development  of  the  basic  nonlinear  theory 
with  respect  to  the  wave  amplitude;  if  the  development  is  carried  out 
formally  to  second  order  it  turns  out  that  the  pressure  fluctuates 
with  an  amplitude  that  does  not  die  out  with  the  depth,  but  depends 
on  the  square  of  the  amplitude.  (For  progressing  waves,  this  is  not 
true. )  In  addition,  the  second  order  pressure  variation  has  a  frequency 
which  is  double  the  frequency  of  the  linear  standing  wave.  (It  is  not 
hard  to  see  in  a  general  way  how  this  latter  nonlinear  effect  arises 
mathematically.  In  the  Bernoulli  law,  the  nonlinear  term  of  the  form 
0y  +  0%  would  lead,  through  an  iteration  process  starting  with 
0  =  Aemy  cos  mx  cos  at,  to  terms  involving  cos2  at  and  thus  to 
harmonics  with  the  double  frequency.)  It  happens  that  seismic  waves 
in  the  earth  of  very  small  amplitudes  — called  microseisms  — and  of 
periods  of  from  3  to  10  seconds  are  observed  by  sensitive  seismo- 
graphs; these  waves  seem  unlikely  to  be  the  result  of  earthquakes  or 
local  causes;  rather,  a  close  connection  between  microseisms  and  dis- 
turbed weather  conditions  over  the  ocean  was  noticed.  However,  since 
it  was  thought  that  surface  waves  in  the  ocean  lead  to  pressure 
variations  which  die  out  so  rapidly  in  the  depth  that  they  could  not 
be  expected  to  generate  observable  waves  in  the  earth,  it  was  thought 
unlikely  that  storms  at  sea  could  be  a  cause  for  microseisms.  The  result 
of  Miche  stated  above  was  invoked  by  Longuet-Higgins  and  Ursell 
[L.  14]  in  1948  to  revive  the  idea  that  storms  at  sea  can  be  the  origin 
of  microseisms.  (See  also  the  paper  of  1950  by  Longuet-Higgins 
[L.13].)  In  addition,  Bernard  [B.8]  had  collected  evidence  in  1941 
indicating  that  the  frequency  of  microseisms  near  Casablanca  was 
just  double  that  of  sea  waves  reaching  the  coast  nearby;  the  same 
ratio  of  frequencies  was  noticed  by  Deacon  [D.  6]  with  respect  to 
microseisms  recorded  at  Kew  and  waves  recorded  on  the  north  coast 
of  Cornwall.  Further  confirmation  of  the  correlation  between  sea 
waves  and  microseisms  is  given  in  the  paper  of  Darby  shire  [D.  4]. 


138  WATER   WAVES 

A  reasonable  explanation  for  the  origin  of  microseisms  thus  seems  to  be 
available.  Of  course,  this  explanation  presupposes  that  standing 
waves  are  generated,  but  Longuet-Higgins  has  shown  that  the  needed 
effects  are  present  any  time  that  two  trains  of  progressing  waves 
moving  in  opposite  directions  are  superimposed,  and  it  is  not  hard  to 
imagine  that  such  things  would  occur  in  a  storm  area — for  example, 
through  the  superposition  of  waves  generated  in  different  portions  of 
a  given  storm  area.  It  might  be  added  that  Cooper  and  Longuet- 
Higgins  [C.3]  have  carried  out  experiments  which  confirm  quantita- 
tively the  validity  of  the  Miche  theory  of  nonlinear  standing  waves.  It 
is  perhaps  also  of  interest  to  refer  to  a  paper  by  Danel  [D.2]  in  which 
standing  waves  of  large  amplitude  with  sharp  crests  are  discussed. 

In  Chapter  6  some  references  will  be  made  to  interesting  studies 
concerning  the  location  of  storms  at  sea  as  determined  by  observa- 
tions on  shore  of  the  long  waves  which  travel  at  relatively  high  speeds 
outward  from  the  storm  area  (cf.  the  paper  by  Deacon  [D.6]). 

The  general  problem  of  predicting  the  character  of  wave  conditions 
along  a  given  shore  is,  of  course,  interesting  for  a  variety  of  reasons, 
including  military  reasons  (see  Bates  [B.6],  for  example).  Methods 
for  the  forecasting  of  waves  and  swell,  and  of  breakers  and  surf  are 
treated  in  two  pamphlets  [U.I,  2]  issued  by  the  U.S.  Hydrographic 
Office. 

A  necessary  preliminary  to  forecasting  studies,  in  general,  is  an 
investigation  of  ways  and  means  of  recording,  analyzing,  and  repre- 
senting mathematically  the  surface  of  the  ocean  as  it  actually  occurs 
in  nature.  Among  those  who  have  studied  such  questions  we  mention 
here  Seiwell  [S.9,  10]  and  Pierson  [P.IO'J.  The  latter  author  concerns 
himself  particularly  with  the  problem  of  obtaining  mathematical  re- 
presentations of  the  sea  surface  which  are  on  the  one  hand  sufficiently 
accurate,  and  on  the  other  hand  not  so  complicated  as  to  be  practically 
unusable.  The  surface  of  the  open  sea  is,  in  fact,  usually  extraordinar- 
ily complicated.  Figure  5.0.4  is  a  photograph  of  the  sea  (taken  from 
the  paper  by  Pierson)  which  bears  this  out.  Pierson  first  tries  re- 
presentations employing  the  Fourier  integral  and  comes  to  the  con- 
clusion that  such  representations  would  be  so  awkward  as  to  preclude 
their  use.  (In  Chapter  6  we  shall  have  an  opportunity  to  see  that  it  is 
indeed  not  easy  to  discuss  the  results  of  such  representations  even  for 
motions  generated  in  the  simplest  conceivable  fashion— by  applying 
an  impulse  at  a  point  of  the  surface  when  the  water  is  initially  at  rest, 
for  example.)  Pierson  then  goes  on  to  advocate  a  statistical  approach 


WAVES    ON    SLOPING    BEACHES    AND    PAST    OBSTACLES 


139 


to  the  problem  in  which  various  of  the  important  parameters  are 
assumed  to  be  distributed  according  to  a  Gaussian  law.  These  deve- 
lopments are  far  too  extensive  for  inclusion  in  this  book -—besides,  the 


Fig.  5.6.4.  Surface  waves  on  the  open  sea 

author  is,  by  temperament,  more  interested  in  deterministic  theories 
in  mechanics  than  in  those  employing  arguments  from  probability 
and  statistics,  while  knowing  at  the  same  time  that  such  methods  are 
very  often  the  best  and  most  appropriate  for  dealing  with  the  com- 
plex problems  which  arise  concretely  in  practice.  It  would,  however, 
seem  to  the  author  to  be  likely  that  any  mathematical  representations 
of  the  surface  of  the  sea— whether  by  the  Fourier  integral  or  any  other 
integrals— would  of  necessity  be  complex  and  cumbersome  in  propor- 
tion to  the  complexity  of  that  surface  and  the  degree  to  which  details 
are  desired. 

Before  leaving  this  subject,  it  is  of  interest  to  examine  another  pho- 
tograph of  waves  given  by  Pierson  [P.10],  and  shown  in  Fig.  5.6.5. 
Near  the  right  hand  edge  of  the  picture  the  wave  crests  of  the  pre- 
dominant system  are  turned  at  about  45°  to  the  coast  line,  and  they 
are  broken  rather  than  continuous;  such  wave  systems  are  said  to  be 
short-crested.  About  half-way  toward  shore  it  is  seen  that  these 
waves  have  arranged  themselves  more  nearly  parallel  to  the  coast 
(indicating,  of  course,  that  the  water  has  become  shallower)  and  at 
the  same  time  the  crests  are  longer  and  less  broken  in  appearance, 


140 


WATER   WAVES 


Fig.  5.6.5.  Aerial  photograph  over  Oracoke 


WAVES    ON    SLOPING    BEACHES    AND    PAST    OBSTACLES  141 

though  no  single  one  of  them  can  be  identified  for  any  great  distance. 
Near  the  shore,  the  wave  crests  are  relatively  long  and  nearly  parallel 
to  it.  On  the  photograph  a  second  train  of  waves  having  a  shorter 
wave  length  and  smaller  amplitude  can  be  detected;  these  waves  are 
traveling  almost  at  right  angles  to  the  shore  (they  are  probably  caused 
by  a  breeze  blowing  along  the  shore)  and  they  are  practically  not 
diffracted.  Each  of  the  two  wave  trains  appears  to  move  as  though 
the  other  were  not  present:  the  case  of  a  linear  superposition  would 
thus  seem  to  be  realized  here.  One  observes  also  that  there  is  a  shoal, 
as  evidenced  by  the  crossed  wave  trains  and  the  white-water  due  to 
breaking  over  the  shoals. 

We  pass  next  to  a  brief  discussion  of  a  few  problems  in  which  our 
emphasis  is  on  the  methods  of  solution,  which  are  different  from  those 
employed  in  the  preceding  chapters  of  Part  II.  The  first  such  problem 
to  be  discussed  employs  what  is  called  the  Wicner-Hopf  method  of 
solving  certain  types  of  boundary  problems  by  means  of  an  ingenious, 
though  somewhat  complicated,  procedure  which  utilizes  an  integral 
equation  of  a  special  form.  This  method  has  been  used,  as  was  men- 
tioned in  the  introduction  to  this  chapter,  by  Heins  [11.12,  13]  and  by 
Keller  and  Weitz  [K.9]  to  solve  the  dock  problem  and  other  problems 
having  a  similar  character  with  respect  to  the  geometry  of  the  domains 
in  which  the  solution  is  sought.  However,  it  is  simpler  to  explain  the 
underlying  ideas  of  the  method  by  treating  a  different  problem,  i.e. 
the  problem  of  diffraction  of  waves  around  a  vertical  half-plane  —  in 
other  words,  Sommerfeld's  diffraction  problem,  which  was  treated  by 
a  different  method  in  the  preceding  section.  We  outline  the  method, 
following  the  presentation  of  Karp  [K.3].  The  mathematical  formula- 
tion of  the  problem  is  as  follows.  A  solution  q>(a\  y)  of  the  reduced 
wave  equation 


(5.6.1)  VV  +  &V  =  0 

is  to  be  found  subject  to  the  boundary  condition 

(5.6.2)  <py  =  0         for  y  =  0,         x  >  0 

and  regular  in  the  domain  excluding  this  ray  (cf.  Fig.  5.6.6).  In  addi- 
tion, a  solution  in  the  form 

(5.6.3)  <P=<PQ+Vi 
with  (jpQ  defined  by 

(5.6.4)  <p0  =  elk<*  cos  e°  +  v  sin  V,         0  <  00  <  27r, 


142 


WATER   WAVES 


and  with  <p±  prescribed  to  die  out  at  oo  is  wanted.  In  other  words,  a 
plane  wave  comes  from  infinity  in  a  direction  determined  by  the  angle 
00,  and  the  scattered  wave  caused  by  the  presence  of  the  screen,  and 


Fig.  5.6.6.  Diffraction  around  a  screen 

given  by  <pv  is  to  be  found.  It  is  a  peculiarity  of  the  Wienor-IIopf 
method—  not  only  in  the  present  problem  but  in  other  applications  to 
diffraction  problems  as  well—  that  the  constant  k  is  assumed  to  be  a 
complex  number  (rather  than  a  real  number,  as  in  the  preceding 
section  )  given,  say,  by  k  =  k±  +  ik^  with  &2  small  and  positive.  With 
this  stipulation  it  is  possible  to  dispense  with  conditions  on  (p±  of  the 
radiation  type  at  oo,  and  to  replace  them  by  boundedness  conditions. 
We  employ  a  Green's  function  in  order  to  obtain  a  representation 
of  the  solution  in  the  form  of  an  integral  equation  of  the  type  to  which 
the  Wiener-Hopf  technique  applies.  In  the  present  case  the  Green's 
function  G(x9  y\  #0,  y0)  is  defined  as  that  solution  of  (5.6.1)  in  the 
whole  plane  which  has  a  logarithmic  singularity  at  the  point  (#0,  yQ) 
and  dies  out  at  oo  (here  the  fact  that  k  is  complex  plays  a  role).  This 
function  is  well-known;  it  is,  in  fact,  the  Hankel  function 
of  the  first  kind: 


(5.6.5)     G(x,  y;x0,  y0)  = 


4 


(k[(x  -  #0)2  +  (y  - 


The  next  step  is  to  apply  Green's  formula  to  the  functions  9?  and  G  in 
the  domain  bounded  by  the  circle  C2  and  the  curves  marked  Cx  in 
Fig.  5.6.6.  Because  of  the  fact  that  G  is  symmetric,  has  a  logarithmic 
singularity,  and  that  99  and  G  both  satisfy  (5.6.1),  it  follows  by  argu- 
ments that  proceed  exactly  as  in  potential  theory  in  similar  cases  that 
?,  y)  can  be  represented  in  the  form 


WAVES    ON    SLOPING    BEACHES    AND    PAST   OBSTACLES  143 


(5.6.6)      <p(x,  y) 


f    r  i  dG 

=         fo>]  5- 
Jo          d?/o 


+  exp  {ik(x  cos  00  +  y  sin  00)} 


when  the  radius  of  the  circle  C2  is  allowed  to  tend  to  oo,  and  the  boun- 
dary condition  (5.6.2),  the  regularity  conditions,  and  conditions  at  oo 
are  used.  (A  mild  singularity  at  the  edge  x  =  0,  y  =  0  of  the  screen 
must  also  be  permitted.)  The  symbol  [9]  under  the  integral  sign  re- 
presents the  jump  in  <p  across  the  screen,  which  is  of  course  not  known 
in  advance.  The  object  of  the  Wiener-Hopf  technique  is  to  determine 
\y]  by  using  the  integral  equation  (5.6.6);  once  this  is  done  (5.6.6) 
yields  the  solution  (p(x,  y).  The  first  step  in  this  direction  is  to  differen- 
tiate both  sides  of  (5.6.6)  with  respect  to  y,  then  set  y  --=  0  and 
confine  attention  to  positive  values  of  x\  in  view  of  the  boundary 
condition  (5.6.2)  we  obtain  in  this  way  the  integral  equation 

/•» 

(5.6.7)  0  :-  ik  sin  0QeikxrQS°o  +        \(p(.rQ)]K(x  —  x^dx^         x  >  0. 

Jo 

The  kernel  K(x  —  #0)  of  the  integral  equation  is  given  by 

d2G 

(5.6.8)  ^-*o)-  v,- 

oyoy0  „  „„,»<) 

Equation  (5.6.7)  is  a  typical  example  of  an  integral  equation  solvable 
by  the  Wiener-Hopf  technique;  its  earmarks  are  that  the  kernel  is  a 
function  of  (x  x0)  and  the  range  of  integration  is  the  positive  real 
axis.  , 

The  starting  point  of  the  method  is  the  observation  that  the  integral 
in  (5.6.7)  is  strongly  reminiscent  of  the  convolution  type  of  integral 
in  the  theory  of  the  Fourier  transform.  In  fact,  if  the  limits  of  integra- 
tion in  (5.6.7)  were  from  —  oo  to  +  oo  and  the  equation  were  valid 
for  all  values  of  x,  it  could  be  solved  at  once  by  making  use  of  the 
convolution  theorem.  This  theorem  states  that  if 


/(<*)  =  f  °°  fM  exP  {—  ioucQ}dccQ  and  K(cn)  =  f  °°  K(x0)  exp  {— 

J  —00  J  —00 

—  i.e.  if  /  and  AT  are  the  Fourier  transforms  of  /  and  k  (cf.  Chapter  6)  — 


then  /(a)^(a)  =  f*    i(xQ)K(x  —  x^dx^  in  other  words,  the  transform 

J  —  -OC 

of  the  integral  on  the  right  is  the  product  of  the  Fourier  transforms 
of  the  function  f(xQ)  and  K(x0)  (cf.  Sneddon  [S.ll],  p.  24).  Conse- 
quently if  (5.6.7)  held  in  the  wider  domain  indicated,  it  could  be 
used  to  yield 


144  WATER   WAVES 


0  -  *(a)  +  [<p(a)]tf  (a), 

with  /t(a)  the  transform  of  the  nonhomogeneous  term  in  the  integral 
equation.  This  relation  in  turn  defines  the  transform  [<p(a)j  of 
[(p(x0)]  since  /&(a)  and  K(<x.)  are  the  transforms  of  known  functions, 
and  hence  [<P(MQ)]  itself.  We  are,  of  course,  not  in  a  position  to  proceed 
at  once  in  this  fashion;  but  the  idea  of  the  Wiener- Hopf  method  is  to 
extend  the  definitions  of  the  functions  involved  in  such  a  way  that 
one  can  do  so.  To  this  end  the  following  definitions  are  made 

g(x)  =  0,         x  >  0;         /(a?0)  -  \<p],         XQ  >  0 

(5.6.9)  -  h(x)  =  0,         x  <  0;         f(x0)  =  0,  XQ  <  0 

h(x)  =  ik  sin  6Qeikx  cos  \  x    >  0. 

Equation  (5.6.7)  can  now  be  replaced  by  the  equivalent  equation 

(5.6.10)  g(x)  =  h(x)  +  f  *  f(x0)K(x  —  x0)dKQ,          —  oo  <  j  <  oo. 

J— 00 

Here  g(x)  is  unknown  for  x  <  0  and  f(xQ)— the  function  we  seek-  is, 
of  course,  unknown  for  x0  >  0;  thus  we  have  only  one  equation  for 
two  unknown  functions.  Nevertheless,  both  functions  can  be 
determined  by  making  use  of  complex  variable  methods  applied  to 
the  Fourier  transform  of  (5.6.10);  we  proceed  to  outline  the  method. 
We  have,  to  begin  with,  from  (5.6.10): 

(5.6.11)  g(a)-Ma)  + /(a)jT(a), 
with  A(a)  and  /?(a)  known  functions  given  by 


(5.6.12)  fe(oc)  = 


k  sin  00 


a  —  k  cos  00' 

(5.6.13)  #(a)  =  ^-(k*  -a2)*. 

The  equation  (5.6.11)  is  next  shown  to  be  valid  in  a  strip  of  the  com- 
plex a-plane  which  contains  the  real  axis  in  its  interior.  We  omit  the 
details  of  the  discussion  required  to  establish  this  fact;  it  follows  in  an 
elementary  way  from  the  assumption  that  the  constant  k  has  a  posi- 
tive imaginary  part,  and  from  the  conditions  of  regularity  and  boun- 
dedness  imposed  on  the  solution  99  of  the  basic  problem.  K(<x.)  is 
factored*  in  the  form  (i/2)JSL(oO  •  JP+(a)  with  J?_(a)  =  (k  -  a)1/2, 
K+(QL)  =  (A;  +  a)1/2  with  K__  and  K+  regular  in  lower  and  upper  half- 

*  Such  a  manipulation  occurs  in  general  in  using  this  technique;  usually  a 
continued  product  expansion  of  the  transform  of  the  kernel  is  required. 


WAVES    ON    SLOPING    BEACHES    AND    PAST   OBSTACLES  145 

planes,  respectively.  The  equation  (5.6.11)  can  then  be  expressed, 
after  some  manipulation,  in  the  form 

(5.6.14)       _ 


(k  +  a)1/2       a  -  k  cos  00  [(*  +  a)1/2       (k  +  fccos  i 

~~  (A:  +  k  cos  00)1/2(a  -  k  cos  0^)       ¥  '  ~~  a 

where  the  symbols  g+  and  /_  refer  to  the  fact  that  g(oc)  and  /(a)  can  be 
shown  to  be  regular  in  upper  and  lower  half-planes  of  the  complex 
oc-plane,  respectively,  each  of  which  overlaps  the  real  axis.  In  fact,  the 
entire  left  side  of  (5.6.14)  is  regular  in  such  an  upper  half-plane,  and 
similarly  for  the  right  hand  side  in  a  lower  half-plane.  Thus  the  two 
sides  of  the  equation  define  a  function  which  is  regular  in  the  entire 
plane,  or,  in  other  words,  each  side  of  the  equation  furnishes  the 
analytic  continuation  of  the  function  defined  by  the  other  side. 
Finally,  it  is  rather  easy  to  show,  by  studying  the  behavior  of  g(oc) 
and  /(a)  at  oo,  that  the  entire  function  thus  defined  tends  uniformly 
to  zero  at  oo;  it  is  therefore  identically  zero.  Thus  (5.6.14)  defines  both 
g(a)  and  /(a)  since  they  can  be  obtained  by  equating  both  sides  se- 
parately to  zero.  Thus  g(x)  and  /(#)  arc  determined,  and  the  problem 
is,  in  principle,  solved. 

The  Wiener-Hopf  method  is,  evidently,  a  most  amusing  and  in- 
genious procedure.  However,  it  also  has  somewhat  the  air  of  a  tour  de 
force  which  uses  a  good  many  tools  from  function  theory  (while 
the  problem  itself  can  be  solved  very  nicely  without  going  into  the 
complex  domain  at  all,  as  we  have  seen  in  the  preceding  section)  and 
it  also  employs  the  artificial  device  of  assuming  a  positive  imaginary 
part  for  the  wave  number  k.  (This  brings  with  it,  we  observe  from 
(5.6.4),  that  while  the  primary  wave  dies  out  as  x  -»•  +  oo,  it  be- 
comes exponentially  infinite  as  x  ->  —  oo.)  In  addition,  the  problem 
of  diffraction  by  a  wedge,  rather  than  by  a  plane  barrier,  can  not  be 
solved  by  the  Wiener-Hopf  method,  but  yields  readily  to  solution  by 
the  simple  method  presented  in  the  preceding  section.  The  author 
hazards  the  opinion  that  problems  solvable  by  the  Wiener-Hopf 
technique  will  in  general  prove  to  be  solvable  more  easily  by  other 
methods— for  example,  by  more  direct  applications  of  complex  in- 
tegral representations,  perhaps  along  the  lines  used  to  solve  the 
difficult  mixed  boundary  problem  treated  in  section  5.4  above. 

We  mention  next  two  other  papers  in  which  integral  equations  are 


146  WATER   WAVES 

employed  to  solve  interesting  water  wave  problems.  The  first  of  these 
is  the  paper  by  Kreisel  [K.19]  in  which  two-dimensional  simple 
harmonic  progressing  waves  in  a  channel  of  finite  depth  containing 
rigid  reflecting  obstacles  are  treated.  Integral  equations  are  obtained 
by  using  an  appropriate  Green's  function;  Kreisel  then  shows  that 
they  can  be  solved  by  an  iteration  method  provided  that  the  domain 
occupied  by  the  water  does  not  differ  too  much  from  an  infinite  strip 
with  parallel  sides.  (Roseau  [R.9]  has  solved  similar  problems  for 
certain  domains  which  are  not  restricted  in  this  way. )  It  is  remarkable 
that  Kreisel  is  able  to  obtain  in  some  important  cases  good  and  useablc 
upper  and  lower  bounds  for  the  reflection  and  transmission  coeffi- 
cients. References  have  already  been  made  to  the  papers  by  John 
[J.5]  on  the  motion  of  floating  bodies.  In  the  second  of  these  papers 
the  problem  of  the  creation  of  waves  by  a  prescribed  simple  harmonic 
motion  of  a  floating  body  is  formulated  as  an  integral  equation.  This 
integral  equation  does  not  fall  immediately  into  the  category  of  those 
which  can  be  treated  by  the  Fredholm  theory;  in  fact,  its  theory  has 
a  number  of  interesting  and  unusual  features  since  it  turns  out  that 
the  homogeneous  integral  equation  has  non-trivial  solutions  which, 
however,  are  of  such  a  nature  that  the  nonhomogeneous  problem 
nevertheless  always  possesses  solutions. 

Various  problems  concerning  the  effect  of  obstacles  on  waves,  and 
of  the  wave  motions  created  by  immersed  oscillating  bodies,  have 
been  treated  in  a  series  of  notable  papers  by  Ursell  [U.3,  4,  5  and 
U.8,  9, 10].  Ursell  usually  employs  the  method  of  expansions  in  terms 
of  orthogonal  functions,  or  representations  by  integrals  of  the  Fourier 
type,  as  tools  for  the  solution  of  the  problems. 

Finally,  it  should  be  mentioned  that  the  approximate  variational 
methods  devised  by  Schwinger  [S.5]  to  treat  difficult  problems  in  the 
theory  of  electromagnetic  waves  can  also  be  used  to  treat  problems  in 
water  waves  (cf.  Keller  [K.7]).  A  notable  feature  of  Schwinger's 
method  is  that  it  is  a  technique  which  concentrates  attention  on  the 
quantities  which  are  often  of  the  greatest  practical  importance,  i.e. 
the  reflection  and  transmission  coefficients,  and  determines  them, 
moreover,  without  solving  the  entire  problem.  Rubin  [R.13]  has  for- 
mulated the  problem  of  the  finite  dock— which  has  so  far  defied  all 
efforts  to  obtain  an  explicit  integral  representation  for  its  solution  — 
as  a  variational  problem  of  a  somewhat  unconventional  type,  and 
proved  the  existence,  on  the  basis  of  this  formulation,  of  solutions  be- 
having at  oo  like  progressing  waves. 


WAVES    ON    SLOPING    BEACHES    AND    PAST   OBSTACLES  147 

An  interesting  type  of  problem  which  might  well  have  been  dis- 
cussed at  length  in  this  book  is  the  problem  of  internal  waves.  This 
refers  to  the  occurrence  of  gravity  waves  at  an  interface  between  two 
liquids  of  different  density.  Such  problems  are  discussed  in  Lamb 
[L.3],  p.  370.  The  case  of  internal  waves  in  media  with  a  continuous 
variation  in  density  has  considerable  importance  also  for  tidal  motions 
in  both  the  atmosphere  (cf.  Wilkes  [W.2])  and  the  oceans  (cf. 
Fjeldstad  [F.4]). 


SUBDIVISION  B 

MOTIONS  STARTING  FROM  REST.  TRANSIENTS. 

CHAPTER  6 

Unsteady  Motions 

6.1.    General  formulation  of  the  problem  of  unsteady  motions 

In  the  region  occupied  by  the  water  we  seek,  as  usual,  a  harmonic 
function  0(x,  y,  z;  t)  which  satisfies  appropriate  boundary  conditions 
and,  in  addition,  appropriate  conditions  prescribed  at  the  initial  in- 
stant /  —  0.  At  the  free  surface  we  have  the  boundary  conditions 

(6.1.1)  -#i+i?«  =  0      \ 

^  1  for  y  =  0,         t  >  0 

(0.1.2)  *t+gn=--v\ 

in  terms  of  the  vertical  elevation  rj(<x,  z;  t)  of  the  free  surface  and  the 
pressure  p(x,  z;  t)  prescribed  on  the  surface.  As  always  in  mechanics, 
a  specific  motion  is  determined  only  when  initial  conditions  at  the 
time  t  -  0  are  given  which  furnish  the  position  and  velocity  of  all 
particles  in  the  system.  This  would  mean  prescribing  appropriate 
conditions  on  0  throughout  the  fluid  at  the  time  t  =  0,  but  since  we 
shall  assume  0  to  be  a  harmonic  function  at  t  =  0  as  well  as  for  t  >  0 
it  is  fairly  clear  that  conditions  prescribed  at  the  boundaries  of  the 
fluid  only  will  suffice  since  0  is  then  determined  uniquely  throughout 
its  domain  of  definition  in  terms  of  appropriate  boundary  conditions.* 
As  initial  conditions  at  the  free  surface,  for  example,  we  might  there- 
fore take 

(6.1.3)  r,(x,z;0)  =  MX,  z)   \  =  Q 

(6.1.4)  a 


with  /!  and  /2  arbitrary  functions  characterizing  the  initial  elevation 
and  vertical  velocity  of  the  free  surface. 

In  water  wave  problems  it  is  of  particular  interest  to  consider  cases 

*  We  shall  see  later  on  (sections  6.2  and  6.9)  that  the  solutions  are  indeed 
uniquely  determined  when  the  initial  conditions  are  prescribed  only  for  the 
particles  at  the  boundary  of  the  fluid. 

149 


150  WATER   WAVES 

in  which  the  motion  of  the  water  is  generated  by  applying  an  impul- 
sive pressure  to  the  surface  when  the  water  is  initially  at  rest.  To 
obtain  the  condition  appropriate  for  an  initial  impulse  we  start  from 
(6.1.2)  and  integrate  it  over  the  small  time  interval  0  <S  t  ^  r.  The 
result  is 


(6.1.5)  pdt  =  —  Q0(x,  0,  *;  r)  —  eg  \ 

Jo  Jo 


r)dt, 


since  0(x,  y,  z;  0)  can  be  assumed  to  vanish.  One  now  imagines  that 
r  ->  0+  while  p  ->  oo  in  such  a  way  that  the  integral  on  the  left  tends 
to  a  finite  value—  the  impulse  /  per  unit  area.  Since  it  is  natural  to 
assume  that  77  is  finite  it  follows  that  the  integral  on  the  right  vanishes 
as  r  ->  0+,  and  we  have  the  formula 

(6.1.6)  /  =  —  Q0(x,  0,  z;  0+) 

for  the  initial  impulse  per  unit  area  at  the  free  surface  in  terms  of  the 
value  of  0  there.  If/  is  prescribed  on  the  free  surface  (together  with 
appropriate  conditions  at  other  boundaries),  it  follows  that 
0(x,  y,  z;  0^  )  can  be  determined,  or,  in  other  words,  the  initial  velocity 
of  all  particles  is  known. 

It  is  also  useful  to  formulate  the  initial  condition  on  0  at  the  free 
surface  appropriate  to  the  case  in  which  the  water  is  initially  at  rest 
under  zero  pressure,  but  has  an  initial  elevation  rj(x9  z;  0).  The  condi- 
tion is  obtained  at  once  from  (6.1.2);  it  is 

(6.1.7)  0t(x,  0,  z;  0+)  -  -  grj(x,  z;  0+), 

since  p  =  0  for  t  >  0.  Prescribing  the  initial  position  and  velocity  of 
the  free  surface  is  thus  equivalent  to  prescribing  the  initial  values  of 
0  and  its  first  time  derivative  0t.  From  now  on  the  notation  0+  will 
not  be  used  in  formulating  initial  conditions—  instead  we  shall  simply 
write  0  instead  of  0+. 

6.2.  Uniqueness  of  the  unsteady  motions  in  bounded  domains 

It  is  of  some  interest  to  consider  the  uniqueness  of  the  unsteady 
motions,  for  one  thing  because  of  the  unusual  feature  pointed  out 
in  the  preceding  section:  it  is  sufficient  to  prescribe  the  initial  position 
and  Velocity,  not  of  all  particles,  but  only  of  those  on  the  boundary. 
A  uniqueness  proof  based  on  the  law  of  conservation  of  energy  will 
be  given. 

To  this  end,  consider  the  motion  of  a  bounded  volume  of  water  con- 
fined to  a  vessel  with  fixed  sides  but  having  a  free  surface  (cf.  Fig. 


UNSTEADY    MOTIONS 


151 


6.2.1).  In  Chapter  1  we  have  already  discussed  the  notion  of  energy 
and  its  time  rate  of  change  with  the  following  results.  For  the  energy 
E  itself  we  have,  obviously: 


h(x,z) 


Fig.  6.2.1.  Water  contained  in  a  vessel 


(6.2.1)     E(t)  -  Q  JJJ 


+ 


+ 


y]  dx  dy  dz. 


Here  R  refers  to  the  volume  occupied  by  the  water  at  any  instant. 
The  x,  z-planc  is,  as  usual,  taken  in  the  plane  of  equilibrium  of  the 
free  surface,  and  y  =  rj(x9  z;  t)  and  y  =  h(x9  z)  are  assumed  to  be  the 
equations  of  the  free  surface  and  of  the  containing  vessel,  respectively. 
The  expression  for  E  can  now  be  written  in  the  form 


(6.2.2)     E(t) 


1  tf  <« 


+  02Z)  dx  dy  dz 


By  S  is  meant  the  projection  on  the  #,  s-plane  of  the  free  surface  and 
the  containing  vessel.  In  Chapter  1  the  following  expression  for  the 
rate  of  change  of  the  energy  E  was  derived: 


(6.2.3) 


dE       ff 
—  =  JJ  [Q& 


-  vn)  -  pvn]   dS. 


152  WATER   WAVES 

By  R  is  meant  the  boundary  surface  of  R,  while  vn  means  the  normal 
component  of  the  velocity  of  R.  It  is  essential  for  our  uniqueness 
proof  to  observe  that  in  the  special  case  in  which  p  =  0  on  the  free 
surface  we  have 

dE 
(6.2.4)  —  =  0,         E  =  const. 

This  follows  at  once  from  the  fact  that  vn  =  0n  =  0  on  the  fixed  part 
of  the  boundary,  while  vn  =  0n  and  p  =  0  on  the  free  surface. 

So  far  no  use  has  been  made  of  the  fact  that  we  consider  only  a 
linear  theory  based  on  the  assumption  of  small  oscillations  about  the 
equilibrium  position.  Suppose  now  that  the  initial  position  and  velo- 
city of  the  water  particles  has  been  prescribed,  or,  as  we  have  seen  in 
the  preceding  section  that  rj(x9  z;  0)  and  0(x9  y,  z;  0)  are  given  func- 
tions: 

(      ^(*>  *:  °)  = /i(*»  *) 
I  0(x,  y,  z;  0)  =  f2(x9  y,  z). 

We  proceed  next  in  the  customary  way  that  one  uses  to  prove  unique- 
ness theorems  in  linear  problems.  Suppose  that  rjl9  019  and  r\^  02 
are  two  solutions  of  the  initial  value  problem.  Then  0  =  ^  —  02 
and  r\  =  ijl  —  r]2  arc  functions  which  satisfy  all  of  the  conditions 
originally  imposed  on  0t  and  17 f  except  that  fl  and  /2  in  (6.2.5)  would 
now  both  vanish,  and  the  free  surface  pressure  would  also  vanish 
(cf.  (6.1.2)  and  (6.1.7)).  (Here  the  linearity  of  our  problem  is  used  in 
an  essential  way. )  It  follows  therefore  that  dE/dt  =  0,  and  E  =  const, 
when  applied  to  0  and  77,  as  we  have  seen.  But  at  the  initial  instant 
rf  =  0  and  0  =  0,  so  that 


(0.2.6)  E  =  -  —         htdxdz, 

5 

from  (6.2.2)  as  applied  to0  =  <t>l  —  02  and  17  =  rjl  —  r]2.  Consequent- 
ly we  have  the  result 

(6.2.7)     fff  (0*  +  01  +  01)  dx  dy  dz  +  g  ff  ^  dx  dz  =  0, 

j  j  j  j  j 

R  S 

and  this  sum  obviously  vanishes  only  if  grad  0  =  0  and  rj  =  0— in 
other  words  it  follows  that  0l  =  02  (except  for  an  unessential  addi- 
tive constant),  rjl  =  rj29  and  the  uniqueness  of  the  solution  of  the 
initial  value  problem  is  proved. 


UNSTEADY   MOTIONS  153 

The  proof  given  here  applies  only  to  a  mass  of  water  occupying  a 
bounded  region.  Nevertheless,  it  seems  clear  that  the  uniqueness  of 
the  solution  of  the  initial  value  problem  is  to  be  expected  if  the  water 
fills  an  unbounded  region,  provided  that  appropriate  assumptions 
concerning  the  behavior  of  the  solution  at  oo  are  made.  In  the  follow- 
ing, a  variety  of  such  cases  will  be  treated  by  making  use  of  the  tech- 
nique of  the  Fourier  transform  and,  although  no  explicit  discussion 
of  the  uniqueness  question  will  be  carried  out,  it  is  well-known  that 
uniqueness  theorems  (of  a  somewhat  restricted  character,  it  is  true) 
hold  in  such  cases  provided  only  that  appropriate  conditions  at  oo 
are  prescribed.  Recently  these  uniqueness  questions  have  been  treated 
by  Kotik  [K.I  7]  and  Finkelstein  [F.  3].  The  latter,  for  example,  proves 
the  uniqueness  of  unsteady  motions  in  unbounded  domains  in 
which  rigid  obstacles  occur,  and  both  writers  obtain  their  uniqueness 
theorems  by  imposing  relatively  weak  conditions  at  infinity.  In  sec.  9 
of  this  chapter  the  theory  devised  by  Finkelstein  will  be  discussed. 

6.3.   Outline  of  the  Fourier  transform  technique 

As  indicated  above,  the  solutions  of  a  series  of  problems  of  unsteady 
motions  in  unbounded  regions  as  determined  through  appropriate 
initial  conditions  will  be  carried  out  by  using  the  method  of  the 
Fourier  transform.  The  basis  for  the  use  of  this  method  is  the  fact  that 
special  solutions  0  of  our  free  surface  problems  are  given—  in  the 
case  of  two-dimensional  motion  in  water  of  infinite  depth,  for  example 
-b 


(6.3.1)  0(a:9  y\  t)  =  e™*  sin  (at  +  a)  cos  m(x  -  r) 
with 

(6.3.2)  a2  =  gw 

and  for  arbitrary  values  of  a  and  r.  From  these  solutions  it  is  possible 
to  build  up  others  by  superposition,  for  example,  in  the  form 


(6.3.3)  0(x,  y\  t)  =        h(a)e™v  sin  (at  +  <x.)dm 

Joo 
f(r)  cos  m(x  —  T)  dr, 
—  00 

in  which  h(a)  and  f(r)  are  arbitrary  functions.  This  in  turn  suggests 
that  the  Fourier  integral  theorem  could  be  used  in  order  to  satisfy 
given  initial  conditions,  since  this  theorem  states  that  an  arbitrary 


154  WATER   WAVES 

function  f(x)  defined  for  —  oo  <  x  <  oo  can  be  represented  in  the 
form 

i  r°°     r°° 

(6.3.4)  f(x)  =  -        da          /(^)  cos  a(^  —  x)  dr\ 

n  J  o        J  ~oo 

provided  only  that  f(x)  is  sufficiently  regular  (for  example,  that  J(x) 
is  piecewise  continuous  with  a  piecewise  continous  derivative  is  more 
than  sufficient)  and  that  f(x)  is  absolutely  integrable: 

(6.3.5)  r   |  /(a?)  |  do?  <  oo. 

J  —00 

Indeed,  we  see  that  if  we  set  h(a)  =  1/n  and  a  =  yt/2  in  (6.3.3)  we 
would  have  exactly  the  integral  in  (6.3.4)  for  t  —  0  and  y  =  0,  and 
hence  0(x9  0;  0)  would  reduce  to  the  arbitrarily  given  function  f(x). 
Thus  a  solution  would  be  obtained  for  an  arbitrarily  prescribed  initial 
condition  on  0. 

It  would  be  perfectly  possible  to  solve  the  problems  treated  below 
by  a  direct  application  of  (6.3.4),  and  this  is  the  course  followed  by 
Lamb  [L.3]  in  his  Chapter  IX.  Actually  the  problems  were  solved 
first  by  Cauchy  and  Poisson  (in  the  early  part  of  the  nineteenth  cen- 
tury), who  derived  solutions  given  by  integral  representations  before 
the  technique  of  the  Fourier  integral  was  known.  It  might  be  added 
that  these  problems  were  considered  so  difficult  that  they  formed  the 
subject  of  a  prize  problem  of  the  Academic  in  Paris. 

We  prefer,  in  treating  these  problems,  to  make  use  of  the  technique 
of  the  Fourier  transform  (following  somewhat  the  presentation  given 
by  Sneddon  [S.ll  ],  Chapter  7)  since  the  building  up  of  the  solution  to 
fulfill  the  prescribed  conditions  then  takes  place  quite  automatically. 
However,  the  method  is  based  entirely  upon  (6.3.4)  and  thus  also 
requires  for  its  validity  that  the  functions  f(x)  to  which  the  technique 
is  applied  should  be  represeritable  by  the  Fourier  integral.  This  is  a 
restriction  of  a  non-trivial  character:  for  example,  the  basically  im- 
portant solutions  given  by  (6.3.1)  are  not  representable  by  the  Fou- 
rier integral. 

It  is  useful  to  express  the  Fourier  integral  in  a  form  different  from 
(6.3.4).  We  write 


f( 


1  f°°  f* 

x)  =  -  lim          i(n)dri      cos  s(r]  —  x)ds. 

n  f->oo  J  -QO  J  0 


UNSTEADY   MOTIONS  155 


But  since   I     cos  s(rj  —  x)ds  =  2     cos  s(rj  —  -  x)ds  and 

f  *  f  ^ 

sin  S(TI  —  x)ds  =  0,    we   may    write         cos  s(rj  —  x)ds  = 

J  —5  JO 

f  * 

%         exp  [is(x  —  rj)}  ds,  and  hence 

J  —  s 

T          /*  GO  /*00 

(6.3.6)  /(a?)  =  —        e™*ds\      f(r))e^s  dr\. 

2^J~00  J-00 

We  now  set 

(6.3.7)  t(s) 


and  call  J(s)  the  Fourier  transform  of  /(#).  It  follows  at  once  from 
(6.3.6)  that  the  original  function  f(x)  is  obtained  from  its  transform 
l(s)  by  the  inversion  formula 


In  our  differential  equation  problems  it  will  be  essential  to  express 
the  Fourier  transform  of  the  derivatives  of  a  function  in  terms  of  the 
transform  of  the  function  itself.  Consider  for  this  purpose  the  trans- 
form of  dnfldxn  and  integrate  by  parts  (which  requires  that  dnf/dxn 
be  continuous): 


d»f 


s         -  —    e 


i 

• 
J 


If  the  (n  —  l)-st  derivative  is  to  possess  a  transform  it  must  tend  to 
zero  at  ^  oo  and  hence  we  have 


(6.3.9) 


dxn  dxn~l 


that  is,  the  transform  of  the  n-th  derivative  is  (is)  times  the  transform 
of  the  (n  —  l)-st  derivative.  By  repeated  application  of  this  formula 
we  obtain  the  result 


156  WATER   WAVES 


(6.8.10)  -        =  (isrl 

provided  that  f(x)  and  its  first  n  derivatives  are  continuous  and 
that  all  of  these  functions  possess  transforms. 

A  rigorous  justification  of  the  transform  technique  used  in  the 
following  for  solving  problems  involving  partial  differential  equations 
is  not  an  entirely  trivial  affair  (sec,  for  example,  Courant-Hilbert 
[C.10],  vol.  2,  p.  202  ff.).  Such  a  justification  could  be  given,  but  we 
shall  not  carry  it  out  here.  Indeed,  it  would  be  reasonable  to  take  the 
attitude  that  one  may  proceed  quite  formally  provided  that  one  veri- 
fies a  posteriori  that  the  solutions  obtained  in  this  way  really  satisfy 
all  conditions  of  the  problem.  This  is  usually  not  too  difficult  to  do, 
and,  since  the  relevant  uniqueness  theorems  are  available,  this  course 
is  perfectly  satisfactory. 

6.4.  Motions  due  to  disturbances  originating  at  the  surface 

We  wish  to  determine  first  the  motion  in  two  dimensions  due  to  the 
application  of  an  impulse  over  a  segment  of  the  surface  —a  ^x  ^  a 
at  t  =  0  when  the  water  is  at  rest  in  the  equilibrium  position.  We 
suppose  the  depth  h  of  the  water  to  be  constant  and  that  it  extends  to 
infinity  in  the  horizontal  direction.  The  velocity  potential  0(cc,  y;  t) 
must  satisfy  the  following  conditions.  It  must  be  a  solution  of  the 
Laplace  equation: 

(6.4.1)  &xx+0yy  =  0,      -oo<o?<oo,      -A^j/^0,     *^0, 
satisfying  the  boundary  conditions 

(6.4.2)  0tt  +  g$y  =  0;         y  =  0,         t  >0 
and 

(6.4.3)  0y  =  0,         y  =  -  h,         t  ^  0. 

The  first  of  these  conditions  states  that  the  pressure  on  the  free 
surface  is  zero  for  t  >  0.  As  initial  conditions  we  have,  in  view  of 
(6.1.6),  (6.1.7),  and  the  assumed  physical  situation: 

(6.4.4)  0(x,  0;  0)  =  --  /(a?), 

Q 

(6.4.5)  0t(x,  0;  0)  =  0, 

with  I(x)  the  impulse  per  unit  area  applied  to  the  free  surface.  In 


UNSTEADY   MOTIONS  157 

addition,  we  must  impose  conditions  at  oo.  These  are  that  0  and 
its  first  two  derivatives  with  respect  to  x,  y,  and  t  should  tend  to  zero 
at  oo  in  such  a  way  that  all  of  these  functions  possess  Fourier  trans- 
forms with  respect  to  x.  This,  in  particular,  requires  that  I(x)  in 
(6.4.4)  should  vanish  at  oo.  Actually  we  consider  only  the  special 
case  in  which 

T/   .        (  P  =  const.,   |  x  |  <  a 

(6.4.6)  I(x)  =  ' 

I  0,  |  x  |  >  a, 

i.e.  the  case  in  which  a  uniform  impulse  is  applied  to  the  segment 
|  x  |  <  a,  the  remainder  of  the  surface  being  left  undisturbed. 

The  solution  0(x,  y;  t)  will  now  be  determined  by  applying  the 
Fourier  transform  in  x  to  the  relations  (6.4.1)  — (6.4.5)  with  the  object 
(as  always  in  such  problems)  of  obtaining  a  simpler  problem  for  the 
transform  0(s,  y;  t)  =  (p(s,  y;  t).  Once  the  transform  99  has  been  found 
by  solving  the  latter  problem  the  inversion  formula  yields  the  solution 
0.  We  begin  by  applying  the  transform  to  (6.4.1),  i.e.  by  multiplying 
by  e~i8X  and  integrating  over  the  interval  —  oo  <  x  <  oo;  the  result  is 

(6.4.7)  -  s*<p(89  y;  t)  +  <pyy(s,  y;  t)  -  0 

in  view  of  (6.3.10)  and  the  assumed  behavior  of  0  at  oo.  (Clearly,  it  is 
also  necessary  to  suppose  that  the  operation  of  differentiating  0 
twice  with  respect  to  y  can  be  interchanged  with  the  operation  of 
integrating  0  over  the  infinite  interval.)  This  step  already  achieves 
one  of  the  prime  objects  of  the  approach  using  a  transform:  the  trans- 
form cp  satisfies  an  ordinary  differential  equation  instead  of  the  partial 
differential  equation  satisfied  by  0.  The  general  solution  of  (6.4.7)  is 

(6.4.8)  p(s,  y,  t)  =  A(s\  t)e  \*\v  +  B(s;  t)e~^v 

in  terms  of  the  arbitrary  "constants"  A(s;  t)  and  B(s;  t).  It  is  a  simple 
matter  to  find  the  appropriate  special  solution  that  also  satisfies  the 
bottom  condition  (6.4.3),  and  from  it  to  continue  (just  as  is  done  in 
what  follows)  in  such  a  way  as  to  find  the  solution  for  water  of  uniform 
depth.  However,  we  prefer  to  take  the  case  of  infinite  depth  and  to 
replace  (6.4.3)  by  the  condition  that  0V  -+  0  when  y  ->  —  oo.  The 
transform  <p  then  also  must  have  this  property  so  that  we  obtain  for 
<p($,  y\  t)  in  this  case  the  solutions 

(6.4.9)  <p(s,  !/;*)  =  A(s;  t)e\*\*. 


158  WATER   WAVES 

The  transform  is  next  applied  to  the  free  surface  condition  (6.4.2)  to 
obtain 

(6.4.10)  cpit  +  g(jpy  =  0,         y  =  0,         t  >  0 

and  upon  insertion  of  <p($,  0;  t)  from  (6.4.9)  we  find  for  A(s;  t)  the 
differential  equation 

(6.4.11)  Ati+g\8\A=0,         t>0. 

Finally,  the  initial  conditions  must  be  taken  into  account.  The  trans- 
form of  (6.4.5)  leads,  evidently,  to  the  condition  At(s;  0)  =  0,  and 
the  solution  of  (6.4.11)  satisfying  this  condition  is 

(6.4.12)  A(s\  t)  =  a(s)  cos  (Vg\s\t) 

with  a(s)  still  to  be  determined  by  using  (6.4.4).  From  (6.4.4)  we 
have  q?(s,  0;  0)  =  —  (l/g)/(s)  in  which  I(s)  is,  of  course,  the  transform 
of  I(x)  as  given  by  (6.4.4);  hence  a(s)  =  —  (I/Q)!(S)  and  we  have  for 
3>(s,  y;  t)  =  (f(s,  y;  t)  the  result 


(6.4.13)  0(8,  y;  *)  -  — 7(*)*W*cos  (Vg|*|  t). 

The  inversion  formula  (6.3.8)  then  leads  immediately  to  the  solution 

I          /»oo  

(6.4.14)  0(x,  y;  t)  =  —  — =        I(s)e^veisx  cos  (Vg\s\  t)  ds. 

eV2nJ  -a. 


In  our  special  case  (cf.  (6.4.6))  we  have  for 


a  2 Pa     sin  sa 

cos  sx  dx  =  •— :=-  *  —    —  ! 


-  P     Ca  2P   Cc 

T(*)  =  -7=\      r4~te  =  -—\ 

V2nJ  -a  V2n  Jo 

and  hence  finally  for  <Z>(#,  t/;  t)  the  solution 

2Pa  f00  sin  sa    qil  y    /—  4X , 

(6.4.15)     0(a?,  y;  t)  = *  v  cos  ^  cos  ( Vg*  0*. 

^rp  Jo      sa 

as  one  can  readily  verify.  For  the  free  surface  elevation  we  have 
(from  (6.1.2)): 


cos 


1          —  2  Pa         f  °°  sin  sa 

(6.4.16)  ri(x;  t)  =  —  -  #e=  -  pz  lim         --  e* 
S          nQVg  «->oJo      ^ 

sin  (Vgst)Vs  ds. 


UNSTEADY   MOTIONS  159 

One  observes  that  the  integrals  converge  well  for  all  y  <  0  because  of 
the  exponential  factor  esy%  i.e.  everywhere  except  possibly  on  the  free 
surface.  These  formulas  can  now  be  used  to  obtain  the  solution  for  the 
case  of  an  impulse  concentrated  on  the  surface  at  x  =  0;  one  need  only 
suppose  that  a  -r  0  while  P  ->  oo  in  such  a  way  that  the  total  im- 
pulse 2Pa  tends  to  a  finite  limit.  For  a  unit  total  impulse  we  would 
then  obviously  obtain  for  0  and  77  the  formulas: 

1    /•»  _ 

(6.4.17)  0(x,  y;  t)  =  —  —       es*  cos  sx  cos  (Vgs  t)  ds, 

KQ  Jo 

If*  _ 

(6.4.18)  w(a?;  0  =  —  ---  7-  lim         esv  cos  sx  sin  (  Vgs  t)\/s  ds. 

KQV8  v^o  Jo 


(We  define  r\(x\  t)  as  a  limit  for  y  ->  0  since  the  integral  obviously 
diverges  for  y  =  0.  This  would,  however,  not  be  necessary  in  (6.4.16).) 
By  operating  in  the  same  way,  one  can  easily  obtain  the  solutions 
corresponding  to  the  case  of  an  initial  elevation  of  the  free  surface  at 
time  t  =  0,  with  no  impulse  applied.  The  only  difference  would  be  that 
0  in  (6.4.4)  would  be  assumed  to  vanish  while  0t  in  (6.4.5)  would  be 
different  from  zero.  We  simply  give  the  result  of  such  a  calculation, 
but  only  for  the  limit  case  in  which  the  initial  elevation  is  concen- 
trated at  the  origin.  For  0  and  r\  the  formulas  are: 


VS  f  *  /—      ds 

(6.4.19)  0(x,  y;  t)  =  —  -^        e8*  cos  sx  sin  (  Vgs  t)  —  > 

n  J0  ys 

If00  _ 

(6.4.20)  rj(x\  t)  =  -  lim        e*v  cos  sx  cos  (Vgs  t)  ds. 

™  v-*0  Jo 

There  is  no  difficulty  in  treating  problems  having  cylindrical  sym- 
metry that  are  exactly  analogous  to  the  above  two-dimensional  cases. 
In  these  cases  also  one  could  begin  with  the  solutions  having  symmetry 
of  this  type  that  are  simple  harmonic  in  the  time  (cf.  Chapter  3): 


(6.4.21  )  0(r,  j/;  t)  —  eM 

with  a2  —  gm  (for  water  of  infinite  depth).  Here  the  quantity  r  is  the 
distance  Vx*  +  z2  from  the  j/-axis,  and  J0(mr)  is  the  Bessel  function 
of  order  zero  that  is  regular  at  the  origin.  One  could  now  build  up  more 
complicated  solutions  by  superposition  of  these  solutions  and  satisfy 
given  initial  conditions  by  using  the  Fourier-Bessel  integral.  This  is 
the  method  followed  by  Lamb  [L.3],  p.  429.  Instead  of  this  procedure, 


160  WATER   WAVES 

one  could  make  use  of  the  Hankel  transform  in  a  fashion  exactly  anal- 
ogous to  the  Fourier  transform  procedure  used  above  (cf.  Sneddon 
[S.ll],  p.  290,  and  Hinze  [H.15]).  We  content  ourselves  here  with 
giving  the  result  for  the  velocity  potential  0(r,  y;  t)  and  the  surface 
elevation  rj(r;  t)  due  to  the  application  of  a  concentrated  unit  impulse 
at  the  origin  at  t  =  0: 

1     /•»  _ 

(6.4.22)  <P(r,y;t)  =   —  -—         esyJ<>(sr)  cos  (Vgst)s  ds, 

2nQJo 

_  1  /»oo  __ 

(6.4.23)  n(r;  t)  =  -  -  T  lim        e»  J0(sr)  sin  (Vgs  t)s*12  ds. 

ZnQVg  *-»o  Jo 


*-»o 

Naturally  we  want  to  discuss  the  character  of  the  motions  furnished 
by  the  above  relations,  and  in  doing  so  we  come  upon  a  fact  that  holds 
good  in  all  problems  of  this  type:  it  is  a  comparatively  straightforward 
matter  to  obtain  an  integral  representation  for  the  solution,  but  not 
always  an  easy  matter  to  carry  out  the  details  of  the  discussion  of  its 
properties.  The  reason  for  this  is  not  far  to  seek  —it  is  due  to  the  fact 
that  the  solutions  are  given  in  terms  of  an  integral  over  an  integrand 
which  is  oscillatory  in  character  and  which  changes  rather  rapidly 
over  even  small  intervals  of  the  integration  variable  for  important 
ranges  in  the  values  of  the  independent  variables.  Hence  even  a  nu- 
merical integration  would  not  be  easy  to  carry  out.  The  fact  is  that  the 
motions  are  really  of  a  complicated  nature,  as  we  shall  see,  and  hence 
a  mathematical  description  of  them  can  be  expected  to  present  some 
difficulties.  Indeed,  the  phenomena  under  consideration  here  arc 
analogous  to  the  refraction  and  diffraction  phenomena  of  physical 
optics  and  thus  depend  on  intricate  interference  effects,  which  are 
further  complicated  in  the  present  instances  by  the  fact  that  the  wave 
motions  are  subject  to  dispersion,  as  we  have  seen  in  Chapter  3. 

Some  insight  into  the  nature  of  the  solutions  furnished  by  our  for- 
mulas can  be  obtained  by  expanding  the  integrands  in  power  series 
and  integrating  term  by  term  (cf.  Lamb.  [L.3],  p.  385).*  The  result 
for  r/(x;  t)  as  given  by  (6.4.20),  for  example,  is  found  to  be  (for  x  >  0): 


It  is  clear  that  there  is  a  singularity  for  x  =  0,  as  one  would  expect. 

*  The  subsequent  discussion  in  this  section  follows  closely  the  presentation 
given  by  Lamb. 


UNSTEADY   MOTIONS  161 

The  series  converges  for  all  values  of  the  dimensionless  quantity 
gt2l2x,  but  practically  the  series  is  useful  only  for  small  values  of 
gt2/2x,  i.e.  for  small  values  of  t,  or  large  values  of  x.  One  observes  also 
that  any  particular  "phase"  of  the  disturbance—  such  as  a  zero  of  rj> 
for  example—  must  propagate  with  a  constant  acceleration,  since  any 
such  phase  is  clearly  associated  with  a  specific  constant  value  of  the 
quantity  gt2/2x. 

It  is  in  many  respects  more  useful  to  find  an  asymptotic  represen- 
tation for  the  motion  valid  in  the  present  case  for  large  values  of 
the  quantity  gt2/2x,  for  which  the  power  series  are  not  very  useful 
because  of  their  slow  convergence.  Indeed,  the  asymptotic  represen- 
tation yields  all  of  the  qualitative  features  contained  in  the  exact 
solution  (6.4.24),  and  is  also  accurate  even  for  rather  small  values  of 
the  quantity  gt2/2x  (cf.  Sneddon  [S.ll],  p.  287).  For  this  purpose  it 
happens  to  be  rather  easy  to  work  out  an  asymptotic  development  of 
the  solution  that  is  valid  for  large  values  of  gt2/2x9  and  this  we  proceed 
to  do,  following  Lamb.  We  write  (6.4.19)  in  the  form 


(6 


-Iff00    °2y        /a2*          \ 
.4.25)        0(x,y;t)  =  ---          e*     sin  I  —  +  at]  da 

n     Uo  \  S  I 

r   °*y    .     (a*x          \       \ 

—        eg     sin  I  —  —  at]  da  } 

Jo  .\g  I        \ 

making  use  of  a  =  Vgs,  2oda  =  gd#.  New  quantities  |  and  o>  are 
introduced  in  (6.4.25)  by  the  relations 


from  which 

-co    -     g 

The  expression  (6.4.25)  is  thus  readily  found  to  take  the  form 


2gl/2    /»o> 

(6.4.26)  0(x,  0;  t)  -  -^       sin  (|2  -  o>2)  d£ 

n%  '  Jo 

where  f  is  introduced  as  new  variable  of  integration  and  y  is  assumed 
to  vanish.  The  corresponding  free  surface  elevation  is  given  by 


162 


WATER   WAVES 


(6.4.27) 


1 

g' 


as  one  readily  verifies.  In  order  to  study  the  last  expression  we  con- 
sider the  integral 

(6.4.28)         I     c  dc  ==::    I      c  dc  —   I     6  dc* 

i  i  I 

JO  */0  Jcu 

It  is  well  known  that 

(6.4.29) 

while  the  second  contribution  can  be  treated  as  follows: 

<-"2>  dt 


f 

Jo 


(6.4.30) 


f  *  c«*-"»  d!;  =  I  (  °°  - 
Jo)  2Jco2  V* 


i  r  °°      i  r°°     a  ~i 

=  _     H  ^<'—*>  +  -         <  "i  ^^-|2)^ 

^  L  L»    2  J^  J 


through  introduction  of  t  —  £2  as  new  variable,  and  an  integration 
by  parts.  We  show  next  that  the  final  integral  is  of  the  order  cw"1. 
as  follows: 


|  3  [ 

\       t    Z6  ut       ^    I        t 

Jet)2  Jw2 

r°° 

Jo)2 


dt  = 


Upon  considering  the  real  parts  of  (6.4.28),  (6.4.29),  (6.4.30),  and 
inserting  in  (6.4.27)  one  finds 


,)  =  -^  (g)»  [cos  (g  -  J 


(6.4.81) 


in  which  the  function  O(co~l)  refers,  as  one  readily  verifies,  to  a  term 
which  behaves  like  (g/2/4#)~1/2.  Consequently,  if  o>2  =  g^2/4<r  is  suffi- 
'ciently  large,  we  may  assume  for  the  free  surface  elevation  due  to  a 
concentrated  surface  elevation  at  x  =  0  and  t  =  0  the  approximate 
expression 


UNSTEADY   MOTIONS  163 

1     /g*2U         /g*2      n\ 

(6.4.32)  ,<«.„-,_  (!-)*„(£--_). 

By  continuing  the  integration  by  parts,  as  in  (6.4.30),  it  would  be 
possible  to  obtain  approximations  valid  up  to  any  order  in  the  quanti- 
ty co"1  =  (gJ2/^)-*1'2,  but  such  an  expansion  would  not  be  convergent; 
it  is  rather  an  asymptotic  expansion  correct  within  a  certain  order  in 
co"1  when  an  appropriate  finite  number  of  terms  in  the  expansion  is 
taken.  Expansions  of  this  type  are—  as  in  other  branches  of  mathe- 
matical physics—  very  useful  in  many  of  our  problems  and  we  shall 
have  many  other  occasions  to  employ  them. 

The  case  of  a  concentrated  point  impulse  applied  at  x  —  0  at  the 
time  t  —  0  can  be  treated  in  exactly  the  same  manner  as  the  case 
just  considered:  one  has  only  to  begin  with  the  solution  (6.4.17)  in- 
stead of  (6.4.19),  and  proceed  along  similar  lines.  In  particular,  the 
approximate  solution  valid  (to  the  same  order  in  a)~l)  for  large  values 
of  g/2/4a?  can  be  obtained;  the  result  for  the  free  surface  elevation  is 

-  2     /g*2\3/2    .    /gt2      x\ 

(6.4.33)  ,j(x;  t)  ~  --  -—    y-        sin    ^-  -  -  . 

'  \AtxJ  \±x       4/ 


The  method  used  to  derive  these  asymptotic  formulas  is  rather 
special:  it  cannot  be  very  easily  used  to  study  the  cylindrical  waves 
given  by  (6.4.23),  for  example.  We  turn,  therefore,  in  the  next  section 
to  the  derivation  of  asymptotic  approximations  in  all  of  these  cases 
by  the  application  of  Kelvin's  method  of  stationary  phase.  After- 
wards, the  motions  themselves  will  be  discussed  in  section  6.6  on  the 
basis  of  the  approximate  formulas. 

6.5.  Application  of  Kelvin's  method  of  stationary  phase. 

The  integrals  of  section  6.4  can  all  be  put  into  the  form 

(6.5.1)  I(k)  =  f  \(£,k)eik*W  d£ 

j  (i 

without  much  difficulty,  and  this  is  a  form  peculiarly  suited  to  an 
approximate  treatment  valid  for  large  values  of  the  real  constant  k. 
In  fact,  Kelvin  seems  to  have  been  led  to  the  approximate  method 
known  as  the  method  of  stationary  phase  through  his  interest  in 
problems  concerning  gravity  waves,  in  particular  the  ship  wave 
problem.  The  general  idea  of  the  method  of  approximation  is  as  fol- 


164  WATER  WAVES 

lows.  When  k  is  large  the  function  exp  (ik<p(!;)}  oscillates  rapidly  as  £ 
changes,  unless  (p(£)  is  nearly  constant,  so  that  the  positive  and  ne- 
gative contributions  to  the  value  of  I(k)  largely  cancel  out,  provided 
that  y>(f  ,  k)  is  not  a  rapidly  oscillating  function  of  £  when  k  is  large. 
Hence  one  might  expect  the  largest  contributions  to  the  integral  to 
arise  from  the  neighborhoods  of  those  points  in  the  interval  from  a 
to  b  at  which  <p(£  ),  the  phase  of  the  oscillatory  part  of  the  integral, 
varies  most  slowly,  i.e.,  from  neighborhoods  of  the  points  where 
(p'(£  )  =  o.  This  indeed  turns  out  to  be  the  case.  In  section  6.8  it  will 
be  shown  that 

(6.5.2)     /(*)  =  2  V(«r.  *) 


By  0  (  1  /&2/3)  we  mean  a  function  which  tends  to  zero  like  l/&2/3  as  k  ->  oo. 
In  these  expressions  the  sums  are  taken  over  all  the  zeros  ocr  of  q/(£) 
in  the  interior  of  the  interval  a  ^  £  ^  b  at  which  9p"(ar)  ^  0  and  over 
the  zeros  as  of  q>'($)  at  which  <p"(ocj  =  0  but  <p'"(ocj  ^  0.  The  sign 
of  the  term  ±  jr/4  in  the  first  sum  should  be  taken  to  agree  with  the 
sign  of  9/'(ar).  The  relation  (6.5.2)  is  valid  if  y(f,  k)  andg?(£)  arc  ana- 
lytic functions  of  £  in  a  ?g  {  <£  69  and  if  the  only  stationary  points  of 
9?(|)are  such  that  9?"(£)  and  <p'"(l;)  do  not  vanish  simultaneously.* 
We  proceed  to  obtain  the  approximate  solution  (6.4.32)  obtained 
in  the  previous  section  once  more  by  this  method.  The  motion  of  the 
water  was  to  be  determined  for  the  case  of  an  elevation  of  the  water 
surface  concentrated  at  a  point;  the  formula  for  the  velocity  potential 
was  put  in  the  form  (cf.  (6.4.25)): 


(6 


1  f  f00    *L        /o*x  \ 

.5.3)          0(x,  y;  t)  =  —  -  ]         e  o  sin  I  —  +  at]  do 

n  [Jo  \  S  J 

f00   *1         /o*x          \       ] 

—        e  ^  sin  1  —  —  at\  da  \ 

Jo  \  g  /       J 

*  If  a  zero  of  <p'(£  )  of  still  higher  order  should  occur,  then  terms  of  other  types 
would  appear,  and  the  error  would  die  out  less  rapidly  in  A:.  It  should  also  be 
noted  that  the  coefficient  function  y  of  section  6.8  is  assumed  to  be  independent 
of  k,  which  is  not  true  in  some  of  the  examples  to  follow.  However,  it  is  not 
difficult  to  see  that  the  proof  of  section  6.8  can  be  modified  quite  easily  in  such 
a  way  as  to  include  all  of  our  cases. 


UNSTEADY    MOTIONS  165 

This  can  in  turn  be  put  in  the  form 

Iff00  f  *  \ 

(6.5.4)  0(x9  y;  t)  =  --          em*  ei(mx+°t}  da  -        emy  ei(mx~at}  da 

n  Uo  Jo  ) 

with  m  —  a2fg.  It  is  understood  that  the  imaginary  part  only  is  to  be 
taken  at  the  end.  It  is  convenient  to  introduce  a  new  dimensionless 
variable  of  integration  as  follows: 

2x 

(6.5.5)  f  =  —  or, 

in  terms  of  which  (6.5.4)  is  readily  found  to  take  the  form 

(6.5.6)  0(x,  yi  t)  --=  -  -'    (  |     <"**  *'*  <**+*>  rff  -  f    e™»  e'*&~**>  d£\ 

'-XX  I  Jo  Jo  ) 

with 

op 

(6.5.7)  A:--  — 
v  4cT 

as  a  dimensionless  parameter.  The  quantity  m  is  of  course  also  a 
function  of  f,  and  exp  {m(g)y}  plays  the  role  of  the  function  ^(£)  in 
(6.5.1).  When  the  parameter  A:  is  large,  we  may  approximate  the 
solution  by  using  (6.5.2).  For  the  phases  </;(£)  we  have 

(0.5.8)  ?(£)-£2±2£ 

with  stationary  points  given  by 

(0.5.0)  ?/(£)  =  2f  ±2  =  0, 

and  we  see  that  |  ~  1  is  the  only  such  point  in  the  interval  0  <  f  <  oo 
over  which  the  integrals  are  taken.  Consequently  only  the  second  inte- 
gral in  (6.5.6)  possesses  a  point  of  stationary  phase,  and  at  this  point 
we  have 

(6.5.10)  ?/'(l)  =  2,         ?(!)=  -  1- 

We  obtain  therefore  from  (6.5.2)  the  approximate  formula 


(6.5.11  )  <t>(x,  y\  t)  c^      — 

"  nx 

as  one  readily  verifies,  and  this  formula  is  a  good  approximation  for 
large  values  of  k  —  gt*/&x.  We  can  also  calculate  the  free  surface  eleva- 


166  WATER   WAVES 

tion  77  in  the  same  way  from  rj  =  —  (l/g)&t  lv=o»*  the  result  is  easily 
found  to  be 

(6.5.12)  ^.^ 

just  as  before  (cf.  (6.4.32)). 

For  the  case  of  a  concentrated  impulse  the  method  of  stationary 
phase  as  applied  to  (6.4.17)  or  (6.4.18)  leads  to  the  following  approxi- 
mation valid  once  again  for  large  values  of  gt2/4>x: 

(6.5.13)  ^^^ 

and  this  coincides  with  the  result  given  in  (6.4.33). 

In  the  case  of  an  impulse  distributed  over  a  segment  one  obtains 
from  (6.4.16)  the  result 

2P          .    gt*a   .     /{>t*      n\ 

(6.5.14)  „(,;  0  =  -  --^  «n  ^  sm  ^  -    -J, 

valid  for  large  values  of  gt2/4tx.* 

For  the  ring  waves  furnished  by  (6.4.23)  the  asymptotic  formula  is 

gt*  gt* 

(6.5.15)  ,(r;  I)  =  -       jff  ™  —. 


To  obtain  this  formula  it  is  necessary  to  replace  the  Bessel  function 
J0(sr)  in  (6.4.23)  by  its  integral  representation 

2  ftt/2 

cos  (sr  cos  /?)  dp 


and  then  apply  the  method  of  stationary  phase  twice  in  succession. 
Since  such  a  procedure  is  discussed  later  on  in  dealing  with  the 
simplified  ship  wave  problem  (cf.  Chapter  8.1),  we  omit  a  discussion 
of  it  here,  except  to  remark  that  the  approximate  formula  (6.5.15)  is 
valid  for  any  r  =£  0  and  g*2/4r  sufficiently  large. 

*  It  may  seem  strange  that  this  formula  indicates  that  x  =  0  is  a  singular 
point  for  r),  while  x  =  0  is  not  singular  in  the  exact  formula  (6.4.16).  This  comes 
about  through  the  introduction  of  the  new  variable  (6.5.5)  and  the  parameter  k 
in  (6.5.7)  which  were  used  to  convert  the  original  integral  to  the  form  (0.5.1). 
However,  the  validity  of  the  formula  (6.5.2)  is  assured,  as  one  can  see  from 
section  6.8,  only  if  x  96  0. 


UNSTEADY   MOTIONS  167 

6.6.  Discussion  of  the  motion  of  the  free  surface  due  to  disturbances 
initiated  when  the  water  is  at  rest 

We  proceed  to  discuss  the  motions  of  the  water  surface  in  accord- 
ance with  the  results  given  in  the  preceding  section.  The  general 
character  of  the  motion  is  well  given  by  the  approximate  formulas, 
and  we  shall  therefore  confine  our  discussion  to  them.  We  observe 
first  that  the  oscillatory  factors  in  the  four  approximate  formulas 
(6.5.12)  — (6.5.15)  do  not  differ  essentially,  but  the  slowly  varying 
nonoscillatory  factors  are  different  in  the  various  cases:  (a)  at  a  fixed 
point  on  the  water  surface  the  disturbance  increases  in  amplitude 
linearly  in  t  in  the  case  of  an  initial  elevation  concentrated  at  a  point 
(cf.  (6.5.12)),  while  for  a  fixed  time  the  amplitude  becomes  large  for 
small  x  like  #-3/2;  (b)  in  the  case  of  an  initial  impulse  concentrated 
at  a  point  the  amplitude  increases  quadratically  in  t  at  a  fixed  point, 
while  for  a  fixed  time  the  amplitude  increases  like  ^~5/2  for  small 
x.  (In  these  limit  cases  the  approximate  formulas  are  valid  for  x  ^  0, 
since  the  only  other  requirement  is  that  the  quantity  gfifax  should  be 
large.)  The  behavior  of  these  solutions  near  x  —  0  is  not  very  sur- 
prising since  there  is  a  singularity  there.  The  behavior  at  any  fixed 
point  x  as  t  ->  oo  is,  however,  somewhat  startling:  the  amplitude  is 
seen  to  grow  large  without  limit  as  the  time  increases  in  both  of  these 
cases.  This  rather  unrealistic  result  is  a  consequence  of  the  fact  that 
the  singularity  at  the  origin  is  very  strong.  If  the  initial  disturbance 
were  finite  and  spread  over  an  area,  the  amplitude  of  the  resulting 
motion  would  always  remain  bounded  with  increasing  time,  as  one 
could  show  by  an  appeal  to  the  general  behavior  of  Fourier  trans- 
forms.* This  fact  is  well  shown  in  the  special  case  of  a  distributed 
impulse,  as  we  see  from  (6.5.14),  which  is  valid  for  all  x  ^=  0  and  large 
/:  the  amplitude  remains  bounded  as  t  ->  oo. 

The  general  character  of  the  waves  generated  by  a  point  disturbance 
is  indicated  schematically  in  the  accompanying  figures  which  show 
the  variation  in  surface  elevation  at  a  fixed  point  x  when  the  time  in- 
creases, and  at  a  fixed  time  for  all  x.  These  figures  are  based  on  the 
formula  (6.5.12)  for  the  case  of  an  initial  elevation;  the  results  for  the 
case  of  initial  impulse  would  be  of  the  same  general  nature. 

*  It  is  also  a  curious  fact  that  the  motion  given  by  (6.5.14)  for  the  case  of 
an  impulse  over  a  segment  requires  infinite  energy  input,  since  the  amplitude 
at  any  fixed  point  does  not  tend  to  y.ero.  For  the  case  of  an  initial  elevation  confined 
to  a  segment,  however,  the  wave  amplitude  would  die  out  with  increasing  time. 


168 


WATER   WAVES 


It  is  worth  while  to  discuss  the  character  of  the  motion  furnished 
by  (6.5.12)  in  still  more  detail.  It  has  already  been  remarked  that  any 
particular  phase— such  as  a  zero,  or  a  maximum  or  minimum  of  r\— is 
of  necessity  propagated  with  an  acceleration  since  each  such  phase  is 
associated  with  a  particular  constant  value  of  the  quantity  g*2/4#:  if 
the  phase  is  fixed  by  setting  g*2/4#  =  c,  then  this  phase  moves  in 
accordance  with  the  relation  x  =  g*2/4c.  The  formula  (6.5.12)  holds 
only  where  the  quantity  gt2/4>x  is  large,  and  hence  the  individual  pha- 
ses are  accelerated  slowly  in  the  region  of  validity  of  this  formula;  or, 


Fig.  6.6.1a,b  Propagation  of  waves  due  to  an  initial  elevation 


UNSTEADY   MOTIONS  169 

in  other  words,  the  phases  move  in  such  regions  at  nearly  constant 
velocity.  Also,  for  not  too  great  changes  in  x  or  t  the  waves  behave 
very  nearly  like  simple  harmonic  waves  of  a  certain  fixed  period  and 
wave  length.  This  can  be  seen  as  follows.  Suppose  that  we  vary  t 
alone  from  /  ==  tQ  to  t  •=  tQ  +  At.  We  may  write  for  the  phase  <p: 


as  one  readily  verifies.  Thus  if  At/tQ  is  small,  i.e.  if  the  change  At  in  the 
time  is  small  compared  with  the  total  lapse  of  time  since  the  motion 
was  initiated,  we  have  for  the  change  in  phase: 


Consequently  the  period  T  =  At  of  the  motion  corresponding  to  the 
change  Aq>  —  2n  in  the  phase  is  given  approximately  by  the  formula 

(6.6.1)  T 

0 

The  accuracy  of  this  formula  is  good,  as  we  know,  if  T/tQ  c^L 

is  small,  and  this  is  the  case  since  g/§M#o  *s  always  assumed  to  be  large. 

Thus  the  period  at  any  fixed  point  varies  slowly  in  the  time.  In  the 

same  way  one  finds  for  the  local  wave  length  A  the  approximate 

formula 


(6.6.2)  A 

go 

by  varying  with  respect  to  <r  alone,  and  this  is  also  easily  seen  to  be 
accurate  if  gt%/4xQ  is  large.  Thus  for  a  fixed  position  x  the  period  and 
wave  length  both  vary  slowly,  and  they  decrease  as  the  time  increases, 
while  for  a  fixed  time  the  same  quantities  increase  with  x,  as  is  borne 
out  by  the  figures  shown  above. 

It  is  of  considerable  interest  next  to  compute  the  local  phase  velo- 
city—the velocity  of  a  zero  of  77,  for  example—  from  gt2/4x  =  c 
when  x  and  t  vary  independently;  the  result  is 

dx       2x 

(6-6-3>  ¥  =  T 

for  the  velocity  of  any  phase;  thus  for  fixed  x  the  phases  move  more 


170  WATER   WAVES 

slowly  as  the  time  increases,  but  for  fixed  t  more  rapidly  as  x  increases 
—that  is,  the  waves  farther  away  from  the  source  of  the  disturbance 
move  more  rapidly,  and  they  are  also  longer,  as  we  know  from  (6.6.2). 
The  wave  pattern  is  thus  drawn  out  continually,  and  the  waves  as 
they  travel  outward  become  longer  and  move  faster.  The  last  fact  is 
not  too  surprising  since  the  waves  in  the  vicinity  of  a  particular  point 
have  essentially  the  simple  character  of  the  sine  or  cosine  waves  of 
fixed  period  that  we  have  studied  earlier,  and  such  waves,  as  we  have 
seen  in  Chapter  3,  propagate  with  speeds  that  increase  with  the  wave 
length.  All  of  the  above  phenomena  can  be  observed  as  the  result  of 
throwing  a  stone  into  a  pond;  though  the  motion  in  this  case  is  three- 
dimensional  it  is  qualitatively  the  same,  as  one  can  see  by  comparing 
(6.5.15)  with  (6.5.12). 

There  is  another  way  of  looking  at  the  whole  matter  which  is 
prompted  by  the  last  observations.  Apparently,  the  disturbance  at 
the  origin  acts  like  a  source  which  emits  waves  of  all  wave  lengths  and 
frequencies.  But  since  our  medium  is  a  dispersive  medium  in  which 
the  propagation  speed  of  a  particular  phase  increases  with  its  wave 
length,  it  follows  that  the  disturbance  as  a  whole  tends  with  increasing 
time  to  break  up  into  separate  trains  of  waves  each  of  which  has  ap- 
proximately the  same  wave  length,  since  waves  whose  lengths  differ 
move  with  different  velocities.  However,  it  would  be  a  mistake  to 
think  that  such  wave  trains  or  groups  of  waves  themselves  move  with 
the  phase  speed  corresponding  to  the  wave  length  associated  with  the 
group.  If  one  fixes  attention  on  the  group  as  a  whole  rather  than  on 
an  individual  wave  of  the  group,  the  velocity  of  the  group  will  be  seen 
to  differ  from  that  of  its  component  waves.  The  phase  velocity  for  the 
present  case  can  be  obtained  in  terms  of  the  local  wave  length  readily 
from  the  equation  (6.6.3)  by  expressing  its  right  hand  side  in  terms 
of  the  local  wave  length  through  use  of  (6.6.2);  the  result  is 

dx       2x       i  In  ; 

(6.6.4)  —  =  —  =  l/^L. 

V          '  *          *          V  271 

On  the  other  hand,  the  position  x  of  a  group  of  waves  of  fixed  wave 
length  A  at  time  t  is  given  closely  by  the  formula 


(6.6.5)  «  = 

as  we  see  directly  from  (6.6.2),  so  that  the  velocity  of  the  group  is 


UNSTEADY   MOTIONS  171 

which  is,  evidently,  just  half  the  phase  speed  of  its  com- 
ponent waves.  In  other  words,  the  component  waves  in  a  particular 
group  move  forward  through  the  group  with  a  speed  twice  that  of  the 
group. 

Finally,  we  observe  that  these  results  are  in  perfect  accord  with  the 
discussion  in  Chapter  3  concerning  the  notions  of  phase  and  group 
velocity.  The  phase  speed  c  for  a  simple  harmonic  wave  of  wave  length 
A  in  water  of  infinite  depth  is  given  (cf.  (3.2.3)!),  by  c  =  VgA/2jr, 
and  this  is  also  the  phase  speed  of  the  waves  whose  wave  length  is 
A  — as  we  sec  from  (6.6.4).  We  have  also  defined  in  section  3.4  the 
notion  of  group  velocity  for  simple  harmonic  waves  in  water  of  infinite 
depth,  and  found  it  to  be  just  half  the  phase  velocity.  The  kinematic 
definition  of  the  group  velocity  given  in  section  3.4  was  obtained  by 
the  superposition  of  trains  of  simple  harmonic  waves  of  slightly  differ- 
ent wave  length  and  amplitude,  while  in  the  present  case  the  waves 
arc  the  result  of  a  superposition  of  waves  of  all  wave  lengths  and 
periods.  However,  the  principle  of  stationary  phase,  which  furnishes 
the  approximate  solution  studied  here,  in  effect  says  that  the  main 
motion  in  certain  regions  is  the  result  of  the  superposition  of  waves 
whose  wave  lengths  and  amplitudes  differ  arbitrarily  little  from  a 
certain  given  value.  The  results  of  the  analysis  in  the  present  case  are 
thus  entirely  consistent  with  the  analysis  of  section  3.4. 

At  any  time,  therefore,  the  surface  of  the  water  is  covered  by  groups 
of  waves  arranged  so  that  the  groups  having  waves  of  greater  length 
are  farther  away  from  the  source.  These  groups,  therefore,  tend  to 
separate,  as  one  sees  from  (6.6.4).  The  waves  in  a  given  group  do  not 
maintain  their  amplitude,  however,  as  the  group  proceeds:  one  sees 
readily  from  (6.5.12)  in  combination  with  (6.6.2)  that  their  amplitude 
is  proportional  to  l/^/x  for  waves  of  fixed  length  L 

The  above  interpretations  of  the  results  of  the  basic  theory  are  all 
borne  out  by  experience.  Figure  6.6.2  shows  a  time  sequence  of  photo- 
graphs of  waves  (given  to  the  author  by  Prof.  J.  W.  Johnson  of  the 
University  of  California  at  Berkeley)  created  by  a  disturbance 
concentrated  in  a  small  area:  the  decrease  in  wave  length  at  a  fixed 
point  with  increasing  time,  the  increase  in  the  wave  lengths  near  the 
front  of  the  outgoing  disturbance  as  the  time  increases,  the  general 
drawing  out  of  the  wave  pattern  with  time,  the  occurrence  of  well- 
defined  groups,  etc.  are  well  depicted. 

An  interesting  development  in  oceanography  has  been  based  on  the 
theory  developed  in  the  present  section.  Deacon  [D.  6,  7]  and  his 


172 


WATER    WAVES 


Fig.  0.6.2.   Waves  due  to  a  concentrated  disturbance 


UNSTEADY    MOTIONS 


173 


Fig.  0.6/2.  (Continued) 


174  WATER   WAVES 

associates  have  carried  out  studies  which  correlate  the  occurrence  of 
storms  in  the  Atlantic  with  the  long  waves  which  move  out  from  the 
storm  areas  and  reach  the  coast  of  Cornwall  in  a  relatively  short  time. 
By  analyzing  the  periods  of  the  swell,  as  determined  from  actual 
wave  records,  it  has  been  possible  to  identify  the  swell  as  having  been 
caused  by  storms  whose  location  is  known  from  meteorological  obser- 
vations. Aside  from  the  interest  of  researches  of  this  kind  from  the 
purely  scientific  point  of  view,  it  is  clear  that  such  hindcasts  could,  in 
principle,  be  turned  into  methods  of  forecasting  the  course  of  storms 
at  sea  in  areas  lacking  meteorological  observations. 

6.7.  Waves  due  to  a  periodic  impulse  applied  to  the  water  when 
initially  at  rest.  Derivation  of  the  radiation  condition  for  purely 
periodic  waves 

In  section  3  of  Chapter  4  we  have  solved  the  problem  of  two-dimen- 
sional waves  in  an  infinite  ocean  when  the  motion  was  a  simple 
harmonic  motion  in  the  time  that  was  maintained  by  an  application 
of  a  pressure  at  the  surface  which  was  also  simple  harmonic  in  the 
time.  In  doing  so,  we  were  forced  to  prescribe  radiation  conditions 
at  oo— effectively,  conditions  requiring  the  waves  to  behave  like  out- 
going progressing  waves  at  oo— in  order  to  have  a  complete  formula- 
tion of  the  problem  with  a  uniquely  determined  solution.  It  was 
remarked  at  the  time  that  a  different  approach  to  the  problem  would 
be  discussed  later  on  which  would  require  the  imposition  of  bounded- 
ness  conditions  alone  at  oo,  rather  than  the  much  more  specific  radia- 
tion condition.  In  this  section  we  shall  obtain  the  solution  worked  out 
in  4.3  without  imposing  a  radiation  condition  by  considering  it  as  the 
limit  of  an  unsteady  motion  as  the  time  tends  to  infinity.  However,  it 
has  a  certain  interest  to  make  a  few  remarks  about  the  question  of 
radiation  conditions  in  unbounded  domains  from  a  more  general  point 
of  view  (cf.  [S.  21]). 

In  wave  propagation  problems  for  what  will  be  called  here,  ex- 
ceptionally, the  steady  state,  i.e.,  a  motion  that  is  simple  harmonic  in 
the  time,  it  is  in  general  not  possible  to  characterize  uniquely  the 
solutions  having  the  desired  physical  characteristics  by  imposing  only 
boundedness  conditions  at  infinity.  It  is,  in  fact,  as  we  have  seen  in 
special  cases,  necessary  to  impose  sharper  conditions.  In  the  simplest 
case  in  which  the  medium  is  such  as  to  include  a  full  neighborhood  of 
the  point  at  infinity  that  is  in  addition  made  up  of  homogeneous  matter, 


UNSTEADY    MOTIONS  175 

the  correct  radiation  condition  is  not  difficult  to  guess.  It  is  simply 
that  the  wave  at  infinity  behaves  like  an  outgoing  spherical  wave 
from  an  oscillatory  point  source,  and  such  a  condition  is  what  is 
commonly  called  the  radiation,  or  Sommerfeld,  condition.  Among 
other  things  this  condition  precludes  the  possibility  that  there  might 
be  an  incoming  wave  generated  at  infinity— which,  if  not  ruled  out, 
would  manifestly  make  a  unique  solution  of  the  problem  impossible. 

If  the  refracting  or  reflecting  obstacles  to  the  propagation  of  waves 
happen  to  extend  to  infinity— for  example,  if  a  rigid  reflecting  wall 
should  happen  to  go  to  infinity —it  is  by  no  means  clear  a  priori  what 
conditions  should  be  imposed  at  infinity  in  order  to  ensure  the  unique- 
ness of  a  simple  harmonic  solution  having  appropriate  properties 
otherwise.*  A  point  of  view  which  seems  to  the  author  reasonable  is 
that  the  difficulty  arises  because  the  problem  of  determining  simple 
harmonic  motions  is  an  unnatural  problem  in  mechanics.  One  should  in 
principle  rather  formulate  and  solve  an  initial  value  problem  by 
assuming  the  medium  to  be  originally  at  rest  everywhere  outside  a 
sufficiently  large  sphere,  say,  and  also  assume  that  the  periodic 
disturbances  are  applied  at  the  initial  instant  and  then  maintained 
with  a  fixed  frequency.  As  the  time  goes  to  infinity  the  solution  of  the 
initial  value  problem  will  tend  to  the  desired  steady  state  solution 
without  the  necessity  to  impose  any  but  boundedness  conditions  at 
infinity.** 

The  steady  state  problem  is  unnatural  — in  the  author's  view,  at 
least— -because  a  hypothesis  is  made  about  the  motion  that  holds 
for  all  time,  while  Newtonian  mechanics  is  basically  concerned  with 
the  prediction  — in  a  unique  way,  furthermore— of  the  motion  of  a 
mechanical  system  from  given  initial  conditions.  Of  course,  in  me- 
chanics of  continua  that  are  unbounded  it  is  necessary  to  impose  con- 
ditions at  oo  not  derivable  directly  from  Newton's  laws,  but  for  the 
initial  value  problem  it  should  suffice  to  impose  only  boundedness 
conditions  at  infinity.  In  sec.  6.9.  the  relevant  uniqueness  theorem  for 
the  special  case  to  be  considered  later  is  proved. 

*  For  a  treatment  of  the  radiation  condition  in  such  cases  see  Rellich  [R.7], 
John  [J.5],  and  Chapter  5.5. 

**  The  formulation  of  the  usual  radiation  condition  is  doubtlessly  motivated 
by  an  instinctive  consideration  of  the  same  sort  of  hypothesis  combined  with  the 
feeling  that  a  homogeneous  medium  at  infinity  will  have  no  power  to  reflect 
anything  back  to  the  finite  region.  Evidently,  we  also  have  in  mind  here  only 
cases  in  which  no  free  oscillations  having  finite  energy  occur  —  if  such  modes  of 
oscillation  existed,  clearly  no  uniqueness  theorems  of  the  type  we  have  in  mind 
could  be  derived. 


176  WATER   WAVES 

If  one  wished  to  be  daring  one  might,  on  the  basis  of  these  remarks, 
formulate  the  following  general  method  of  obtaining  the  appropriate 
radiation  condition:  Consider  any  convenient  problem  in  which  the 
part  of  the  domain  outside  a  large  sphere  is  maintained  intact  and 
initially  at  rest.  (In  other  words,  one  might  feel  free  to  modify  in  any 
convenient  way  any  bounded  part  of  the  medium.)  Next  solve  the 
initial  value  problem  for  an  oscillatory  point  source  placed  at  any 
convenient  point.  Afterwards  a  passage  to  the  limit  should  be  made  in 
allowing  the  time  t  to  approach  oo,  and  after  that  the  space  variables 
should  be  allowed  to  approach  infinity.  The  behavior  at  the  far  distant 
portions  of  the  domain  should  then  furnish  the  appropriate  radiation 
conditions  independent  of  the  constitution  of  the  finite  part  of  the 
domain.  It  might  be  worth  pointing  out  specifically  that  this  is  a  case 
in  which  the  order  of  the  two  limit  processes  cannot  be  interchanged: 
obviously,  if  the  time  /  is  first  held  fixed  while  the  space  variables  tend 
to  infinity  the  result  would  be  that  the  motion  would  vanish  at  oo, 
and  no  radiation  conditions  could  be  obtained. 

The  writer  would  not  have  set  down  these  remarks  —  which  are  of  a 
character  so  obvious  that  they  must  also  have  occurred  to  many 
others—  if  it  were  not  for  two  considerations.  Every  reader  will  doubt- 
lessly have  said  to  himself:  "That  is  all  very  well  in  principle,  but  will 
it  not  be  prohibitively  difficult  to  carry  out  the  solution  of  the  initial 
value  problem  and  to  make  the  subsequent  passages  to  the  limit?" 
In  general,  such  misgivings  are  probably  all  too  well  founded.  How- 
ever, the  problem  concerning  water  waves  to  be  treated  here  happens 
to  be  an  interesting  special  case  in  which  (1)  the  indicated  program 
can  be  carried  out  in  all  detail,  and  (2)  it  is  slightly  easier  to  solve  the 
initial  value  problem  than  it  is  to  solve  the  steady  state  problem  with 
the  Sommerfeld  condition  imposed. 

We  restrict  ourselves  to  two-dimensional  motion  in  an  x,  z/-plane, 
with  the  y-  axis  taken  vertically  upward  and  the  #-axis  in  the  originally 
undisturbed  horizontal  free  surface.  The  velocity  potential  (p(x,  y\  t) 
is  a  harmonic  function  in  the  lower  half-plane: 

(6.7.1)  (pxx  +  q>yv  =  0,         y  <  0,         t  >  0. 

The  free  surface  boundary  conditions  are  (cf.  (6.1.1),   (6.1.2)): 


(6.7.2) 


UNSTEADY   MOTIONS  177 

As  usual,  r\  =  r\(x\  t)  represents  the  vertical  displacement  of  the  free 
surface  measured  from  the  #-axis,  and  p  —  p(x;  t)  represents  the 
pressure  applied  on  the  free  surface.  We  suppose  that  9?  and  its  first 
and  second  derivatives  tend  to  zero  at  oo  for  any  given  time  t— in  fact 
that  they  tend  to  zero  in  such  a  way  that  Fourier  transforms  exist— 
but  we  do  not,  in  accordance  with  our  discussion  above,  make  any 
more  specific  assumptions  about  the  behavior  of  our  functions  as 
/  ->  oo.  At  the  time  t  —  0  we  prescribe  the  following  initial  conditions 

(6.7.4)  <p(x,  0;  0)  ==  <pt(x,  0;  0)  =  0, 

which  state  (cf.  (6.1.6),  (6.1.7))  that  the  free  surface  is  initially  at 
rest  in  its  horizontal  equilibrium  position. 

In  what  follows  we  consider  only  the  special  case  in  which  the  sur- 
face pressure  p(x;  t)  is  given  by 

(6.7.5)  .  p(x;  t)  =  d(x)eimt,     t  >  0 

in  which  d(x)  is  the  Dirac  d-function.  We  have  not  made  explicit  use 
of  the  d-function  until  now,  but  we  have  used  it  implicitly  in  section 
6.1  in  dealing  with  concentrated  impulses.  It  is  to  be  interpreted  in 
the  same  way  here,  i.e.  as  a  symbol  for  a  limit  process  in  which  the 
pressure  is  first  distributed  over  a  segment  the  length  of  which  is 
considered  to  grow  small  while  the  total  pressure  is  maintained  at  the 
constant  value  one.  By  inserting  this  expression  for  p  in  (6.7.3)  and 
eliminating  the  quantity  rj  by  making  use  of  (6.7.2)  the  free  surface 
condition  is  obtained  in  the  form 

i(*>  .  * 

(6.7.6)  gVy  +  Vii  -  -  —  6(x)ewt9  t  >  0. 

Our  problem  now  consists  in  finding  a  solution  <fj(x,  y\  t)  of  (6.7.1) 
which  behaves  properly  at  oo,  and  which  satisfies  the  free  surface 
condition  (6.7.6)  and  the  initial  conditions  (6.7.4). 

We  proceed  to  solve  the  initial  value  problem  by  making  use  of  the 
Fourier  transform  applied  to  the  variable  x.  The  result  of  transforming 
(6.7.1)  is 

(6.7.7)  -s*$  +  ^,,  =  0, 

in  which  <p(s,  y\  t)  is  the  transform  of  y(x,  y\  t)  and  use  has  been  made 
of  the  conditions  at  oo.  The  bounded  solutions  of  (6.7.7)  for  y  <  0, 
*  >  0  are  all  of  the  form 


178  WATER   WAVES 

(6.7.8)  y(s,  y;  t)  =  A(s;  t)e'v. 

The  transform  is  now  applied  to  the  boundary  condition  (6.7.6), 
with  the  result: 

(6.7.9)  g<pv  +  fc,  =  -  -4=  ^7  «"*,  for  j,  =  0, 


and  on  substitution  of  <p(s,  0;  t)  from  (6.7.8)  we  find 

1      ico 

(6.7.10)  Att+g,A  =  -^7<"". 

The  initial  conditions  (6.7.4)  now  furnish  for  A (s;  t)  the  conditions 

(6.7.11)  ^(*;0)  =  ^(*;0)  =  0. 

The  solution  of  (6.7.10)  subject  to  the  initial  conditions  (6.7.11)  is 

i  fco  r*  £ta>(**T)        ,— 

(6.7.12)  A(s;  t)= =:—       — -_^~  sin  Vgs  rdr. 

VZn  Q  Jo    Vgs 

Finally,  we  insert  the  last  expression  for  A(s;  t)  in  (6.7.8)  and  apply 
the  inverse  transform  to  obtain  the  following  integral  representation 
for  our  solution  <p(x9y9t): 

ia)  f00  /•^uoU-t) 

(6.7.13)  qp(x,  y;  t)  =  —  —        e*y  cos  sx ^_—  sin  Vgs  rdrds. 

0rcJo  Jo    Vgs 

The  fact  that  the  solution  is  an  even  function  of  x  has  been  used  here. 
Our  object  now  is  to  study  the  behavior  of  this  solution  as  /  ->  oo. 
Since  y  is  negative  (we  do  not  discuss  here  the  limit  as  y  ->  0, 
i.e.  the  behavior  on  the  free  surface)  the  integral  with  respect  to  s 
converges  well  and  there  is  no  singularity  on  the  positive  real  axis  of 
the  complex  $-plane.  However,  the  passage  to  the  limit  t  ->  oo  is  more 
readily  carried  out  by  writing  the  solution  in  a  different  form  in 
which  a  singularity  — a  pole,  in  fact— then  appears  on  the  real  axis 
of  the  0-plane.  (It  seerns,  indeed,  likely  that  such  an  occurrence  would 
be  the  rule  in  any  considerations  of  the  present  kind  since  the  limit 
function  as  t  ->  oo  would  not  usually  be  a  function  having  a  Fourier 
transform,  and  one  could  expect  that  the  limit  function  would  some- 
how appear  as  a  contribution  in  the  form  of  a  residue  at  a  pole. )  It  is 


UNSTEADY   MOTIONS 


179 


convenient  to  deform  the  path  of  integration  in  the  $-plane  into  the 
path  L  indicated  in  the  accompanying  figure.  The  path  L  lies  on  the 


s-  plane 


Fig.  0.7.1.  Path  of  integration  in  s-plane 

positive  real  axis  except  for  a  semicircle  in  the  upper  half-plane  cen- 
tered at  the  point  s  =  o)2/g.  By  Cauchy's  integral  theorem  this  leaves 
the  function  tp  given  in  (6.7.13)  unchanged. 

We  now  replace  sin  Vgs  r  in  (6.7.13)  by  exponentials  and  carry 
out  the  integration  on  i  to  obtain 


(6.7.14)     p(a?,0;0  - 


—~\e 
*Q    JL 


cos  sx 


2  Vgs  Vgs 


2  Vgs    Vgs 


I  "  "« 


cb. 


We  wish  now  to  consider  the  three  items  in  the  bracket  separately, 
and,  as  we  see,  two  of  them  do  indeed  have  a  singularity  at  s  =  o>2/g 
which  is  by-passed  through  our  choice  of  the  path  L.  The  first  two 
items  arc  rather  obviously  the  result  of  the  initial  conditions  and 
hence  could  be  expected  to  pro\ide  transients  which  die  out  as 
t  ->  oo.  This  is  in  fact  the  case,  as  can  be  seen  easily  in  the  following 
way :  That  branch  of  \/s  is  taken  which  is  positive  on  the  positive  real 
axis,  and  we  operate  always  in  the  right  half-plane.  If,  in  addition, 
s  is  in  the  upper  half-plane  it  follows  that  i(Vl&s  db  <*>)  has  its  real  part 
negative  (o>  being  real).  Consider  now  the  contribution  furnished  by 
the  uppermost  item  in  the  square  brackets.  Since  the  exponential  has 
a  negative  real  part  on  the  semi-circular  portion  of  the  path  L  it  is 
clear  that  as  t  -+  +  oo  this  part  of  the  path  makes  a  contribution 
that  tends  to  zero.  The  remaining  portions  of  L,  which  lie  on  the  real 


180  WATER    WAVES 

axis,  are  then  readily  seen  to  make  contributions  which  die  out 
like  1/t:  this  can  be  seen  easily  by  integration  by  parts,  for  example, 
or  by  application  of  known  results  about  Fourier  transforms.  The 
middle  item  in  the  square  brackets  has  no  singularity  on  the  real 
axis,  so  that  the  path  L  can  be  taken  entirely  on  the  real  axis;  thus, 
in  accordance  with  the  remarks  just  made  concerning  the  similar 
situation  for  the  first  item,  it  is  clear  that  this  contribution  also  dies 
out  like  1/t.  Thus  for  large  t  we  obtain  the  following  asymptotic  re- 
presentation for  99: 


(6.7.15) 

L  Ss  ~ 

Actually,  the  right  hand  side  is  the  solution  of  the  steady  state 
problem—  as  obtained,  for  example,  in  the  paper  of  Lamb  [L.2]  and 
by  a  different  method  by  us  in  section  4.3  (although  in  a  different  form) 
—when  the  condition  at  oo  is  the  radiation  condition  stating  that  cp 
behaves  like  an  out-going  progressing  wave.  The  steady  state  solution 
as  obtained  in  section  4.3  actually  was  a  little  more  awkward  to  obtain 
directly  through  use  of  the  radiation  condition  than  it  was  to  obtain 
the  solution  (6.7.13)  of  the  initial  value  problem.  In  particular,  the 
asymptotic  behavior  of  an  integral  representation  had  to  be  investi- 
gated in  the  former  case  also  before  the  radiation  condition  could  be 
used.  Thus  we  have  seen  in  this  special  case  that  the  radiation  condi- 
tion can  be  replaced  by  boundedness  conditions  (in  the  space  varia- 
bles, that  is)  if  one  treats  an  appropriate  initial  value  problem  instead 
of  the  steady  state  problem. 

Even  though  not  strictly  necessary  —  since  (6.7.15)  is  known  to 
furnish  the  desired  steady  state  solution—  it  is  perhaps  of  interest  to 
show  directly  that  the  right  hand  side  of  (6.7.15)  has  the  behavior 
one  expects  for  an  out-going  progressing  wave  when  x  ->  +  °o.  The 
procedure  is  the  same  as  that  used  in  discussing  (6.7.14):  The  factor 
cos  sx  is  replaced  by  exponentials  to  obtain 

(6.7.16) 

By  the  same  argument  as  above  one  sees  that  the  first  integral  makes 
a  contribution  that  tends  to  zero  as  x  ->  +  oo.  The  second  integral 
is  treated  by  deforming  the  path  L  over  the  pole  s  =  o>2/g  into  a  path 
M  which  consists  of  the  positive  real  axis  except  for  a  semi-circle 
in  the  lower  half-plane.  The  contribution  of  the  second  integral  then 


fn        fir    p*v  eisx  1     C   esv  e~lKX      "I 

<p(x,y;t)~-e*«\-±-\    J_L_  cfc  +  _  I    ?  -—  -<fr   . 
0         \*™JL&-<»*      ^2mJLg*-a>*     ] 


UNSTEADY   MOTIONS  181 

consists  of  the  residue  at  the  pole  plus  the  integral  over  the  path  M  . 
But  the  contribution  of  the  latter  integral  is,  once  more,  seen  to  tend 
to  zero  as  x  ->  +  oo  because  of  the  factor  e~i8X.  Thus  y(x,  y;  t)  behaves 
for  large  x  as  follows: 

o>2  /o>2  \ 

(I)         —  V       —  t[   —  X—(Ot  ] 

(6.7.17)  <p(x,  y)~  —  —  eg    e    \g        ', 

SQ 

and  this  does  in  fact  represent  a  progressing  wave  in  the  positive 
^-direction  which,  in  addition,  has  the  wavelength  2jrg/o>2  appropriate 
to  a  progressing  sine  wave  with  the  frequency  co  in  water  of  infinite 
depth. 

6.8.  Justification  of  the  Method  of  Stationary  Phase 

In  section  6.5  the  method  of  stationary  phase  was  used  (and  it  will 
be  used  again  later  on)  to  obtain  approximations  of  an  asymptotic 
character  for  the  solutions  of  a  variety  of  problems  when  these 
solutions  are  given  by  means  of  integrals  of  the  form 


(6.8.1) 


C 

= 

J  a 


and  the  object  is  to  obtain  an  approximation  valid  when  the  real 
constant  k  is  large.  Since  we  make  use  of  such  approximate  formulas 
in  so  many  important  cases,  it  seems  worth  while  to  give  a  mathema- 
tical justification  of  the  method  of  stationary  phase,  following  a  pro- 
cedure due  to  Poincard.  The  presentation  given  here  is  based  upon  the 
presentation  given  by  Copson  [C.5]. 

PoineareS's  proof  requires  the  assumption  that  <p(z)  and  y>(z)  are 
regular  analytic  functions  of  the  complex  variable  z  in  a  domain 
containing  the  segment  S:  a  <£  x  ^  b  of  the  real  axis  in  its  interior. 
(In  what  follows,  we  assume  a  and  b  to  be  finite,  but  an  extension  to 
the  case  of  infinite  limits  would  not  be  difficult.)  In  addition  <p(z)  is 
assumed  to  be  real  when  z  is  real.  These  conditions  are  more  restric- 
tive than  is  necessary  for  the  validity  of  the  final  result.  For  example, 
the  function  y  might  also  depend  on  &,  provided  that  y(x,  k)  is  not 
strongly  oscillatory,  or  singular,  for  large  values  of  k.  The  assumption 
of  analyticity  is  also  not  indispensable.  However,  these  generaliza- 
tions would  complicate  both  the  formulation  and  proof  of  our  theorem 
without  changing  their  essentials;  consequently  we  do  not  consider 
them  here. 


182  WATER   WAVES 

It  will  be  shown  that  the  main  contributions  to  I(k)  arise  from  the 
points  of  S  near  those  values  of  x  for  which  (p'(x)  =  0—  that  is,  near 
the  points  of  stationary  phase.  The  term  of  lowest  order  in  the  asymp- 
totic development  of  I(k)  with  respect  to  k  will  then  be  obtained  on 
the  basis  of  this  observation.  Kelvin  himself  offered  a  heuristic  argu- 
ment (cf.  sec.  5  above)  indicating  why  such  a  procedure  should  yield 
the  desired  result. 

Since  <p'  (z)  is  regular  in  the  domain  containing  S,  it  follows  that  its 
zeros  are  isolated.  Hence  S  can  be  divided  into  a  finite  number  of 
segments  on  which  (p(z)  has  either  one  stationary  point  or  no  stationary 
point.  We  shall  show  first  that  the  contribution  to  I(k)  from  a  segment 
containing  no  stationary  point  is  of  order  l/k.  Next  it  will  be  shown 
that  a  segment  containing  any  given  point  of  stationary  phase  can  be 
found  such  that  the  contribution  to  the  integral  furnished  by  the 
segment  is  of  lower  order  than  1  /k9  and  a  formula  for  this  contribution 
will  be  derived.  It  turns  out  that  this  contribution  of  lowest  order 
is  independent  of  the  length  of  the  segment  containing  the  point 
of  stationary  phase,  provided  only  that  the  segment  has  been  chosen 
short  enough.  Once  these  facts  have  been  proved,  it  is  clear  that  the 
lowest  order  contributions  to  the  integral  arc  to  be  found  by  adding 
the  contributions  arising  at  each  of  the  points  of  stationary  phase. 

Suppose,  then,  that  (p(x)  has  no  stationary  point  on  a  segment 
c  ^  x  ^  d  of  S.  We  may  write 


-  f 
Jc 


dx  -       -  ^X]-  — 


since  <p'(x)  ^Qinc^x^dby  hypothesis.  Integration  by  parts 
then  leads  to  the  result 


ik<p'(d)  ik(p'(c) 

d 
with  tp^x)  =  —  (y/y)-       Since 


_.  i  ( 

ikjc 


C 
<£ 

Jc 


dx 


because  of  the  fact  that  kq>(x)  is  real,  it  follows  that  the  integral  in  the 
above  expression  is  bounded.  Thus  /x  is  indeed  of  order  I/A:,  as  stated 
above.  It  might  be  noted  that  this  argument  really  does  not  require 
the  analyticity  of  <p  and  \p9  but  only  that  the  integrands  be  integrable 
and  that  integration  by  parts  may  be  performed.  Infinite  limits  for  the 
integrals  could  also  be  permitted  if  cp(x)  and  \p(x)  behave  appropriately 
at  oo. 


UNSTEADY   MOTIONS  183 

Suppose  now  that  <p(x)  has  one  stationary  value  at  x  =  a  in  the 
segment  a  —  el  ^  x  ^  a  +  £i>  sl  >  09  i.e.,  <p'(#  )  vanishes  only  at 
x  =  a  in  this  interval.  Suppose,  in  addition,  that  the  second  deriva- 
tive y"  (x)  does  not  vanish  at  x  —  a,  and  indeed  is  positive  there: 
<p"(<x.)  >  0.  (The  case  in  which  9/'(a)  is  negative  and  the  more  critical 
case  in  which  9?"(a)  =  0  will  be  discussed  later.)  We  shall  show  that 
a  positive  number  e  ^  sl  exists  such  that 


(6.8.2)    /,(*)  =  (*)«'*«  dx  = 


oe-, 

In  other  words,  we  shall  show  that  a  fixed  segment  of  length  %e  con- 
taining a  exists  such  that  its  contribution  to  /  is  independent  of  e  and 
is  of  order  l/\/k,  with  an  error  of  order  I/A;. 

To  prove  these  statements  we  begin  by  introducing  new  variables 
as  follows: 

(6.8.3)  x  =  a  +  u,         <p(x)  =  y(a)  +  w(u). 

Consider  first  the  integral  /2(&,  e^: 


(6.8.4)  /a(fc,  cx)  -  ^^W         ^fcM)(tt)  ^(a  +  M)  du  =  ^^(a)  J. 

J-«i 
It  is  convenient  to  write  the  integral  J  as  the  sum  of  two  terms: 

f°  rei 

(6.8.5)  J  ==         f«^(«i)  y(a  +  MI)  duj  +       elkw(u*>  y(a  +  ^2)  ^2 

J-ej  Jo 

-  Ji   +  t/2. 

Since  99(0?)  has  a  minimum  at  x  —  a,  it  follows  that  w(ul)  is  a  positive 
monotonic  function  in  the  interval  —  e1  ^  %  ^  0,  and  likewise 
w(t42)  in  the  interval  0  ^  i/2  ^  £x.  Hence  we  may  introduce  a  new 
integration  variable  t,  which  is  furthermore  real,  in  each  of  the  in- 
tegrals, defined  as  follows: 

5  —  w(ul)  in  —  e1  ^  %  5^  0,     and 


(6,8.6)  . 

\t2  =--  w(u2)  in  0  ^  u2  <^ 

In  each  interval  t  is  taken  as  the  positive  square  root.  The  integrals 
«/!  and  J2  become,  as  one  readily  sees: 


(6.8.7) 


184  WATER   WAVES 


with  tfj  ==  Vw(—  fii),  and  t2  =  Vwfa).  The  functions  %(£)»  w2(tf)  are 
solutions  of  w(ui)  =  t2.  For  w(u)  we  have  the  power  series  develop- 
ment 

(6.8.8)         w(u)  =  <p(oc  +  u)  —  9?(a)  =  a9u2  +  a3w3  +  .  .  . 

since  w?(0)  =  w'(Q)  =  0  (cf.  (6.8.3));  in  addition  2a2  =  g/'(«)  >  °> 
by  assumption.  We  suppose  that  this  series  converges  in  a  circle  which 
contains  the  entire  interval  —  £2  ^  u  ^  £2  in  its  interior,  with 
e2  <  fii-  Since  J2  =  w>(w2)  we  may  write 


for  0  <S  u2  ^  £         and 
—  £2  = 


Since  a2  7^  0  we  may  express  the  square  roots  as  power  series  in  ui 
and  then  invert  the  series  to  obtain  u^  and  u2  as  power  series  in  t,  as 
follows: 


(6.8.10) 

u,  =       c,«  +  c2<2  +  .  .  .. 


with  Cj  =  +  V/2/9?"(a).  Hence  we  may  write 


in  which  P^(t)  and  P2(0  are  convergent  power  series.  It  may  be  that 
these  series  do  not  converge  up  to  the  values  ^  and  t2  of  the  upper  li- 
mits of  the  above  integrals  Jl  and  J2  in  (6.8.7).  In  that  case  we  simply 
assume  the  length  of  the  segment  is  taken  to  be  still  less  than  2e2  so 
that  the  inversion  of  the  series  (6.8.8)  is  permissible  and  the  series 
Pi(t)  and  P2(0  converge  up  to  appropriate  values  ?x  and  £2.  It  is 
clear  that  numbers  ?x  and  12  with  these  properties  exist.  The  integrals 
Jt  and  J2  may  now  be  written  in  the  form* 


(6.8.11) 


J2 


=  f  h  e™*  {^(a)  +  tP^t)}  dt,       and 

j°_ 

-  clV(a)  pV*'8<tt  +  (t*e™*tPt(t)dt  =  J3  +  J4. 
jo  jo 


*  The  requirement  of  analyticity  for  (p  and  y  is  used  to  permit  this  simple 
introduction  of  t  as  variable  of  integration.  However,  the  existence  of  a  finite 
number  of  derivatives  would  clearlv  have  sufficed. 


UNSTEADY   MOTIONS 


185 


We  proceed  to  study  the  integrals  J3  and  «/4.  Upon  introducing  6  —  kt2 
as  new  variable  in  J3  we  obtain 


_  ci^(a)  r  »e__  dQ 

2  \/k  J  o     \/0 


But  we  may  write 

piQ 

(6.8.12)  —  d8 


(*Jcta  pi®  /*oo  piQ  /*oo    piQ 

—.  de  =  \     —  dB  -  \     — 

Jo     V^  Jo    \/^  J  kt2  \/v 


by  a  known  formula.  The  last  integral  can  now  be  shown  to  be  of 
order  l/\/&  by  integration  by  parts,  as  follows: 


The  first  contribution  on  the  right  hand  side  is  obviously  of  order 
since  12  is  a  fixed  number;  as  for  the  second,  we  have 


pv 

L-dB 


and  hence  the  second  contribution  is  also  of  order  l/\/k.  Thus  for  J3 
we  have  the  result 


(6.8.13) 


J    = 


The  integral  ./4  is  first  integrated  by  parts  to  obtain 


t 

Jo 


-  P,(0) 


and  hence 

(6.8.14)         |  J4  |  ^  1 


|  P,(0)  | 


P't(t)  |  di 


and  the  right  hand  side  is  thus  of  order  1/fc.  The  integral  Jl  can  ob- 
viously be  treated  in  the  same  way  as  J2  and  with  an  exactly  analogous 


186  WATER   WAVES 

result;  consequently  we  have  from  (6.8.13),  and  (6.8.14)  for  the  integral 
given  in  (6.8.4)  the  result 


(6.8.15)          !,<*,  e3)  =  /,(*)  =  y(a)  «'**w+i    +  0 


once  £3  has  been  chosen  small  enough.  One  observes  how  it  comes 
about  that  the  lowest  order  term  is  independent  of  the  values  of  ?x  and 
?2,  and  hence  of  the  length  of  the  segment:  the  entire  argument  re- 
quires only  that  ?x  and  £2  be  any  fixed  positive  numbers  since  one  needs 
only  the  fact  that  the  products  kl\  and  kl\  grow  large  with  k. 

If  (p  (x)  had  been  assumed  to  have  a  maximum  at  x  —  a,  with 
9/'(<x)  <  0,  the  only  difference  would  be  that  —  k<p"(u.)  and  —  rc/4 
would  appear  in  the  final  formula  instead  of  +  k(p"(<x.)  and  +  vt/4. 
Consequently,  in  all  cases  in  which  <p"  (a)  ^  0  we  have 

(6.8.16)  I,(k)  =  y(«)  (—^L 

U  \<p   (a) 

and  the  sign  of  the  term  w/4  should  agree  with  the  sign  of  <p"(oc). 
Finally,  in  case  <p"(oc)  =  0,  but  9?'"(oc)  ^  0  it  is  not  difficult  to 
derive  the  appropriate  asymptotic  formula  for  /(ft).  In  fact,  the  steps 
are  nearly  identical  with  those  taken  just  now  for  the  case  <p"(oc)  ^  0. 
One  introduces  x  =  a  +  u,  <p(x)  =  y(a)  +  ^(w)  as  before  and  then 
makes  use  of  power  series  in  the  variable  t  defined  by  23  =  w(u  )  in  the 
same  way  as  above.  The  result  is,  for  e  sufficiently  small: 

(6.8.17)  I2(k)  =  f  a+V**<*)  y(x)  dx 


=  f  a+ 


where  F(\)  refers  to  the  gamma  function.  Hence  the  contribution 
arising  from  the  stationary  point  is  now  of  a  different  order  of  magni- 
tude, i.e.,  of  order  1/&1/3  instead  of  I/ft1/2.  This  fact  is  of  significance 
in  the  case  of  the  ship  wave  problem  which  will  be  treated  later. 

Naturally  the  lowest  order  terms  in  I(k)  consist  of  a  sum  of  terms 
furnished  by  the  contributions  of  all  of  the  points  of  stationary  phase 
in  the  interval  S.  It  is  important  enough  to  bear  repetition  that  if  no 
such  points  exist,  then  I(k)  is  in  general  of  order  1/fc. 

In  case  a  stationary  point  falls  at  an  end  point  x  =  a  or  x  =  b  of 
the  interval  of  integration,  one  sees  readily  that  the  contribution 
furnished  by  such  a  point  to  I(k)  is  the  same  as  that  given  above  in 


UNSTEADY   MOTIONS  187 

case  <p"  ^  0  except  that  a  factor  1/2  would  appear  in  the  final  result. 
On  the  other  hand,  if  g/'  =  0  but  <p'"  ^  0  at  an  end  point,  then  the 
contribution  differs  in  phase  as  well  as  in  the  numerical  factor  from 
the  contribution  given  above  in  (6.8.17). 

6.9.     A  time-dependent  Green's  function.  Uniqueness  of  unsteady 
motions  in  unbounded  domains  when  obstacles  are  present 

In  sec.  6.2  above  the  uniqueness  of  unsteady  wave  motions  for 
water  confined  to  a  vessel  of  finite  dimensions  was  proved.  More  gener- 
al results  have  been  obtained  by  Kotik  [K.17],  Kamp£  de  Feriet  and 
Kotik  [K.l],  and  Finkelstein  [F.3]  with  regard  to  such  uniqueness 
questions.  In  the  present  section  a  rather  general  uniqueness  theorem 
will  be  proved,  following  the  methods  of  Finkelstein,  who,  unlike  the 
other  authors  mentioned,  obtains  uniqueness  theorems  when  obstacles 
arc  present  in  the  water.  The  essential  tool  for  this  purpose  is  a  time- 
dependent  Green's  function,  which  is  in  itself  of  interest  and  worth 
while  discussing  for  its  own  sake  quite  apart  from  its  use  in  deriving 
uniqueness  theorems.  With  the  aid  of  such  a  function,  for  example,  all 
of  the  problems  solved  in  the  preceding  sections  can  be  solved  once 
more  in  a  different  fashion,  and  still  other  and  more  complicated  un- 
solved problems  can  be  reduced  to  solving  an  integral  equation,  as 
we  shall  see. 

We  shall  derive  the  time-dependent  Green's  function  in  question 
for  the  case  of  three-dimensional  motion  in  water  of  infinite  depth, 
although  there  would  be  no  difficulty  to  obtain  it  in  other  cases  as 
well.  The  Green's  function  G  in  question  is  required  to  be  a  harmonic 
function  in  the  variables  (x,  y,  z)  with  a  singularity  of  appropriate 
character  at  a  certain  point  (£,  77,  £)  which  is  introduced  at  the  time 
t  =  r  and  maintained  thereafter;  thus  G  depends  upon  f  ,  77,  f  ;  r  and 
x,  y,  z;  t:  G  =  G(£  ,  r],  f  ;  r  \  x,  y,  z;  t).  In  fact,  G  is  the  velocity  poten- 
tial which  yields  the  solution  of  the  following  water  wave  problem: 
A  certain  disturbance  is  initiated  at  the  point  (£,  77,  £)  at  the  time 
t  —  r.  The  pressure  on  the  free  surface  of  the  water  is  assumed  to  be 
zero  always,  and  at  the  time  t  —  r  the  water  is  assumed  to  have  been 
at  rest  in  its  equilibrium  position.  Since  G  is  a  harmonic  function  in 
x,  y,  z  it  is  reasonable  to  expect  that  the  correct  singularity  to  impose 
at  the  point  (£,  rj,  £)  in  order  that  it  should  have  the  properties  one 
likes  a  Green's  function  to  have  is  that  it  behaves  there  like  I/R,  with 


R  ^  V^-r-(i?»)*  -Hfa).  Thus  G  should  satisfy  the 


188  WATER   WAVES 

following  conditions:  It  should  be  a  solution  of  the  Laplace  equation 

(6.9.1)  Gxx  +  Gyy  +  GZZ  =  0         for  -  oo  <  y  <  0,         t  ^  T, 
satisfying  the  free  surface  condition 

(6.9.2)  Gtt  +  gGy  =  0,         y  =  0. 

At  oo  we  require  G,  Gt  and  their  first  derivatives  to  be  uniformly 
bounded  at  any  given  time  t.  (Actually,  they  will  be  seen  to  tend  to 
zero  at  oo.)  At  the  point  £,  77,  f  we  require 

(6.9.3)  G  —  —  to  be  bounded. 

R 

As  initial  conditions  at  the  time  t  —  r  we  have  (cf.  sec.  6.1) 

(6.9.4)  G  =  Gt  =  0         for  t  =  r,         y  =  0. 

As  we  shall  see  later  on,  these  conditions  determine  G  uniquely. 
We  proceed  to  construct  the  function  G  explicitly.  As  a  first  step 
we  set 

(6.9.5)  G(f,  17,  £;  T  |  v,  y,  *;  t)  =  A(£,  17,  f  |  *,  y,  a)  + 

5(1, *?,  C;  T  |  a?,  /y,  z; «) 
with  ^f  defined  by 

(6.9.6)  A  =  1  -  -L  with  «'  =  V(f  - V)»Tfo  "+2/F+  G  ^^- 

/t          /v 

Thus  ^4  contains  the  prescribed  singularity,  and  we  may  require  B 
to  be  regular.  Since  A  is  a  harmonic  function,  it  follows  that  B  is 
harmonic;  in  addition,  B  satisfies  the  free  surface  condition 


(6.9.7)     Btt+gBy  = 

He  —  x)-  -t-  r)'  -t-  (t,  —  z)"]°" 

d  i 

at  «  =  0' 


— r 

9*7  [(I  -  *)2  +  *?2  +  (C  ~  *)2]1/2 

as  one  can  readily  verify.  To  determine  B  from  this  and  the  other  con- 
ditions arising  from  those  imposed  on  G  it  would  be  possible  to  employ 
the  Hankel  transform  in  exactly  the  same  way  as  the  Fourier  trans- 
form was  used  in  preceding  sections.  However,  it  seems  better  in  the 
present  case  to  proceed  directly  by  using  the  special,  but  well-known, 
Hankel  transform  for  the  function  e-b*/s  (cf.,  for  example,  Sneddon 
[S.ll],  p.  528);  this  yields  the  formula 


UNSTEADY   MOTIONS  189 

i          r°° 

(6.9.8)  - =        e~bs  J0(as)  ds, 

Va2  +  b*      Jo 

valid  for  b  >  0.  By  means  of  this  formula  the  right  hand  side  of  (6.9.7) 
can  be  written  in  a  different  form  to  yield 

(6.9.9)  Btt  +  gBy  =  2g  JL  f  V<  J0(sr)  ds  =  2g  f  *«>*•  JQ(sr)  ds 

fyJo  Jo 

at  y  =  0 

valid  for  r\  <  0  and  with 


(6.9.10)  r  =  V(S  -  x)*  +  (C  -  *)2. 

Since  B  is  a  harmonic  function  in  x9  y,  z,  it  would  seem  reasonable  to 
seek  it  among  functions  of  the  form 

(6.9.11)  B  =  I™  sT(t,  s)e(*+ti'  J0(sr)  ds, 

Jo 

which  are  harmonic  functions.  The  free  surface  condition  (6.9.9)  will 
now  be  satisfied,  as  one  can  easily  sec,  if  T(t)  satisfies  the  differential 
equation 

(6.9.12)  Ttt  +g*T  =  2g. 

The  function  T  is  now  uniquely  determined  from  (6.9.12)  and  the 
initial  conditions  T  =  Tt  =  0  for  t  =  T  derived  from  (6.9.4);  the 
result  is 

(6.9.13)  T(t,  s)  =  2  LT 


Thus  we  have  for  G  the  function 
(6.9.14) 


+  2  f  "VtH-*)  [i  _  cos 
Jo 

and  it  clearly  satisfies  all  of  the  conditions  prescribed  above,  except 
possibly  the  conditions  at  oo,  which  we  shall  presently  investigate  in 
some  detail  because  of  later  requirements.  Before  doing  so,  however, 
we  observe  the  important  fact  that  G  is  symmetrical  not  only  in  the 
space  variables  f  ,  77,  £  and  #,  y,  z9  but  also  in  the  time  variables  r  and 
/,  i.e.  that 


190  WATER  WAVES 

(6.9.15)     G(£,  r),  f  ;  r  |  a?,  y,  *;  t)  =  G(x,  y,z\t\  f,  ??,  f  ;  r)  and 


We  turn  next  to  the  discussion  of  the  behavior  of  G  at  oo.  Consider 
first  the  function  A  ==  l/R  —  1/i?'.  This  function  evidently  will 
behave  at  oo  like  a  dipole;  hence  if  a  represents  distance  from  the 
origin  it  follows  that  A  and  its  radial  derivative  Aa  behave  as  follows 
for  large  a: 

A    ~1(T2 


On  the  free  surface  where  y  —  0  we  have 
A  =  0  for  y  =  0, 


l/a3         for  y  =  0  and  large  a. 

To  determine  the  behavior  of  B—  i.e.  of  the  integral  in  (6.9.14)—  we 
expand  [1  —  cos  Vgs  (r  —  t)]  in  a  power  series  in  r  —  t  and  write 

(6.9.18)       B  =  2 


It  is  clearly  legitimate  to  integrate  term-wise  for  y  negative.  The 
formula  (6.9.8)  can  be  expressed  in  the  form 


(6.9.19) 


4=  (* 
R      Jo 


and  from  it  we  obtain 


=  r  $n 


with  //  =  cos  0,  by  a  well-known  formula  for  spherical  harmonics. 
It  follows,  since  Pn(//)  are  bounded  functions,  that  the  leading  term 
in  the  asymptotic  expansion  of  B  arises  from  the  first  term  in  the 
square  bracket.  Hence  the  behavior  of  B  is  seen  from  (6.9.20)  for  the 
case-  n  =  I  to  be  given  by 

(6.9.21)  l?~l/(r2, 

for  a  large  and  any  fixed  values  of  r  and  t.  The  derivative  By  is  seen, 
also  from  (6.9.20),  to  behave  like  I/a3  and  the  derivative  Br  also  can 
be  seen  to  behave  like  I/a3;  thus  the  radial  derivative  Ba  behaves  in 
the  same  way  and  we  have 

(6.9.22)  Ba  ~  I/a3,         Bv  ~  I/a3. 


UNSTEADY   MOTIONS  191 

Summing  up,  we  have  for  the  Green's  function  G  the  following  behavior 

at  oo: 

G  ~  I/a2 


(6.9.23) 


Ga  ~  I/a3 

Gy  ~  I/a3. 


All  of  these  conditions  hold  uniformly  for  any  fixed  finite  ranges  in 
the  values  of  r  and  t. 

We  turn  next  to  the  consideration  of  a  water  wave  problem  of  very 
general  character,  as  follows.  The  space  y  <  0  is  filled  with  water  and 
in  addition  there  are  immersed  surfaces  St  of  finite  dimensions  having 
a  prescribed  motion  (which,  of  course,  must  of  necessity  be  a  motion  of 
small  amplitude  near  to  a  rest  position  of  equilibrium).  The  pressure 
on  the  free  surface  Sf  is  prescribed  for  all  time,  and  the  initial  position 
and  velocity  of  the  particles  on  the  free  surface  and  the  immersed 
surfaces  are  given  at  the  time  t  =  0.  At  infinity  the  displacement  and 
velocity  of  all  particles  are  assumed  to  be  bounded.  The  resulting 
motion  can  be  described  for  all  times  t  >  0  in  terms  of  a  velocity 
potential  0(x9  y,  z;  t)  which  satisfies  conditions  of  the  kind  studied  in 
the  first  section  of  this  chapter;  these  conditions  are: 

(6.9.24)  V2,V(Z0  =  0 

in  the  region  R  consisting  of  the  half  space  y  <  0  exterior  to  the  im- 
mersed surfaces  S^  On  the  free  surface  the  condition 

(6.9.25)  0tt+g0v=  ~-Pt=:P(x909z;t)9         t  >  0,         y  =  0 

Q 

is  prescribed,  with  p  the  given  surface  pressure  (cf.  (6.1.1 )  and  (6.1.2)). 
At  the  equilibrium  position  of  the  immersed  surfaces  the  condition 

(6.9.26)  0n  =  V  on  Si9         t  ^  0, 

with  V  the  normal  velocity  of  Si9  is  prescribed.  The  initial  position  of 
St  at  t  =  0  is,  of  course,  assumed  known,  and  for  the  initial  conditions 
otherwise  we  know  (cf.  6.1)  that  it  suffices  to  prescribe  0  and  0t  on 
the  free  surface  at  t  —  0: 

(6.9.27)  j^O^O)^*,*) 

I  #«(*,0,*;0)  =  /,(*,*). 

At  oo  we  assume  that  0,  0t  and  their  first  derivatives  are  uniformly 
bounded. 


192 


WATER   WAVES 


We  proceed  now  to  set  up  a  representation  for  the  function  0  by 
using  the  Green's  function  obtained  above.  In  case  there  are  no  im- 
mersed surfaces  this  representation  furnishes  an  explicit  solution  of 
the  problem,  and  in  the  other  cases  it  leads  to  an  integral  equation  for 
it.  In  all  cases,  however,  a  uniqueness  theorem  can  be  obtained. 

To  carry  out  this  program  we  begin,  in  the  usual  fashion,  by  applying 
Green's  formula  to  the  Green's  function  G  and  to  0t  (rather  than  0) 
in  a  sphere  centered  at  the  origin  of  radius  a  large  enough  to  include 
the  immersed  surfaces  and  the  singular  point  (|,  77,  £)  of  the  Green's 
function  minus  a  small  sphere  of  radius  e  centered  at  the  singular  point. 
Since  G  and  0t  are  both  harmonic  functions  and  G  behaves  like  l/R 
at  the  singular  point,  it  follows  by  the  usual  arguments  in  potential 
theory  that  0t(x9  y>  z;  t)  is  obtained  in  the  form  of  a  surface  integral, 
as  follows: 


(6.9.28) 


0t(x,  y, 


=  -?-    (T 

4rc  JJ 


(G0tn  ~  &tGn)  dS. 


The  symmetry  of  G  has  been  used  at  this  point.  The  integration  varia- 
bles are  f,  77,  £.  Even  though  6?  depends  on  the  difference  t  —  r  the 
integral  in  (6.9.28)  depends  only  on  t;  that  is,  only  the  singular  part  of 
the  behavior  of  G  matters  in  applying  Green's  formula,  and  the  re- 
sulting expression  for  0t  depends  only  on  the  time  at  which  0t  and 
0tn  are  measured.  The  surface  integral  is  taken  over  the  boundary 
of  the  region  just  described  (cf.  Fig.  6.9.1),  and  n  is  the  normal  taken 


Jk  y 


sfl 


Fig.6.9.1.  Domain  for  application  of  Green's  formula 

outward  from  the  region.  The  boundary  is  composed  of  three  different 
parts:  the  portion  of  the  sphere  Sa  of  radius  a  lying  below  the  plane 
y  =  0,  the  part  Sf  of  the  plane  y  =  0  cut  out  by  the  sphere  Sa9  and 


UNSTEADY    MOTIONS 


193 


the  immersed  surfaces  St  (which  might  possibly  cut  out  portions  of  the 
plane  y  =  0). 

It  is  important  to  show  first  of  all  that  the  contribution  to  the 
surface  integral  provided  by  Sa  tends  to  zero  as  a  ->  oo,  and  that  the 
integral  over  Sf  exists  as  a  ->  oo.  The  second  part  is  readily  shown: 
The  integrand  to  be  studied  is  G0ty  —  0tGy.  From  the  symmetry 
of  G  and  (6.9.23)  we  see  that  the  above  integrand  behaves  like  I/a2 
for  large  a  since  0ty  and  0t  are  assumed  to  be  uniformly  bounded  at 
oo ;  hence  the  integral  converges  uniformly  in  t  and  r  for  any  fixed 
ranges  of  these  variables.  To  show  that  the  integral  of  G0to  —  0tGa 
over  Sa  tends  to  zero  for  a  ->  oo  requires  a  lengthier  argument.  Con- 
sider first  the  term  0tGa.  Since  Ga  behaves  like  I/a3  for  large  a  while 
0t  is  bounded,  it  is  clear  that  the  integral  of  this  term  behaves  like 
I/a  and  hence  tends  to  zero  as  a  ->  oo.  The  integral  over  the  remaining 
term  is  broken  up  into  two  parts,  as  follows: 


(6.9.29)  !(0iaGdS  =  f  "f* 

JJ  Jo   Jinl: 


(a  12) +6 

+ 


,Ga*sinOded(o 

>iaGat  sin  0  dd  do). 


f2n  f(nl2)+6 

I  * 

JO    Jnj2 


The  integrations  arc  carried  out  in  polar  coordinates,  and  d  is  a  small 
angle  (cf.  Fig.  6.9.2);  the  second  integral  represents  the  contribution 
from  a  thin  strip  of  the  sphere  Sa  adjacent  to  the  free  surface.  Since 


Fig.  6.9.2.  The  sphere  Sa 

0ta  is  bounded  and  G  behaves  like  I/a2  for  large  a,  it  is  clear  that  the 
absolute  value  of  the  second  contribution  (i.e.  that  from  the  thin  strip) 
can  be  made  less  than  e/2,  say,  if  d  is  chosen  small  enough.  Once  d  has 
been  fixed,  it  can  be  seen  that  the  contribution  of  the  remaining  part 


194  WATER  WAVES 

of  Sa  can  also  be  made  less  than  e/2  in  absolute  value  if  a  is  taken  large 
enough.  If  this  is  once  shown  it  is  then  clear  that  the  integral  in  ques- 
tion vanishes  in  the  limit  as  a  ->  oo.  The  proof  of  this  fact  is,  however, 
not  difficult:  we  need  only  observe  that  0t  is  by  assumption  bounded 
at  oo  and  it  is  a  well-known  fact*  that  0ta  then  tends  to  zero  uniformly 
like  1  1  a  along  any  ray  from  the  origin  which  makes  an  angle  ^  6  with 
the  plane  y  =  0.  Thus  the  integrand  in  the  first  term  of  (6.9.29) 
behaves  like  I/a  and  it  therefore  can  be  made  arbitrarily  small  by 
taking  a  sufficiently  large.  Thus  for  0t  we  now  have  the  representation 

(6.9.30)  0t(x,  y,  z;  t)  =  i-  [((G0tr)  -  0, 


Si 

in  which  it  is,  of  course,  understood  that  any  parts  of  the  plane  y  —  0 
cut  out  by  St  are  omitted  in  the  first  integral.  The  next  step  is  to 
integrate  both  sides  of  (6.9.30)  with  respect  to  t  from  0  to  r.  The  result 
is 

(6.9.31)        0(x9  y,  z;  r)  -  0(x9  y,  z;  0) 

•  *  If  [J 

»7=*0  S, 

=  1    ff  [  (CO*,  +  l  0tGt)  T  -  f^G,  +  l0ttGt)  dt\  d£  d'C  +  I 

to  JJ  L  g  o      Jo  S  J 

»7=0 

when  Gtt  +  gGy  =  0  for  y  =  0  is  used  (cf.  (6.9.2))  and  /  is  intended 
as  a  symbol  for  the  integral  over  S^  We  have  G  =  G  t  =  0  for  t  =  r; 
while  for  t  =  0  we  have  0t  =  /2,  and  0y  \  <=s0  uniquely  determined  by  /j** 
from  the  conditions  (6.9.27).  In  addition,  we  know  that  0y+(l/g)0tt 
=  (  l/g)P  for  *  >  0  from  (6.9.25).  It  follows  that  (6.9.31  )  can  be  written 
in  the  form 

*  One  way  to  prove  it  is  to  use  the  Poisson  integral  formula  expressing  <&t 
at  any  point  in  terms  of  its  values  on  the  surface  of  a  sphere  centered  at  the  point 
in  question.  Differentiation  of  this  formula  yields  for  any  first  derivative  of  <Pt 
an  estimate  of  the  form  M/6  where  M  depends  only  on  the  bound  for  3>t  on  the 
sphere  and  b  is  the  radius  of  the  sphere.  Finally,  since  our  domain  for  6  >  (;r/2)-h<5 
contains  spheres  of  arbitrarily  large  radius  at  points  arbitrarily  far  from  the 
origin,  the  result  we  need  follows. 

**  Since  <P(xt  t/,  z;  0)  is  harmonic,  it  is  uniquely  determined  by  its  boundary 
values  on  y'—  0  and  the  boundedness  conditions  at  oo. 


UNSTEADY   MOTIONS  195 


(6.9.32)         0(x,y,z;r)  -  4>(x,  y,  z;  0) 
1 


»;=0 


-  /2G,)  I       +  ("-  GtPdt\  di-  dl 
S          |«-o      Jog  J 


We  now  sec  that  if  there  are  no  immersed  surfaces  *S\  an  explicit 
solution  0(x9  y,  z\  r)  is  given  at  once  by  (6.9.32)  in  terms  of  the  initial 
conditions,  which  fix  0y  |f=0  and  /2,  and  the  condition  on  the  free 
surface  pressure  fixing  P  —  in  fact,  our  general  argument  shows  that 
every  solution  having  the  required  properties  is  representable  in  this 
form.  Consequently,  the  uniqueness  theorem  is  proved  for  these  cases. 
In  particular,  the  Green's  function  constructed  above  is  therefore 
uniquely  determined  since  its  regular  part,  B,  satisfies  the  conditions 
imposed  above  on  0. 

In  case  there  are  immersed  surfaces  present  the  equation  (6.9.32) 
does  not  yield  the  solution  0,  but  it  docs  yield  an  integral  equation 
for  it  in  the  following  way  (which  is  the  standard  way  of  obtaining 
an  integral  equation  for  a  harmonic  function  satisfying  various 
boundary  conditions):  One  goes  back  to  the  derivation  of  (6.9.30), 
but  considers  that  the  singularity  is  at  a  point  (x,  y,  z)  of  *S\.  If  St  is 
sufficiently  smooth  (and  we  assume  that  it  is)  the  equation  (6.9.30) 
still  holds,  except  that  the  factor  l/4jt  is  replaced  by  I/2n,  and  0  is 
then  of  course  given  only  on  St.  The  integration  on  t  from  0  to  r  is 
once  more  performed,  and  an  equation  analogous  to  (6.9.32)  is  ob- 
tained; it  can  be  written  in  the  form 


(6.9.33)     0(x,  i/,  *;  r)  -  F(x,  y,  z;  r)  -  ~  JJ  ["  JT 


dS 


with  F  a  known  function  obtained  by  adding  together  what  corres- 
ponds to  the  first  two  integrals  on  the  right  hand  side  of  (6.9.32). 
As  we  see,  this  is  an  integral  equation  for  the  determination  of 
0(x,  y,  z-9  r)  on  S{.  If  it  were  once  solved,  the  value  of  0  on  St  could 
be  used  in  (6.9.32)  to  furnish  the  values  of  0  everywhere. 

We  may  make  use  of  (6.9.32)  to  obtain  our  uniqueness  theorem  in 
the  following  fashion.  Suppose  there  were  two  solutions  0l  and  $2. 
Set  0  =  0l  —  02.  Then  0  satisfies  all  of  the  conditions  imposed  on 
01  and  02  except  that  the  nonhomogeneous  boundary  conditions  and 


196  WATER   WAVES 

initial  conditions  are  now  replaced  by  homogeneous  conditions,  i.e. 
/2  =  P  =  0;  fl  =  0  and  hence  0y  \toQ  =  0  since  0(x9  y,  z;  0)  is  a 
harmonic  function  which  vanishes  for  y  =  0,  and  0tn  =  0  since 
0n  =  0  on  St.  Thus  for  0  we  would  have  the  integral  representation: 

(6.9.34)  0(x,  y,  z;  r)  =  -  i-  ff  f  f**^ 

5, 

Since  Gn  behaves  at  oo  like  I/or3  (cf.  (6.9.23))  and  values  of0t  on  the 
bounded  surfaces  Si  are  alone  in  question,  it  follows  that  0  also  be- 
haves like  I/a3  at  oo  for  any  fixed  r  since  the  surfaces  Si  are  bounded. 
The  derivatives  of  0  could  also  be  shown  to  die  out  at  oo  at  least  as 
rapidly  as  I/a3  since  the  derivatives  of  G  could  be  shown  to  have  this 
property—  for  example,  by  proceeding  in  the  fashion  used  to  obtain 
(6,9.23). 

As  a  consequence  the  following  function  of  t  (essentially  the  energy 
integral)  exists:* 

(6.9.35)  E(t)  =  -  [IT  [01  +01+  0\]  dxdydz  +  -L  jT#f  dxdz. 


s 


, 


Differentiation  of  both  sides  with  respect  to  t  yields 
(6.9.36)     E'(t)  =         [(*«).*.  +  (**),(*)»  +  (*«),(*).]  dxdydz 


R 


-      &t&tt 


g 

s 


S, 


by  application  of  Green's  first  formula,  with  B  —  St  +  Sf  the 
boundary  of  R.  But  0n  =  0  on  S^  and  0W  =  0y  =  —  ( 3  /g)<P«  on  S/. 
It  follows  therefore  that  E'(t)  =  0  and  E  =  const.  But  0  =  0  at 
t  =  0  and  hence  JB  =  0  from  (6.9.85).  It  follows  that  0X9  0y,  0Z  are 
identically  zero,  and  0  thus  also  vanishes  identically.  Hence  0±  =  02 
and  our  uniqueness  theorem  is  proved. 


*  It  should  perhaps  be  noted  that  the  energy  integral  for  the  original  motions 
need  not,  and  in  general  will  not  exist,  since  the  velocity  potential  and  its  derivati- 
ves are  required  only  to  be  bounded  at  oo. 


SUBDIVISION  C 


Waves  on  a  Running  Stream.  Ship  Waves 


In  this  concluding  section  of  Part  II  made  up  of  Chapters  7,  8,  and 
9,  we  treat  problems  which  involve  small  disturbances  on  a  running 
stream  with  a  free  surface;  that  is,  motions  which  take  place  in  the 
neighborhood  of  a  uniform  flow,  rather  than  in  the  neighborhood  of 
the  state  of  rest,  as  has  been  the  case  in  all  of  the  preceding  chapters 
of  Part  II.  In  Chapter  7  the  classical  problems  concerning  steady 
two-dimensional  motions  in  water  of  uniform  (finite  or  infinite)  depth 
are  treated  first.  It  is  of  considerable  interest,  however,  to  consider 
also  unsteady  motions  (which  seem  to  have  been  neglected  hitherto) 
both  because  of  their  intrinsic  interest  and  because  such  a  study 
throws  some  light  on  various  aspects  of  the  problems  concerning 
steady  motions.  In  Chapter  8  the  classical  ship  wave  problem,  in 
which  the  ship  is  idealized  as  a  disturbance  concentrated  at  a  point 
on  the  surface  of  a  running  stream,  is  studied  in  considerable  detail. 
In  particular,  a  method  of  justifying  the  asymptotic  treatment  of  the 
solution  through  the  repeated  use  of  the  method  of  stationary  phase  is 
given,  and  the  description  of  the  character  of  the  waves  for  both 
straight  and  curved  courses  is  carried  out  at  length.  Finally,  in  Chapter 
1)  the  problem  of  the  motion  of  a  ship  of  given  hull  shape  is  treated 
under  very  general  conditions:  the  ship  is  assumed  to  be  a  rigid  body 
having  six  degrees  of  freedom  and  to  move  in  the  water  subject  only 
to  the  propeller  thrust,  gravity,  and  the  pressure  of  the  water,  while 
the  motion  of  the  water  is  not  restricted  in  any  way. 


197 


CHAPTER  7 


Two-dimensional  Waves  on  a  Running  Stream 
in  Water  of  Uniform  Depth 

As  indicated  in  Fig.  7.0.1  we  consider  waves  created  in  a  channel 

'  y 


y=  -h 


Fig.  7.0.1.  Waves  on  a  running  stream 

of  constant  depth  h,  when  the  stream  has  uniform  velocity  U  in  the 
positive  ^-direction  in  the  undisturbed  state.  Such  a  uniform  flow  can 
readily  be  seen  to  fulfill  the  conditions  derived  in  Chapter  1  for  a 
potential  flow  with  y  =  0  as  a  free  surface  under  constant  pressure. 
We  assume  that  the  motions  arising  from  disturbances  created  in  the 
uniform  stream  have  a  velocity  potential  0(x,  y\  t)9  and  we  set 

(7.0.1)  0(x9  y;  t)  =  Ux  +  <p(x,  y;  t)9   -  oo  <  x  <  oo,  -  h<  y  <  r). 

Since  </>(#,  y;  t)  is  a  harmonic  function  of  x  and  y  it  follows  that 
<p(x9  y\  t)  is  also  harmonic: 

(7.0.2)  W  =  <>• 


The  function  <p(x,  y;t)  is  assumed  to  yield  a  small  disturbance  on  the 
running  stream,  and  we  interpret  this  to  mean  that  <p  and  its  deriva- 
tives are  all  small  quantities  and  that  quadratic  and  higher  order 
terms  in  them  can  be  neglected  in  comparison  with  linear  terips.  We 
assume  also  that  the  vertical  displacement  y  =  rj(x;  t)  of  the  free 


198 


TWO-DIMENSIONAL   WAVES  199 

surface,  as  measured  from  the  undisturbed  position  y  —  0,  is  also  a 
small  quantity  of  the  same  order  as  <p(x,  y;  t).  Under  these  circum- 
stances the  dynamic  free  surface  condition  as  given  by  Bernoulli's 
law  (cf.  (1.4.6))  and  the  kinematic  free  surface  condition  (cf.(  1.4.5)) 
take  the  forms 


(7.0.3)      L  +  gr,  +  <pt  +  Upx  +  -  I/*  =  0 


at  y  =  0, 


1  O      I  •  I      V  •  I      M/  I  f- 

Q  2 

(7.0.4)  r,t 

when  quadratic  terms  in  <p  and  77  are  neglected  and  an  unessential 
additive  constant  is  ignored  in  (7.0.3).*  At  the  same  time,  it  is  proper 
and  consistent  in  such  an  approximation  to  satisfy  the  free  surface 
conditions  at  y  =  0  instead  of  at  the  displaced  position  y  —  r\.  (The 
reason  for  this  is  explained  in  Chapter  2 —actually  only  for  the  case 
[7  =  0,  but  the  discussion  would  be  the  same  in  the  present  case.) 
At  the  bottom  y  =  —  h  we  have  the  condition 

(7.0.5)  (pv  =  0         at  y  =  —  h. 

In  case  the  channel  has  infinite  depth  we  replace  (7.0.5)  with 

(7.0.5)'        y  and  its  derivatives  up  to  second  order  are  bounded  at 

y  =  —  oo. 

In  addition  to  the  conditions  (7.0.2)  to  (7.0.5)  it  is  necessary  also  to 
postulate  conditions  at  x  =  ±  <x>  and,  unless  the  motion  to  be  studied 
is  a  steady**  motion  with  <p  independent  of  t,  it  is  also  necessary  to 
impose  initial  conditions  at  the  time  t  =  0.  The  cases  to  be  treated  in 
the  remainder  of  this  chapter  differ  with  respect  to  these  various 
types  of  conditions,  and  we  shall  formulate  them  as  they  are  needed. 

7.1.  Steady  motions  in  water  of  infinite  depth  with  p  =  0  on  the 
free  surface 

If  the  disturbance  potential  <p  is  independent  of  t,  and  if  p  =  0  on 
the  free  surface  it  follows  that  99(0?,  y)  satisfies  the  conditions 

*  It  is  perhaps  worth  noting  explicitly  that  it  would  be  inappropriate  to 
assume  that  U,  the  velocity  of  the  stream,  is  a  small  quantity  of  the  same  order 
as  r]  and  (p:  to  do  so  would  lead  to  the  elimination  of  the  terms  in  U  and  the 
resulting  theory  would  not  differ  from  that  of  the  preceding  chapters. 

**  In  this  chapter  the  term  "steady  motion"  is  used  in  the  customary  way  to 
describe  a  flow  which  is  the  same  at  each  point  in  space  for  all  times.  In  the 
preceding  chapters  we  have  sometimes  used  this  term  (in  conformity  with  esta- 
blished custom  in  the  literature  dealing  with  wave  propagation)  in  a  different 
sense. 


200  WATER   WAVES 


(7.1.1)  VV  =  0,         _  oo  <y  ^0, 

J72 

(7.1.2)  <pv  +  —  <pxx  =  o,     y  =  °- 

s 

In  addition,  we  require  that 

(7.1.3)         9?  and  its  derivatives  up  to  second  order  are  bounded  at  oo, 

though  this  condition  is  more  restrictive  than  is  necessary.  The  second 
of  these  conditions  was  obtained  from  (7.0.3)  and  (7.0.4)  by  differen- 
tiating (7.0.3)  and  eliminating  77. 

It  is  interesting  to  find  all  functions  q>(x,  y)  satisfying  these  condi- 
tions, and  it  is  easy  to  do  so  following  the  same  arguments  as  were 
used  in  Chapter  3.1.  Using  (7.1.1)  we  may  re-write  (7.1.2)  in  the  form 

U* 

(7.1.4)  v*-—v,i,  =  °>      y  =  o- 

o 

(This  of  course  makes  use  of  the  fact  that  <p  is  harmonic  for  y  =  0, 
which  we  assume  to  be  true.  One  could  easily  show,  in  fact,  that  the 
free  surface  condition  (7.1,2)  permits  an  analytic  continuation  of  9? 
over  y  =  0,  so  that  9?  is  actually  harmonic  in  a  domain  including 
y  =  0  in  its  interior.)  We  observe  that  (7.1.4)  is  the  same  condition 
on  <pv  as  was  imposed  on  the  function  called  <p  in  Chapter  3,  and  we 
proceed  as  we  did  there  by  introducing  a  harmonic  function  y(x9  y) 
through 

U2 

(7.1.5)  V  =<Pv  -  —<f>yy      ,       2/^0- 

e 

This  function  vanishes  on  y  —  0,  and  can  therefore  be  continued 
analytically  by  reflection  into  the  upper  half  plane.  Since  (p  and  its 
derivatives  were  assumed  to  be  bounded  in  the  lower  half  plane,  it 
follows  that  \p  is  bounded  in  the  entire  plane  and  hence  by  Liouville's 
theorem  it  is  a  constant;  hence  \p  vanishes  identically  since  \p  =  0 
for  y  =  0.  Thus  we  have  for  y>y  a  differential  equation  given  by  (7.1.5) 
with  \f  =  0,  and  it  has  as  its  only  solutions  the  functions 


(7.1.6)  9>v  =  c( 

Since  <py  is  also  a  harmonic  function,  it  follows  that  c(x)  is  a  solution 
of  the  differential  equation 

/  a  \  2 

c  =  0. 


TWO-DIMENSIONAL   WAVES  201 

Hence  <p  is  given  by 

(7.1.8)  <p(x,  y)  =  Aeu*vcos  /-L  *  +  aj  +  c^x) 

with  A  and  a  constants  and  c^x)  an  arbitrary  function  of  x.  By  making 

d2c 

use  of  (7.1.2),  however,  one  finds  that  —  -  =  0,  and  hence  that 

dx2 

c±  ==  const,  since  cp  is  bounded  at  oo.  There  is  no  loss  of  generality  in 
taking  cl  =  0.  The  only  solutions  of  our  problem  are  therefore  given 
by 

(7.1.9)  <p(x,  y)  =  Aeu*vcos  UL  +  A 


Thus  the  only  steady  motions  satisfying  our  conditions,  aside  from 
a  uniform  flow,  are  periodic  in  x  with  the  fixed  wave  length  A  given  by 

772 

(7.1.10)  A  =  2n  —  . 

g 

The  amplitude  and  phase  of  the  motions  are  arbitrary.  If  we  were  to 
observe  these  waves  from  a  system  of  coordinates  moving  in  the  x- 
direction  with  the  constant  velocity  E7,  we  would  see  a  train  of  pro- 
gressing waves  given  by 

<p  =  Aemv  cos  m  (x  -\~  Ut) 
with 

g        2n 
m  =  —  =  —  . 
C/2        A 

These  waves  are  identical  with  those  already  studied  in  Chapter  3 
(cf.  sec.  3.2).  The  phase  speed  of  these  waves  would  of  course  be  the 
velocity  U  and  the  wave  length  A  would,  as  it  should,  satisfy  the  rela- 
tion (3.2.8)  for  waves  having  this  propagation  speed.  In  other  words, 
the  only  waves  we  find  are  identical  (when  observed  from  a  coordinate 
system  moving  with  velocity  U  )  with  the  progressing  waves  that  are 
simple  harmonic  in  the  time  and  which  have  such  a  wave  length  that 
they  would  travel  at  velocity  U  in  still  water. 

7.2.  Steady  motions  in  water  of  infinite  depth  with  a  disturbing 
pressure  on  the  free  surface 

The  same  hypotheses  are  made  as  in  the  previous  section,  except 
that  we  assume  the  pressure  on  the  free  surface  to  be  a  function 


202 


WATER   WAVES 


over  the  segment  —  a  ^  x  ^  a  and  zero  otherwise,  as  indicat- 
ed in  Fig.  7.2.1.  The  free  surface  condition,  as  obtained  from  (7.0.8) 


p«0 


-o 


-i-a 


U 
Fig.  7.2.1.  Pressure  disturbance  on  a  running  stream 

and  (7.0.4)  by  eliminating  rj  and  assuming  r\  and  9?  to  be  independent 
of  t,  is  now  given  by 


(7.2.1) 


4.  JL  m    =  —  J^l ,         on  z/  =  0, 
^  U^v  UQ 


as  one  readily  verifies.  We  prescribe  in  addition  that  <p  and  its  first 
two  derivatives  are  bounded  at  oo. 

The  solutions  <p  of  our  problems  are  conveniently  derived  by  intro- 
ducing the  analytic  function  f(z)  of  the  complex  variable  z  =  x  +  iy 
whose  real  part  is  9?: 

(7.2.2)  /(a)  =  <p(x,  y)  +  iy(x9  y). 

Since  yy  =  — -  yx,  the  condition  (7.2.1)  can  be  put  in  the  form 


(7.2.3) 


0  p 

—  ~i  V  =  —  ~r-  +  const.,        on  y  =  0, 


and  the  constant  can  be  taken  as  zero  without  loss  of  generality,  since 
adding  a  constant  to  p  can  not  affect  the  motion. 

We  consider  now  only  the  case  in  which  the  surface  pressure  p  is  a 
constant  p  =  p0  over  the  segment  \x\  ^  a,  and  zero  otherwise.  Since 
this  surface  pressure  is  discontinuous  at  x  =  ±  a,  it  is  necessary  to 
admit  a  singularity  at  these  points;  we  shall  see  that  a  unique  solution 
of  our  problem  is  obtained  if  we  require  that  q>  is  bounded  at  these 
points  while  <px  and  <py  behave  like  1/r1"*,  e  >  0,  with  r  the  distance 


TWO-DIMENSIONAL   WAVES  203 

from  the  points  x  =  ±  a  on  the  free  surface.  (This  singularity  is 
weaker  than  the  logarithmic  singularity  of  <p  appropriate  at  a  source 
or  sink.) 

In  terms  of  f(z),  the  free  surface  condition  (7.2.3)  clearly  can  be  put 
in  the  form 


(7.2.4) 


(ifz  -  A)  /  = 

^  ' 


for  Jm  a  =  0. 


,  \x\  >  a 


The  device  of  applying  the  boundary  condition  in  this  form  seems  to 
have  been  used  first  by  Keldysh  [K.21].  We  now  introduce  the  ana- 
lytic function  F(z)  defined  in  the  lower  half  plane  by  the  equation 

(7.2.5)  F(*)  =  »y.-  A/. 

This  function  has  the  following  properties:  1)  Its  imaginary  part  is 
prescribed  on  the  real  axis.  2)  The  first  derivatives  of  its  imaginary 
part  are  bounded  at  oo,  since  the  first  two  derivatives  of  99  are  assumed 
to  have  this  property  and  hence  fzz  and  fz  are  bounded  in  view  of  the 
Cauchy-Riemann  equations.  3)  Near  z  =  ±o  its  imaginary  part  be- 
haves like  l/\z  ^r  a\l~e,  £  >  0,  as  one  readily  sees.  It  is  now  easy  to 
show  that  F(z)  is  uniquely  determined,*  within  an  additive  real  con- 
stant, as  follows:  Let  G  —  JPX  —  F2  be  the  difference  of  two  functions 
satisfying  these  three  conditions.  Jm  G  then  vanishes  on  the  entire 
real  axis,  except  possibly  at  the  points  (±  #>  0),  and  G  can  therefore 
be  continued  as  a  single-valued  function  into  the  whole  plane  except 
at  the  points  (±  a,  0).  However,  the  singularity  prescribed  at  the 
points  (±  a,  0)  is  weaker  than  that  of  a  pole  of  first  order,  and  hence 
the  singularities  at  these  points  are  removable.  Since  the  first  deriva- 
tives of  Jm  G  are  bounded  at  oo,  it  follows  from  the  Cauchy-Riemann 
equations  that  Gz  is  bounded  at  oo.  Hence  Gz  is  constant,  by  Liouville's 
theorem,  and  G  is  the  linear  function:  G  =  cz  +  d.  Since  Jm  G  =  0 
on  the  real  axis,  it  follows  that  c  and  d  are  real  constants.  However, 
a  term  of  the  form  cz  +  d  on  the  left  hand  side  of  (7.2.5)  leads  to  a 

a 
term  of  the  form  OLZ  +  0,  with  —  —  a  =  c,  in  the  solution  of  this 

*  In  Chapter  4,  the  function  F(z)  given  by  (4.3.10)  had  a  real  part  which 
satisfied  identical  conditions  except  that  the  condition  2)  is  slightly  more  restric- 
tive in  the  present  case. 


204  WATER   WAVES 

equation  for  /(*),  and  since  f(z)  is  assumed  to  be  bounded  at  oo.  it 
follows  that  c  =  0. 

We  have  here  the  identical  situation  that  has  been  dealt  with  in 
sec.  3  of  Chapter  4,  except  that  it  was  the  real  part  of  the  function 
F(z),  rather  than  the  imaginary  part,  that  was  prescribed  on  the  real 
axis,  and  we  can  take  over  for  our  present  purposes  a  number  of  the 
results  obtained  there.  The  function  F(z),  now  known  to  be  uniquely 
determined  within  an  additive  real  constant,  is  given  by 

(7.2.6)  *(, 


, 

UQTt  Z  +  a 

which  differs  from  F(z)  as  given  by  (4.3.12)  essentially  only  in  the 
factor  i—  as  it  should.  In  any  case,  one  can  readily  verify  that  F(z) 
satisfies  the  conditions  imposed  above.  We  take  that  branch  of  the 
logarithm  that  is  real  for  z  real  and  \z\  >  a,  and  specify  a  branch  cut 
starting  at  z  =  —  a  and  going  to  oo  along  the  positive  real  axis.  The 
equation  (7.2.5)  is  now  an  ordinary  differential  equation  for  the  func- 
tion f(z)  which  we  are  seeking. 

The  differential  equation  (7.2.5)  has,  of  course,  many  solutions, 
and  this  means  that  the  free  surface  condition  and  the  boundcdness 
conditions  at  oo  and  at  the  points  (±  a,  0)  are  not  sufficient  to  ensure 
that  a  unique  solution  exists.  In  fact,  it  is  clear  that  the  non-vanishing 
solution  of  the  homogeneous  problem  found  in  the  preceding  section 
could  always  be  added  to  the  solution  of  the  problem  formulated  up 
to  now.  A  condition  at  oo  is  needed  similar  to  the  radiation  condition 
imposed  in  the  analogous  circumstances  in  Chapter  4.  In  the  present 
case,  the  solution  can  be  made  unique  by  requiring  that  the  dis- 
turbance created  by  the  pressure  over  the  segment  |  x  \  fS  a  should 
die  out  on  the  upstream  side  of  the  channel,  i.e.  at  x  —  —  oo.  The  only 
justification  for  such  an  assumption—  aside  from  the  fact  that  it 
makes  the  solution  unique—  is  based  on  the  observation  that  one  never 
sees  anything  else  in  nature.*  In  sec.  7.4  we  shall  give  a  more  satis- 
factory discussion  of  this  point  which  is  based  on  studying  the  un- 
steady flow  that  arises  when  the  motion  is  created  by  a  disturbance 
initiated  at  the  time  t  =  0,  and  the  steady  state  is  obtained  in  the 
limit  as  t  ->  oo.  In  this  formulation,  the  condition  that  the  motion 

*  Lamb  [L.3],  p.  399,  makes  use,  once  more,  of  the  device  of  introducing 
dissipative  forces  of  a  very  artificial  character  which  then  lead  to  a  steady  state 
problem  with  a  unique  solution  when  only  boundedness  conditions  are  prescribed 
at  oo. 


TWO-DIMENSIONAL   WAVES  205 

should  die  out  on  the  upstream  side  is  not  imposed;  instead,  it  turns 
out  to  be  satisfied  automatically. 

A  solution  of  the  differential  equation  (7.2.5)  (in  dimensionless 
form)  has  been  obtained  in  Chapter  4  (cf.  (4.3.13))  which  has  exactly 
the  properties  desired  in  the  present  case;  it  is: 

(7.2.7)  /(*)  =  -2±-e-&['     &  log  t-^  dt, 

t  +  a 


The  path  of  integration  (cf.  Fig.  4.3.1)  comes  from  iao  along  the  posi- 
tive imaginary  axis  and  encircles  the  origin  in  such  a  way  as  to  leave 
it  and  the  point  (—  a,  0)  to  the  left.  That  (7.2.7)  yields  a  solution  of 
(7.2.5)  is  easily  checked.  One  can  also  verify  easily  that  q>  =  &ef(z) 
satisfied  all  of  the  boundary  and  regularity  conditions,  except  perhaps 
the  condition  at  oo  on  the  upstream  side.  In  Chapter  4,  however,  it 
was  found  (cf.  (4.3.15))  that  f(z)  behaves  at  oo  as  follows: 


O  /—  \ 


for  Ste  z  <  0, 

, 


Thus  /(s)  dies  out  as  x  ->  —  oo,  but  there  are  in  general  waves  of 
nonzero  amplitude  far  downstream,  i.e.  at  x  =  +  oo.  The  uniquely 
determined  harmonic  function  9?  =  9te  f(z)  is  now  seen  to  satisfy  all 
conditions  that  were  imposed. 

The  waves  at  x  =  +  oo  are  identical  (within  a  term  of  order  I/a) 
with  the  steady  waves  that  we  have  found  in  the  preceding  section  to 
be  possible  when  the  stream  is  subject  to  no  disturbance  (cf.  (7.1.9)), 
and  the  wave  far  downstream  has  the  wave  length  A  =  2nU2/g. 
However,  we  observe  the  curious  and  interesting  fact  (pointed  out  by 
Lamb  [L.3],  p.  404)  that  this  wave  may  also  vanish:  clearly  if 
ga/U2  =  nn,  n  —  1,  2,  .  .  .,  99  =  Ste  f(z)  vanishes  downstream  as  well 
as  upstream,  and  this  occurs  whenever  2a/A  is  an  integer,  i.e.  whenever 
the  length  of  the  segment  over  which  the  disturbing  pressure  is  applied 
is  an  integral  multiple  of  the  wave  length  of  a  steady  wave  in  water  of 
velocity  U  (with  no  disturbance  anywhere).  This  in  turn  gives  rise  to 
the  observation  that  there  exist  rigid  bodies  of  such  a  shape  that  they 
create  only  a  local  disturbance  when  immersed  in  a  running  stream: 
one  need  only  calculate  the  shape  of  the  free  surface—  which  is,  of 
course,  a  streamline—  for  ga/U2  =  nn,  take  a  rigid  body  having  the 


206  WATER  WAVES 

shape  of  a  segment  of  this  surface  and  put  it  into  the  water.  (Involved 
here  is,  as  one  sees,  a  uniqueness  theorem  for  problems  in  which  the 
shape  of  the  upper  surface  of  the  liquid,  rather  than  the  pressure,  is 
prescribed  over  a  segment,  but  such  a  theorem  could  be  proved  along  the 
lines  of  the  uniqueness  proof  of  the  analogous  theorem  for  simple  har- 
monic waves  given  by  F.  John  [J.5].  )  This  fact  has  an  interesting  physi- 
cal consequence,  i.e.,  that  such  bodies  are  not  subject  to  any  wave  re- 
sistance (by  which  we  mean  that  the  resultant  of  the  pressure  forces 
on  the  body  has  no  horizontal  component)  while  in  general  a  resistance 
would  be  felt.  This  can  be  seen  as  follows:  Observe  the  motion  from 
a  coordinate  system  moving  with  velocity  U  in  the  ^-direction.  All 
forces  remain  the  same  relative  to  this  system,  but  the  wave  at  +  oo 
would  now  be  a  progressing  wave  simple  harmonic  in  the  time  and 
having  the  propagation  speed  —  U9  while  at  —  oo  the  wave  ampli- 
tude is  zero.  Thus  if  we  consider  two  vertical  planes  extending  from 
the  free  surface  down  into  the  water,  one  far  upstream,  the  other  far 
downstream  we  know  from  the  discussion  in  Chapter  3.3  that  there  is 
a  net  flow  of  energy  into  the  water  through  these  planes  since  energy 
streams  in  at  the  right,  but  no  energy  streams  out  at  the  left  since 
the  wave  amplitude  at  the  left  is  zero.  Consequently,  work  must  be 
done  on  the  water  by  the  disturbance  pressure  and  this  work  is  done 
at  the  rate  RU  =  F,  where  R  represents  the  horizontal  resistance 
and  F  the  net  energy  flux  into  the  water  through  two  planes  contain- 
ing the  disturbing  body  between  them.  Thus  if  .F  =  0—  which  is  the 
case  if  the  wave  amplitude  dies  out  downstream  as  well  as  upstream— 
then  R  =  0.  This  result  might  have  practical  applications.  For  exam- 
ple, pontoon  bridges  lead  to  motions  which  are  approximately  two- 
dimensional,  and  hence  it  might  pay  to  shape  the  bottoms  of  the 
pontoons  in  such  a  way  as  to  decrease  the  wave  resistance  and  hence 
the  required  strength  of  the  moorings.  However,  such  a  design  would 
yield  an  optimum  result,  as  we  have  seen,  only  at  a  definite  velocity  of 
the  stream;  in  addition,  the  wave  resistance  is  probably  small  com- 
pared with  the  resistance  due  to  friction,  etc.,  except  in  a  stream 
flowing  with  high  velocity. 

We  conclude  this  section  by  giving  the  solution  of  the  problem  of 
determining  the  waves  created  in  a  stream  when  the  disturbance  is 
concentrated  at  a  point,  i.e.  in  the  case  in  which  the  length  2a  of  the 
segment  over  which  the  pressure  p0  is  applied  tends  to  zero  but 
lim  2p^a  =  P0.  The  desired  solution  is  obtained  at  once  from  (7.2.7); 


it  is: 


(7.2.9)  /(«)  =  -      e 


TWO-DIMENSIONAL  WAVES  207 

ig. 


f*      I     igt 
\       L  &*  dt. 

J  ioo  t 


This  solution  behaves  like  I/*  far  upstream  and  like  (—  2P0/C7g) 
exp  {—  igz/U2}  far  downstream.  Note  that  the  amplitude  downstream 
does  not  vanish  for  any  special  values  of  U  in  this  case.  It  is  perhaps 
also  of  interest  to  observe  that  f(z)  behaves  near  the  origin  like  i  log  z, 
and  hence  the  singularity  at  the  point  of  disturbance  has  the  character 
of  a  vortex  point;  we  recall  that  the  singularity  in  the  analogous  case 
of  the  waves  created  by  an  oscillatory  point  source  that  were  studied  in 
Chapter  4  had  the  character  of  a  source  point,  since  f(z)  behaved  like 
log  z  rather  than  like  ilogz  (cf.  4.8.28)),  with  a  strength  factor 
oscillatory  in  the  time.  When  one  thinks  of  the  physical  circumstances 
in  these  two  different  cases  one  sees  that  the  present  result  fits  the 
physical  intuition. 

7.3.  Steady  waves  in  water  of  constant  finite  depth 

In  water  of  constant  finite  depth  the  circumstances  are  more  com- 
plicated, and  in  several  respects  more  interesting,  than  in  water  of 
infinite  depth.  This  is  already  indicated  in  the  simplest  case,  in  which 
the  free  surface  pressure  is  assumed  to  be  everywhere  zero  and  the 
motion  is  assumed  to  be  steady.  In  this  case  we  seek  a  function  <p(x,  y) 
satisfying  the  conditions  (7.0.2)  to  (7.0.5),  with  <pt  and  rjt  both  iden- 
tically zero.  The  boundary  conditions  are  thus 

U2 

(7.3.1)  <py  H  --  (pxx  =  0,         y  =  0, 

g 
and 

(7.3.2)  <p,  =  0,         y  ==  -  h. 

A  harmonic  function  which  satisfies  these  conditions  is  given  by: 

(7.8.3)  (p(x9  y)  =  A  cosh  m(y  +  h)  cos  (mx  +  a) 

with  A  and  a  arbitrary  constants,  and  m  a  root  of  the  equation 

(7.8.4)  g^tanhmfe 

gh  mh 

The  condition  (7.8.4)  ensures  that  the  boundary  condition  on  the 
free  surface  is  satisfied,  as  one  can  easily  verify.  It  is  very  important 
for  the  discussion  in  this  and  the  following  section  to  study  the  roots 


208 


WATER   WAVES 


of  the  equation  (7.8.4).  The  curves  £  =  tanh  f  and  f  =  (U*/gh)  f  are 
plotted  in  Fig.  (7.3.1).  The  roots  of  (7.3.4)  are  of  course  furnished  by 
the  intersections  f  =  mh  of  these  curves.  One  observes:  1)  m  =  0  is 
always  a  root;  2)  there  are  two  real  roots  different  from  zero  if  U2/gh<l ; 


Fig.  7.8.1.  Roots  of  the  transcendental  equation  (U2/gh  <  1) 

8)  there  are  no  real  roots  other  than  zero  if  U2/gh  ^  1;  4)  if  U2/gh  =  1 
the  function  U2m  —  g  tanh  mh  vanishes  at  m  =  0  like  m3;  5)  since 
tan  if  =  i  tanh  £,  it  follows  that  (7.8.4)  has  infinitely  many  pure 
imaginary  roots  no  matter  what  value  is  assigned  to  U2/gh. 

On  the  basis  of  this  discussion  of  the  roots  of  (7.3.4)  we  therefore 
expect  that  no  motions  other  than  the  steady  flow  with  no  surface 
disturbance  (for  which  <p  =  const. )  will  exist  unless  U2/gh  <  1 .  These 
waves  are  then  seen  to  have  the  wave  length  appropriate  for  simple 
harmonic  waves  of  propagation  speed  c  =  U  in  water  of  depth  h,  as 
can  be  seen  from  (8.2.1),  (3.2.2),  and  (3.2.8).  It  is  possible  to  give  a 
rigorous  proof  of  this  uniqueness  theorem— which  holds  when  no  con- 
ditions at  oo  other  than  boundedness  conditions  are  imposed— by 
making  use  of  an  appropriate  Green's  function,  or  by  making  use  of 
the  method  devised  by  Weinstein  [W.7]  for  simple  harmonic  waves  in 
water  of  finite*  depth,  but  we  will  not  do  so  here. 

More  interesting  problems  arise  when  we  suppose  that  steady  waves 
$re  created  by  disturbances  on  the  free  surface,  or  perhaps  also  on  the 
bottom.  Mathematically  this  means  that  a  nonhomogeneous  boundary 
condition  would  replace  one,  or  perhaps  both,  of  the  homogeneous 


TWO-DIMENSIONAL  WAVES  209 

boundary  conditions  (7.3.1)  and  (7.3.2).  In  addition,  as  we  infer  from 
the  discussion  of  the  preceding  section,  it  is  also  necessary  in  general 
to  prescribe  a  condition  of  "radiation"  type  at  oo  in  addition  to  boun- 
dedness  conditions,  and  an  appropriate  such  condition  is  that  the 
disturbance  should  die  out  upstream.  In  the  present  problem,  how- 
ever, the  additional  parameter  furnished  by  the  depth  of  the  water 
leads  to  some  peculiarities  that  are  conditioned  in  part  by  the  differ- 
ence in  behavior  of  the  solutions  of  the  homogeneous  problem  in  their 
dependence  on  the  parameter  U*/gh:  Since  the  only  solution  of  the 
homogeneous  problem  in  the  case  U2/gh  S>  1  is  q>  =  0,  one  expects 
that  the  solution  of  the  nonhomogeneous  problem  will  be  uniquely 
determined  in  this  case  without  the  necessity  of  prescribing  a  radiation 
condition  at  oo.  However,  if  U*/gh  <  1  it  is  clear  that  the  nonhomo- 
geneous problem  can  not  have  a  unique  solution  unless  a  condition  — 
such  as  that  requiring  the  disturbance  to  die  out  upstream— is 
imposed  that  will  rule  out  the  otherwise  possible  addition  of  the  non- 
vanishing  solution  of  the  homogeneous  problem.  These  cases  have  been 
worked  out  (cf.  Lamb  [L.3],  p.  407)  with  the  expected  results,  as 
outlined  above,  for  U*/gh  >  1  and  U2/gh  <  1,  but  the  known  re- 
presentations of  these  solutions  for  the  steady  state  make  the  wave 
amplitudes  large  for  U2jgh  =  1  and  \x\  large. 

We  shall  not  solve  these  steady  state  problems  directly  here  be- 
cause the  peculiarities— not  to  say  obscurities— indicated  above  can 
all  be  clarified  and  understood  by  re-casting  the  formulation  of  the 
problem  in  a  way  that  has  already  been  employed  in  the  previous 
chapter  (cf.  sec.  6.7)).  The  basic  idea  (cf.  Stoker  [S.22])  is  to  abandon 
the  formulation  of  the  problem  in  terms  of  a  steady  motion  in  favor  of 
a  formulation  involving  appropriate  initial  conditions  at  the  time 
t  =  0,  and  afterwards  to  make  a  passage  to  the  limit  in  the  solutions 
for  the  unsteady  motion  by  allowing  the  time  to  tend  to  oo.  As  was 
indicated  in  sec.  6.7,  the  advantage  of  such  a  procedure  is  that  the 
initial  value  problem,  being  the  natural  dynamical  problem  in  New- 
tonian mechanics  (while  the  steady  state  is  an  artificial  problem),  has 
a  unique  solution  when  no  conditions  other  than  boundedness  con- 
ditions are  imposed  at  oo.  If  a  steady  state  exists  at  all,  it  should  then 
result  upon  letting  t  ->  oo,  and  the  limit  state  would  then  automati- 
cally have  those  properties  at  oo  which  satisfy  what  one  calls  radia- 
tion conditions,  and  which  one  has  to  guess  at  if  the  steady  state 
problem  is  taken  as  the  starting  point  of  the  investigation. 

We  shall  proceed  along  these  lines  in  the  next  section  in  attacking 


210  WATER   WAVES 

the  problem  of  the  waves  created  in  a  stream  of  uniform  depth  when 
a  disturbance  is  created  in  the  undisturbed  uniform  stream  at  the 
time  t  =  0.  The  subsequent  unsteady  motion  will  be  determined  when 
only  boundedness  conditions  are  imposed  at  oo.  It  will  then  be  seen 
that  the  behavior  of  the  solutions  as  t  -*  oo  is  indeed  as  indicated 
above,  i.e.  the  waves  die  out  at  infinity  both  upstream  and  down- 
stream when  U2/gh  >  1,  that  they  die  out  upstream  but  not  down- 
stream when  U2/gh  <  1.  One  might  be  inclined  to  say:  "Well,  what  of 
it,  since  one  guessed  the  correct  condition  on  the  upstream  side  any- 
way?" However,  we  now  get  a  further  insight,  which  we  did  not 
possess  before,  i.e.  that  for  U2/gh  =  1  there  just  simply  is  no  steady 
state  when  t  ->  oo  although  a  uniquely  determined  unsteady  motion 
exists  for  every  given  value  of  the  time  t .  In  fact  it  will  be  shown  that 
the  disturbance  potential  becomes  infinite  like  J2/3  at  all  points  of  the 
fluid  when  t  ->  oo  and  U2/gh  =  1,  and  that  the  velocity  also  becomes 
infinite  everywhere  when  t  ->  oo. 

7.4.  Unsteady  waves  created  by  a  disturbance  on  the  surface  of  a 
running  stream 

The  boundary  conditions  on  the  disturbance  potential  <p(x,  y\  t)  at 
the  free  surface  (cf.  Fig.  7.0.1  and  equations  (7.0.3)  and  (7.0.4))  are 

v  U2 

(7.4.1)  11  +  gr,  +  <pt  +  U<px  +  —  -  0, 

Q  * 

(7.4.2)  ty  +  Ur,x  -<py  =  0, 

to  be  satisfied  at  y  =  0  for  all  times  t  >  0.  The  quantity  p  =  p(x;  t) 
is  the  pressure  prescribed  on  the  free  surface.  At  the  bottom  y  =  —  h 
we  have,  of  course,  the  condition 

(7.4.3)  <py  =  0,         t  ^  0. 

At  the  initial  instant  t  =  0  we  suppose  the  flow  to  be  the  undisturbed 
uniform  flow,  and  hence  we  prescribe  the  initial  conditions: 

(7.4.4)  q>(x,  y;  0)  =  ^(x;  0)  =  p(x;  0)  =  0. 

From  (7.4.1),  which  we  assume  to  hold  at  t  =  0,  we  thus  have  the 
condition 

(7.4.5)  (f>t(x,  y;  0)  =  0. 

Finally,  we  prescribe  the  surface  pressure  p  for  t  >  0: 

(7.4.6)  p  =  p(x)9         t  >  0. 


TWO-DIMENSIONAL   WAVES  211 

(The  surface  pressure  is  thus  constant  in  time.)  At  oo  we  make  no 
assumptions  other  than  boundedness  assumptions.  We  shall  not 
formulate  these  boundedness  conditions  explicitly:  instead,  they  are 
used  implicitly  in  what  follows  because  of  the  fact  that  Fourier  trans- 
forms in  x  for  —  oo  <  x  <  oo  are  applied  to  q>  and  p  and  their 
derivatives.  Of  course,  this  means  that  these  quantities  must  not  only 
be  bounded  but  also  must  tend  to  zero  at  oo,  and  this  seems  reasonable 
since  the  initial  conditions  leave  the  water  undisturbed  at  oo. 

We  have,  therefore,  the  problem  of  finding  the  surface  elevation 
r)(x;  t)  and  the  velocity  potential  <p(x,  y;  t)  in  the  strip  —  h  ^  y  5^  0, 
—  oo  <  x  <  oo,  which  satisfy  the  conditions  (7.4.1)  to  (7.4.6).  We 
begin  the  solution  of  our  problem  by  eliminating  the  surface  elevation 
77  from  the  first  two  boundary  conditions  to  obtain: 

(7.4.7)  <ptt  +  t/Vxx  +  W<pxt  +g<py=--  px,         at  y  =  0. 

Q 

The  Fourier  transform  with  respect  to  x  is  now  applied  to  (pxx  +<pvv  =  0 
to  yield  (cf.  sec.  6.3): 

(7.4.8)  ¥„  -  s*y  =  0, 

where  the  bar  over  <tp  refers  to  the  transform  ip  =  (p(s,  y;  t)  of  <p.  From 
(7.4.3)  we  have  q>y  =  0  for  y  =  —  h;  hence  ip,  in  view  of  (7.4.8)  must 
be  of  the  form 

(7.4.9)  v(*>  y>  0  =  A(*>  ')  cosh  s(y  +  A)> 

with  A(s;  t)  a  function  to  be  determined.  The  transform  is  next  applied 
to  (7.4.7)  with  the  result: 


_  _          _  _  _ 

(7.4.10)     Vii  +  2isU<pt  +  g<pv  -  UWcp  =  --  p,         at  y  =  0, 

e 

and  this  yields,  from  (7.4.9)  for  y  ==  0,  the  differential  equation 

isUp 


(7.4.11)    AU  +  2isUAt  +  fe*  tanh  sh  -  s2U2]A  =  — 


Q  cosh  sh 


Here  p(s)  is  of  course  the  transform  of  p(x).  As  initial  conditions  at 
t  =  0  for  A($;  t)  we  have  from  (7.4.4)  and  (7.4.5)  the  conditions  (again 
in  conjunction  with  (7.4.9)): 

(7.4.12)  A(a\  0)  =  At(si  0)  =  0. 

The  function  A(s;t)  is  then  easily  found;  it  is 


212  WATER  WAVES 


(7.4.13)    A(s;  t)  =      UVp 


Q  cosh  sh 


s2U2  —  gs  tanh  sh 

j  e-it  (sU  +  Vgs  tanh  sh) 


2  Vgs  tanh  sh  sU  +  Vgs  tanh  sh 

1  e-  it  (sU  -  V 


2  A/gs  tanh  5/i  st7  —  Vgs  tanh 


The  solution  y(x>  y;  t)  of  our  problem  is  of  course  now  obtained  by 
inverting  <p(s,  y;  t): 

I      f00 
(7.4.14)    w(x,  y;  t)  ==  —  —          A(s;  t)  cosh  s(y  +  h)  elsx  ds. 

V%n  J  -oo 


The  path  of  integration  is  the  real  axis.  One  finds  easily  that  the 
integrand  behaves  for  large  s  like  e^v/s9  since  the  denominators  of 
the  terms  in  the  square  brackets  in  (7.4.13)  behave  like  s2,  the  ratio 
cosh  s(y  +  A)/cosh  sh  behaves  like  0W  v  for  large  s,  and  p(s)  tends  to 
zero  at  oo  in  general.  Since  y  is  negative  (cf.  Fig.  7.0.1)  it  is  clear  that 
the  integral  converges  uniformly.  (We  omit  a  discussion  of  the  be- 
havior on  the  free  surface  corresponding  to  y  =  0,  although  such  a 
discussion  would  not  present  any  real  difficulties.)  Upon  examining 
the  function  A(s;  t)  in  (7.4.13)  it  might  seem  that  it  has  singularities 
at  zeros  of  the  denominators  (and  such  zeros  can  occur,  as  we  shall  see) 
but  in  reality  one  can  easily  verify  that  the  function  has  no  singulari- 
ties when  the  three  terms  in  the  square  brackets  are  taken  together— 
or,  as  one  might  also  put  it,  any  singularities  in  the  individual  terms 
cancel  each  other.  Thus  the  solution  given  by  (7.4.14)  is  a  regular 
harmonic  function  in  the  strip  —  h  ^  y  <  0  for  all  time  t,  or,  in 
other  words,  a  motion  exists  no  matter  what  values  are  given  to  the 
parameters.  In  addition,  the  fact  that  the  integral  exists  ensures  that 
9?  (and  also  its  derivatives)  tends  to  zero  for  any  given  time  when 
\x\  ->  oo—  this  is  the  content  of  the  so-called  Riemann-Lebesgue 
theorem.  This  means  that  the  amplitude  of  the  disturbance  dies  out 
at  infinity  at  any  given  time  /—a  not  unexpected  result  since  a  certain 
time  must  elapse  before  any  appreciable  effects  of  a  disturbance  are 
felt  at  a  distance  from  the  seat  of  the  disturbance.* 

However,  we  know  from  our  earlier  discussion  (and  from  everyday 

*  *  It  should  be  pointed  out  once  more  that  disturbances  propagate  at  infinite 
speed  since  our  medium  is  incompressible.  Each  Fourier  component,  however, 
propagates  with  a  finite  speed. 


TWO-DIMENSIONAL   WAVES  213 

observation  of  streams,  for  that  matter)  that  as  t  ->  oo  it  may  happen 
that  a  disturbance  also  propagates  downstream  as  a  wave  with  non- 
vanishing  amplitude.  Our  main  interest  here  is  to  study  such  a  passage 
to  the  limit.  It  is  clear  that  one  cannot  accomplish  such  a  purpose 
simply  by  letting  t  ->  oo  in  (7.4.14),  since,  for  one  thing,  the  transform 
y>  of  <p  cannot  exist  if  <p  does  not  tend  to  zero  at  oo.  What  we  wish  to  do 
is  to  consider  the  contributions  of  the  separate  items  in  the  brackets 
in  (7.4.13),  and  to  avoid  any  singularities  caused  by  zeros  in  their 
denominators  by  regarding  A  (s;  t)  as  an  analytic  function  in  the  neigh- 
borhood of  the  real  axis  of  a  complex  $-plane  and  deforming  the  path 
of  integration  in  (7.4.14)  by  Cauchy's  integral  theorem  in  such  a  way 
as  to  avoid  such  singularities.  One  can  then  study  the  limit  situation 
as  t  ->  oo. 

In  carrying  out  this  program  it  is  essential  to  study  the  separate 
terms  defining  the  function  A(s;  t)  given  by  (7.4.13).  To  begin  with, 

we  observe  that  the  function  Vgs  tanh  sh  can  be  defined  as  an  analytic 
and  single-  valued  function  in  a  neighborhood  of  the  real  axis  since 
the  function  s  tanh  sh  has  a  power  series  development  at  s  =  0  that 
is  valid  for  all  s  and  begins  with  a  term  in  s2,  and,  in  addition,  the 
function  has  no  real  zero  except  s  =  0.  Once  the  function  Vgs  tanh  sh 
has  been  so  defined,  it  follows  that  each  of  the  terms  in  (7.4.13)  is  an 
analytic  function  in  a  strip  containing  the  real  axis  except  at  real  zeros 
of  the  denominators.  It  is  important  to  take  account  of  these  zeros, 
as  we  have  already  done  in  sec.  7.3.  For  our  present  purposes  it  is 
useful  to  consider  the  function 


sh 


/U2  \ 

(7.4.15)      W(s)  =  gs  [  —  .  sh  —  tanh  sh  )  =  s2U2  —  gs  tanh 

\gh      _    /  _ 

=  (sU  +  Vgs  tanh  sh)(sU  —  Vgs  tanh  sh) 

=  /+(')/-(*)• 

With  reference  to  Fig.  7.3.1  above  and  the  accompanying  discussion, 
one  sees  that  there  are  at  most  three  real  zeros  of  the  function  W(s): 
s  =  0  is  in  all  cases  a  root,  and  there  exist  in  addition  two  other  real 
roots  if  the  dimensionless  parameter  gh/U2  is  greater  than  unity.  Also, 
it  is  clear  that  if  gh/U2  =£  1  the  origin  is  a  double  root  of  W(s),  but  is 
a  quadruple  root  if  gh/U2  =  1.  In  case  gh/U2  >  I  the  real  roots  ±  ft 
of  W(s)  are  simple  roots.  (It  might  be  noted  in  passing  that  W(s)  has 
infinitely  many  pure  imaginary  zeros  ±i/?n,  n  =  1,  2,  .  .  ..) 

It  follows  at  once  that  if  we  deform  the  path  of  integration  in 


214 


WATER  WAVES 


(7.4.14)  from  the  real  axis  to  the  path  P  shown  in  Figure  7.4.1  we  can 
consider  separately  the  contributions  to  the  integral  furnished  by  each 
of  the  three  items  in  the  square  brackets  in  (7.4.13),  since  the  separate 


+  13 


Fig.  7.4.1.  The  path  P  in  the  s-plane 

integrals  would  then  exist.  This  we  proceed  to  do,  except  that  we  pre- 
fer to  consider  the  velocity  components  q>9  and  <py  of  the  disturbance 
rather  than  9?  itself.  For  q>x  we  write* 

(7.4.16)  9>«  =  ?if)  +  V®, 

with  <pW  and  <p®  defined  (in  accordance  with  (7.4.13)  and  (7.4.14))  as 
follows: 


(7.4.17) 


7.4.18)   ?W  - 


p(s)s2  cosh  s(y  +  h) 


>t'«x  , 


-u 


W(s)  cosh  sh 
p(s)s*  cosh  s(y  +  h) 


[e   ttf+S    __ e   ltf~(S] 


cosh  sh  \/gs  tanh  sh 

The  functions  W(s),  f-(s),  and  f+(s)  have  been  defined  in  (7.4.15). 
Evidently  the  notation  <p^\  (p®  has  been  chosen  to  point  to  the  fact 
that  <p^  should  yield  the  steady  part  of  the  motion  while  q>®  should 
furnish  "transients"  which  die  out  as  t  ->  oo.  This  is  indeed  the  case, 
as  we  now  show,  at  least  when  the  parameter  gh/U2  is  not  equal  to 
unity,  its  critical  value. 

Consider  first  the  case  gh/U*  <  1.  In  this  case  there  are  no  singu- 
larities on  the  real  axis,  even  at  the  origin  (cf.  (7.4.18)),  since  /+  and 
/_  vanish  to  the  first  power  and  i/gs  tanh  sh  vanishes  to  the  first 
power  also  at  s  =  0.  Since  p(s)  is  regular  at  s  =  0  and  $2  occurs  in  the 
numerator  of  the  integrand  our  statement  follows.  Consequently  the 
'path  P  can  be  deformed  back  again  into  the  real  axis.  In  this  case  the 

*  The  discussion  would  differ  in  no  essential  way  for  <pv  instead  of  <px. 


TWO-DIMENSIONAL   WAVES  215 

behavior  of  <p(£  for  large  t  can  be  obtained  by  the  principle  of  station- 
ary phase  (cf.  sec.  6.8).  In  the  present  case  the  functions  f+(s)  and 
/_(s)  have  non-vanishing  first  derivatives  for  all  s,  and  consequently 
<p(£  ->  0  at  least  like  l/t  since  there  are  no  points  where  the  phase  is 
stationary.  (Here  and  in  what  follows  no  attempt  is  made  to  give  the 
asymptotic  behavior  with  any  more  precision  than  is  necessary  for 
the  purposes  in  view. )  As  t  ->  oo  therefore  we  obtain  the  steady  state 
solution  cpW.  The  behavior  of  q>W  for  \x\  large  is  also  obtained  at 
once:  one  sees  that  the  integrand  in  (7.4.17)  has  no  singularities  in 
this  case  also,  and  it  follows  at  once  from  the  Riemann-Lebesgue 
theorem  that  9?^  ->  0  as  \x\  -»  oo.  Thus  a  steady  state  exists,  and 
it  has  the  property  that  the  disturbances  die  out  both  upstream  and 
downstream. 

We  turn  next  to  the  more  complicated  case  in  which  gh/U2  >  1. 
The  integrand  for  q>^  has  no  singularity  at  the  origin,  but  it  has 
simple  poles  at  s  =  db  P  (cf.  Figure  7.4.1)  furnished  by  simple  zeros 
of  f_(s)  at  these  points.  Again  we  show  that  (pW  ->  0  as  t  ->  oo. 
Consider  first  the  contribution  of  the  semicircles  at  s  =  ±  /?.  (Since 
s  =  0  is  not  a  singularity,  we  deform  the  path  back  into  the  real  axis 
there.)  In  the  lower  half-plane  near  s  =  ±  P  one  sees  readily  that 
f-(s)  has  a  negative  imaginary  part,  and  thus  the  exponent  in 
exp  {—  itf_(s)}  has  a  negative  real  part,  since  /_($)  is  real  on  the  real 
axis  and  its  first  derivative  f_(s)  is  positive  there  (so  that  /_($)  be- 
haves like  c(s  ^f  /?)  with  c  a  positive  constant).  Thus  for  any  closed 
portion  of  the  semicircles  which  excludes  the  end-points  the  contribu- 
tion to  the  integral  tends  to  zero  as  t  ->  oo,  and  hence  also  for  the 
whole  of  the  semicircles.  On  the  straight  parts  of  the  path  the  prin- 
ciple of  stationary  phase  can  be  used  again  to  show  that  <p$  ->  0  as 
t  ->  oo.  In  fact,  this  function  behaves  like  l/^/t  since  one  can  easily 
verify  that  /_(s)  has  exactly  two  points  of  stationary  phase,  i.e.  two 
points  ±  j80  where  /-(±  A>)  =  °  and  /"(±  /30)  ^  0.  (The  point  s  =  00 
lies  between  the  origin  and  the  point  s  =  0  where  /_($)  vanishes.) 
Thus  the  steady  state  is  again  given  by  <p^.  However,  unlike  the  pre- 
ceding case,  the  steady  state  does  not  furnish  a  motion  which  dies 
out  both  upstream  and  downstream.  This  can  be  seen  as  follows. 
Consider  first  the  behavior  upstream,  i.e.  for  x  <  0.  On  the  semicircu- 
lar parts  of  the  path  P  in  the  lower  half-plane  we  see  that  the  expo- 
nent in  ei8X  in  (7.4.17)  has  a  negative  real  part,  and  therefore  by  the 
same  argument  as  above,  these  parts  of  P  make  contributions  which 
vanish  as  x  -»  —  oo.  The  straight  parts  of  P  also  make  contributions 


216  WATER  WAVES 

which  vanish  for  large  x  (either  positive  or  negative),  by  the  Riemann- 
Lebesgue  theorem.  Thus  the  disturbance  vanishes  upstream.  On  the 
downstream  side,  i.e.  for  x  >  0,  we  cannot  conclude  that  the  semi- 
circular parts  of  P  make  vanishing  contributions  for  large  x  since  the 
exponent  in  ei8X  now  has  a  positive  real  part.  We  therefore  make  use 
of  the  standard  procedure  of  deforming  the  path  P  through  the  poles 
at  s  =  ±  /?  and  subtracting  the  residues  at  these  poles.  It  is  clear  that 
the  semicircles  in  the  upper  half-plane  yield  vanishing  contributions 
to  <pM  when  x  ->  +  <x>:  the  argument  is  the  same  as  was  used  above. 
This  leads  to  the  following  asymptotic  representation  (obtained  from 
the  contributions  at  the  poles),  valid  for  x  large  and  positive: 

(7.4.19,  ^.r.  00)= 


.. 

Q  cosh 

Here  W'(ft)  ^  0  is  the  value  of  the  derivative  of  W  (cf.  (7.4.15))  at 
s  =  /?,  and  the  fact  that  W(/J)  is  an  odd  function  has  been  used.  In 
particular,  if  the  surface  pressure  p(x)  were  given  by  the  delta  func- 
tion p(x)  =  d(x),  i.e.  if  the  disturbance  were  caused  by  a  concentrated 
pressure  point  at  the  origin,  (7.4.19)  would  yield 

,~,    ^  /  x 

(7.4.19),  ?.(*,„;  «,).= 

since  the  transform  of  d(x)  is  l/\/2n.  Another  interesting  special 
case  is  that  in  which  p(x)  is  a  constant  p0  over  the  interval  —  a  ^  x 
^  a  and  zero  over  the  rest  of  the  free  surface.  In  this  case  p  = 
(2p0/  \/2n)  (  sin  sa  )/s  and  q>x  behaves  for  large  positive  x  and  t  as  follows  : 


x          0      cosh  fi(y  +  h)   .    D     .    0 
(7A19),     „.(«,  y;  oo)  -  «n  ^a  sm  /to. 


This  yields  the  curious  result  (mentioned  above)  that  under  the  pro- 
per circumstances  the  disturbance  may  die  out  downstream  as  well  as 
upstream;  it  will  in  fact  do  so  if  (ia  =  nn,  i.e.  if  the  length  2a  of  the 
segment  over  which  the  disturbing  pressure  is  applied  is  an  integral 
multiple  of  the  wave  length  at  oo—  which  is,  in  turn,  fixed  by  the 
velocity  U  and  the  depth  h. 

Finally  we  consider  the  critical  case  gh/U*  =  1,  and  begin  by  dis- 
cussing the  behavior  of  the  time  dependent  terms  in  <p  as  t  ->  oo.  For 
this  purpose  it  is  convenient  to  deal  first  with  the  time  derivative  of 
this  function: 


TWO-DIMENSIONAL   WAVES  217 


(7.4.80)    *,= 


P  cosh  $A  \/g$  tanh 


The  integrand  has  no  singularities  on  the  real  axis  and  consequently 
the  path  P  can  be  deformed  into  the  real  axis.  Thus  the  principle  of 
stationary  phase  can  be  employed  once  more.  Since  the  derivative  of 
f+(s)  =  sU  +  \/gs  tanh  sh  evidently  does  not  vanish  for  any  real  s 
while  the  derivative  of  /_(s)  has  one  zero  at  s  =  0,  it  follows  that  the 
leading  term  in  the  asymptotic  development  of  9?^  for  large  t  arises 
from  the  term  exp  {—  itf__(s)}.  Since,  in  addition,  £'(0)  =  0  but  /'"(O) 
^  0  we  have  (cf.  sec.  6.8): 

(7.4.21  )     0f  -  Ap(0).  -  ,         A  =  const.  ^  0. 


Since  p(0)  is  in  general  different  from  zero,  it  follows  that  <pf^  behaves 
like  r1/3  and  hence  that  q>(t)  becomes  infinite  everywhere  (for  all  x 
and  y,  that  is)  like  t*/3  as  t  ->  oo.*  Thus  a  steady  state  does  not  exist 
if  one  considers  it  to  be  the  limit  as  t  ->  oo.  It  might  be  thought  that 
the  existence  in  practice  of  dissipative  forces  could  lead  to  the  vanish- 
ing of  the  transients  and  thus  still  leave  the  steady  state  (p^  as  given 
by  (7.4.17)  as  a  representation  of  the  final  motion.  That  is,  however, 
also  not  satisfactory  since  <p^  becomes  unbounded  for  x  large  when 
gh/U2  =  1:  at  the  origin  there  is  a  pole  of  order  two  since  W($)  be- 
haves like  s4  and  consequently  the  term  isx  in  the  power  series  for  eisx 
leads  to  a  contribution  from  this  pole  which  is  linear  in  x.  In  linear 
theories  based  on  assuming  small  disturbances  one  is  reconciled  to 
singularities  and  infinities  at  isolated  points,  but  hardly  to  arbitrarily 
large  disturbances  in  whole  regions.  All  of  this  suggests  that  the 
reasonable  attitude  to  take  in  these  circumstances  is  that  the  linear 
theory,  which  assumes  small  disturbances,  fails  altogether  for  flows 
at  the  critical  speed  U*/gh  =  I  and  that  one  should  go  over  to  a  non- 


*  It  might  seem  odd  that  we  have  chosen  to  discuss  the  function  <p^  rather 
than  the  function  <p®  (as  we  did  in  the  other  cases).  The  reason  is  that  the  asymp- 
totic behavior  of  q>^  is  not  easily  obtained  directly  by  the  method  of  stationary 
phase  in  the  present  case  since  the  coefficient  of  the  leading  term  in  this  develop- 
ment would  be  zero.  However,  one  could  show  (by  using  Watson's  lemma,  for 
example,  which  yields  the  complete  asymptotic  expansion  of  the  integral)  that 

q>®  behaves  like  f~^3,  and  hence  that  q>®  behaves  like  *^3. 


218  WATER  WAVES 

linear  theory  in  order  to  obtain  reasonable  results  from  the  physical 
point  of  view.  In  Chapter  10.9,  which  deals  with  the  solitary  wave  (an 
essentially  nonlinear  phenomenon),  we  shall  see  that  such  a  steady 
wave  exists  for  flows  with  velocities  in  the  neighborhood  of  the 
critical  value. 


CHAPTER  8 


Waves  Caused  by  a  Moving  Pressure  Point.  Kelvin's 
Theory  of  the  Wave  Pattern  Created  by  a  Moving  Ship 

8.1.  An  idealized  version  of  the  ship  wave  problem.  Treatment  by  the 
method  of  stationary  phase 

The  peculiar  pattern  of  the  waves  created  by  objects  moving  over 
the  surface  of  the  water  on  a  straight  course  has  been  noticed  by 
everyone:  the  disturbance  follows  the  moving  object  unchanged  in 
form  and  it  is  confined  to  a  region  behind  the  object  that  has  the  same 
v-shape  whether  the  moving  object  is  a  duck  or  a  battleship.  An  ex- 
planation and  treatment  of  the  phenomenon  was  first  given  by 
Kelvin  [K.ll],  and  this  work  deserves  high  rank  among  the  many 
imaginative  things  created  by  him.  As  was  mentioned  earlier,  Kelvin 
invented  his  method  of  stationary  phase  as  a  tool  for  approximating 
the  solution  of  this  particular  problem,  and  it  is  indeed  a  beautiful 
and  strikingly  successful  example  of  its  usefulness. 

It  should  be  stated  at  once  that  there  is  no  notion  in  this  and  the 
next  following  section  of  solving  the  problem  of  the  waves  created  by 
an  actual  ship  in  the  sense  that  the  shape  of  the  ship's  hull  is  to  be 
taken  into  account;  such  problems  will  be  considered  in  the  next 
chapter.  For  practical  purposes  an  analysis  of  the  waves  in  such  cases 
is  very  much  desired,  since  a  fraction— even  a  large  fraction  if  the 
speed  of  the  ship  is  large— of  the  resistance  to  the  forward  motion  of  a 
ship  is  due  to  the  energy  used  up  in  maintaining  the  system  of  gravity 
waves  which  accompanies  it.  The  problem  has  of  course  been  studied, 
in  particular,  in  a  long  series  of  notable  papers  by  Havelock,*  but  the 
difficulties  in  carrying  out  the  discussion  in  terms  of  parameters  which 
fix  the  shape  of  the  ship  are  very  great.  Indeed,  a  more  or  less  com- 
plete discussion  of  the  solution  to  all  orders  of  approximation  even  in 
the  very  much  idealized  case  to  be  studied  in  the  present  chapter,  is 
by  no  means  an  easy  task— in  fact,  such  a  complete  discussion,  along 

References  to  some  of  these  papers  will  be  given  in  the  next  chapter. 

219 


220  WATER   WAVES 

lines  quite  different  from  those  of  Kelvin,  has  been  carried  out  only 
rather  recently  by  A.  S.  Peters  [P.  4]  (cf.  also  the  earlier  paper  by 
Hogner  [H.16]).  However,  we  shall  follow  Kelvin's  procedure  here  in 
a  general  way,  but  with  many  differences  in  detail. 

The  problem  we  have  in  mind  to  discuss  as  a  primitive  substitute 
for  the  case  of  an  actual  ship  is  the  problem  of  the  surface  waves 
created  by  a  point  impulse  which  moves  over  the  surface  of  the  water 
(assumed  to  be  infinite  in  depth).  We  shall  take  the  solution  of  section 
6.5  for  the  wave  motion  due  to  a  point  impulse  and  integrate  it  along 
the  course  of  the  "ship"—  in  effect,  the  surface  waves  caused  by  the 
ship  are  considered  to  be  the  cumulative  result  of  impulses  delivered 
at  each  point  along  its  course.  The  result  will  be  an  integral  represen- 
tation for  the  solution,  in  the  form  of  a  triple  integral,  which  can  be 
discussed  by  the  method  of  stationary  phase.  However,  it  is  necessary 
to  apply  the  method  of  stationary  phase  three  times  in  succession,  and 
if  this  is  not  done  with  some  care  it  is  not  clear  that  the  approximation 
is  valid  at  all;  or  what  is  perhaps  equally  bad  from  the  physical  point 
of  view,  it  may  not  be  clear  where  the  approximation  can  be  expected 
to  be  accurate.  Thus  it  seems  worth  while  to  consider  the  problem  with 
some  attention  to  the  mathematical  details;  this  will  be  done  in  the 
present  section,  and  the  interpretation  of  the  results  of  the  approxima- 
tion will  be  carried  out  in  the  next  section  (which,  it  should  be  said, 
can  be  read  pretty  much  independently  of  the  present  section). 

From  section  6.4  the  vertical  displacement*  r)(x,  y,  z;  t)  of  the  water 
particles  due  to  a  point  impulse  applied  on  the  surface  at  the  origin 
and  at  the  time  t  =  0  can  be  put  in  the  form 

I        /*oo  /»w/2 

(8.1.1)    77(0?,  t/,  z;t)  =  —  -  -  I    asmat'emvmdm         cos(mrcos/J)d/J 


in  which  a2  =  gm  and  r2  =  x2  +  z2.  We  have  replaced  the  Bessel 
function  J0(mr)  by  its  integral  representation 

2  f*/2 
J0(mr)  =  —         cos  (mr  cos  ft)  df} 


for  reasons  which  will  become  clear  in  a  moment.  As  we  have 
indicated,  our  intention  is  to  sum  up  the  effect  of  such  impulses 
as  the  "ship"  moves  along  its  course  C.  The  notations  to  be  used  for 

*  Actually,  we  have  considered  only  the  displacement  of  the  free  surface  in 
that  section,  but  it  is  readily  seen  that  (8.1.1)  furnishes  the  vertical  displacement 
of  any  points  in  the  water. 


WAVE  PATTERN  CREATED  BY  A  MOVING  SHIP 


221 


this  purpose  are  indicated  in  Figure  8.1.1,  which  is  to  be  considered  as 
a  vertical  projection  of  the  free  surface  on  any  plane  y  =  const.  The 
course  of  the  ship  is  given  in  terms  of  a  parameter  t  by  the  relations 


(8.1.2) 


0  ^  t  ^  T, 


and  t  is  assumed  to  mean  the  time  required  for  the  ship  to  travel 
from  any  point  Q(xv  z^  on  its  course  to  its  present  position  at  the 


kP(x,z) 

Fig.  8.1.1.  Notation  for  the  ship  wave  problem 

origin.  We  seek  the  displacement  of  the  water  at  (x,  y,  z)  when  the 
ship  is  at  the  origin;  it  is  therefore  determined  by  the  integral 

(8.1.3)    r](x,y,z) 

I      /»r  /•»  pjr/2 

= k(t)  dt       a  sin  at  emym  dm        cos  (mr  cos 

%ngQ  Jo  Jo  Jo 


In  this  formula  k(t)  represents  the  strength  of  the  impulse,  which  we 
might  reasonably  assume  to  be  constant  if  the  speed  of  the  ship  is 
constant;  this  constant  is  therefore  the  only  parameter  at  our  disposal 
which  might  serve  to  represent  the  effect  of  the  volume,  shape,  etc. 
of  a  ship's  hull.  We  write  the  last  relation  in  the  form 

rooji/2 
(8.1.4)  ri(x,y>z)~K  [[(amemv[ei(at~mrcos®  +  ei(at+mrcos®]  d@  dm  dt 

ooo 

with  the  understanding  that  the  imaginary  part  of  the  integral  is  to  be 
taken.  (K  is  a  constant  the  value  of  which  is  not  important  for  the 


222  WATER   WAVES 

discussion  to  follow.)  It  should  be  noted  that  r2  =  (x  —  x^2  + 
(z  —  3X)2.  Since  y  <  0,  the  integral  converges  strongly  because  of  the 
exponential  factor. 

One  of  the  puzzling  features  (to  the  author,  at  least)  of  existing 
treatments  of  the  problem  by  the  method  of  stationary  phase  is  that 
it  is  not  made  clear  what  parameter  is  large  in  the  exponentials  as  the 
method  is  applied  to  each  of  the  three  integrals  in  turn,  so  that  one  is 
not  quite  sure  whether  there  might  not  be  an  inconsistency.  The 
ntatter  is  easily  clarified  by  introduction  of  appropriate  dimensionless 
quantities,  as  follows  (cf.  Figure  8.1.1): 

(8.1.5) 

x  =  R  cos  a,  xl  =  Rl  cos  ax ,         z  =  R  sin  a,  z±  —  Rl  sin  al5 


r  =  R  V(A  cos  ax  —  cos  a)2  +  (Asinax  —  sin  a)2  =  R  •  /, 

r  =  ct/R,        R,/R  =  A,        *  =  ^  ,         m  =  ^-l2. 

4c2  4r2 

Here  the  quantity  c  represents  the  speed  of  the  ship  in  its  course.  It 
should  be  noted  that  x,  y,  and  z  are  held  fixed—  they  represent  the 
point  at  which  the  displacement  is  to  be  observed—,  but  that  xl9  zl 
(and  hence  Rl  and  o^),  and  r  all  depend  on  t.  We  have  also  introduced 
a  new  variable  of  integration  f  ,  replacing  m,  which  depends  on  t.  The 
Jacobian  9(ra,  t)/d(£,  r)  has  the  value  gt2Rg/(2cr2)  and  hence  vanishes 
only  for  t  =  0.  In  terms  of  the  new  quantities  the  integral  (8.1.4)  is 
found  to  take  the  form: 

(8.1.6)     r)(x,y,z) 

TO  oojz/2 


m-,3T5£4     xr^y  f 
w  ^ 


000 

where  TO  =  cT//e. 

Again  we  remark  that  the  integral  converges  uniformly  for  y  <  0. 
However,  the  integrand  has  a  singularity  if  the  point  (x,  y,  z)  happens 
to  be  vertically  under  a  point  on  the  course  of  the  ship:  in  such  a  case 
we  have  R  =  Rl  (i.e.  A  ==  1  ),  and  a  =  alf  so  that  I  =  0  for  a  certain 
value  r  T£  0  in  the  interval  0  fS  r  ^  TO,  Because  of  the  exponential 
factor,  the  integral  continues  to  exist,  however.  Indeed,  one  sees 
from  (8.1.4)  that  taking  r  =  0  does  not  make  the  integrand  singular; 


WAVE  PATTERN  CREATED  BY  A  MOVING  SHIP        228 

the  fact  that  a  singularity  crops  up  in  (8.1.6)  arises  from  our  choice 
of  the  variable  |  which  replaces  ra.  This  disadvantage  caused  by  intro- 
duction of  the  new  variables  is  much  more  than  outweighed  by  the 
fact  that  we  now  can  see  that  the  approximation  by  the  method  of 
stationary  phase  depends  only  on  one  parameter,  i.e.  the  parameter 
x  =  gjR/4r2  in  the  exponentials.  We  can  expect  the  use  of  the  method 
of  stationary  phase  to  yield  an  accurate  result  if  this  parameter  is 
large,  and  that  in  turn  is  certainly  the  case  if  R  is  large,  i.e.  for  points 
not  too  near  the  vertical  axis  through  the  present  location  of  the  ship. 

The  application  of  the  method  of  stationary  phase  to  the  integral 
in  (8.1.6)  can  now  be  justified  by  an  appeal  to  the  arguments  used  in 
section  6.8.  In  doing  so,  the  multiple  integral  is  evaluated  by  inte- 
grating with  respect  to  each  variable  in  turn;  at  the  same  time,  the 
integrands  are  replaced  by  their  asymptotic  representations  as  fur- 
nished by  the  method  of  stationary  phase.  One  need  only  observe,  in 
verifying  the  correctness  of  such  a  procedure,  that  the  integrands 
remain,  after  each  integration,  in  a  form  such  that  the  arguments  of 
that  section  apply— in  particular  that  they  remain  analytic  functions 
of  their  arguments  provided  only  that  points  (x,  y,  z)  on  or  under  the 
ship's  course  are  avoided*— and  that  an  asymptotic  series  can  be  in- 
tegrated termwise.  It  is  not  difficult  to  see  that  the  contributions  to 
YI(X,  y,  z)  of  lowest  order  in  \\x  are  made  by  arbitrarily  small  domains 
containing  in  their  interiors  a  point  where  the  derivatives  9?^,  <p^9  <pr  of 
the  phase  9?  =  (2£  —  £2  cos  ^)r2/l(r)  vanish  simultaneously. 

Even  for  points  on  the  ship's  course  the  argument  of  section  6.8 
will  still  hold  provided  that  no  stationary  point  of  the  phase  <p  occurs 
for  a  value  of  r  such  that  l(r)  =  0:  the  reason  for  this  is  that  the 
assumption  of  analyticity  was  used  in  section  6.8  only  to  treat  a 
neighborhood  of  a  point  of  stationary  phase,  while  for  other  segments 
of  the  field  of  integration  only  the  assumptions  of  integrability  and 
the  possibility  of  integration  by  parts  are  needed.  It  happens  that  the 
cases  to  be  treated  later  on  are  such  that  l(r)  does  not  vanish  at  any 
points  of  stationary  phase,  and  hence  for  them  the  asymptotic 
approximation  is  valid  also  for  points  on  the  ship's  course. 

There  is  one  further  mathematical  point  to  be  mentioned.  The 

/•  & 
*  In  section  6.8  the  integrals  studied  were  of  the  form       y(x)  exp  (ikq>(x)}  dx, 

/•b  Jo 

while  here  the  integral  is  of  the  form       y>(x,  k)  exp  (ifop(x)}  dx.   However,  one 

Ja 

can  verify  that  the  argument  used  in  section  6.8  can  easily  be  generalized  to 
include  the  present  case. 


224  WATER   WAVES 

above  discussion  requires  that  we  take  y  <  0,  and  it  is  not  entirely 
clear  that  the  passage  to  the  limit  y  ->  0  is  legitimate  in  the  approxi- 
mate formulas,  so  that  the  validity  of  the  discussion  might  be  thought 
open  to  question  for  points  on  the  free  surface.  Indeed,  it  would  appear 
to  be  difficult  to  justify  such  a  limit  procedure  for  the  integral  in 
(8.1.1),  for  instance,  since  it  certainly  does  not  converge  if  we  set 
y  =  0  since  the  integrand  then  does  not  even  approach  zero  as  m  ->•  oo. 
However,  this  is  a  consequence  of  dealing  with  a  point  impulse.  If  we 
had  assumed  as  model  for  our  ship  a  moving  circular  disk  of  radius  a 
over  which  a  constant  distribution  of  impulse  is  taken,  the  result  for 
the  vertical  displacement  due  to  such  a  distributed  impulse  applied 
at  t  =  0  could  be  shown  to  be  given  by 

/*00 

rj(x,  y,  z\  t)  =  Kl        a  sin  at  •  ^mvJr0(mr)J1(ma)  dm 
j  o 

with  Ji(ma)  the  Bessel  function  of  order  one  and  Kl  a  certain  constant. 
This  integral  converges  uniformly  for  y  ^  0,  as  one  can  see  from  the 
asymptotic  behavior  of  JQ(mr)  and  J^(ma).  Consequently  rj(x,  y,  z;  t) 
is  continuous  for  y  ~  0.  On  the  other  hand,  if  the  radius  a  of  the  disk 
is  small  the  result  cannot  be  much  different  from  that  for  the  point 
impulse.  Thus  we  might  think  of  the  results  obtained  in  the  next 
section,  which  start  with  the  formula  (8.1.1 )  for  a  point  impulse,  as  an 
approximation  on  the  free  surface  to  the  case  of  an  impulse  distributed 
over  a  disk  of  small  radius. 

It  has  already  been  mentioned  that  the  problem  under  discussion 
here  has  been  treated  by  A.  S.  Peters  [P.4]  by  a  different  method. 
Peters  obtains  a  representation  for  the  solution  based  on  contour 
integrals  in  the  complex  plane,  which  can  then  be  treated  by  the 
saddle  point  method  to  obtain  the  complete  asymptotic  development 
of  the  solution  with  respect  to  the  parameter  x  defined  above,  while 
we  obtain  here  only  the  term  of  lowest  order  in  such  a  development. 
However,  the  methods  used  by  Peters  lead  to  rather  intricate  deve- 
lopments. 

8.2.  The  classical  ship  wave  problem.  Details  of  the  solution 

In  the  preceding  section  we  have  justified  the  repeated  application 
of  the  method  of  stationary  phase  to  obtain  an  approximate  solution 
for  the  problem  of  the  waves  created  when  a  point  impulse  moves  over 
the  surface  of  water  of  infinite  depth.  In  particular,  it  was  seen  that 
the  approximation  obtained  in  that  way  is  valid  at  all  points  on  the 


WAVE    PATTERN    CREATED    BY   A   MOVING   SHIP 


225 


surface  of  the  water  not  too  near  to  the  position  of  the  "ship"  at  the 
instant  when  the  motion  is  to  be  determined  (provided  only  that  a 
certain  condition  is  satisfied  at  points  on  the  ship's  course).  In  this 
section  we  carry  out  the  calculations  and  discuss  the  results,  returning 
however  to  the  original  variables  since  no  gain  in  simplicity  would  be 
achieved  from  the  use  of  the  dimensionless  variables  of  the  preceding 
section. 

Kelvin  carried  out  his  solution  of  the  ship  wave  problem  for  the 
case  of  a  straight  line  course  traversed  at  constant  speed.  Up  to  a 
certain  point  there  is  no  difficulty  in  considering  more  general  courses 


\P(x,z) 
Fig.  8.2.1.  Notation  for  the  ship  wave  problem 

for  the  ship.  In  Figure  8.2.1  we  indicate  the  course  C  as  any  curve 
given  in  terms  of  a  parameter  t  by  the  equations 


(8.2.1) 


x   = 


for  0  <,  t  <  T. 


The  parameter  t  is  taken  to  represent  the  time  required  for  the  ship 
to  pass  from  any  point  (#15  2X)  to  its  present  position  at  the  origin  O, 
but  it  is  convenient  to  take  t  =  0  to  correspond  to  the  origin  so  that 
the  point  (xv  yx)  moves  backward  along  the  ship's  course  as  t  increases. 
The  shape  of  the  waves  on  the  free  surface  is  to  be  determined  at  the 
moment  when  the  ship  is  at  the  origin.  The  #-axis  is  taken  along  the 
tangent  to  the  course  C,  but  is  taken  positive  in  the  direction  opposite 
to  the  direction  of  travel  of  the  ship,  Since  we  have  taken  t  =  0  at  the 
origin  the  parameter  t  in  (8.2.1)  is  really  the  negative  of  the  time;  as 
a  consequence  the  tangent  vector  t  to  C  at  a  point  Q(xl9  t/t)  as  given  by 


is  in  the  direction  opposite  to  that  of  the  ship  in  traversing  the  course 


226  WATER   WAVES 

C.  The  speed  c(t)  of  the  ship  is  the  length  of  the  vector  t  and  is  given  by 

•» 

The  point  P(x,  z)  is  the  point  at  which  the  amplitude  of  the  surface 
waves  is  to  be  computed;  it  is  located  by  means  of  the  vector  r: 

(8.2.3)  r  =  (x  —  xl9  z  —  %). 

The  angle  0  indicated  on  the  figure  is  the  angle  (^  n)  between  the 
vectors  r  and  —  t. 

As  we  have  stated  earlier,  the  surface  elevation  r)(%9  z)  at  P(x,  z) 
is  to  be  determined  by  integrating  the  elevations  due  to  a  point  im- 
pulse moving  along  C.  The  effect  of  an  impulse  at  the  point  Q  is  ass- 
sumed  to  be  given  by  the  approximate  formula  (6.5.15),  in  which, 
however,  we  omit  a  constant  multiplier  which  is  unessential  for  the 
discussion  to  follow: 

—  t3        tf? 

(8.2.4)  rj(r;t)~  ---  sin  2_. 

r4  4r 

In  other  words,  we  assume  that  the  formula  (8.1.1)  for  the  surface 
elevation  r)  has  been  approximated  by  two  successive  applications 
of  the  method  of  stationary  phase.  This  formula  yields  the  effect 
at  time  t  and  at  a  point  distant  r  from  the  point  where  the  impulse 
was  applied  at  the  time  t  =  0;  it  therefore  applies  in  the  present  situa- 
tion with 

(8.2.5)  r*  =  (x  -  ^)2  +  (z  -  ^)2, 

since  t  does  indeed  represent  the  length  of  time  elapsed  since  the 
"ship"  passed  the  point  Q  on  its  way  to  its  present  position  at  O.  The 
integrated  effect  of  all  the  point  impulses  is  therefore  given  by 


dt, 


CTl3        et2 

(8.2.6)  n(x9  z)  =  A;0      -  sin  51 

Jo  r4        4r 

with  fc0  a  certain  constant.  For  points  on  the  ship's  course,  where 
r  =  0  for  some  value  t  =  t0  in  the  interval  0  rg  t  ^  T,  this  integral 
evidently  does  not  exist.  However,  it  has  been  shown  in  the  preceding 
section  that  neighborhoods  of  such  points  can  be  ignored  in  calculating 
rj  approximately  provided  that  they  are  not  points  of  stationary  phase. 
This  condition  will  be  met  in  general,  and  hence  we  may  imagine  that 
a  small  interval  about  a  point  where  r(t0)  =  0  has  been  excluded  from 


WAVE  PATTERN  CREATED  BY  A  MOVING  SHIP        227 

the  range  of  integration  in  case  we  wish  the  wave  amplitude  at  a  point 
on  the  ship's  course.  We  write  the  integral  in  the  form 

fr 

(8.2.7)  ri(xt  z)  =       V(*)*w)  dt, 

Jo 

and  take  the  imaginary  part.  The  function  \p(t)  and  the  phase  <p(t) 
are  given  by 

(8.2.8)  y(t)  =  kQt*/r* 

(8.2.9)  <p(t)  =  gt*/4r. 

We  proceed  to  make  the  calculations  called  for  in  applying  the 
stationary  phase  method.  In  the  integral  given  by  (8.2.7)  no  large 
parameter  multiplying  the  phase  is  put  explicitly  in  evidence;  how- 
ever, from  the  discussion  of  the  preceding  section  we  know  that  the 
approximation  will  be  good  if  the  dimensionless  quantity  gJ?/4c2, 
with  R  the  distance  from  the  ship,  is  large.  It  could  also  be  verified 
that  (8.2.6)  would  result  if  the  integrations  in  (8.1.6)  on  ft  and  £  were 
first  approximated  by  stationary  phase  followed  by  a  re-introduction 
of  the  original  variables.  We  therefore  begin  by  calculating  dq>/dt: 


(8.2.10)  = 

v  '  dt       4\r 

Hence  the  condition  of  stationary  phase,  dtpjdt  =  0,  leads  to  the  im- 
portant relation 

(8.2.11)  <^  =  *. 

dt        t 

The  quantity  dr/dt  is  next  calculated  for  the  ship's  course  using 
(8.2.5);  we  find  (cf.  Figure  8.2.1): 


=  —  r  •  t  —  cr  cos  0, 
in  which  c(t)  is  once  more  the  speed  of  the  ship.  Thus 

dr 

(8.2.13)  —  =  ccos0, 

dt 

which  is  a  rather  obvious  result  geometrically.  Combining  (8.2.11) 
and  (8.2.13)  yields  the  stationary  phase  condition  in  the  form 

(8.2.14)  r  = 


228 


WATER   WAVES 


We  recall  once  more  the  significance  of  this  relation:  for  a  fixed  point 
P(#,  y)  it  yields  those  points  Qf  on  C  which  are  the  sole  points  effective 
(within  the  order  of  the  approximation  considered)  in  creating  the 
disturbance  at  P— the  contributions  from  all  other  points  being,  in 
effect,  cancelled  out  through  mutual  interference.  It  is  helpful  to  intro- 
duce the  term  influence  points  for  the  points  fy  determined  in  this 
way  relative  to  a  point  P  at  which  the  surface  elevation  of  the  water 
is  to  be  calculated. 

The  last  observation  makes  it  possible  to  draw  an  interesting  con- 
clusion at  once  from  (8.2.14),  which  can  be  interpreted  in  the  following 
way  (cf.  Figure  8.2.2):  At  point  Q  the  speed  c  of  the  ship  and  /  are 


Fig.  8.2.2.  Points  influenced  by  a  given  point  Q 

known.  The  relation  (8.2.14)  then  yields  the  polar  coordinates  (r,  0), 
with  respect  to  Q,  of  all  points  P  for  which  Q  is  the  influence  point  in 
the  sense  of  the  stationary  phase  approximation.  Such  points  P 
evidently  lie  on  a  circle  with  a  diameter  tangent  to  the  course  C  of 
the  ship  at  Q,  and  Q  is  at  one  end  of  the  diameter.  The  center  of  the 
circle  is  located  on  the  tangent  line  from  Q  in  the  direction  toward 
which  the  ship  moves  (i.e.  in  the  direction  —  t).  We  repeat  that  the 
points  P  on  the  circle  just  described  are  the  only  points  for  which  Q 
is  a  point  of  stationary  phase  of  the  integral  (8.2.7),  and  consequently 
the  contribution  of  the  impulse  applied  at  Q  vanishes  (within  the 
order  considered  by  us)  for  all  points  except  those  on  the  circle.  It 
now  becomes  obvious  that  the  disturbance  created  by  the  ship  does 
not  affect  the  whole  surface  of  the  wpter,  since  only  those  points  are 


WAVE  PATTERN  CREATED  BY  A  MOVING  SHIP 


229 


affected  which  lie  on  one  or  more  of  the  circles  of  influence  of  all  points 
Q  on  the  ship's  course.  In  other  words,  the  surface  waves  created  by 
the  moving  ship  will  be  confined  to  the  region  covered  by  all  the  in- 
fluence circles,  and  thus  to  the  region  bounded  by  the  envelope  of  this 
one-parameter  family  of  curves.  This  makes  it  possible  to  construct 
graphically  the  outline  of  the  disturbed  region  for  any  given  course 
traversed  at  any  given  speed:  one  need  only  draw  the  circles  in  the 
manner  indicated  at  a  sufficient  number  of  points  Q  and  then  sketch 
the  envelope.  Two  such  cases,  one  of  them  a  straight  course  traversed 
at  constant  speed,  the  other  a  circular  course,  are  shown  in  Figure 
8.2.3.  In  the  case  of  the  straight  course  it  is  clear  that  the  envelope 


(a)  (b) 

Fig.  8.2.3.  Region  of  disturbance  (a)  Circular  course  (b)  Straight  course 

is  a  pair  of  straight  lines;  the  disturbance  is  confined  to  a  sector  of 
semi-angle  r  given  by  r  =  arc  sin  1/3  =  19°28',  as  one  readily  sees 
from  Figure  8.2.3.  This  is  already  an  interesting  result:  it  says  that 
the  waves  following  the  ship  not  only  are  confined  to  such  a  sector 
but  that  the  angle  of  the  sector  is  independent  of  the  speed  of  the 
ship  as  long  as  the  speed  is  constant.  If  the  speed  were  not  constant 
along  a  straight  course,  the  region  of  disturbance  would  be  bounded 
by  curved  lines,  and  its  shape  would  also  change  with  the  time.  It  is, 
of  course,  not  true  that  the  disturbance  is  exactly  zero  outside  the 
region  of  disturbance  as  we  have  defined  it  here;  but  rather  it  is 
small  of  a  different  order  from  the  disturbance  inside  that  region. 
The  observations  of  actual  moving  ships  bear  out  this  conclusion  in 
a  quite  startling  way,  as  one  sees  from  Figures  8.2.4  and  8.2.5. 
The  discussion  of  the  region  of  disturbance  has  furnished  us  with  a 
certain  amount  of  interesting  information,  but  we  wish  to  know  a  good 
deal  more.  In  particular,  we  wish  to  determine  the  character  of  the 


230  WATER   WAVES 

wave  pattern  created  by  the  ship  and  the  amplitude  of  the  waves. 
For  these  purposes  a  more  thoroughgoing  analysis  is  necessary,  and  it 
will  be  carried  out  later. 

In  the  special  case  of  a  straight  course  traversed  at  constant  speed 
it  is  possible  to  draw  quite  a  few  additional  conclusions  through  fur- 
ther discussion  of  the  condition  (8.2.14)  of  stationary  phase.  In  the 
above  discussion  we  asked  for  the  points  P  influenced  by  a  given 
point  Q  on  the  ship's  course.  We  now  reverse  the  question  and  ask  for 


Fig.  8.2.4.  Ships  in  a  straight  course 


WAVE   PATTERN   CREATED   BY   A   MOVING   SHIP  231 

the  location  of  all  influence  points  Q*  that  correspond  to  a  given  point 
P.  This  question  can  be  answered  in  our  special  case  by  making  an- 
other simple  geometrical  construction  (cf.  Lamb  [L.3],  p.  435),  as 
indicated  in  Figure  8.2.6.  In  this  figure  O  represents  the  location  of 


Fig.  8.2.5a.  A  ship  in  a  circular  course 

the  ship,  P  the  point  for  which  the  influence  points  are  to  be  deter- 
mined. The  construction  is  made  as  follows:  A  circle  through  P  with 
center  on  OP  and  diameter  half  the  length  of  OP  is  constructed;  its 
intersections  with  the  ship's  course  are  denoted  by  Sl  and  S2.  From 
the  latter  points  lines  are  drawn  to  P  and  segments  orthogonal  to 
them  at  P  are  drawn  to  their  intersections  Qt  and  Q2  on  the  ship's 
course.  The  points  Ql  and  Q2  are  the  desired  influence  points.  To  prove 
that  the  construction  yields  the  desired  result  requires  only  a  verifica- 
tion that  P  does  indeed  lie  on  the  influence  circles  determined  by  the 
points  Q!  and  Q2  in  the  manner  explained  above.  Consider  the  point 


282 


WATER   WAVES 


Q19  for  example.  Since  the  angle  SlPQl  =  90°,  it  follows  that  a 
circle  with  S^  as  diameter  contains  the  point  P.  The  segments  RS1 
and  PQl  are  parallel  since  both  are  at  right  angles  to  S^;  by  con- 
sidering the  triangle  OPQa  one  now  sees  that  the  segment  OSl  is  just 


Fig.  8.2.5b.  Ships  in  curved  courses 

half  the  length  of  OQ1?  and  that  is  all  that  is  necessary  to  show  that  the 
circle  having  S^  as  diameter  is  the  influence  circle  for  Qr  Thus  there 
are  in  general  two  influence  points  or  no  influence  points,  the  latter 
case  corresponding  to  points  P  outside  the  influence  region;  the  tran- 
sition occurs  when  P  is  on  the  boundary  of  the  region  of  influence 
(i.e.  when  the  circle  of  Figure  8.2.6  having  PR  as  diameter  is  tangent 
to  the  course  OQ2  of  the  ship),  and  one  sees  that  in  this  limit  case  the 
two  influence  points  Qx  and  Q2  coalesce.  Consequently  one  might  well 
expect  that  the  amplitude  of  the  waves  at  the  boundary  of  the  region 
of  disturbance  will  be  higher  than  at  other  places,  and  this  phenome- 
non is  indeed  one  of  the  prominent  features  always  observed  physical- 


WAVE    PATTERN    CREATED    BY   A   MOVING    SHIP 


233 


ly.  In  addition,  the  direction  of  the  curves  of  constant  phase— a  wave 
crest,  or  trough,  for  example— can  be  determined  graphically  by  the 
above  construction:  one  expects  these  curves  to  be  orthogonal  to  the 


Fig.  8.2.5c.  Aircraft  carriers  maneuvering  (from  Life  Magazine) 

lines  PQl  and  PQ2  drawn  back  from  a  point  P  to  each  of  the  points  of 
influence  corresponding  to  P.  That  this  is  indeed  the  case  will  be  seen 
later,  but  it  is  evidently  a  consequence  of  the  fact  that  the  wave  at  P 
is  the  sum  of  two  circular  waves,  one  generated  at  Qi  and  the  other  at 


234 


WATER   WAVES 


Q2-  Thus  we  see  that  the  wave  pattern  behind  the  ship  is  made  up  of 
two  different  trains  of  waves— another  fact  that  is  a  matter  of  com- 
mon observation  and  which  is  well  shown  in  Figures  8.2.4  and  8.2.5. 
We  have  been  able  to  draw  a  considerable  number  of  interesting  and 


Fig.  8.2.6.  Influence  points  corresponding  to  a  given  point 

basic  conclusions  of  a  qualitative  character  through  use  of  the  condi- 
tion of  stationary  phase  (8.2.14).  We  proceed  next  to  study  analytic- 
ally the  shape  of  the  disturbed  water  surface  by  determining  the 
curves  of  constant  phase,  and  later  on  by  determining  the  amplitude 
of  the  waves.  To  calculate  the  curves  of  constant  phase  it  is  convenient 
to  express  the  basic  condition  (8.2.14)  of  stationary  phase  in  other 
forms  through  introduction  of  the  following  quantity  a,  which  has  the 
dimension  of  length: 

2c2          c2t2 

(8.2.15)  a  =  — w  =  — . 

g  2r 

From  (8.2.14)  one  then  finds 

(8.2.16)  ct  =  a  cos  0,  and 

(8.2.17)  r  =  £acos20, 

as  equivalent  expressions  of  the  stationary  phase  condition. 

It  would  be  possible  to  calculate  the  curves  of  constant  phase  for 
any  given  course  of  the  ship.  We  carry  this  out  for  the  case  of  a  cir- 
cular course  (this  case  has  been  treated  by  L.  N.  Sretenski  [S.15] )  and 
a  straight  course  traversed  at  constant  speed.  The  notation  for  the  case 
of  the  circular  course  is  indicated  in  Figure  8.2.7,  which  should  be  com- 
pared with  Figure  8.2.1.  For  the  past  position  (xv  zl )  of  the  ship  we  have 


WAVE    PATTERN   CREATED    BY   A   MOVING   SHIP 


285 


•P(x,z) 

Fig.  8.2.7.  Case  of  a  circular  course 


{xl  ==  R  sin  a 
z1  =  R(l  —  cos  a) 


(8.2.18) 

with 

(8.2.19)  a  =  ct/R. 

Here  R  is  the  turning  radius  of  the  ship,  t  the  time  required  for  it  to 
travel  from  Q  to  O,  and  c  is  the  constant  speed  of  the  ship.  The  coor- 
dinates of  the  point  P,  where  the  disturbance  created  by  the  ship  is 
to  be  found,  are  given  by 

{x  =  xl  —  r  cos  (a  +  6 ) 
z^=zl-r^n  (a  +  0) 

in  which  r  and  0  are  the  distance  and  angle  noted  on  the  figure.  In 
these  equations  we  replace  xl  and  ^  from  (8.2.18)  and  make  use  of 
(8.2.17)  to  obtain 


(8.2.20) 


a 


x  —  R  sin  a  —  -  cos2  0  cos  (a  +  6) 
2 


(8.2.21) 

We  wish  to  find  the  locus  of  points  (x*  z)  such  that  the  phase  <p  remains 


z  =  R(l  —  cos  a)  —  -  cos2  6  sin  (a  +0). 

A 


236 


WATER   WAVES 


fixed,  i.e.  such  that  the  quantity  a  in  (8.2.15)  is  constant  (cf.  the 
remarks  following  (8.2.9)).  It  is  convenient  to  introduce  the  dimen- 
sionless  parameter  K  through 

(8.2.22)  K  =  a/R. 

One  then  finds  that  the  angle  a  (cf.  (8.2.19))  is  given  by 

(8.2.23)  a  =  HCOS0, 

through  use  of  (8.2.16).  In  terms  of  these  quantities  the  relations 
(8.2.21)  can  be  put  in  the  following  dimensionless  form: 


(8.2.24) 


x/R  =  sin  (x  cos  6)  —  -  cos2  6  cos  (6  +  K  cos  0) 

z/R  =  1  —  cos  (x  cos  0)  —  -  cos2  0  sin  (0  +  «  cos  0). 

2 


These  equations  furnish  the  curves  of  stationary  phase  in  terms  of  0 
as  parameter.  Each  fixed  value  of  x  furnishes  one  such  curve,  since 
fixing  K  (for  a  fixed  turning  radius  R)  is  equivalent  to  fixing  the  phase 
(p.  In  Figure  8.2.8  a  few  curves  of  constant  phase,  as  well  as  the 


Fig.  8.2.8.  Wave  crests  for  a  circular  course 

outline  of  the  region  of  disturbance,  as  calculated  from  (8.2.24),  arc 
shown;  the  successive  curves  differ  by  2n  in  phase.  These  curves  should 
be  compared  with  the  photographs  of  actual  cases  given  in  Figures 
8.2.4  and  8.2.5.  One  sees  that  the  wave  pattern  is  given  correctly  by 
the  theory,  at  least  qualitatively.  The  agreement  between  theory  and 
observation  is  particularly  striking  in  view  of  the  manner  in  which 
the  action  of  a  ship  has  been  idealized  as  a  moving  pressure  point.  In 
particular  there  are  two  distinct  sets  of  waves  apparent,  in  conformity 
with  the  fact  that  we  expect  each  point  in  the  disturbed  region  to 


WAVE   PATTERN   CREATED   BY   A   MOVING   SHIP 


237 


correspond  to  two  influence  points:  one  set  which  seems  to  emanate 
from  the  ship's  bow,  and  another  set  which  is  arranged  roughly  at 
right  angles  to  the  ship's  course.  These  two  systems  of  waves  are  called 
the  diverging  and  the  transverse  systems,  respectively. 

From  (8.2.24)  we  can  obtain  the  more  important  case  of  the  ship 
waves  for  a  straight  course  by  letting  R  ->  oo  while  x  ->  0  in  such  a 
way  that  Rx  ->  a  (cf.  (8.2.22)).  The  result  is 


(8.2.25) 


x  =  -  (2  cos  0  —  cos3  0) 

z  —  —  --  cos2  0  sin  0 
2 


for  the  curves  of  constant  phase.  In  Figure  8.2.9  the  results  of  cal- 
culations from  these  equations  are  shown.  These  should  once  more  be 
compared  with  Figure  8.2.4,  which  shows  an  actual  case.  Again  the 
agreement  is  striking  in  a  qualitative  way.  Actually,  the  agreement 


Fig.  8.2.9.  Wave  crests  for  a  straight  course 

would  be  still  better  if  the  two  systems  of  waves —the  diverging  and 
transverse  systems— had  been  drawn  in  Figure  8.2.9  with  a  relative 
phase  difference:  the  photograph  indicates  that  the  crests  of  the  two 
systems  do  not  join  with  a  common  tangent  at  the  boundary  of  the 
region  of  disturbance.  We  shall  see  shortly  that  a  closer  examination 
of  our  approximate  solution  shows  the  two  systems  of  waves  to  have  a 
phase  difference  there.  It  is  worth  while  to  verify  in  the  present  case  a 
general  observation  made  earlier,  i.e.  that  the  curves  of  constant  phase 


238 


WATER   WAVES 


are  orthogonal  to  the  lines  drawn  back  to  the  corresponding  influence 
points.  One  finds  from  (8.2.25): 


(8.2.26) 


_  =  _  -(3sin20  -  I)sin0 
dd  2 

—  =  -(3sin20  -  I)cos0. 
dd        2 


Hence  dz/dx  =  —  I/tan  0,  which  (cf.  Figure  8.2.10)  means  that  the 
curves  of  constant  phase  are  indeed  orthogonal  to  the  lines  drawn  to 


z> 


Fig.  8.2.10.  Construction  of  curves  of  constant  phase 

the  influence  points.  The  values  0  =  0*  at  which  3  sin2  0  —  1  —  0 
are  singular  points  of  the  curves;  they  correspond  to  points  P  at  the 
boundary  of  the  influence  region  where  the  influence  points  Ql  and 
Q2  coincide.  Evidently  there  are  cusps  at  these  points.  One  sees  also 
that  the  diverging  set  of  waves  (for  z  >  0,  say)  is  obtained  when  0 
varies  in  the  range  0*  ^  0  fg  jr/2,  while  the  transverse  waves  corres- 
pond to  values  of  0  in  the  range  0^0^  0*.  In  addition,  we  observe 
that  to  any  point  on  the  ship's  course  there  corresponds  (for  0  —  0°) 
only  one  influence  point  (of  type  Q2)  and  it  does  not  coincide  with  the 
point  P.  (One  sees,  in  fact,  that  the  diverging  wave  does  not  occur  on 
the  ship's  course.)  This  is  a  fact  that  is  needed  to  justify  the  applica- 
tion of  the  method  of  stationary  phase  to  points  on  the  ship's  course, 
as  we  have  remarked  earlier  in  this  section  (cf.  also  the  preceding 
section). 

In  order  to  complete  our  discussion  we  must  consider  the  amplitude 
of  the  surface  waves,  as  given  by  our  approximation,  as  well  as  the 
shape  of  the  curves  of  constant  phase.  To  this  end  we  must  calculate 
y>  and  d2(p/dt2  (and  even  d3y>/dt*)  for  such  values  of  t  as  satisfy  the 


WAVE    PATTERN   CREATED   BY   A   MOVING   SHIP 


239 


stationary  phase  condition  dq>/dt  —  0,  as  we  know  from  the  discussion 
of  section  6.5  and  section  6.8.  From  (8.2.10)  we  find  easily 

(8.2.27)  ^?  =  £  (l  -  —  - 
v  dt2        2r  \         2r  d 

in  view  of  (8.2.11).  We  shall  also  need  the  value  of  d3<p/dt3  at  points 
such  that  dqp/dt  =  d2(p/dt2  =  0;  it  is  readily  found  to  be  given  by 

d3q>  gt2  d3r 

(8.2.28)  — £  =  —  —2  —  - 

We  wish  to  express  our  results  in  terms  of  the  parameter  0  instead  of 
t.  Since  drjdt  —  c  cos  6  from  (8.2.13)  we  have 


(8.2.29) 


.    n 

—  —  —  c  smO  — 
d*2  dt 


with  c,  the  speed  of  the  ship,  now  assumed  to  be  constant.  In  order 
to  calculate  dO/dt  we  introduce  the  angles  /?  and  r  indicated  in  Figure 
8.2.11.  We  have  0  =  n  —  (/?  +  T),  and  hence 


2> 


-t 


P(x,z) 


Fig.  8.2.11.  The  angles  ft  and  T 


in  which  s  refers  to  the  arc  length  of  C.  But  dr/ds  —-  1/18,  with  12  the 
radius  of  curvature  of  C;  and  since  /?  =  arc  tan  (z  —  z^)l(x  —  x±)  we 
find 


8.2.31 


sin  9 


d8  If/  v  d%\       /  v  dB\\       gin 

/  -  -  -      «  -  ^    -i  -  (s  -  z,)  -1  ^     =  — 
ds  r2  L  ^  dsjr 


240  WATER   WAVES 

since  the  quantity  in  the  square  brackets  is  the  vector  product  of  r 
and  t/|t|.  The  expression  for  d2qp/dt2  given  by  (8.2.27)  can  now  be 
expressed  in  terms  of  6  and  r  as  follows: 


(8.2.32)  nZ  =  .L  I  1  -  2  tan2  0  -  -  sin  01 
V           '  JJ°        2r  [  R          J 

-  3  sin2  0  -  —  sin  0  (1  -  sin2  0) 

~  /t 

2p . 

COS2  0 

as  one  can  easily  verify.  The  quantity  a  is  defined  by  (8.2.15),  and 
the  relation  (8.2.17),  in  addition  to  those  immediately  above,  has 
been  used.  The  points  on  the  boundary  of  the  region  of  disturbance 
could  be  determined  analytically,  as  follows:  the  set  of  all  influence 
points  is  the  one-parameter  family  of  circles  given  by  dy/dt  =  %(x,z,t) 
=  0,  and  the  region  of  disturbance  is  bounded  by  the  envelope  of 
these  circles,  i.e.  by  the  points  at  which  d2(p/dt2  =  d%/dt  =  0  in 
addition  to  jf  =  0.  In  the  case  of  a  straight  course  traversed  at  con- 
stant speed,  for  example,  we  .see  from  (8.2.32)  for  R  =  oo  that  0 
then  has  the  value  0*  given  by  1  —  3  sin2  0  =  0— a  result  found  above, 
where  the  value  0  =  0*  also  was  seen  to  characterize  cusps  on  the  loci 
of  constant  phase.  From  the  form  of  the  relation  (8.2.32)  one  can  con- 
clude that  the  only  courses  for  which  the  pattern  of  waves  behind  the 
ship  follows  it  without  change  (i.e.  follows  it  like  a  rigid  body)  are 
those  for  which  R  =  const.;  and  thus  only  the  straight  and  the  cir- 
cular courses  have  this  property. 

Finally,  we  have  to  consider  the  amplitude  j\(x,  z)  of  the  waves  given 
by  our  approximate  solution.  The  contribution  of  a  point  tQ  of  sta- 
tionary phase  to  (8.2.7)  is  given  by  (cf.  (6.5.2)): 

(8.2.33)  ,(«,  »)  =  y(r,  0)  I ?L_\*  X*M)±i) 

\   |?>"(r»   0)1    / 

in  which  (r,  0)  are  polar  coordinates  which  locate  the  point  of  sta- 
tionary phase  on  the  course  C  relative  to  the  point  (#,  z)  (cf.  Figure 
8.2.1).  The  sign  of  the  term  ±  n/4t  is  to  be  taken  the  same  as  that  of 
q>"  =  d2q?/dt2.  In  principle,  the  surface  elevation  can  be  calculated  for 
any  course,  but  the  results  are  not  very  tractable  except  for  the 
simplest  case;  i.e.  a  straight  course.  We  confine  our  discussion  of 
amplitudes,  therefore,  to  this  case  in  what  follows.  From  (8.2.32)  we 
have 


WAVE  PATTERN  CREATED  BY  A  MOVING  SHIP        241 


(8.2.34) 


dt*        2r  \      cos 


rin'OX 
s2  0       / 


We  know  that  there  are  two  values  of  0—  call  them  8l  and  02—  at  each 
point  in  the  disturbed  region  for  which  dcpjdt  =  0:  one  belonging  for 
0  ^  Ol  ^  0*  ~  arc  sin  l/\/8  to  the  transverse  system,  the  other  for 
0*  ^S  02  <  nl%  to  the  diverging  system  of  waves.  In  the  former  case 
d2<p/dt2  is  positive;  in  the  latter  case  negative.  (At  the  boundary  of 
the  region  of  disturbance,  where  99"  =  0,  the  formula  (8.2.33)  is  not 
valid,  as  we  know.  This  case  will  be  dealt  with  later.)  For  points  in 
the  interior  of  the  region  of  disturbance  we  have,  therefore, 

(8.2.35)  ri(x,  z) 


V|?"(r2,02)| 

Since  ri  =  £  ctt  cos  0t>  rt-  =  ^ai  cos2  0^,  at-  =  2c^p€/g  =  C2t2/2ri9  and 
y>  =  fc^J/r}  (cf.  (8.2.15),  (8.2.8))  at  the  points  of  stationary  phase, 
we  may  write  (8.2.35)  in  the  form 

(8.2.36)         \ 


I  V  1 1  -  3  sin2  0X  | 

sec3  02 
aj  V|!~3  sin202| 

The  two  systems  of  waves  are  thus  seen,  as  was  stated  above,  to  have 
a  relative  phase  difference  of  n/2  at  any  point  where  ax  =  «2.  Their 
amplitudes  die  out  like  l/\/at  on  going  away  from  the  ship,  and  that 
means  that  they  die  out  like  the  inverse  square  root  of  the  distance 
from  the  ship.  The  wave  amplitudes  of  both  systems  of  waves  become 
infinite  according  to  these  formulas  for  0  =  0*,  i.e.  for  points  at  the 
boundary  of  the  disturbed  region,  but  the  asymptotic  formula  (8.2.33) 
is  not  valid  at  such  points  since  <p"  =  0  there.  We  shall  consider  these 
points  in  a  moment.  The  diverging  system  also  has  infinite  amplitude 
for  02  =  n/29  but  this  corresponds  to  the  origin,  and  the  infinite  am- 
plitude there  results  from  our  assumption  of  a  moving  point  impulse 
as  a  model  for  our  ship. 

To  determine  the  amplitude  of  the  waves  along  the  boundary  of  the 
disturbed  region,  we  must  calculate  the  value  of  ds(p/dt3  at  such  points 
in  order  to  evaluate  the  appropriate  term  in  (6.5.2).  (The  problem  of 


242  WATER   WAVES 

the  character  of  the  waves  in  this  region  has  been  treated  by  Hogner 
[H.13].)  By  differentiating  (8.2.29)  after  replacing  dO/dt  by  c  sin  6/r 
(cf.  (8.2.30)  and  (8.2.31)  for  R  =  oo),  one  finds  readily 

(&MT)  £  -  -  *5=»-*». 

and  from  (8.2.28)  in  combination  with  r  =  \ct  cos  0,  r  —  £a  cos2  0: 

dzw       4>gc  sin2  0 

(8.2.38)  —  =  — . 

V  '  dt*         a2   cos*0 

The  amplitude  of  the  waves  along  the  boundary  of  the  disturbed 
region  is  given  by  (cf.  (6.5.2)): 


with  all  functions  evaluated  for   0  =  0*  =  arc  sin  l/\/f<*- 
result  is 

(8.2.40)  77  ~  -i  exp 


The  quantity  A^  is  a  certain  constant.  We  observe  that  the  wave  am- 
plitudes now  die  out  like  I/a1/3  instead  of  like  I/a1/2,  as  they  do  in  the 
interior  of  the  disturbed  region;  i.e.  the  wave  amplitudes  are  now  of  a 
different,  and  higher,  order  of  magnitude.  As  we  have  seen  in  all  of 
our  illustrations  of  ship  waves,  the  wave  amplitudes  are  quite  notice- 
ably higher  along  the  boundary  of  the  disturbed  region.  The  phase 
also  differs  now  by  n/4>  from  the  former  values.  On  some  of  the  photo- 
graphs (cf.  especially  Fig.  8.2.4),  there  is  some  evidence  of  a  rather 
abrupt  change  of  phase  in  the  region  of  the  boundary,  though  it  may 
be  that  one  should  interpret  this  effect  as  due  rather  to  the  finite 
dimensions  of  the  actual  ship,  which  then  acts  as  though  several 
moving  point  sources  were  acting  simultaneously. 

In  the  treatment  of  the  present  problem  by  A.  S.  Peters  [P.4] 
mentioned  in  the  preceding  section,  the  complete  asymptotic  develop- 
ment of  the  solution  was  obtained. 

The  above  developments  hold  only  for  the  case  of  a  point  impulse 
moving  on  the  surface  of  water  of  infinite  depth.  It  has  some  interest 
to  point  out  that  there  are  considerable  differences  in  the  results  if 
the  depth  of  the  water  is  finite.  Havelock  [H.8]  has  carried  out  the 
approximation  to  the  solution  by  the  method  of  stationary  phase  for 


WAVE    PATTERN    CREATED    BY    A    MOVING    SHIP 


243 


the  case  of  constant  finite  depth,  with  the  following  general  results: 
1 )  If  the  speed  c  of  the  ship  and  the  depth  h  satisfy  the  inequality 
c2/gh  <  1,  the  general  pattern  of  the  waves  is  much  the  same  as  for 
water  of  infinite  depth  except  that  the  angle  of  the  sector  within 
which  the  main  part  of  the  disturbance  is  found  is  now  larger  than 
for  water  of  infinite  depth.  2)  If  c2/gk  >  I  holds,  the  system  of  trans- 
verse waves  no  longer  occurs,  but  the  diverging  system  is  found. 


Fig.  8.2.12.  Speed  boat  in  shallow  water 

Figure  8.2.12  is  a  photograph  of  a  speed  boat  creating  waves,  presum- 
ably in  shallow  water,  in  view  of  the  difference  in  the  wave  pattern 
when  compared  with  Fig.  8.2.4.  Finally,  if  c2/gh  =  1  (i.e.  for  the  case 
of  the  critical  speed),  the  method  of  stationary  phase  yields  no  rea- 
sonable results;  that  this  should  be  so  is  perhaps  to  be  understood  in 
the  light  of  the  discussion  of  the  corresponding  two-dimensional 
problem  in  Chapter  7.4. 


CHAPTER  9 

The  Motion  of  a  Ship,  as  a  Floating  Rigid  Body,  in  a 

Seaway 

9.1.  Introduction  and  summary 

The  purpose  of  this  chapter  is  to  develop  a  mathematical  theory  for 
the  motion  of  a  ship,  to  be  treated  as  a  freely  floating  rigid  body  under 
the  action  of  given  external  forces  (a  propeller  thrust,  for  example), 
under  the  most  general  conditions  compatible  with  a  linear  theory  and 
the  assumption  of  an  infinite  ocean.*  This  of  course  requires  the 
amplitude  of  the  surface  waves  to  be  small  and,  in  general,  that  the 
motion  of  the  water  should  be  a  small  oscillation  near  its  rest  position 
of  equilibrium;  it  also  requires  the  ship  to  have  the  shape  of  a  thin 
disk  so  that  it  can  have  a  translatory  motion  with  finite  velocity  and 
still  create  only  small  disturbances  in  the  water.  In  addition,  the  mo- 
tion of  the  ship  itself  must  be  assumed  to  consist  of  small  oscillations 
relative  to  a  motion  of  translation  with  constant  velocity.  Within 
these  limitations,  however,  the  theory  presented  is  quite  general  in 
the  sense  that  no  arbitrary  assumptions  about  the  interaction  of  the 
ship  with  the  water  are  made,  nor  about  the  character  of  the  coupling 
between  the  different  degrees  of  freedom  of  the  ship,  nor  about  the 
waves  present  on  the  surface  of  the  sea:  the  combined  system  of  ship 
and  sea  is  treated  by  using  the  basic  mathematical  theory  of  the 
hydrodynamics  of  a  non- turbulent  perfect  fluid.  For  example,  the 
theory  presented  here  would  make  it  possible  in  principle  to  deter- 
mine the  motion  of  a  ship  under  given  forces  which  is  started  with 
arbitrary  initial  conditions  on  a  sea  subjected  to  given  surface  pres- 
sures and  initial  conditions,  or  on  a  sea  covered  with  waves  of  pre- 
scribed character  coming  from  infinity. 

It  is  of  course  well  known  that  such  a  linear  theory  for  the  non- 
turbulent  motion  of  a  perfect  fluid,  complicated  though  it  is,  still  does 
not  contain  all  of  the  important  elements  needed  for  a  thoroughgoing 
discussion  of  the  practical  problems  involved.  For  example,  it  ignores 

*  The  presentation  of  the  theory  given  here  is  essentially  the  same  as  that 
given  in  a  report  of  Peters  and  Stoker  [P.7]. 

245 


246 


WATER   WAVES 


the  boundary-layer  effects,  turbulence  effects,  the  existence  in  general 
of  a  wake,  and  other  important  effects  of  a  non-linear  character.  Good 
discussions  of  these  matters  can  be  found  in  papers  of  Lunde  and  Wig- 
ley  [L.I 8],  and  Havelock  [H.7],  Nevertheless,  it  seems  clear  that  an 
approach  to  the  problem  of  predicting  mathematically  the  motion  of 
ships  in  a  seaway  under  quite  general  conditions  is  a  worthwhile  enter- 
prise, and  that  the  problem  should  be  attacked  even  though  it  is 
recognized  at  the  outset  that  all  of  the  important  physical  factors  can 
not  be  taken  into  account.  In  fact,  the  theory  presented  here  leads  at 
once  to  a  number  of  important  qualitative  statements  without  the 
necessity  of  producing  actual  solutions— for  example,  we  shall  see 
that  certain  resonant  frequencies  appear  quite  naturally,  and  in 
addition  that  they  can  be  calculated  solely  with  reference  to  the  mass 
distribution  and  the  given  shape  of  the  hull  of  the  ship.  Interesting 
observations  about  the  character  of  the  coupling  between  the  various 
degrees  of  freedom,  and  about  the  nature  of  the  interaction  between 
the  ship  and  the  water,  are  also  obtained  simply  by  examining  the 
equations  which  the  theory  yields. 

In  order  to  describe  the  theory  and  results  to  be  worked  out  in 
later  sections  of  this  chapter,  it  is  necessary  to  introduce  our  notation 
and  to  go  somewhat  into  details.  In  Fig.  9.1.1  the  disposition  of  two  of 
the  coordinate  systems  used  is  indicated.  The  system  (X,  Y,  Z)  is  a 

AY 


-*.  X 


'2  * 

Fig.  9.1.1.  Fixed  and  moving  coordinate  systems 

system  fixed  in  space  with  the  X,  Z-plane  in  the  undisturbed  free 
surface  of  the  water  and  the  F-axis  vertically  upward.  A  moving 
system  of  coordinates  (x,  y,  z)  is  introduced;  in  this  system  the  #,  z- 
plane  is  assumed  to  coincide  always  with  the  X,  Z-plane,  and  the 
t/-axis  is  assumed  to  contain  the  center  of  gravity  (abbreviated  to  e.g. 
in  the  following)  of  the  ship.  The  course  of  the  ship  is  fixed  by  the 
motion  of  the  origin  of  the  moving  system,  and  the  #-axis  is  taken  along 


THE   MOTION   OF  A   SHIP   IN   A   SEAWAY 


247 


the  tangent  to  the  course.  It  is  then  convenient  to  introduce  the 
speed  s(t)  of  the  ship  in  its  course:  the  speed  s(t)  is  simply  the  magni- 
tude of  the  vector  representing  the  instantaneous  velocity  of  this  point. 
At  the  same  time  we  introduce  the  angular  speed  co(t)  of  the  moving 
system  relative  to  the  fixed  system:  one  quantity  fixes  this  rotation 
because  the  vertical  axes  remain  always  parallel.  The  angle  oc(£) 
indicated  in  Fig.  9.1.1  is  defined  by 


(9.1.1) 


0)(t)  (It, 


implying  that  t  —  0  corresponds  to  an  instant  when  the  #-axis  and 
JT-axis  are  parallel.  In  order  to  deal  with  the  motion  of  the  ship  as  a 
rigid  body  it  is  convenient,  as  always,  to  introduce  a  system  of  coor- 
dinates fixed  in  the  body.  Such  a  system  (#',  y\  z')  is  indicated  in 
Fig.  D.I. 2.  The  #',  t/'-plane  is  assumed  to  be  in  the  fore-and-aft  plane 


(a) 


(b) 


Fig.  0.1.  2a,  b.  Another  moving  coordinate  system 

of  symmetry  of  the  ship's  hull,  and  the  ?/'-axis  is  assumed  to  contain 
the  e.g.  of  the  ship.  The  moving  system(tr',  j/',  z')  is  assumed  to  coin- 
cide with  the  (iT,  j/,  z)  system  when  the  ship  and  the  water  are  at  rest 
in  their  equilibrium  positions.  The  e.g.  of  the  ship  will  thus  coincide 
with  the  origin  of  the  (#',  j/',  z')  system  only  in  case  it  is  at  the  level 
of  the  equilibrium  water  line  on  the  ship;  we  therefore  introduce  the 
constant  y'c  as  the  vertical  coordinate  of  the  e.g.  in  the  primed  coor- 
dinate system. 

The  motion  of  the  water  is  assumed  to  be  given  by  a  velocity  poten- 
tial 0(X,  F,  Z;  t)  which  is  therefore  to  be  determined  as  a  solution 
of  Laplace's  equation  satisfying  appropriate  boundary  conditions  at 
the  free  surface  of  the  water,  on  the  hull  of  the  ship,  at  infinity,  and 
also  initial  conditions  at  the  time  t  =  0.  The  boundary  conditions  on 
the  hull  of  the  ship  clearly  will  depend  on  the  motion  of  the  ship, 


248  WATER   WAVES 

which  in  its  turn  is  fixed,  through  the  differential  equations  for  the 
motion  of  a  rigid  body  with  six  degrees  of  freedom,  by  the  forces  acting 
on  it  —including  the  pressure  of  the  water— and  its  position  and  veloc- 
ity at  the  time  t  =  0.  As  was  already  stated,  no  further  restrictive 
assumptions  except  those  needed  to  linearize  the  problem  are  made. 
Before  discussing  methods  of  linearization  we  interpolate  a  brief 
discussion  of  the  relation  of  the  theory  presented  here  to  that  of  other 
writers  who  have  discussed  the  problem  of  ship  motions  by  means  of 
the  linear  theory  of  irrotational  waves.  The  subject  has  a  lengthy 
history,  beginning  with  Michell  in  1898,  and  continuing  over  a  long 
period  of  years  in  a  sequence  of  notable  papers  by  Havelock,  starting 
m  1909.  This  work  is,  of  course,  included  as  a  special  case  in  what  is 
presented  here.  Extensive  and  up-to-date  bibliographies  can  be  found 
in  the  papers  by  Weinblum  [W.3]  and  Lunde  [L.19].  Most  of  this  work 
considers  the  ship  to  be  held  fixed  in  space  while  the  water  streams 
past;  the  question  of  interest  is  then  the  calculation  of  the  wave 
resistance  in  its  dependence  on  the  form  of  the  ship.  Of  particular 
interest  to  us  here  are  papers  of  Krylov  [K.20],  St.  Denis  and  Wein- 
blum [S.I],  Pierson  and  St.  Denis  [P.9]  and  Haskind  [H.4],  all  of 
whom  deal  with  less  restricted  types  of  motion.  Krylov  seeks  the 
motion  of  the  ship  on  the  assumption  that  the  pressure  on  its  hull 
is  fixed  by  the  prescribed  motion  of  the  water  without  reference  to 
the  back  effect  on  the  motion  of  the  water  induced  by  the  motion 
of  the  ship.  St.  Denis  and  Weinblum,  and  Pierson  and  St.  Denis, 
employ  a  combined  theoretical  and  empirical  approach  to  the  prob- 
lem which  involves  writing  down  equations  of  motion  of  the  ship 
with  coefficients  which  should  be  in  part  determined  by  model  ex- 
periments; it  is  assumed  in  addition  that  there  is  no  coupling  be- 
tween the  different  degrees  of  freedom  involved  in  the  general  mo- 
tion of  the  ship.  Haskind  attacks  the  problem  in  the  same  degree 
of  generality,  and  under  the  same  general  assumptions,  as  are  made 
here;  in  the  end,  however,  Haskind  derives  his  theory  completely  only 
in  a  certain  special  case.  Haskind 's  theory  is  also  not  the  same  as  the 
theory  presented  here,  and  this  is  caused  by  a  fundamental  difference 
in  the  procedure  used  to  derive  the  linear  theory  from  the  underlying, 
basically  nonlinear,  theory.  Haskind  develops  his  theory  by  assuming 
that  he  knows  a  priori  the  relative  orders  of  magnitude  of  the  various 
quantities  involved.  The  problem  is  attacked  in  this  chapter  by  a 
formal  development  with  respect  to  a  small  parameter  (essentially  a 
thickness-length  ratio  of  the  ship);  in  doing  so  every  quantity  is 


THE   MOTION   OF   A   SHIP   IN   A   SEAWAY  249 

developed  systematically  in  a  formal  series  (for  a  similar  type  of 
discussion  see  F.  John  [J.5]).  In  this  way  a  correct  theory  should  be 
obtained,  assuming  the  convergence  of  the  series— and  there  would 
seem  to  be  no  reason  to  doubt  that  the  series  would  converge  for 
sufficiently  small  values  of  the  parameter.  Aside  from  the  relative 
safety  of  such  a  method— purchased,  it  is  true,  at  the  price  of  making 
rather  bulky  calculations  — it  has  an  additional  advantage,  i.e.,  it 
makes  possible  a  consistent  procedure  for  determining  any  desired 
higher  order  corrections.  It  is  not  easy  to  compare  Haskind's  theory 
in  detail  with  the  theory  presented  here.  However,  it  can  be  stated 
that  certain  terms,  called  damping  terms  by  Haskind,  are  terms  that 
would  be  of  higher  order  than  any  of  those  retained  here.  A  more 
precise  statement  on  this  point  will  be  made  later. 

One  of  the  possible  procedures  for  linearizing  the  problem  begins 
by  writing  the  equation  of  the  hull  of  the  ship  relative  to  the  coordinate 
system  fixed  in  the  ship  in  the  form 

(9.1.2)  z'  -  ±  0h(x'9  y')9         z'>0, 

with  ft  a  small  dimensionless  parameter.*  This  is  the  parameter  with 
respect  to  which  all  quantities  will  be  developed.  In  particular,  the 
velocity  potential  0(X,  F,  Z;  /;  ft)  =.  <p(x*  y,  z;  t;  ft)  is  assumed  to 
possess  the  development 

(9.1.8)  <p(x9  y,  z;  t;  ft)  =  fafa  y,  z;  t)  +  ft*<p2(x,  &*;<)+•••  • 
The  free  surface  elevation  r](x9  z;  t;  ft)  and  the  speed  s(t;  ft)  and  angu- 
lar velocity  a>(t;  ft)  (cf.  (9.1.1 ))  are  assumed  to  have  the  developments 

(9.1.4)  ri(x,  z\  t;  ft)  =  /%(*,  z;  t)  +  ft*r]2(x.  z;t)  +  ...  9 

(9.1.5)  s(t;  ft)  ==  So(t)  +  ftSl(t)  +  ...  , 

(9.1.6)  o>(f;  ft)  =  o>0(0  +  ftco^t)  +  .  .  .  . 

Finally,  the  vertical  displacement  yc(t)  of  the  center  of  gravity  and 
the  angular  displacements**  01,  02,  03  of  the  ship  with  respect  to  the 
#,  t/,  and  z  axes  respectively  are  assumed  given  by 

*  It  is  important  to  consider  other  means  of  linearization,  and  we  shall  discuss 
some  of  them  later.  However,  it  should  be  said  here  that  the  essential  point  is 
that  a  linearization  can  be  made  for  any  body  having  the  form  of  a  thin  disk: 
it  is  not  at  all  essential  that  the  plane  of  the  disk  should  be  assumed  to  be  vertical, 
as  we  have  done  in  writing  equation  (9.1.2). 

**  Since  we  consider  only  small  displacements  of  the  ship  relative  to  a  uniform 
translation,  it  is  convenient  to  assume  at  the  outset  that  the  angular  displacement 
can  be  given  without  ambiguity  as  a  vector  with  the  components  0lf  0f,  08  relative 
to  the  #,  t/,  2-coordinate  system. 


250  WATER    WAVES 


(9.1.7)  OM  ft)  =  ftO^t)  +  £»0<i(0  +  •  •  -  >         i  =  1,2,3, 

(9.1.8)  yc(t;  ft)  -  y'e  =  ftVi(t) 


These  relations  imply  that  the  velocity  of  the  water  and  the 
elevation  of  its  free  surface  are  small  of  the  same  order  as  the  "slender- 
ness  parameter"  ft  of  the  ship.  On  the  other  hand,  the  speed  s(t)  of  the 
ship  is  assumed  to  be  of  zero  order.  The  other  quantities  fixing  the 
motion  of  the  ship  are  assumed  to  be  of  first  order,  except  for  co(t), 
but  it  turns  out  in  the  end  that  coQ(t)  vanishes  so  that  a)  is  also  of  first 
order.  The  quantity  y'c  in  (9.1.8)  was  defined  in  connection  with  the 
description  of  Fig.  9.1.2;  it  is  to  be  noted  that  we  have  chosen  to 
express  all  quantities  with  respect  to  the  moving  coordinate  system 
(x,  y,  z)  indicated  in  that  figure.  The  formulas  for  changes  of  coordi- 
nates must  be  used,  and  they  also  are  to  be  developed  in  powers  of 
ft;  for  example,  the  equation  of  the  hull  relative  to  the  (x,  y,  z)  co- 
ordinate system  is  found  to  be 


2l 


x  -  ftOu(y  -  y'c)  -  fth(x,  y)  +  .  .  .  =  o 


after  developing  and  rejecting  second  and  higher  order  terms  in  ft. 
In  marine  engineering  there  is  an  accepted  terminology  for  describ- 
ing the  motion  of  a  ship;  we  wish  to  put  it  into  relation  with  the  no- 
tation just  introduced.  In  doing  so,  the  case  of  small  deviations  from 
a  straight  course  is  the  only  one  in  question.  The  angular  displace- 
ments are  named  as  follows:  Ol  is  the  rolling,  02  +  a  is  the  yawing,  and 
03  is  the  pitching  oscillation.  The  quantity  ft#i(t)  in  (9.1.5)  is  called 
the  surge  (i.e.,  it  is  the  small  forc-arid-aft  motion  relative  to  the  finite 
speed  s0(t)  of  the  ship,  which  turns  out  to  be  necessarily  a  constant), 
while  yc  —  y'c  fixes  the  heave.  In  addition  there  is  the  sidewisc  dis- 
placement dz  referred  to  as  the  sway;  this  quantity,  in  lowest  order, 
can  be  calculated  in  terms  of  sQ(t)  and  the  angle  a  defined  by  (9.1.1) 
in  terms  of  co(t)  as  follows: 


(9.1.9)  dz 


f* 

*,,«  S0J^( 


since  a>Q(t)  turns  out  to  vanish. 

In  one  of  the  problems  of  most  practical  interest,  i.e.  the  problem 
of  a  ship  that  has  been  moving  for  a  long  time  (so  that  all  transients 
have  disappeared)  under  a  constant  propeller  thrust  (considered  to  be 
simply  a  force  of  constant  magnitude  parallel  to  the  keel  of  the  ship) 


THE   MOTION    OF   A   SHIP    IN   A   SEAWAY  251 

into  a  seaway  consisting  of  a  given  system  of  simple  harmonic  progres- 
sing waves  of  given  frequency,  one  expects  that  the  displacement  com- 
ponents would  in  general  be  the  sum  of  two  terms,  one  independent  of 
the  time  and  representing  the  displacements  that  would  arise  from 
motion  with  uniform  velocity  through  a  calm  sea,  the  other  a  term 
simple  harmonic  in  the  time  that  has  its  origin  in  the  forces  arising 
from  the  waves  coming  from  infinity.  On  account  of  the  symmetry  of 
the  hull  only  two  displacements  of  the  first  category  would  differ 
from  zero:  one  the  vertical  displacement,  i.e.  the  heave,  the  other  the 
pitching  angle,  i.e.  the  angle  03.  The  latter  two  displacements  apparent- 
ly are  referred  to  as  the  trim  of  the  ship.  In  all,  then,  there  would  be 
in  this  case  nine  quantities  to  be  fixed  as  far  as  the  motion  of  the  ship 
is  concerned:  the  amplitudes  of  the  oscillations  in  each  of  the  six 
degrees  of  freedom,  the  speed  sQ9  and  the  two  quantities  determining 
the  trim.  A  procedure  to  determine  all  of  them  will  next  be  outlined. 
We  proceed  to  give  a  summary  of  the  theory  obtained  when  the 
scries  (9.1.2)  to  (9.1.8)  are  inserted  in  all  of  the  equations  fixing  the 
motion  of  the  system,  which  includes  both  the  differential  equations 
and  the  boundary  conditions,  and  any  functions  involving  ft  are  in 
turn  developed  in  powers  of  ft.  For  example,  one  needs  to  evaluate  (px 
on  the  free  surface  y  —  rj  in  order  to  express  the  boundary  conditions 
there;  one  calculates  it  as  follows  (using  (9.1.3)  and  (9.1.4)): 

(9.1.10)    Vm(x9  TI,  z;  t;  ft)  =  ft[<pl9(x,  0,  a;  *)  +  rpp^(x,  0,  z;  t)  +  ...] 


x,  0,  z;  t)  +  ^fotfW*  °'  z'>  0 


We  observe  the  important  fact—  to  which  reference  will  be  made 
later  —  that  the  coefficients  of  the  powers  of  ft  are  evaluated  at  y  =  0, 
i.e.  at  the  undisturbed  equilibrium  position  of  the  free  surface  of  the 
water.  In  the  same  way,  it  turns  out  that  the  boundary  conditions 
for  the  hull  of  the  ship  arc  automatically  to  be  satisfied  on  the  vertical 
longitudinal  mid-section  of  the  hull.  The  end  result  of  such  calcula- 
tions, carried  out  in  such  a  way  as  to  include  all  terms  of  first  order  in 
ft  is  as  follows:  The  differential  equation  for  (p±  is,  of  course,  the  La- 
place equation: 

(9.1.11)  Vix* 


in  the  domain  y  <  0,  i.e.  the  lower  half-space,  excluding  the  plane 
area  A  of  the  x,  t/-plane  which  is  the  orthogonal  projection  of  the 


252  WATER   WAVES 

hull  (cf.  Fig.  9.1.  2b),  in  its  equilibrium  position,  on  the  x9  r/-plane. 
The  boundary  conditions  on  9^  are 


(9  1  12) 

02i)  -  K  +  02i)<*  +  Ou(y  ~  2/c)»  on     A- 

in  which  A+  and  .^__  refer  to  the  two  sides  z  =  0+  and  2  =  0_  of  the 
plane  disk  A.  The  boundary  conditions  on  the  free  surface  are 

(9.1.13) 

The  first  of  these  results  from  the  condition  that  the  pressure  vanishes 
on  the  free  surface,  the  second  arises  from  the  kinematic  free  surface 
condition.  If  sQ9  col9  021,  and  0n  were  known  functions  of  t,  these  boun- 
dary conditions  in  conjunction  with  (9.1.11)  and  appropriate  initial 
conditions  would  serve  to  determine  the  functions  cpl  and  rjl  uniquely; 
i.e.  the  velocity  potential  and  the  free  surface  elevation  would  be 
known.  Of  course,  the  really  interesting  problems  for  us  here  are  those 
in  which  the  quantities  s0,  col9  021,  and  Oll9  referring  to  the  motion  of 
the  ship,  are  not  given  in  advance  but  are  rather  unknown  functions 
of  the  time  to  be  determined  as  part  of  the  solution  of  the  boundary 
problem.  In  principle,  one  method  of  approach  would  be  to  apply  the 
Laplace  transform  with  respect  to  the  time  t  to  (9.1  .11),  (9.1.12),  and 
(9.1.13)—  of  course  taking  account  of  initial  conditions  at  the  time 
/  =  0—  and  then  to  solve  the  resulting  boundary  value  problem 
for  the  transform  q>i(x9  y9  z;  a)  regarding  s0  and  the  transforms  oi^cr), 
02i((T)>  a]Qd  fliifa)  as  parameters.  However,  for  the  purposes  of  this 
introduction  it  is  better  to  concentrate  on  the  most  important  special 
case  (already  mentioned  above)  in  which  the  ship  has  a  motion  of 
translation  with  uniform  speed  combined  with  small  simple  harmonic 
oscillations  of  the  ship  and  the  sea  having  the  same  frequency.*  In 
this  case  we  write  the  velocity  potential  ^(tf,  j/,  z;  t)9  the  surface 
elevation  rjl9  and  the  other  dependent  quantities  in  the  form 

(<?!(#,  y9  z;  t)  =  \pt(x9  y9  z)  +  ^(x9  y9  z)eiat 
^(a?,  z;  t)  =  //„(*,  z)  +  //!<#,  z)eM 
a*!  -  Q^\  0n  -  eneM9  0ai  -  e^eM. 

The  functions  y;0  and  ^  are  of  course  both  harmonic  functions.  We 
expect  the  functions  g^  and  rjl  to  have  time-independent  components 

*  It  can  be  seen,  however,  that  the  discussion  which  follows  would  take  much 
the  same  course  if  more  general  motions  were  to  be  assumed. 


THE    MOTION    OF   A   SHIP    IN    A    SEAWAY  253 

due  to  the  forward  motion  of  the  ship;  certainly  they  would  appear 
in  the  absence  of  any  oscillatory  components  due,  say,  to  a  wave  train 
in  the  sea.  Upon  insertion  of  these  expressions  in  equations  (9.1.12), 
and  (9.1.13)  we  find  for  y>0  the  conditions: 

on  A±9 

at  y  =  0, 


and  for  y>i  the  conditions 

!  —  y)lz  —  —  s0021  -\-(Ql-\-io02l)x  —  io&n(y—y')  on  A± 
. 
—  g//!  +  *oVi.r  —  *aVi  -'-  Q  I 

We  observe,  in  passing,  that  y>0  satisfies  the  same  boundary  conditions 
as  in  the  classical  Michcll-Havelock  theory.  A  little  later  we  shall  see, 
in  fact,  that  the  wave  resistance  is  indeed  independent  of  all  compo- 
nents of  the  motion  of  the  ship  (to  lowest  order  in  /3,  that  is)  except  its 
uniform  forward  motion  with  speed  s0,  and  that  the  wave  resistance 
is  determined  in  exactly  the  same  way  as  in  the  Michell-Havelock 
theory.  We  continue  the  description  of  the  equations  which  determine 
the  motion  of  the  ship,  and  which  arise  from  developing  the  equations 
of  motion  with  respect  to  ft  and  retaining  only  the  terms  of  order  /? 
and  /ft2.  (We  observe  that  it  is  necessary  to  consider  terms  of  both 
orders.)  In  doing  so  the  mass  M  of  the  ship  is  given  by  M  =  M^9 
with  Ml  a  constant,  since  we  assume  the  average  density  of  the  ship 
to  be  finite  and  its  volume  is  of  course  of  order  0.  The  moments  of 
inertia  are  then  also  of  order  (3.  The  propeller  thrust  is  assumed  to  be 
a  force  of  magnitude  T  acting  in  the  ^'-direction  and  in  the  x'9  y'- 
plane  at  a  point  whose  vertical  distance  from  the  e.g.  is  —  /;  the  thrust 
T  is  assumed  to  be  of  order  /?2,  since  the  mass  is  of  order  /?  and  accelera- 
tions are  also  of  order  /?.*  The  propeller  thrust  could  also,  of  course, 
be  called  the  wave  resistance. 

The  terms  of  order  /?  yield  the  following  conditions: 

*  We  have  in  mind  problems  in  which  the  motion  of  the  ship  is  a  small  deviation 
from  a  translatory  motion  with  uniform  finite  speed.  If  it  were  desired  to  study 
motions  with  finite  accelerations  —  as  would  be  necessary,  for  example,  if  the 
ship  were  to  be  considered  as  starting  from  rest  —  it  would  clearly  be  necessary 
to  suppose  the  development  of  the  propeller  thrust  T  to  begin  with  a  term  of  first 
order  in  /?,  since  the  mass  of  the  ship  is  of  this  order.  In  that  case,  the  motion 
of  the  ship  at  finite  speed  and  acceleration  would  be  determined  independently 
of  the  motion  of  the  water:  in  other  words,  it  would  be  conditioned  solely  by  the 
inert  mass  of  the  ship  and  the  thrust  of  order  0. 


254  WATER   WAVES 

(9.1.15)  *0  =  0, 

(9.1.16)  2eg  f    fihdA  -  MJg, 

J  A 

(9.1.17)  f    xfihdA  =  0, 

J  A 

(9.1.18)  f     [(Vli  -  Wlc)]+  <L4  -  0, 

J  x 

(9.1.19)  f     [a(Vlf  -  *tf>lx)]*  cW  -  0, 

J  A 

(9.1.20)  f     [i/^,  -  8<fflx)-]+  dA  =  0. 

J  A 

The  symbol  [  ]*  occurring  here  means  that  the  jump  in  the  quan- 
tity in  brackets  on  going  from  the  positive  to  the  negative  side  of  the 
projected  area  A  of  the  ship's  hull  is  to  be  taken.  The  variables  of  in- 
tegration are  x  and  y.  The  equation  (9.1.15)  states  that  the  term  of 
order  zero  in  the  speed  is  a  constant,  and  hence  the  motion  in  the 
^-direction  is  a  small  oscillation  relative  to  a  motion  with  uniform 
velocity.  (This  really  comes  about  because  we  assume  the  propeller 
thrust  T  to  be  of  order  /?2.)  Equation  (9.1.16)  is  an  expression  of  the 
law  of  Archimedes:  the  rest  position  of  equilibrium  must  be  such  that 
the  weight  of  the  ship  just  equals  the  weight  of  the  water  it  displaces. 
Equation  (9.1.17)  expresses  another  law  of  equilibrium  of  a  floating 
body,  i.e.  that  the  center  of  buoyancy  should  be  on  the  same  vertical 
line  as  the  center  of  gravity  of  the  ship.  The  remaining  three  equations 
(9.1.18),  (9.1.19),  and  (9.1.20)  in  the  group  serve  to  determine  the  dis- 
placements 0U,  021,  and  cov  which  occur  in  the  boundary  condition 
(9.1.12)  for  the  velocity  potential  9^.  In  the  special  case  we  consider 
(cf.  (9.1.14))  we  observe  that  these  three  equations  would  determine 
the  values  of  the  constants  Q19  0n,  and  0zl  (the  complex  amplitudes 
of  certain  displacements  of  the  ship)  which  occur  as  parameters  in  the 
boundary  conditions  for  the  harmonic  function  tpi(x,  y,  z)  given  in 
(9.1.14)!. 

We  are  now  able  to  draw  some  interesting  conclusions.  Once  the 
speed  s0  is  fixed,  it  follows  that  the  problem  of  determining  the  har- 
monic function  (jp1  is  completely  formulated  through  the  equations 
(9.1.14),  (9.1.14)0,  (9.1.14)!,  and  (9.1.18)  to  (9.1.20)  inclusive  (to- 
gether with  appropriate  conditions  at  oo).  In  other  words,  the  motion 
of  the  water,  which  is  fixed  solely  by  <pv  is  entirely  independent  of  the 


THE    MOTION   OF   A   SHIP   IN   A   SEAWAY  255 

pitching  displacement  031(0»  the  heave  yi(t)9  and  the  surge  s^t),  i.e. 
of  all  displacements  in  the  vertical  plane  except  the  constant  forward 
speed  SQ.  A  little  reflection,  however,  makes  this  result  quite  plausible: 
Our  theory  is  based  on  the  assumption  that  the  ship  is  a  thin  disk 
disposed  vertically  in  the  water,  whose  thickness  is  a  quantity  of 
first  order.  Hence  only  finite  displacements  of  the  disk  parallel  to 
this  vertical  plane  could  create  oscillations  in  the  water  that  are  of 
first  order.  On  the  other  hand,  displacements  of  first  order  of  the  disk 
at  right  angles  to  itself  will  create  motions  in  the  water  that  are  also 
of  first  order.  One  might  seek  to  describe  the  situation  crudely  in  the 
following  fashion.  Imagine  a  knife  blade  held  vertically  in  the  water. 
Up-and-down  motions  of  the  knife  evidently  produce  motions  of  the 
water  which  arc  of  a  quite  different  order  of  magnitude  from  motions 
produced  by  displacements  of  the  knife  perpendicular  to  the  plane  of 
its  blade.  Stress  is  laid  on  this  phenomenon  here  because  it  helps  to 
promote  understanding  of  other  occurrences  to  be  described  later. 
The  terms  of  second  order  in  /?  yield,  finally,  the  following  conditions: 

(9.1.21) 

A/!*!  -  —  p         hx[(<Pit  -  s0<plxY  +  (<plt 
J  A 

(9.1.22) 

^i^~%f     (2/i+^31)Mr-p     hy[(q'it 
JL  JA 

(9.1.23) 

'si^ai^ -2eg°3i|    (y-Vc)h*A-lQSyA   xhdx 

-2gg031f  ,r2Mr+/T 

J  L 

-Q\    \^hv-(y~yfc)hx][((plt~s^lx)+  +  ((plt- 

J  A 

We  note  that  integrals  over  the  projected  water-line  L  of  the  ship  on 
the  vertical  plane  when  in  its  equilibrium  position  occur  in  addition 
to  integrals  over  the  vertical  projection  A  of  the  entire  hull.  The 
quantity  /31  arises  from  the  relation  /  =  /?/31  for  the  moment  of 
inertia  /  of  the  ship  with  respect  to  an  axis  through  its  e.g.  parallel 
to  the  s'-axis.  The  equation  (9.1.21 )  determines  the  surge  sv  and  also 
the  speed  s0  (or,  if  one  wishes,  the  thrust  T  is  determined  if  SQ  is 


256  WATER  WAVES 

assumed  to  be  given).  Furthermore,  the  speed  s0  is  fixed  solely  by  T 
and  the  geometry  of  the  ship's  hull.  This  can  be  seen,  with  reference 
to  (9.1.14)  and  the  discussion  that  accompanies  it,  in  the  following 
way:  The  term  y0(#,  y,  z)  in  (9.1.14)  is  the  term  in  q>^  that  is  indepen- 
dent of  t.  It  therefore  determines  T  upon  insertion  of  ^  in  (9.1.21). 
This  term,  however,  is  obtained  by  finding  the  harmonic  function 
y0  as  a  solution  of  the  boundary  problem  for  \pQ  formulated  in  (9.1.14)0. 
In  fact,  the  relation  between  s0  and  T  is  now  seen  to  be  exactly  the 
same  relation  as  was  obtained  by  Michell.  (It  will  be  written  down  in 
a  later  section. )  In  other  words,  the  wave  resistance  depends  only  on 
the  basic  translatory  motion  with  uniform  speed  of  the  ship,  and  not 
at  all  on  its  small  oscillations  relative  to  that  motion.  If,  then,  effects 
on  the  wave  resistance  due  to  the  oscillation  of  the  ship  are  to  be 
obtained  from  the  theory,  it  will  be  necessary  to  take  account  of  higher 
order  terms.  Once  the  thrust  T  has  been  determined  the  equations 
(9.1.22)  and  (9.1.23)  form  a  coupled  system  for  the  determination  of 
yl  and  031,  since  9^  and  0n  have  presumably  been  determined  previous- 
ly. Thus  our  system  is  one  in  which  there  is  a  considerable  amount  of 
cross-coupling.  It  might  also  be  noted  that  the  trim,  i.e.  the  constant 
values  of  yl  and  031  about  which  the  oscillations  in  these  degrees  of 
freedom  occur  are  determined  from  (9.1.22)  and  (9.1.23)  by  the  time- 
independent  terms  in  these  equations —  including,  for  example,  the 
moment  IT  of  the  thrust  about  the  e.g. 

We  proceed  to  the  discussion  of  other  conclusions  arising  from  our 
developments  and  concerning  two  questions  which  recur  again  and 
again  in  the  literature.  These  issues  center  around  the  question:  what 
is  the  correct  manner  of  satisfying  the  boundary  conditions  on  the 
curved  hull  of  the  ship?  Michell  employed  the  condition  (9.1.12), 
naturally  with  0n  =  021  =  coj  =  0,  on  the  basis  of  the  physical  argu- 
ment that  s0hx  represents  the  component  of  the  velocity  of  the  water 
normal  to  the  hull,  and  since  the  hull  is  slender,  a  good  approximation 
would  result  by  using  as  boundary  condition  the  jump  condition 
furnished  by  (9.1.12).  Havelock  and  others  have  usually  followed  the 
same  practice.  However,  one  finds  constant  criticism  of  the  resulting 
theory  in  the  literature  (particularly  in  the  engineering  literature) 
because  of  the  fact  that  the  boundary  condition  is  not  satisfied  at  the 
actual  position  of  the  ship's  hull,  and  various  proposals  have  been 
made  to  improve  the  approximation.  This  criticism  would  seem 
to  be  beside  the  point,  since  the  condition  (9.1.12)  is  simply  the  con- 
sequence of  a  reasonable  linearization  of  the  problem.  To  take  account 


THE    MOTION    OF    A    SHIP   IN    A   SEAWAY 


257 


of  the  boundary  condition  at  the  actual  position  of  the  hull  would,  of 
course,  be  more  accurate —but  then,  it  would  be  necessary  to  deal 
with  the  full  nonlinear  problem  and  make  sure  that  all  of  the  essential 
correction  terms  of  a  given  order  were  obtained.  In  particular,  it 
would  be  necessary  to  examine  the  higher  order  terms  in  the  condi- 
tions at  the  free  surface—after  all,  the  conditions  (9.1.13),  which  are 
also  used  by  Michell  and  Havelock  (and  everyone  else,  for  that 
matter),  are  satisfied  at  y  =  0  and  not  on  the  actual  displaced  position 
of  the  free  surface.  One  way  to  obtain  a  more  accurate  theory  would 
be,  of  course,  to  carry  out  the  perturbation  scheme  outlined  here  to 
higher  order  terms. 

Still  another  point  has  come  up  for  frequent  discussion  (cf.,  for 
example,  Lunde  and  Wigley  [L.18])  with  reference  to  the  boundary 
condition  on  the  hull.  It  is  fairly  common  in  the  literature  to  refer  to 
ships  of  Miehell's  type,  by  which  is  meant  ships  which  are  slender 
not  only  in  the  fore-and-aft  direction,  but  which  are  also  slender  in 
the  cross-sections  at  right  angles  to  this  direction  (cf.  Fig.  9.1.3)  so 

-  y 


(o)  (b) 

Fig.  9.1.3a,  b.  Ships  with  full  and  with  narrow  mid-sections 

that  hy,  in  our  notation,  is  small.  Thus  ships  with  a  rather  broad 
bottom  (cf.  Fig.  9.1.3a),  or,  as  it  is  also  put,  with  a  full  mid-section, 
arc  often  considered  as  ships  to  which  the  present  theory  does  not 
apply.  On  the  other  hand,  there  are  experimental  results  (cf.  Havelock 
[H.7])  which  indicate  that  the  theory  is  just  as  accurate  for  ships 
with  a  full  mid-section  as  it  is  for  ships  of  Miehell's  type.  When  the 
problem  is  examined  from  the  point  of  view  taken  here,  i.e.  as  a 
problem  to  be  solved  by  a  development  with  respect  to  a  parameter 
characterizing  the  slenderness  of  the  ship,  the  difference  in  the  two 
cases  would  seem  to  be  that  ships  with  a  full  mid-section  should  be 
regarded  as  slender  in  both  draft  and  beam,  (otherwise  no  lineariza- 
tion based  on  assuming  small  disturbances  in  the  water  would  be 


258  WATER   WAVES 

reasonable),  while  a  ship  of  Michell's  type  is  one  in  which  the  draft  is 
finite  and  the  beam  is  small.  In  the  former  case  a  development  dif- 
ferent from  the  one  given  above  would  result:  the  mass  and  moments 
of  inertia  would  be  of  second  order,  for  instance,  rather  than  first 
order.  Later  on  we  shall  have  occasion  to  mention  other  possible  ways 
of  introducing  the  development  parameter. 

We  continue  by  pointing  out  a  number  of  conclusions,  in  addition 
to  those  already  given,  which  can  be  inferred  from  our  equations 
without  solving  them.  Consider,  for  example,  the  equations  (9.1.22) 
and  (9.1.23)  for  the  heave  yl  and  the  pitching  oscillation  031,  and  make 
the  assumption  that 

(9.1.24)  f  xhdx  =  0 

(which  means  that  the  horizontal  section  of  the  ship  at  the  water  line 
has  the  e.g.  of  its  area  on  the  same  vertical  as  that  of  the  whole  ship). 
If  this  condition  is  satisfied  it  is  immediately  seen  that  the  oscillations 
031  and  yl  are  not  coupled.  Furthermore,  these  equations  are  seen  to 
have  the  form 


(9.1.25)  */i  +  AJft  =  p(0 

(9.1.26)  031  +  AJ081  -  q(t) 
with 

(9.1.27)  Af  = 

r~  r  r  ~\ 

%6£  (y  ~  y'  )hdA  -\-       x2hdx 

(9.1.28)  **= t-*-- L _Ji: -=L. 

It  follows  that  resonance*  is  possible  ifp(t)  has  a  harmonic  component 
of  the  form  A  cos  (Aj  +  B)  or  q(t)  a  component  of  the  form 
A  cos  (A2<  +  5):  in  other  words,  one  could  expect  exceptionally 
heavy  oscillations  if  the  speed  of  the  ship  and  the  seaway  were  to  be 
such  as  to  lead  to  forced  oscillations  having  frequencies  close  to  these 
values.  One  observes  also  that  these  resonant  frequencies  can  be 
computed  without  reference  to  the  motion  of  the  sea  or  the  ship: 
the  quantities  A15  A2  depend  only  on  the  shape  of  the  hull.** 

*  The  term  resonance  is  used  here  in  the  strict  sense,  i.e.  that  an  infinite 
amplitude  is  theoretically  possible  at  the  resonant  frequency. 

**  The  equation  (9.1.27)  can  be  interpreted  in  the  following  way:  it  furnishes 
the  frequency  of  free  vibration  of  a  system  with  one  degree  of  freedom  in  which 
the  restoring  force  is  proportional  to  the  weight  of  water  displaced  by  a  cylinder 
of  cross-section  area  2§L  hdx  when  it  is  immersed  vertically  in  water  to  a  depth  yv. 


THE    MOTION   OF   A    SHIP   IN    A    SEAWAY  259 

In  spite  of  the  fact  that  the  linear  theory  presented  here  must  be 
used  with  caution  in  relation  to  the  actual  practical  problems  con- 
cerning ships  in  motion,  it  still  seems  likely  that  such  resonant  fre- 
quencies would  be  significant  if  they  happened  to  occur  as  harmonic 
components  in  the  terms  p(t)  or  q(t)  with  appreciable  amplitudes. 
Suppose,  for  instance,  that  the  ship  is  moving  in  a  sea-way  that 
consists  of  a  single  train  of  simple  harmonic  progressing  plane  waves 
with  circular  frequency  a  which  have  their  crests  at  right  angles  to 
the  course  of  the  ship.  If  the  speed  of  the  ship  is  s0  one  finds  that  the 
circular  excitation  frequency  of  the  disturbances  caused  by  such 
waves,  as  viewed  from  the  moving  coordinate  system  (#,  t/,  z)  that  is 
used  in  the  discussion  here,  is  a  +  *0a2/g,  since  o2/g  is  2n  times  the 
reciprocal  of  the  wave  length  of  the  wave  train.  Thus  if  Ax  or  A2  should 
happen  to  lie  near  this  value,  a  heavy  oscillation  might  be  expected. 
One  can  also  see  that  a  change  of  course  to  one  quartering  the  waves  at 
angle  y  would  lead  to  a  circular  excitation  frequency  a+sQ  cos  y  •  a2/g 
and  naturally  this  would  have  an  effect  on  the  amplitude  of  the  response. 

It  has  already  been  stated  that  the  theory  presented  here  is  closely 
related  to  the  theory  published  by  Haskind  [H.4]  in  1946,  and  it  was 
indicated  that  the  two  theories  differ  in  some  respects.  We  have  not 
made  a  comparison  of  the  two  theories  in  the  general  case,  which  would 
not  be  easy  to  do,  but  it  is  possible  to  make  a  comparison  rather  easily 
in  the  special  case  treated  by  Haskind  in  detail.  This  is  the  special 
case  dealt  with  in  the  second  of  his  two  papers  in  which  the  ship  is 
assumed  to  oscillate  only  in  the  vertical  plane— as  would  be  possible 
if  the  seaway  consisted  of  trains  of  plane  waves  all  having  their  crests 
at  right  angles  to  the  course  of  the  ship.  Thus  only  the  quantities  y^t) 
and  031(/),  which  are  denoted  in  Haskind's  paper  by  £(0  and  y>(t),  are  of 
interest.  Haskind  finds  differential  equations  of  second  order  for  these 
quantities,  but  these  equations  are  not  the  same  as  the  corresponding 
equations  (9.1.22),  (9.1.23)  above.  One  observes  that  (9.1.22)  con- 
tains as  its  only  derivative  the  second  derivative/}^  and  (9.1.23)  con- 
tains as  its  sole  derivative  a  term  with  031;  in  other  words  there  are  no 
first  derivative  terms  at  all,  and  the  coupling  arises  solely  through 
the  undiffcrentiated  terms.  Haskind's  equations  are  quite  different 
since  first  and  second  derivatives  of  both  dependent  functions  occur 
in  both  of  the  two  equations;  thus  Haskind,  on  the  basis  of  his  theory, 
can  speak,  for  example,  of  damping  terms,  while  the  theory  presented 
here  yields  no  such  terms.  On  the  basis  of  the  theory  presented  so  far 
there  should  be  no  damping  terms  of  this  order  for  the  following 


260  WATER   WAVES 

reasons:  In  the  absence  of  frictional  resistances,  the  only  way  in 
which  energy  can  be  dissipated  is  through  the  transport  of  energy  to 
infinity  by  means  of  out-going  progressing  waves.  However,  we  have 
already  given  valid  reasons  for  the  fact  that  those  oscillations  of  the 
ship  which  consist  solely  of  displacements  parallel  to  the  vertical 
plane  produce  waves  in  the  water  with  amplitudes  that  are  of  higher 
order  than  those  considered  in  the  first  approximation.  Thus  no  such 
dissipation  of  energy  should  occur.*  In  any  case,  our  theory  has  this 
fact  as  one  of  its  consequences.  Haskind  [H.4]  also  says,  and  we  quote 
from  the  translation  of  his  paper  (sec  page  59):  "Thus,  for  a  ship 
symmetric  with  respect  to  its  midship  section  .  .  .,  only  in  the  absence 
of  translatory  motion,  i.e.,  for  s0  =  0,  are  the  heaving  and  pitching 
oscillations  independent."  This  statement  does  not  hold  in  our  version 
of  the  theory.  As  one  sees  from  (9.1.22)  and  (9.1.23)  coupling  occurs 

if,  and  only  if,      ochdx  ^  0,  whether  SQ  vanishes  or  not.  In  addition, 


Haskind  obtains  no  resonant  frequencies  in  these  displacements  be- 
cause of  the  presence  of  first-derivative  terms  in  his  equation;  the 
author  feels  that  such  resonant  frequencies  may  well  be  an  important 
feature  of  the  problem.  Thus  it  seems  likely  that  Haskind's  theory 
differs  from  that  presented  here  because  he  includes  a  number  of 
terms  which  are  of  higher  order  than  those  retained  here.  Of  course,  it 
does  not  matter  too  much  if  some  terms  of  higher  order  are  included 
in  a  perturbation  theory,  at  least  if  all  the  terms  of  lowest  order  are 
really  present:  at  worst,  one  might  be  deceived  in  giving  too  much 
significance  to  such  higher  order  terms. 

The  fact  that  the  theory  presented  so  far  leads  to  the  conclusion 
that  no  damping  of  the  pitching,  surging,  and  heaving  oscillations 
occurs  is  naturally  an  important  fact  in  relation  to  the  practical  pro- 
blems. Unfortunately,  actual  hulls  of  ships  seem  in  many  cases  to  be 
designed  in  such  a  way  that  damping  terms  in  the  heaving  and  pitch- 
ing oscillations  are  numerically  of  the  same  order  as  other  terms  in 
the  equations  of  motion  of  a  ship.  (At  least,  there  seems  to  be  experi- 
mental evidence  from  model  studies— see  the  paper  by  Korvin- 
Krukovsky  and  Lewis  [K.I 6]— which  bears  out  this  statement.) 
Consequently,  one  must  conclude  that  either  actual  ships  are  not 

*  It  is,  however,  important  to  state  explicitly  that  there  would  be  damping 
of  the  rolling,  yawing,  and  swaying  oscillations,  since  these  motions  create  waves 
having  amplitudes  of  the  order  retained  in  the  first  approximation,  and  thus 
energy  would  be  carried  off  to  infinity  as  a  consequence  of  such  motions. 


THE   MOTION   OF   A   SHIP   IN   A   SKA  WAY  261 

sufficiently  slender  for  the  lowest  order  theory  developed  here  to  apply 
with  accuracy,  or  that  important  physical  factors  such  as  turbulence, 
viscosity,  etc.,  have  effects  so  large  that  they  cannot  be  safely  neg- 
lected. If  it  is  the  second  factor  that  is  decisive,  rather  than  the  loss 
of  energy  due  to  the  creation  of  waves  through  pitching  and  heaving, 
it  is  clear  that  only  a  basic  theory  different  from  the  one  proposed 
here  would  s6rve  to  include  such  effects.  If,  however,  the  damping 
has  its  origin  in  the  creation  of  gravity  waves  we  need  not  be  entirely 
helpless  in  dealing  with  it  in  terms  of  the  sort  of  theory  contemplated 
here.  It  would  not  be  helpful,  though,  to  try  to  overcome  the  difficulty 
by  carrying  the  development  to  terms  of  higher  order,  for  example, 
even  though  there  would  certainly  then  be  damping  effects  in  pitching 
and  heaving:  such  damping  effects  of  higher  order  could  evidently  not 
introduce  damping  into  the  motions  of  lower  order.  This  is  fortunately 
not  the  only  way  in  which  the  difficulty  can  be  attacked.  One  rather 
obvious  procedure  would  be  to  retain  the  present  theory,  and  simply 
add  damping  terms  with  coefficients  to  be  fixed  empirically,  in  some- 
what the  same  fashion  as  has  been  proposed  by  St.  Denis  and  Wein- 
blum  [S.I],  for  example. 

There  are  still  other  possibilities  for  the  derivation  of  theories  which 
would  include  damping  effects  without  requiring  a  semi-empirical 
treatment,  but  rather  a  different  development  with  respect  to  a  slen- 
derness  parameter.  One  such  possibility  has  already  been  hinted  at 
above  in  the  course  of  the  discussion  of  ships  of  broad  mid-section 
compared  with  ships  of  MichelPs  type.  If  the  ship  is  considered  to  be 
slender  in  both  draft  and  beam  the  waves  due  to  oscillations  of  the 
ship  would  be  of  the  same  order  with  respect  to  all  of  the  degrees  of 
freedom;  a  theory  utilizing  this  observation  is  being  investigated. 
Another  possibility  would  be  to  regard  the  draft  as  small  while  the 
beam  is  finite  (thus  the  ship  is  thought  of  as  a  flat  body  with  a 
planing  motion  over  the  water),  i.e.  to  base  the  perturbation  scheme 
on  the  following  equation  for  the  hull  (instead  of  (9.1.2)): 


and  to  carry  out  the  development  with  respect  to  /?,  This  theory  has 
been  worked  out  in  all  detail,  though  it  has  not  yet  been  published. 
With  respect  to  damping  effects  the  situation  is  now  in  some  respects 
just  the  reverse  of  that  described  above:  now  it  is  the  oscillations  in 
the  vertical  plane,  together  with  the  rolling  oscillation,  that  are 
damped  to  lowest  order,  while  the  yawing  and  swaying  oscillations 


262 


WATER    WAVES 


are  undamped.  It  would  seem  reasonable  therefore  to  investigate  the 
results  of  such  a  theory  for  conventional  hulls  and  make  comparisons 
with  model  experiments.  This  still  does  not  exhaust  all  of  the  possibili- 
ties with  respect  to  various  types  of  perturbation  schemes,  particu- 
larly if  hulls  of  special  shape  are  introduced.  Consider,  for  example, 
a  hull  of  the  kind  used  for  some  types  of  sailing  yachts,  and  shown 
schematically  in  Fig.  9.1.4.  Such  a  hull  has  the  property  that  its  beam 


Fig.  9.1.4.  Cross  section  of  hull  of  a  yacht 

and  draft  are  both  finite,  but  the  hull  cross  section  consists  of  two 
thin  disks  joined  at  right  angles  like  a  T.  In  this  case  an  appropriate 
development  with  respect  to  a  slenderness  parameter  can  also  be 
made  in  regarding  both  disks  as  being  slender  of  the  same  order.  The 
result  is  a  theory  in  which  all  oscillations,  except  the  surge,  would  be 
damped;  this  theory  has  been  worked  out  too  but  not  yet  published. 
It  would  take  up  an  inordinate  amount  of  space  in  this  book  to  deal 
in  detail  with  all  of  the  various  types  of  possible  perturbation  schemes 
mentioned  above.  In  addition,  only  one  of  them  seems  so  far  to  permit 
explicit  solutions  even  in  special  cases,  and  that  is  the  generalization 
of  the  Michell  theory  which  was  explained  at  some  length  above.  Con- 
sequently, only  this  theory  (in  fact,  only  a  special  case  of  it)  will  be 
developed  in  detail  in  the  remainder  of  the  chapter.  In  all  other 
theories,  it  seems  necessary  to  solve  certain  integral  equations  before 
the  motion  of  the  ship  can  be  determined  even  under  the  most  restric- 
tive hypotheses— such  as  a  motion  of  pure  translation  with  no  oscilla- 
tions whatever,  for  example.  Even  in  the  case  of  the  generalized 
Michell  theory  (i.e.  the  case  of  a  ship  regarded  as  a  thin  disk  disposed 
vertically)  an  explicit  solution  of  the  problem  for  the  lowest  order 
approximation  ^  to  the  velocity  potential— in  terms  of  an  integral 


THE   MOTION   OF   A   SHIP   IN   A    SEAWAY  268 

representation,  say— seems  out  of  the  question.  In  fact,  as  soon  as 
rolling  or  yawing  motions  occur,  explicit  solutions  are  unlikely  to  be 
found.  The  best  that  has  been  done  so  far  in  such  cases  has  been  to 
formulate  an  integral  equation  for  the  values  of  9^  over  the  vertical 
projection  A  of  the  ship's  hull;  this  method  of  attack,  which  looks 
possible  and  somewhat  hopeful  for  numerical  purposes  since  the 
motion  of  the  ship  requires  the  knowledge  of  (pl  only  over  the  area  A, 
is  under  investigation.  However,  if  the  motion  of  the  ship  is  confined 
to  a  vertical  plane,  so  that  co1  —  6n  =  02i  —  0>  ft  is  possible  to  solve 
the  problems  explicitly.  This  can  be  seen  with  reference  to  the  bound- 
ary conditions  (9.1.12)  and  (9.1.13)  which  in  this  case  are  identical 
with  those  of  the  classical  theory  of  Michell  and  Havelock,  and  hence 
permit  an  explicit  solution  for  q>±  which  is  given  later  on  in  section 
9.4.  After  9?!  is  determined,  it  can  be  inserted  in  (9.1.21),  (9.1.22),  and 
(9.1.23)  to  find  the  forward  speed  $Q  corresponding  to  the  thrust  T, 
the  two  quantities  fixing  the  trim,  and  the  surge,  pitching,  and  heav- 
ing oscillations.*  In  all,  six  quantities  fixing  the  motion  of  the  ship 
can  be  determined  explicitly.  Only  this  version  of  the  theory  will  be 
presented  in  detail  in  the  remainder  of  the  chapter. 

The  theory  discussed  here  is  very  general,  and  it  therefore  could  be 
applied  to  the  study  of  a  wide  variety  of  different  problems.  For  exam- 
ple, the  stability  of  the  oscillations  of  a  ship  could  be  in  principle 
investigated  on  a  rational  dynamical  basis,  rather  than  as  at  present 
by  assuming  the  water  to  remain  at  rest  when  the  ship  oscillates.  It 
would  be  possible  to  investigate  theoretically  how  a  ship  would  move 
with  a  given  rudder  setting,  and  find  the  turning  radius,  angle  of  heel, 
etc.  The  problem  of  stabilization  of  a  ship  by  gyroscopes  or  other  de- 
vices could  be  attacked  in  a  very  general  way:  the  dynamical  equa- 
tions for  the  stabilizers  would  simply  be  included  in  the  formulation 
of  the  problem  together  with  the  forces  arising  from  the  interactions 
of  the  water  with  the  hull  of  the  ship. 

In  sec.  9.2  the  general  formulation  of  the  problem  is  given;  in 
sec.  9.3  the  details  of  the  linearization  process  are  carried  out  for  the 
case  of  a  ship  which  is  slender  in  beam  (i.e.  under  the  condition 
implied  in  the  classical  Michell-Havelock  theory);  and  in  sec.  9.4  a 
solution  of  the  problem  is  given  for  the  case  of  motion  confined  to  the 
vertical  plane,  including  a  verification  of  the  fact  that  the  wave 
resistance  is  given  by  the  same  formula  as  was  found  by  Michell. 

*  These  free  undamped  vibrations  are  uniquely  determined  only  when  initial 
conditions  are  given. 


264 


WATER   WAVES 


9.2.  General  formulation  of  the  problem 

We  derive  here  a  theory  for  the  most  general  motion  of  a  rigid  body 
through  water  of  infinite  depth  which  is  in  its  turn  also  in  motion  in 
any  manner.  As  always  we  assume  that  a  velocity  potential  exists. 
Since  we  deal  with  a  moving  rigid  body  it  is  convenient  to  refer  the 
motion  to  various  types  of  moving  coordinate  systems  as  well  as  to  a 
fixed  coordinate  system.  The  fixed  coordinate  system  is  denoted  by 
O  —  X,  F,  Z  and  has  the  disposition  used  throughout  this  book:  The 
X,  Z-plane  is  in  the  equilibrium  position  of  the  free  surface  of  the 
water,  and  the  Y-axis  is  positive  upwards.  The  first  of  the  two  moving 
coordinate  systems  we  use  (the  second  will  be  introduced  later)  is 
denoted  by  o  —  x9  y,  z  and  is  specified  as  follows  (cf.  Fig.  9.2.1): 


Fig.  9.2.1.  Fixed  and  moving  coordinate  system 

The  x,  2-plane  coincides  with  the  X,  Z-plane  (i.e.  it  lies  in  the  undis- 
turbed free  surface),  the  y-axis  is  vertically  upward  and  contains  the 
center  of  gravity  of  the  ship.  The  #-axis  has  always  the  direction  of  the 
horizontal  component  of  the  velocity  of  the  center  of  gravity  of  the 
ship.  (If  we  define  the  course  of  the  ship  as  the  vertical  projection  of 
the  path  of  its  center  of  gravity  on  the  X,  Z-plane,  then  our  conven- 
tion about  the  a?-axis  means  that  this  axis  is  taken  tangent  to  the 
ship's  course.)  Thus  if  Rc  =  (Xe,  YC9  Zc)  is  the  position  vector  of  the 
center  of  gravity  of  the  ship  relative  to  the  fixed  coordinate  system 
and  hence  Rc  =  (XC9  yc,  Zc)  is  the  velocity  of  the  e.g.,  it  follows  that 
the  #-axis  has  the  direction  of  the  vector  u  given  by 

(9-2.1)  u  =  X€I  +  ZCK 

with  I  and  K  unit  vectors  along  the  X-axis  and  the  Z-axis.  If  i  is  a 
unit  vector  along  the  <r-axis  we  may  write 
(9.2.2)  s(t)i  =  u, 


THE   MOTION   OF   A   SHIP   IN   A   SEAWAY  265 

thus  introducing  the  speed  s(t)  of  the  ship  relative  to  a  horizontal 
plane.  For  later  purposes  we  also  introduce  the  angular  velocity 
vector  <o  of  the  moving  coordinate  system: 

(9.2.3)  <o  =  a)(t)J, 
and  the  angle  a  (cf.  Fig.  9.2.1)  by 

(9.2.4)  a(0  =  I    aj(r)dr. 

J  o 

The  equations  of  transformation  from  one  coordinate  system  to  the 
other  are 

!X=x  cos  a  +z  sin  a+Jtc  ;x=(X— Xc)  cos  a—  (Z— Zc)  sin  a 
Y=y  ;y=Y 

Z—  —x sin  a  +z  cos  a  -\-Zc;  z=(X—Xc)  sin  a  -\-(Z— Zc)  cos  a. 
By  0(X,  Y,  Z;  t)  we  denote  the  velocity  potential  and  write 

(9.2.6)  0(X,Y,Z;t) 

~  0(x  cos  a  +  2  sin  a  +  Xc,  y,  —  x  sin  a  +  z  cos  a  +  Zc;  t) 
=  <p(x,  y,  z;  t). 

In  addition  to  the  transformation  formulas  for  the  coordinates,  we 
also  need  the  formulas  for  the  transformation  of  various  derivatives. 
One  finds  without  difficulty  the  following  formulas: 

&x  —  <Px  cos  <*  +  y>z  sin  a 

(9.2.7)  0Y  =  <py 

0Z  =  —  (px  sin  a  -f-  q>z  cos  a. 

It  is  clear  that  grad2  0(X,  F,  Z;  t)  =  grad2  <p(x,  y,  z;  t)  and  that  y  is 
a  harmonic  function  in  a?,  y,  z  since  0  is  harmonic  in  X,  F,  Z.  To  cal- 
culate 0t  is  a  little  more  difficult;  the  result  is 

(9.2.8)  0t  =  -  (s  +  coz)(px  +  cox<pz  +  <pt. 

(To  verify  this  formula,  one  uses  0t  —  (fyxi  +  <pvyt  +  <pzzt  +  <Pt  and 
the  relations  (9.2.5)  together  with  s  cos  a  =  Xc>  «  sin  a  =  —  ZC0  The 
last  two  sets  of  equations  make  it  possible  to  express  Bernoulli's  law 
in  terms  of  (p(x,  y,  z;  t);  one  has: 

v  I 

(9.2.9)  -f-  +  gy  +  —  (grad  <p)2  —  (s  +  <oz)<px  +  cox<pz  +  <pt  =  0. 

Q  2 

Suppose  now  that  F(X,  F,  Z;  J)  —  0  is  a  boundary  surface  (fixed 
or  moving)  and  set 

(9.2.10)  F(x  cos  a  +  .  .  .,  j/,  —  x  sin  a  +  .  .  .;  t)  =  f(x,  y,  z;  t), 


266 


WATER  WAVES 


so  that  /(a?,  t/,  z;  t)  =  0  is  the  equation  of  the  boundary  surface  rela- 
tive to  the  moving  coordinate  system.  The  kinematic  condition  to  be 
satisfied  on  such  a  boundary  surface  is  that  the  particle  derivative 
dF/dt  vanishes,  and  this  leads  to  the  boundary  condition 

(9.2.11  )    <pJ9  +  <pyfv  +  <pj,  -  (*  +  a*)/.  +  <*>*/*  +  ft  =  0 

relative  to  the  moving  coordinate  system  upon  using  the  appropriate 
transformation  formulas.  In  particular,  if  y  —  rj(x,  z\  t)  =  0  is  the 
equation  of  the  free  surface  of  the  water,  the  appropriate  kinematic 
condition  is 


(9.2.12)       -  yjj9  +<PV-  <Ptfz  +  (9  +  coztyx  -  0*09,  -  fy  =  0 

to  be  satisfied  for  y  =  77.  (The  dynamic  free  surface  condition  is  of 

course  obtained  for  y  =  ??  from  (9.2.9)  by  setting  p  —  0.) 

We  turn  next  to  the  derivation  of  the  appropriate  conditions,  both 
kinematic  and  dynamic,  on  the  ship's  hull.  To  this  end  it  is  convenient 
to  introduce  another  moving  coordinate  system  0'  —  #',  y'9  z'  which 
is  rigidly  attached  to  the  ship.  It  is  assumed  that  the  hull  of  the  ship 
has  a  vertical  plane  of  symmetry  (which  also  contains  the  center  of 
gravity  of  the  ship);  we  locate  the  x'9  y'  -plane  in  it  (cf.  Fig.  9.2.2)  and 
suppose  that  the  t/'-axis  contains  the  center  of  gravity.  The  o'  —  #', 
t/',  z'  system,  like  the  other  moving  system,  is  supposed  to  coincide 


(a)  (b) 

Fig.  9.2.2a,  b.  Another  moving  coordinate  system 

with  the  fixed  system  in  the  rest  position  of  equilibrium.  The  center 
of  gravity  of  the  ship  will  thus  be  located  at  a  definite  point  on  the 
t/'-axis,  say  at  distance  y'0  from  the  origin  o':  in  other  words,  the  system 
of  coordinates  attached  rigidly  to  the  ship  is  such  that  the  center  of 
gravity  has  the  coordinate  (0,  yc,  0). 

In  the  present  section  we  do  not  wish  in  general  to  carry  out  lineari- 
zations. However,  since  we  shall  in  the  end  deal  only  with  motions 


THE    MOTION   OF   A   SHIP   IN   A   SEAWAY  267 

which  involve  small  oscillations  of  the  ship  relative  to  the  first  moving 
coordinate  system  o  —  x,  y,  z9  it  is  convenient  and  saves  time  and 
space  to  suppose  even  at  this  point  that  the  angular  displacement 
of  the  ship  relative  to  the  o  —  #,  y,  z  system  is  so  small  that  it  can 
be  treated  as  a  vector  8: 

(9.2.13)  8-0^  +0J  +  03k. 

The  transformation  formulas,  correct  up  to  first  order  terms  in  the 
components  0i  of  8,  are  then  given  by: 

'  x9  =  x  +  03(t/  —  yc)  -  02* 
y         y        \y  c        y  §  /    i      i  3 

z'  =  z  +  62?  —  O^y  -  yc) 

with  yc  of  course  representing  the  ^-coordinate  of  the  center  of 
gravity  in  the  unprimed  system.  It  is  assumed  that  yc  —  y'e  is  a  small 
quantity  of  the  same  order  as  Qt  and  only  linear  terms  in  this  quantity 
have  been  retained.  (The  verification  of  (9.2.14)  is  easily  carried  out 
by  making  use  of  the  vector-product  formula  8  =  8  X  r,  for  the 
small  displacement  8  of  a  rigid  body  under  a  small  rotation  8.) 

The  equation  of  the  hull  of  the  ship  (assumed  to  be  symmetrical 
with  respect  to  the  #',  j/'-plane)  is  now  supposed  given  relative  to  the 
primed  system  of  coordinates  in  the  form: 

/a  o  i  K\  yf  —    r    /Yr'    ii'\  y'  >  O 

\*Jȣt,\.*Jj  A      ^T    S  V4^   5  a    />  < 

The  equation  of  the  hull  relative  to  the  o  —  x,  y,  s-system  can  now 
be  written  in  the  form 


(9.2.16)  z  +  Bv-OAy  -  y'e)  -  f  (a,  y)  +  [0#  -  03(j/  -  ^)]C.  (*,y) 

+  [(yc  -  y'e)  -  0i*  +  OrtM*,  y)  =  o,      *'  >  o, 

when  higher  order  terms  in  (yc  —  y'c)  and  0t  are  neglected.  The  left 
hand  side  of  this  equation  could  now  be  inserted  for  /  in  (9.2.11)  to 
yield  the  kinematic  boundary  condition  on  the  hull  of  the  ship,  but 
we  postpone  this  step  until  the  next  section. 

The  dynamical  conditions  on  the  ship's  hull  are  obtained  from  the 
assumption  that  the  ship  is  a  rigid  body  in  motion  under  the  action 
of  the  propeller  thrust  T,  its  weight  Mg,  and  the  pressure  p  of  the 
water  on  its  hull.  The  principle  of  the  motion  of  the  center  of  gravity 
yields  the  condition 

(9.2.17)  M~  (*i  +  yj)  -  J    pn  dS  +  T  -  Mgj. 


268  WATER   WAVES 

By  n  we  mean  the  inward  unit  normal  on  the  hull.  Our  moving 
coordinate  system  o  —  #,  y,  z  is  such  that  di/dt  =  —  cok  and  dj/dt  = 
0,  so  that  (9.2.17)  can  be  written  in  the  form 

(9.2.18)  Msi  -  Mscok  +  Mycj  =  (    pn  dS  +  T  -  Mgj, 

with  p  defined  by  (9.2.9).  The  law  of  conservation  of  angular  momen- 
tum is  taken  in  the  form: 

(9.2.19)  1  (     (R  -  Rc)  X  (R  -  Rc)dm 
dt  J  M 

=   (    p(R  -  Rc)  n  dS  +  (RT  -  Rc)  x  T. 

J  s 

The  crosses  all  indicate  vector  products.  By  R  is  meant  the  position 
vector  of  the  element  of  mass  dm  relative  to  the  fixed  coordinate 
system.  Rc  (cf.  Fig.  9.2.1)  fixes  the  position  of  the  e.g.  and  RT 
locates  the  point  of  application  of  the  propeller  thrust  T.  We  introduce 
r  =  (#,  y,  z)  as  the  position  vector  of  any  point  in  the  ship  in  the 
moving  coordinate  system  and  set 

(9.2.20)  q  =  r  -  ycj, 

so  that  q  is  a  vector  from  the  e.g.  to  any  point  in  the  ship.  The  relation 

(9.2.21)  R  =  Rc  +  (co  +  6)  X  q 

holds,  since  co  +  8  is  the  angular  velocity  of  the  ship;  thus  (9.2.21) 
is  simply  the  statement  of  a  basic  kinematic  property  of  rigid  bodies. 
By  using  the  last  two  relations  the  dynamical  condition  (9.2.19)  can 
be  expressed  in  terms  of  quantities  measured  with  respect  to  the 
moving  coordinate  system  o  —  x9  y,  z,  as  follows: 

(9.2.22)  ~  J    (r  -  ycj)  X  [(co  +  9)  X  (r  -  ycj)]dn 

r 

0(r  —  J/cJ)  X  n  dS  +  (RT  —  Rc)  x  T. 


We  have  now  derived  the  basic  equations  for  the  motion  of  the 
ship.  What  would  be  wanted  in  general  would  be  a  velocity  potential 
<p(x,  y,  z;  t)  satisfying  (9.2.11)  on  the  hull  of  the  ship,  conditions 
(9.2.9)  (with  p  =  0)  and  (9.2.12)  on  the  free  surface  of  the  water; 
and  conditions  (9.2.17)  and  (9.2.22),  which  involve  9?  under  integral 
signs  through  the  pressure  p  as  given  by  (9.2.9).  Of  course,  the  quan- 


THE   MOTION   OF   A   SHIP   IN   A   SEAWAY  269 

tities  fixing  the  motion  of  the  ship  must  also  be  determined  in  such  a 
way  that  all  of  the  conditions  are  satisfied.  In  addition,  there  would 
be  initial  conditions  and  conditions  at  oo  to  be  satisfied.  Detailed 
consideration  of  these  conditions,  and  the  complete  formulation  of  the 
problem  of  determining  y(x,  y,  z;  t)  under  various  conditions  will  be 
postponed  until  later  on  since  we  wish  to  carry  out  a  linearization 
of  all  of  the  conditions  formulated  here. 


9.3.  Linearization  by  a  formal  perturbation  procedure 

Because  of  the  complicated  nature  of  our  conditions,  it  seems  wise 
(as  was  indicated  in  sec.  1  of  this  chapter)  to  carry  out  the  lineari- 
zation by  a  formal  development  in  order  to  make  sure  that  all  terms  of 
a  given  order  are  retained;  this  is  all  the  more  necessary  since  terms 
of  different  orders  must  be  considered.  The  linearization  carried  out 
here  is  based  on  the  assumption  that  the  motion  of  the  water  relative 
to  the  fixed  coordinate  system  is  a  small  oscillation  about  the  rest 
position  of  equilibrium.  It  follows,  in  particular,  that  the  elevation  of 
the  free  surface  of  the  water  should  be  assumed  to  be  small  and,  of 
course,  that  the  motion  of  the  ship  relative  to  the  first  moving  coor- 
dinate system  o  —  x,  y>  z  should  be  treated  as  a  small  oscillation.We 
do  not,  however,  wish  to  consider  the  speed  of  the  ship  with  respect 
to  the  fixed  coordinate  system  to  be  a  small  quantity:  it  should  rather 
be  considered  a  finite  quantity.  This  brings  with  it  the  necessity  to 
restrict  the  form  of  the  ship  so  that  its  motion  through  the  water  does 
not  cause  disturbances  so  large  as  to  violate  our  basic  assumption; 
in  other  words,  we  must  assume  the  ship  to  have  the  form  of  a  thin 
disk.  In  addition,  it  is  clear  that  the  velocity  of  such  a  disk-like  ship 
must  of  necessity  maintain  a  direction  that  does  not  depart  too  much 
from  the  plane  of  the  thin  disk  if  small  disturbances  only  are  to  be 
created.  Thus  we  assume  that  the  equation  of  the  ship's  hull  is  given  by 

(9.3.1)  z'  =ph(x'9y'),        z'  >  0, 

with  ft  a  small  dimensionless  parameter,  so  that  the  ship  is  a  thin 
disk  symmetrical  with  respect  to  the  xl \  j/'-plane,  and  @h  takes  the 
place  of  f  in  (9.2.15).  (It  has  already  been  noted  in  the  introduction 
to  this  chapter  that  this  is  not  the  most  general  way  to  describe  the 
shape  of  a  disk  that  would  be  suitable  for  a  linearization  of  the  type 
carried  out  here. )  We  have  already  assumed  that  the  motion  of  the 
ship  is  a  small  oscillation  relative  to  the  moving  coordinate  system 


270  WATER  WAVES 

o  —  tc,y,  z.  It  seems  reasonable,  therefore,  to  develop  all  our  basic 
quantities  (taken  as  functions  of  x,  y,z;t)  in  powers  of  /?,  as  follows: 

(9.8.2)  y(x,  y,  2;  t;  /?)  =  ^ 

(9.8.8)  17(0,  2;  <;  0)  =  fa 

(9.8.4)  <(<;  0)  =  *0  +  fo  +  /»•«,  +  .  .  ., 

(9.8.5)  «(<;  j8)  =  o>0  +  /toj  +  /Wo,  +  .  .  ., 

(9.8.6)  0,(*;  /?)  =  /30«  +  /W«  +  .  .  .,  1  =  1,  2,  8 


The  first  and  second  conditions  state  that  the  velocity  potential  and 
the  surface  wave  amplitudes,  as  seen  from  the  moving  system,  are 
small  of  order  /3.  The  speed  of  the  ship,  on  the  other  hand,  and  the 
angular  velocity  of  the  moving  coordinate  system  about  the  vertical 
axis  of  the  fixed  coordinate  system,  are  assumed  to  be  of  order  zero. 
(It  will  turn  out,  however,  that  cu0  must  vanish—  a  not  unexpected 
result.)  The  relations  (9.3.6)  and  (9.3.7)  serve  to  make  precise  our 
assumption  that  the  motion  of  the  ship  is  a  small  oscillation  relative 
to  the  system  o  —  x,  t/,  z. 

We  must  now  insert  these  developments  in  the  conditions  derived 
in  the  previous  section.  The  free  surface  conditions  are  treated  first. 
As  a  preliminary  step  we  observe  that 

(9.3.8)     <px(x,  rj,  z;  t;  ft)  -  P[<pl9(x,  0,  z;  t)  +  rff^(x9  0,  z;  t)  +  .  .  .] 


*,  0,  z;  t)  +  P*[fli<pixv(x>  °»  *;  ')  +  ?i«(*»  °»  *5  0] 


with  similar  formulas  for  other  quantities  when  they  are  evaluated 
on  the  free  surface  y  =  rj.  Here  we  have  used  the  fact  that  r\  is  small 
of  order  ft  and  have  developed  in  Taylor  series.  Consequently,  the 
dynamic  free  surface  condition  fory==ri  arising  from  (9.2.9)  with 
p  =  0  can  be  expressed  in  the  form 

(9.3.9)     gfot  +  (3*r)2  +  ...]  +tf«[(grad  ^)2  +  .  .  .] 

-  Oo  +  fo  +  •  •  •  +  *K  +#»!  +  ••  OHM*  + 

+  <P*x)  +  .  .  •] 


and  this  condition  is  to  be  satisfied  for  y  =  0.  In  fact,  as  always  in 


THE   MOTION    OF   A   SHIP   IN   A   SEAWAY  271 

problems  of  small  oscillations  of  continuous  media,  the  boundary 
conditions  are  satisfied  in  general  at  the  equilibrium  position  of  the 
boundaries.  Upon  equating  the  coefficient  of  the  lowest  order  term 
to  zero  we  obtain  the  dynamical  free  surface  condition 

(9.3.10)  —  g^  +  (s0  +  co<p)(plx  —  a)<p(plz  —  (plt  =  0  for  y  =  0, 

and  it  is  clear  that  conditions  on  the  higher  order  terms  could  also  be 
obtained  if  desired.  In  a  similar  fashion  the  kinematic  free  surface 
condition  can  be  derived  from  (9.2.12);  the  lowest  order  term  in  ft 
yields  this  condition  in  the  form: 

(9.3.11)  q>ly  +  (SQ  +  co^)rjlx  -  a)^rjlz  -  i?lt  =  0         for  y  =  0. 

We  turn  next  to  the  derivation  of  the  linearized  boundary  condi- 
tions on  the  ship's  hull.  In  view  of  (9.3.6)  and  (9.3.7),  the  transforma- 
tion formulas  (9.2.14)  can  be  put  in  the  form 

(x'=x +06^-9'.) -fin* 

(9.3.12)  ly'  =  y  -  fa  +  ftQ^z  -  ftB^x 

(z'=z+p02lx-(30n(y-y'c) 

when  terms  involving  second  and  higher  powers  of  ft  are  rejected. 
Consequently,  the  equation  (9.2.16)  of  the  ship's  hull,  up  to  terms  in 
ft2,  can  be  written  as  follows: 

z  +  fi62lx  -  peu(y  -  y'c)  -  fth[x  +  ftQ^y  -  y'c)  -  ftQ^z, 

y  -  Pvi  +  flu*  -  &**]  =  o, 

and,  upon  expanding  the  function  fe,  the  equation  becomes 

(9.3.13)  z  +  ftQ2lx  -  00n(y  -  y'c)  -  ph(x,  y)  +  . . .  =  0, 

the  dots  representing  higher  order  terms  in  ft.  We  can  now  obtain  the 
kinematic  boundary  condition  for  the  hull  by  inserting  the  left  hand 
side  of  (9.3.13)  for  the  function  /  in  (9.2.11);  the  result  is 

(9314)    l*00^0 

l^i*  =  *<>(02i  -  *«) 

when  the  terms  of  zero  and  first  order  only  arc  taken  into  account. 
It  is  clear  that  these  conditions  are  to  be  satisfied  over  the  domain 
A  of  the  x,  t/-plane  that  is  covered  by  the  projection  of  the  hull  on  the 
plane  when  the  ship  is  in  the  rest  position  of  equilibrium.  As  was 
mentioned  earlier,  it  turns  out  that  o>0  =  0,  i.e.,  that  the  angular 
velocity  about  the  t/-axis  must  be  small  of  first  order,  or,  as  it  could 
also  be  put,  the  curvature  of  the  ship's  course  must  be  small  since  the 


272  WATER  WAVES 

speed  in  the  course  is  finite.  The  quantity  s^t)  in  (9.3.4)  evidently 
yields  the  oscillation  of  the  ship  in  the  direction  of  the  #-axis  (the  so- 
called  "surge"). 

It  should  also  be  noted  that  if  we  use  z'  =  —  f}h(x',  y')  we  find, 
corresponding  to  (9.3.14),  that 

9>i*  =  *o(02i  +  hx)  -  K  +  02i)*  +  bu(y  -  y'c). 
This  means  that  A  must  be  regarded  as  two  sided,  and  that  the  last 
equation  is  to  be  satisfied  on  the  side  of  A  which  faces  the  negative 
2-axis.  The  last  equation  and  (9.3.14)  imply  that  <p  may  have  discon- 
tinuities at  the  disk  A. 

The  next  step  in  the  procedure  is  to  substitute  the  developments 
with  respect  to  /?,  (9.3.2)  — (9.3.7),  in  the  conditions  for  the  ship's 
hull  given  by  (9.2.18)  and  (9.2.22).  Let  us  begin  with  the  integral 


1 


pn  dS  which  appears  in  (9.2.18).  In  this  integral  S  is  the  immersed 
s 

surface  of  the  hull,  n  is  the  inward  unit  normal  to  this  surface  and  p 
is  the  pressure  on  it  which  is  to  be  calculated  from  (9.2.9).  With 
respect  to  the  o  —  x  ,  y,  z  coordinate  system  the  last  equations  of  the 
symmetrical  halves  of  the  hull  are 


02"       Z  == 

where 


<Q  q  im 

(9.3.16) 

We  can  now  write 


pn  dS  =       pnt  dS1  +       pn2  dS2 

Js  Js1  Js2 


in  which  nx  and  n2  are  given  by 
H1  +  H      ~  k 


We  can  also  write 


pn  dS  =  -  eg  \   yn  dS  +  \   Pln  dS 

Js  Js  Js 

=  —68]   Vn  ds  +  I    Pinidsi  + 

Js  Jsi  Jsa 

where  pl9  from  (9.2.9),  is  given  by 

(9.3.17)   pl  =  -  0[£(grad  <p)2  -  (s  +  (oz)(px  +  xcoy,  +  yt]. 


THE    MOTION    OF   A   SHIP   IN   A   SEAWAY  273 

If  S0  is  the  hull  surface  below  the  #,  2-plane,  the  surface  area  SQ  —  S 
is  of  order  ft  and  in  this  area  each  of  the  quantities  t/,  Hv  H2  is  of 
order  ft.  Hence  one  finds  the  following  to  hold: 

-  f  yn  dS  =  -  f    yndS  +  (l+  j)O(^)  +  kO(^). 
Js  Js0 

From  the  divergence  theorem  we  have 

-  f   yn  dS  =  VI 

JS9 

where  V  is  the  volume  bounded  by  SQ  and  the  x9  s-plane.  With  an 
accuracy  of  order  /?3,   V  is  given  by 


V=20  f  hdA-  f  ft(yl+0Blx)dB=2ft  f  hdA-2p*  (  ( 
JA  JB  JA  JL 

Here  A  is  the  projection  of  the  hull  on  the  vertical  plane  when  the 
hull  is  in  the  equilibrium  position,  B  is  the  equilibrium  water  line 
area,  and  L  is  the  projection  of  the  equilibrium  water  line  on  the 


If  Wv  W2  are  the  respective  projections  of  the  immersed  surfaces 
Sl9  S2  on  the  x,  j/-plane  we  have 

f  Pln  dS  =  1  (  f    Pl(jr,  y,  HI;  t)HlrdW,-  !    Pl(xt  y,  H2; 

Js  \Jwl  Jwz 

+J  (  f    Pi(*>  *J<  »  i5  0#  ndH'i  -  f    Pi(*>  y>  H* 
U  \YI  J  \\-2 

-k  (  f     Pi(*>  y>  »i;  t)d\\\-  f     Pl(x,  y,  H2;  t)d 

I  J  ^  J  ira 

Neither  Wl  nor  W2  is  identical  with  A.  Each  of  the  differences 
ll\  —  A,  W2  —  A  is,  however,  an  area  of  order  ft.  From  this  and  the 
fact  that  each  of  the  quantities  p,  Hlx,  Hly,  H2X9  H2y  is  of  order  ft,  it 
follows  that 

(9.3.18) 

J  Pln  dS=l  j  f  [Pl(^,  y,  II,;  t)IIlx-Pl(x,  y,  II2;t)H2x]dA+O(p*)\ 

+j(  f  [Pi(^  y,Hi;t)Hly  -Pl  (cr,i/,//2;  t)H2y] 
\JA 

y,  HI;  0-Pite  2/»  W 


274 


WATER  WAVES 


(9.3.19) 


It  was  pointed  out  above  that  <p  may  be  discontinuous  on  A.  Hence 
from  (9.8.17),  (9.8.2),  (9.8.4)  we  write 

.(*,!,,  ffi;<)  = 
(x,y,H»t)  =  t 

Here  the  +  and  —  superscripts  denote  values  at  the  positive  and 
negative  sides  of  the  disk  A  whose  positive  side  is  regarded  as  the  side 
which  faces  the  positive  s-axis.  If  we  substitute  the  developments  of 
HI*,  Hly,  ff2a!,  ff2y,  and  (9.3.19)  in  (9.3.18),  then  collect  the  previous 
results,  we  find 


f  pn  dS=i  I  491  f  [(A.-fln)(t^.-^) 
Js  \      JA 


(9.3.20) 


!  [(hy+Ou)(sQ<plx-<plt)++(hy^^^^^ 
JA 


f 

JA 


f 
The  integral     p(r  —  yc\)  X  ndS  which  appears  in  (9.2.22)  can 

Js 
be  written 

dS 


f  p(r-t/cj)xndS=-eg|  y(r-yc])xn 
Js  Js 


If  we  use  the  same  procedure  as  was  used  above  for  the  expansion  of 


pn  dS  we  find 


-^^ 


f  W«cft.-^)+- 

JA 


THE  MOTION  OF  A  SHIP  IN  A  SEAWAY  275 

(9.3.21) 

8gp  f  xhdA-*QgpOu  f  (y-y'cWA-2Qgpyi  !  xhfa-2QgpoJ  x*hdx 
JA  JA  JL  JL 

+e/?2  f  [^(^+eii)(Wix~9ie)++^(Av-0n)(Wix-^ie)1^ 
JA 

-Qn[(y-y*)(h*-^i)(*^^ 

JA 

We  now  assume  that  the  propeller  thrust  T  is  of  order  fP  and  is 
directed  parallel  to  the  #'-axis:  that  is 

T  =  ]WY 

where  i'  is  the  unit  vector  along  the  #'-axis.  We  also  assume  that  T  is 
applied  at  a  point  in  the  longitudinal  plane  of  symmetry  of  the  ship  / 
units  below  the  center  of  mass.  Thus  we  have  the  relations 


(9.3.22)  T  =  P*Ti  +  0(/J3), 
and 

(9.3.23)  (Rr  -  Rc)  X  T  =  -  JJ  X  T 


The  mass  of  the  ship  is  of  order  /?.  If  we  write  M  —  M^  and  ex- 
pand the  left  hand  side  of  (9.2.18)  in  powers  of  ft  it  becomes 

P/^o  +  MJtit  +  003*)]  +  Jflf^  +0(jJ>)]  - 
(9.3.24)  =  f  pn  dS  +  T  -  Af^gJ. 

The  expansion  of  the  left  hand  side  of  (9.2.22)  gives 
(9.3.25) 


where  /J/31  is  the  moment  of  inertia  of  the  ship  about  the  axis  which  is 
perpendicular  to  the  longitudinal  plane  of  symmetry  of  the  ship  and 
which  passes  through  the  center  of  mass. 

If  we  replace  the  pressure  integrals  and  thrust  terms  in  the  last  two 
equations  by  (9.3.20),  (9.3.21),  (9.3.22),  (9.3.23),  and  then  equate  the 
coefficients  of  like  powers  of  ft  in  (9.3.24)  and  (9.3.25)  we  obtain  the 
following  linearized  equations  of  motion  of  the  ship.  From  the  first 
order  terms  we  find 


276  WATER  WAVES 

(9.8.26)  *„  =  0 

(9.8.27)  2gg  f  phdA^MJg 

JA 

(9.8.28)  I    xphdA=0 

JA 

(9.3.29)  I    [(s<fl>lx-<plt)+-(s(fplx-<plt)-]dA=0 
JA 

(9.8.80)    f   [x(s^lx-(plt)+-x(s0<plx-(plt)-]dA^O 
JA 

(9.3.31)  f  [(y-y'e)(s<fplx-<plt)+-(y-y'e)(s0<plx-<plt)-]dA=o 
JA 

or  by  (9.8.29) 

(9.3.32)  |    [y(s</plx-(f>lt)+-y(s0<jplx-<plt)-]dA=0. 
JA 

From  the  second  order  terms  we  find 
(9.3.33) 

Mi*i=e  f  [(hx-0Zi 
JA 

=6     [ 

JA 

(9.3.34) 

Miy^-lqn 
JL 

+Q      [(hv+Ou)(stf>lx-(plt)++(hy-On)(s0(plx-(plt)-]dA 
JA 

=  -20g     (Vi+*9n)Ma:+Q  \    [hv 
JL  JA 

iJn  =-2pgeSl  f  (y-y'e)hdA-2egyi  ( xhdx-'2eg031  ( 
JA  JL  J 

+6      [y(hv+^i)(^i 
JA 

-e    l(y-yc)(kx-0n 

JA 


THE   MOTION   OF   A    SHIP   IN   A   SEAWAY  277 

or  by  (9.8.30),  (9.3.31) 

. .  r  ,  /•  r 

^ai^ai  —  "-2?^!  I    (y—ye^hdA—ZQgyt  I    xhdx—2QgQ3l  I    x2hdx+lT 

JA  JL  JL 

(9.8.35) 

f 
JA  ° 

Equation  (9.3.26)  states  that  the  motion  in  the  ^-direction  is  a 
small  oscillation  relative  to  a  motion  with  uniform  speed  SQ  =  const. 
Equation  (9.3.27)  is  an  expression  of  Archimedes'  law:  the  rest  position 
of  equilibrium  must  be  such  that  the  weight  of  the  water  displaced 
by  the  ship  just  equals  the  weight  of  the  ship.  The  center  of  buoyancy 
of  the  ship  is  in  the  plane  of  symmetry,  and  equation  (9.8.28)  is  an 
expression  of  the  second  law  of  equilibrium  of  a  floating  body;  namely 
that  the  center  of  buoyancy  for  the  equilibrium  position  is  on  the 
same  vertical  line,  the  t/'-axis,  as  the  center  of  gravity  of  the  ship. 

The  function  <pl  must  satisfy 

<Plxx  +  <Plw  +  <Pizz  =  ° 

in  the  domain  D  —  A  where  D  is  the  half  space  y<  0,  and  A  is  the 
plane  disk  defined  by  the  projection  of  the  submerged  hull  on  the 
x,  y-plane  when  the  ship  is  in  the  equilibrium  position.  We  assume  that 
A  intersects  the  x,  2-plane.  The  boundary  conditions  at  eacli  side  of  A 
are 


(9.3.36) 

=  +  sofa  +  02i)  -  K  +  #21)*  + 

The  boundary  condition  at  y  =  0  is  found  by  eliminating  rjl  from 
(9.3.10)  and  (9.3.11).  Since  co0  =  0  these  equations  are 


—  <Pit  = 

—  <Piv  ~  Mix  +  ^i<  =  ° 
and  they  yield 

(9.3.37)  sfolxx  -  2^9?!^  +  g<fiy  +  <Pw  =  0 

for  y  =  0.  The  boundary  conditions  (9.8.36)  and  (9.8.87)  show  that 
Pi  depends  on  co^t),  On(t)  and  021(0-  The  problem  in  potential  theory 
for  (pl  can  in  principle  be  solved  in  the  form 


without  using  (9.3.29),  (9.3.30),  (9.3.32).  The  significance  of  this  has 


278  WATER   WAVES 

already  been  discussed  in  sec.  9.1  in  relation  to  equations  (9.1.14). 
The  general  procedure  to  be  followed  in  solving  all  problems  was  also 
discussed  there. 

The  remainder  of  this  chapter  is  concerned  with  the  special  case  of 
a  ship  which  moves  along  a  straight  course  into  waves  whose  crests  are 
at  right  angles  to  the  course.  In  this  case  there  are  surging,  heaving 
and  pitching  motions,  but  we  have  dl  =  0,  02  =  0,  co  =  0;  in  addition 
we  note  that  the  potential  function  <p  can  be  assumed  to  be  an  even 
function  of  z.  Under  these  conditions  the  equations  of  motion  are 
much  simpler.  They  are 

(9.8.38)  Mj^gl  hx(s<fpix~<Pit)dA+T 

JA 

(9.3.39)  M^yl=-2Qgy1  \hdx-2eg6^  \xhdx+2e      hy(sQ<plx-<plt)dA 

JL  JL  JA 

(9.3.40)  /3i03i  =  -2eg03if   (y-y'e}hdA-2QgyA  xhdx 

JA  JL 


i  f 

JL 


x*hdx+lT 


+2Q       [xhy-(y-y'c)hx}(s<fplx-(plt)dA. 
JA 


It  will  be  shown  in  the  next  section  that  an  explicit  integral  represen- 
tation can  be  found  for  the  corresponding  potential  function  and  that 
this  leads  to  integral  representations  for  the  surge  sv  the  heave  yl 
and  the  pitching  oscillation  031. 


9.4.  Method  of  solution  of  the  problem  of  pitching  and  heaving  of 
a  ship  in  a  seaway  having  normal  incidence 

In  this  section  we  derive  a  method  of  solution  of  the  problem  of 
calculating  the  pitching,  surging,  and  heaving  motions  in  a  seaway 
consisting  of  a  train  of  waves  with  crests  at  right  angles  to  the  course 
of  the  ship,  which  is  assumed  to  be  a  straight  line  (i.e.,  co  =  0).  The 
propeller  thrust  is  assumed  to  be  a  constant  vector. 

The  harmonic  function  9^  and  the  surface  elevation  iyt  therefore 
satisfy  the  following  free  surface  conditions  (cf.  (9.8.10)  and  (9.8.11), 
with  co0  =  0): 


THE    MOTION   OF   A   SHIP   IN   A    SEAWAY  279 

(9.4.1) 

=  0 

The  kinematic  condition  arising  from  the  hull  of  the  ship  is  (cf. 
(9.3.14)  with  021  =  0n  =  o>!  =  0): 

(9.4.2)  <plz  =  -  V**- 

Before  writing  down  other  conditions,  including  conditions  at  oo, 
we  express  <pl  as  a  sum  of  two  harmonic  functions,  as  follows 

(9.4.3)  9^ (x,  y,  z;  t)  =  ^0(a?,  y,  2)  +  Xl(x,  y,  z;  t). 

Here  Xo  is  a  harmonic  function  independent  of  t  which  is  also  an  even 
function  of  z.  We  now  suppose  that  the  motion  of  the  ship  is  a  steady 
simple  harmonic  motion  in  the  time  when  observed  from  the  moving 
coordinate  system  o  —  x,  y,  z.  (Presumably  such  a  state  would  result 
after  a  long  time  upon  starting  from  rest  under  a  constant  propeller 
thrust.)  Consequently  we  interpret  XQ(X,  J/»  z)  as  the  disturbance 
caused  by  the  ship,  which  therefore  dies  out  at  oo;  while  %i(x9  t/,  z;  t) 
represents  a  train  of  simple  harmonic  plane  waves  covering  the  whole 
surface  of  the  water.  Thus  fa  is  given,  with  respect  to  the  fixed  coor- 
dinate system  O— X,  Y,  Z  by  the  well-known  formula  (cf.  Chapter  8): 

~Y       /a2  \ 

Xl  =  Ce"    sin  lat  +  -X  +yl, 

with  a  the  frequency  of  the  waves.  In  the  o  —  x,  y,  z  system  we  have, 
therefore: 

o  r"   2  /  2\  *n 

(9.4.4)  Xl(x,  y,  z;  t)  =  Ce^  sin     —  x  +  la  +^J  <  +  Y  • 

We  observe  that  the  frequency,  relative  to  the  ship,  is  increased  above 
the  value  a  if  $0  is  positive  —  i.e.  if  the  ship  is  heading  into  the  waves 
—and  this  is,  of  course,  to  be  expected.  With  this  choice  of  £lf  it  is 
easy  to  verify  that  %Q  satisfies  the  following  conditions: 

(9.4.5)  S$XQXX  +  gXov  =  0         at  y  =  0, 
obtained  after  eliminating  r/l  from  (9.4.1),  and 

(9.4.6)  fa,  =  —  s^hx        on  A, 

with  A,  as  above,  the  projection  of  the  ship's  hull  (for  z  >  0)  on  its 
vertical  mid-section.  In  addition,  we  require  that  XQ  ->  0  at  oo. 


280  WATER   WAVES 

It  should  be  remarked  at  this  point  that  the  classical  problem  con- 
cerning the  waves  created  by  the  hull  of  a  ship,  first  treated  by  Michell 
[M.9],  Havelock  [H.7],  and  many  others,  is  exactly  the  problem  of 
determining  #0  from  the  conditions  (9.4.5)  and  (9.4.6).  Afterwards, 
the  insertion  of  ^  =  %Q  in  (9.3.38),  with  ix  =  0,  <plt  =  0,  leads  to  the 
formula  for  the  wave  resistance  of  the  ship—  i.e.  the  propeller  thrust  T 
is  determined.  Since  yl  and  03  are  independent  of  the  time  in  this  case, 
one  sees  that  the  other  dynamical  equations,  (9.3.39)  and  (9.3.40), 
yield  the  displacement  of  the  e.g.  relative  to  the  rest  position  of 
equilibrium  (the  heave),  and  the  longitudinal  tilt  angle  (the  pitching 
angle).  However,  in  the  literature  cited,  the  latter  two  quantities  are 
taken  to  be  zero,  which  implies  that  appropriate  constraints  would 
be  needed  to  hold  the  ship  in  such  a  position  relative  to  the  water. 
The  main  quantity  of  interest,  though,  is  the  wave  resistance,  and  it 
is  not  affected  (in  the  first  order  theory,  at  least)  by  the  heave  and  pitch. 

We  proceed  to  the  determination  of  #0,  using  a  method  different 
from  the  classical  method  and  following,  rather,  a  course  which  it  is 
hoped  can  be  generalized  in  such  a  way  as  to  yield  solutions  in  more 
difficult  cases. 

Suppose  that  we  know  the  Green's  function  G*(f  ,  77,  £;  x,y,z)  such  that 
G*  is  a  harmonic  function  for  rj  <  0,  f  >  0  except  at  (#,  y,  z)  where 
it  has  the  singularity  1/r;  and  G*  satisfies  the  boundary  conditions 
(9.4.7)  G£  +  kG*  =  0  on  77  =  0 

G*  =  0         on  C  =  0 

where  k  =  g/s§«  We  shall  obtain  this  function  explicitly  in  a  moment, 
and  will  proceed  here  to  indicate  how  it  is  used.  Let  27  denote  the  half 
plane  r\  =  0,  £  >  0;  and  let  Q  denote  the  half  plane  f  —  0,  r]  <  0. 
From  Green's  formula  and  the  classical  argument  involving  the 
singularity  1/r  we  have 


Then,  since 


=0, 


THE    MOTION   OF   A   SHIP   IN   A   SEAWAY  281 

we  have  an  explicit  representation  of  the  solution  in  the  form 

or 


(*>  2/>  «)=  -  —  J  j  Xo*G*d£dr), 


(9.4.8)      £,(*,  y,  z)=  A«(£,  ^  )<?*(£,  q,  0;  *,  y,  z 

^ 

upon  using  (9.4.6). 

In  order  to  determine  G*  consider  the  Green's  function  G(£,  r],  C; 
x,y,z)  for  the  half  space  77  <  0  which  satisfies 


on  77  =  0.  This  function  can  be  written  as 


where 


and  g  is  a  potential  function  in  77  <  0  which  satisfies 

.,  a  i 


on  r)  =  0.  The  formula 
_.   d  1 


-^^  -2A; 


(obtained  from  the  well-known  analogous  representation  for  l/r)  in 
which  the  Besscl  function  J0  can  be  expressed  as 

.  ___         2  f  n/2 

-)2]=  -       oos  [p(f-#)  cos  0]  cos  [p(C-s)  sin  0]  d0, 
^Jo 


allows  us  to  write 

4&  r°°  r/2 


cos       ~~a?  cos     cos       ~~ 

0 


4A;  f00  f 

—  — 

^  J  0  J 

for  r\  =  0  and  j/  <  0.  It  is  now  easy  to  see  that 


282  WATER  WAVES 


4Jfe  f  °°  f*/2 

,=  — 

n  Jo  Jo 


gtf  +*&,=  —  pc'Wrt  cos  [p(f-a?)  cos  6]  cos  [p(C-a)  sin  9]  dB  dp 

n  Jo  Jo 

is  a  potential  function  in  77  <  0  which  satisfies  the  boundary  condition. 
An  interchange  of  the  order  of  integration  gives 

dp 


4Jfc  M*  f  * 

=  —       dQ  9te      p  cos  [p(f-*)  sin 
ft  Jo  Jo 


where  &e  denotes  the  real  part.  If  we  think  of  p  as  a  complex  variable, 
the  path  from  0  to  oo  in  the  last  result  can  be  replaced  by  any  equi- 
valent path  L,  to  be  chosen  later: 


4fk  f  ^2  f 

,=  —\     dOMei 

n  Jo  JL 


p  cos  [p(t-*)  sin  0]e»[(m)+«£-)  cos  o]  dp 


Since  the  right  hand  side  of  this  differential  equation  for  g  is  expressed 
as  a  superposition  of  exponentials  in  |  and  rj  it  is  to  be  expected  that 
a  solution  of  it  can  be  found  in  the  form 


L  &p-p2cos20 
provided  the  path  L  can  be  properly  chosen.  The  path  L,  which  will 
be  fixed  by  a  condition  given  below,  must,  of  course,  avoid  the  pole 
p  =  A/cos2  6. 

It  can  now  be  seen  that  the  function  G*(f,  17,  £;  #,  y,  z)  — 
G(f,  rj,  £;  x,  y,  z)  +  G(|,  ?j,  £;  a?,  y,  —  2)  satisfies  all  the  conditions 
imposed  on  the  Green's  function  employed  in  (9.4.8):  the  sum  on  the 
right  has  the  proper  singularity  in  77  <  0,  £  >  0,  it  satisfies  the 
boundary  condition  (9.4.7)  and 

Gc(f,  ??,  C;  x,  y,  z)  +  Gc(£,  ??,  £;  x,  y,  -  z) 
is  zero  at  £  =  0.  Thus  we  have  for  G*  the  representation: 

-r         ' 

LA/-a;i- 


f=o 

8ft  f*/2        -    f  ens  (<nz  sin  0^  ^(v+i?)+<(e-*)  cos  a]  , 

+  ^"Jo  JL 


ft— p  cos2 
The  substitution  of  this  in  (9.4.8)  gives  finally 


)=^  ff  A««, 
2*JJ 


THE   MOTION  OF  A  SHIP  IN  A  SEAWAY 


283 


A  condition  imposed  on  %Q(x,  y,  z)  is  that  #0(#,  y,  z)  -*  0  as  x  ->  +  oo. 
This  condition  is  satisfied  if  we  take  L  to  be  the  path  shown  in  Fig.  9.4.1 . 


(P) 


c/cos20 

>       i — • — i * 


Fig.  9.4.1.  The  path^L  in  the  p-plane,  with  c  =  k 
The  function  g^  is  given  by 


and  therefore  the  important  quantity  s<fplx  —  <plt  is  given  by 

—       ra2jc      I       s  a2\         1 
(9.4.9)    Wu  -  Vlt  -  -  Cce  9  cos  —  +  I  a  +  -±-\t  +  y    +  StfQx. 

If  this  is  substituted  in  the  equation  (9.3.38)  for  the  surge  we  have 


T—  +  |(T+  — 


The  last  equation  shows  that  in  order  to  keep  s1  bounded  for  all  t  we 
must  take  for  T  the  value 


(9.4.10) 
where 


T  =  - 


igp  cos  0 


284  WATER  WAVES 

In  effect,  T  is  determined  by  the  other  time-independent  term  in  the 
equation  of  motion.  Equation  (9.4.10)  gives  the  thrust  necessary  to 
maintain  the  speed  s0,  or  inversely  it  gives  the  speed  s0  which  corres- 
ponds to  a  given  thrust.  The  integral  in  (9.4.10)  is  called  the  wave 
resistance  integral.  As  one  sees,  it  does  not  depend  on  the  seaway.  The 
integral  can  be  expressed  in  a  simpler  form  as  follows. 
The  function  fax(x,  t/,  0)  is  a  sum  of  integrals  of  the  type 

?;*, 


If  an  integral  of  this  type  is  substituted  in  the  wave  resistance  integral 
we  have 


4      A 

say.  This  is  the  same  as 


(I    IJ 


A      A 

and  we  see  that  /  =  0  if 

/(£,??;#,  t/)  =  —f(x,y;£, 
Therefore 


,  r,)/, 

A       A 

where 

/  =  f  "I2d0  &e  (  igP  C°S  °  ^^  °OS  [p(*-~x">  cos  0]  dp 
1     Jo  JL  g-s§pcos20 

Since  3te  I    is  zero  except  for  the  residue  from  the  integration  along 

JL 
the  semi-circular  path  centered  at  the  point 

g    _   & 

*J  cos2  6       cos2  6  ' 
we  find  from  the  evaluation  of  this  residue  that 

/1==  f*!  r  sec8 

*Q    J  0 


6  e^+rt  "^  e  cos  [k(£-x)  cos  6]  d6. 


THE    MOTION   OF   A    SHIP   IN    A   SEAWAY  285 

We  introduce  MichelPs  notation: 


=  ((hx(x,  y)ekv***«  cos  (kx  sec  6)  dxdy 


Q(0)  =      hx(x,  y)ekv  ^  e  sin  (kx  sec  0)  dxdy 


A 

and  can  then  write 


This  is  the  familiar  formula  of  Michcll  for  the  wave  resistance. 
The  surge  is  given  by 


-*°-\      (P2+Q2)sec30d0. 


g 

A 

Hereafter  we  will  suppose  for  simplicity  that  there  is  no  coupling 

between  (9.3.39)  and  (9.3.40),  so  that     xhdx  =  0.    The  substitution 

JL 

of  (9.4.9)  in  (9.3.39)  therefore  gives  the  following  equation  for  the 
heave: 


2e«  J 


\\ 


A 

The  time  independent  part  of  yr  the  heave  component  of  the  trim, 
we  denote  by  r/f;  it  is  given  by 


(9.4.11  )  g     h 

A 

Here  y*  is  the  vertical  displacement  of  the  center  of  gravity  of  the 
ship  from  its  rest  position  when  moving  in  calm  water.  The  integral 
on  the  right  hand  side  of  (9.4.11)  is  even  more  difficult  to  evaluate 
than  the  wave  resistance  integral. 

The  response  to  the  seaway  in  the  heave  component  is  given  by 


286  WATER   WAVES 


/y      —       [~a2T      I          o2\         1 
—  2gCa  \\hye°  cos  — -  -f|cr+s0  —  h+y 


,** 

~~ 


For  the  case  under  consideration,  the  theory  predicts  that  resonance 
in  the  heave  occurs  when 


g 

The  equation  for  the  pitching  angle  is 


I"  f  (y-y'e)hdA+  \  a*hfa\ 
LJA  JL          J 


cos    —  +  L  +  *-t+y    dxdy 

g     \      e 


The  time  independent  part  of  081,  which  we  denote  by  0*j_  is  given  by 

2fig[  f  (y-y'c)hdA+  [x^hdx\Q^ 

LJA  JL  J 


f  [xh.- 

JA 

The  angle  0*!  is  called  the  angle  of  trim;  it  is  the  angular  displacement 
of  a  ship  which  moves  with  the  speed  SQ  in  calm  water. 
The  oscillatory  part  of  the  heave  031  to  the  sea  is 

dxdy 


ff  [xhy-(y-y'e)hx-]  cos  (  —  +  lo+  9^]t+y  \ 

JJ  I   8       \         g  /         ) 


268 

and  we  see  that  the  theory  predicts  resonance  when 


THE   MOTION   OF   A   SHIP   IN   A   SEAWAY  287 

Of  course,  the  differential  equations  for  yl  and  031  permit  also 
solutions  of  the  type  of  free  undamped  oscillations  of  a  definite  fre- 
quency (in  fact,  having  the  resonant  frequencies  just  discussed)  but 
with  arbitrary  amplitudes  which  could  be  fixed  by  appropriate  initial 
conditions.  This  point  has  been  discussed  at  length  in  the  introduction 
to  this  chapter. 


PART  III 


CHAPTER  10 


Long  Waves  in  Shallow  Water 

10.1.  Introductory  Remarks  and  Recapitulation  of  the  Basic  Equations 

The  basic  theory  for  waves  in  shallow  water  has  already  been  de- 
rived at  length  in  Chapter  2  in  two  different  ways:  one  derivation, 
along  conventional  lines,  proceeded  on  the  basis  of  assuming  the 
pressure  to  be  determined  by  the  hydrostatic  pressure  law  p  = 
&Q(n  —  y)  (see  Fig.  10.1.1),  the  other  by  making  a  formal  develop- 
ment in  powers  of  a  parameter  a;  the  two  theories  are  the  same  in 


y  - 


.Free  Surface 


h(x)  >  0 


Bottom 


Fig.  10.1.1.  Long  waves  in  shallow  water 

lowest  order.  With  one  exception,  the  present  chapter  will  make  use 
only  of  the  theory  to  lowest  order  and  consequently  the  derivation  of 
it  given  in  sections  2  and  3  of  Chapter  2  suffices  for  all  sections  oi  this 
chapter  except  section  9. 

We  recapitulate  the  basic  equations.  In  terms  of  the  horizontal 
velocity  component  u  =  u(x,  t),  and  the  free  surface  elevation 
rj  =rj(x9t)  the  differential  equations  (cf.  (2.2.11),  (2.2.12))  are 

(10.1.1)  ut  +uux  =  -gr)x, 

(10.1.2)  [wfo  +  h)]x  =  -  fit. 

291 


292  WATER   WAVES 

It  is  sometimes  useful  and  interesting  to  make  reference  to  the  gas 
dynamics  analogy,  by  introducing  the  "density"  Q  through 

(10.1.3)  Q  =  Q(r)  +  A), 

and  the  "pressure"  p  by  p  =  \  pdy,  which  in  view  of  the  hydrostatic 
pressure  law  yields  the  relation 

(10.1.4)  Pr==/-£8' 

This  is  an  "adiabatic  law"  with  "adiabatic  exponent"  2  connecting 
pressure  and  density.  As  one  sees,  it  is  the  depth  of  the  water,  essen- 
tially, which  plays  the  role  of  the  density  in  a  gas.  In  terms  of  these 
quantities,  the  equations  (10.1.1)  and  (10.1.2)  take  the  form 

(10.1.5)  g(ut  +  uux)  =  -  px  +  gQhX9 

(10.1.6)  (QU)X  =  -  Qt. 

These  equations,  together  with  (10.1.4),  correspond  exactly  to  the 
equations  of  compressible  gas  dynamics  for  a  one-dimensional  flow  if 
hx  =  0,  i.e.  if  the  depth  of  the  undisturbed  stream  is  constant.  It 
follows  that  a  "sound  speed"  or  propagation  speed  c  for  the  pheno- 
mena governed  by  these  equations  is  defined  by  c  =  V  dp/dp,  as  in 
acoustics,  and  this  quantity  in  our  case  has  the  value 


(10.1.7)  c  =       —  =  Vg(7?  +h)9 

as  we  see  from  (10.1.4)  and  (10.1.3).  Later  on,  we  shall  see  that  it  is 
indeed  justified  to  call  the  quantity  c  the  propagation  speed  since  it 
represents  the  local  speed  of  propagation  of  "small  disturbances" 
relative  to  the  moving  stream.  We  observe  the  important  fact  that  c 
(which  obviously  is  a  function  ofx  and  t)  is  proportional  to  the  square 
root  of  the  depth  of  the  water. 

The  propagation  speed  c(x,  t)  is  a  quantity  of  such  importance  that 
it  is  worthwhile  to  reformulate  the  basic  equations  (10.1.1  )  and  (10.1.2) 
with  c  in  place  of  77.  Since  cx  =  (grjx  +  ghx)/2c  and  ct  =  gr)tj2c  one 
finds  readily 

(10.1.8)  ut  +  uux  +  2ccx  -Hx  =  0, 

(10.1.9)  2ct  +  2ucx  +  cux  =  0, 
with 

(10.1.10)  H  =  gh. 


LONG   WAVES   IN   SHALLOW   WATER  .  293 

The  verification  in  the  general  case  that  the  quantity  c  represents 
a  wave  propagation  speed  requires  a  rather  thorough  study  of  certain 
basic  properties  of  the  differential  equations.  However,  if  we  restrict 
ourselves  to  motions  which  depart  only  slightly  from  the  rest  position 
of  equilibrium  (i.e.  the  state  with  rj  =  0,  u  =  0)  it  is  easy  to  verify 
that  the  quantity  c  then  is  indeed  the  propagation  speed.  From  (10.1.7) 
we  would  have  in  this  case  c  =  c0  +  e(x,  t),  with  c0  =  Vgh  and  s  a 
small  quantity  of  first  order.  We  assume  u  and  its  derivatives  also  to 
be  small  of  first  order  and,  in  addition,  take  the  case  in  which  the 
depth  h  is  constant.  Under  these  circumstances  the  equations  (10.1.8) 
and  (10.1.9)  yield 

(10.1.11)  ut  +  2^  =  0, 

(10.1.12)  2et+cQux  =  0 

if  first  order  terms  only  are  retained.  By  eliminating  s  we  obtain  for 
u  the  differential  equation 

(10.1.13)  utt  -  c*uxx  -  0. 

This  is  the  classical  linear  wave  equation  all  solutions  of  which  are 
functions  of  the  form  u  —  u(x  i  c0t)  and  this  means  that  the  motions 
arc  superpositions  of  waves  with  constant  propagation  speed  c0=  Vgfe. 
The  role  of  the  quantity  c  as  a  propagation  speed  (together  with 
many  other  pertinent  facts)  can  be  understood  most  readily  by  dis- 
cussing the  underlying  integration  theory  of  equations  (10.1.8)  and 
(10.1.9)  by  using  what  is  called  the  method  of  characteristics;  we 
turn  therefore  to  a  discussion  of  this  method  in  the  next  section. 


10.2.  Integration  of  the  Differential  Equations  by  the  Method  of 
Characteristics 

The  theory  of  our  basic  differential  equations  (10.1.8)  and  (10.1.9), 
which  are  of  the  same  form  as  those  in  compressible  gas  dynamics, 
has  been  very  extensively  developed  because  of  the  practical  necessity 
for  dealing  with  the  flow  of  compressible  gases.  The  purpose  of  the 
present  section  is  to  summarize  those  features  of  this  theory  which 
can  be  made  useful  for  discussing  the  propagation  of  surface  waves 
in  shallow  water.  In  doing  so,  extensive  use  has  been  made  of  the 
presentation  given  in  the  book  by  Courant  and  Friedrichs  [C.9];  in 
fact,  a  good  deal  of  the  material  in  sections  10.2  to  10.7,  inclusive, 
follows  the  presentation  given  there. 


294  WATER   WAVES 

The  essential  point  is  that  the  partial  differential  equations  (10.1.8) 
and  (10.1.9)  are  of  such  a  form  that  the  initial  value  problems  asso- 
ciated with  them  admit  of  a  rather  simple  discussion  in  terms  of  a 
pair  of  ordinary  differential  equations  called  the  characteristic  differ- 
ential equations.  We  proceed  to  derive  the  characteristic  equations 
for  the  special  case  in  which  [cf.  (10.1.10)] 

(10.2.1)  Hx  =  m  =  const. 

i.e.  the  case  in  which  the  bottom  slope  is  constant.  In  fact,  this  is  the 
only  case  we  consider  in  this  chapter.  If  we  add  equations  (10.1.8) 
and  (10.1.9)  it  is  readily  seen  that  the  result  can  be  written  in  the 
form: 

9     .       .  a 


(10.2.2)  ^-  +  (u+c)  —  ^.(u  +  2c-  mt)  =  0. 

The  expression  in  brackets  is,  of  course,  to  be  understood  as  a  dif- 
ferential operator.  Similarly,  a  subtraction  of  (10.1.9)  from  (10.1.8) 
yields 

id  d  \ 

(10.2.3)  1  _  +  (u  _  c)  —  1  .  (tt  _  2c  -  mt)  =  0. 
[  ot  ox] 

But  the  interpretation  of  the  operations  defined  in  (10.2.2)  and 
(10.2.3)  is  well  known  (cf.  (1.1.3)):  the  relation  (10.2.2),  for  example, 
states  that  the  function  (u  +  2c  —  mt)  is  constant  for  a  point  moving 
through  the  fluid  with  the  velocity  (u  +  c),  or,  as  we  may  also  put  it, 
for  a  point  whose  motion  is  characterized  by  the  ordinary  differential 
equation  dx/dt  =  u  +  c.  Equation  (10.2.3)  can  be  similarly  interpre- 
ted. That  is,  we  have  the  following  situation  in  the  x,  /-plane:  There  are 
two  sets  of  curves,  Cl  and  C2,  called  characteristics,  which  are  the 
solution  curves  of  the  ordinary  differential  equations 


dx 
Cx: 

(10.2.4) 


Cx  :  —  =  u  +  c,  and 


dx 

£     .  =  U  C 

and  we  have  the  relations 

u  +  2c  —  mt  =  Aj  =  const,  along  a  curve  Cl  and 


1  u  —  2c  —  mt  =  &2  =  const,  along  a  curve  C2. 

Of  course  the  constants  k^  and  k2  will  be  different  on  different  curves 
in  general.  It  should  also  be  observed  that  the  two  families  of  charac- 


LONG   WAVES   IN   SHALLOW   WATER  295 

teristics  determined  by  (10.2.4)  are  really  distinct  because  of  the  fact 

that  c  =  Vg(r)  +  h)  ^  0  since  we  suppose  that  rj  >  —  h,  i.e.  that 
the  water  surface  never  touches  the  bottom. 

By  reversing  the  above  procedure  it  can  be  seen  rather  easily  that 
the  system  of  relations  (10.2.4)  and  (10.2.5)  is  completely  equivalent 
to  the  system  of  equations  (10.1.8)  and  (10.1.9)  for  the  case  of  con- 
stant bottom  slope,  so  that  a  solution  of  either  system  yields  a  solu- 
tion of  the  other.  In  fact,  if  we  set  f(x,  t)  =  u  +  2c  — -  mt  and  ob- 
serve that  f(x,  t)  =  &!  =  const,  along  any  curve  x  =  x(t)  for  which 
dxfdt  =  u  +  c  it  follows  that  along  such  curves 

dx 

(10.2.6)  /,  +  /.  —  -  h  +  (u  +  c)f,  =  0. 

In  the  same  way  the  function  g(x,  t)  =  u  —  2c  —  mt  satisfies  the 
relation 

(10.2.7)  ft  +  (u  -  c)gx  =  0 

along  the  curves  for  which  dx/dt  —  u  —  c.  Thus  wherever  the  curve 
families  C1  and  C2  cover  the  r,  J-plane  in  such  a  way  as  to  form  a  non- 
singular  curvilinear  coordinate  system  the  relations  (10.2.6)  and 

(10.2.7)  hold.  If  now  equations  (10.2.6)  and  (10.2.7)  are  added  and 
the  definitions  of  /(#,  t )  and  g(x,  t)  are  recalled  it  is  readily  seen  that 
equation    (10.1.8)   results.    By   subtracting    (10.2.7)    from    (10.2.6) 
equation  (10.1.9)  is  obtained.  In  other  words,  any  functions  u  and  c 
which  satisfy  the  relations   (10.2.4)  and   (10.2.5)  will  also  satisfy 

(10.1.8)  and  (10.1.9)  and  the  two  systems  of  equations  are  therefore 
now  seen  to  be  completely  equivalent. 

As  we  would  expect  on  physical  grounds,  a  solution  of  the  original 
dynamical  equations  (10.1.8)  and  (10.1.9)  could  be  shown  to  be 
uniquely  determined  when  appropriate  initial  conditions  (for  t  =  0, 
say)  are  prescribed;  it  follows  that  a  solution  of  (10.2.4)  and  (10.2. 
5)  is  also  uniquely  determined  when  initial  conditions  are  prescribed 
since  we  know  that  the  two  systems  of  equations  are  equivalent. 

At  first  sight  one  might  be  inclined  to  regard  the  relations  (10.2.4) 
and  (10.2.5)  as  more  complicated  to  deal  with  than  the  original  dif- 
ferential equations,  particularly  since  the  right  hand  sides  of  (10.2.4) 
are  not  known  in  advance  and  hence  the  characteristic  curves  are  also 
not  known:  they  must,  in  fact,  be  determined  in  the  course  of  deter- 
mining the  unknown  functions  u  and  c  which  constitute  the  desired 
solution.  Nevertheless,  the  formulation  of  our  problems  in  terms  of 


296 


WATER   WAVES 


the  characteristic  form  is  quite  useful  in  studying  properties  of  the 
solutions  and  also  in  studying  questions  referring  to  the  appropriate- 
ness of  various  boundary  and  initial  conditions.  It  is  useful  to  begin 
by  describing  briefly  a  method  of  determining  the  characteristics  and 
thus  the  solution  of  a  given  problem  by  a  method  of  successive  approx- 
imation which  at  the  same  time  makes  possible  a  number  of  useful 
observations  and  interpretations  regarding  the  role  played  by  the 
characteristics  in  general.  Let  us  for  this  purpose  consider  a  problem 
in  which  the  values  of  the  velocity  u  and  the  surface  elevation  rj 
(or,  what  amounts  to  the  same  thing,  the  propagation  or  wave  speed 
c  =  Vg(?7  +  h))  are  prescribed  for  all  values  of  a?  at  the  initial  instant 
t  =  0.  We  wish  to  calculate  the  solution  for  t  >  0  by  determining  u 
and  c  through  use  of  (10.2.4)  and  (10.2.5)  and  the  given  initial  condi- 
tions. At  t  =  0  we  assume  that 


(10.2.8) 


u(x,  0)  =  u(x) 
(x,  0)  =  c(x) 

in  which  u(x)  and  ~c(x)  are  given  functions.  We  can  approximate  the 
values  01  u  and  c  for  small  values  of  t  as  follows:  consider  a  scries  of 
points  on  the  #-axis  (cf.  Fig.  10.2.1)  a  small  distance  6x  apart.  At  all 
of  these  points  the  values  of  u  and  c  are  known  from  (10.2.8).  Conse- 
quently the  slopes  of  the  characteristics  Cl  and  C2  at  these  points  are 


t 


Fig.  10.2.1.  Integration  by  finite  differences 

known  from  (10.2.4).  From  the  points  1,  2,  3,  4  straight  line  segments 
with  these  slopes  are  drawn  until  they  intersect  at  points  5,  6,  and  7, 
and  if  dx  is  chosen  sufficiently  small  it  is  reasonable  to  expect  that 


LONG   WAVES    IN    SHALLOW   WATER  297 

the  positions  of  these  points  will  be  good  approximations  to  the  inter- 
sections of  the  characteristics  issuing  from  the  points  1,  2,  3,  4  since 
we  are  simply  replacing  short  segments  of  these  curves  by  their 
tangents.  The  values  of  both  x  and  t  at  points  5,  6,  and  7  are  now 
known— they  can  be  determined  graphically  for  example— and 
through  the  use  of  (10.2.5)  and  the  initial  conditions  we  can  also 
determine  the  approximate  values  of  u  and  c  at  these  points.  For  this 
purpose  we  observe  that  along  any  particular  segment  issuing  from 
the  points  1,  2,  3  or  4  the  values  of  u  +  2c  •—  mt  and  u  —  2c  —  rnt 
are  known  constants  since  the  values  of  u  and  c  are  fixed  by  (10.2.8) 
for  t  =  0;  hence  we  have 

I  along  Cx:  u  +  2c  —  mt  =  u •  -f-  2c,  and 
\  along  C2:  u  —  2c  —  mt  =  u  —  2c. 

At  the  points  5,  6,  and  7  we  know  the  values  of  t  and  hence  (10.2.9) 
furnishes  two  independent  linear  equations  for  the  determination  of 
the  values  of  u  and  c  at  each  of  these  points.  Once  u  and  c  are  known 
at  points  5,  6,  and  7  the  slopes  of  the  characteristics  issuing  from  these 
points  can  be  determined  once  more  from  (10.2.4)  and  the  entire  pro- 
cess can  be  carried  out  again  to  yield  the  additional  points  8  and  9 
and  the  approximate  values  of  u  and  c  at  these  points.  In  this  way 
we  can  approximate  the  values  of  u  and  c  at  the  points  of  a  net  over 
a  certain  region  of  the  x,  <-plane,  and  can  then  obtain  approximate 
values  for  u  and  c  at  any  points  in  the  same  region  either  by  inter- 
polation or  by  refining  the  net  inside  the  region.  It  is  quite  plausible 
and  could  be  proved  mathematically  that  the  above  process  would 
converge  as  dx  ->Q  to  the  unique  solution  of  (10.2.4)  and  (10.2.5) 
corresponding  to  the  given  initial  conditions  for  sufficiently  small 
values  of  t  (i.e.  for  a  region  of  the  #,  /-plane  not  too  far  from  the  #-axis) 
provided  that  the  prescribed  initial  values  of  u  and  c  are  sufficiently 
regular  functions  of  x— for  example,  if  they  have  piecewise  con- 
tinuous first  derivatives. 

It  should  be  clear  that  once  the  characteristics  are  known  the  values 
of  u  and  c  for  all  points  of  the  x9  t  -plane  covered  by  them  are  also 
known,  since  the  constants  A:x  and  k2  in  (10.2.5)  are  known  on  each 
characteristic  through  the  initial  data  and  hence  the  values  of  u  and  c 
for  any  point  (x,  t)  can  be  calculated  by  solving  the  linear  equation 
(10.2.5)  for  the  characteristics  through  that  point.  This  statement  of 
course  implies  that  each  one  of  the  two  families  of  characteristics 
covers  a  region  of  the  x,  /-plane  simply  and  that  no  two  members  of 


298 


WATER   WAVES 


different  families  are  tangent  to  each  other— in  other  words  it  is 
implied  that  the  two  families  of  characteristics  form  a  regular  curvi- 
linear coordinate  system  over  the  region  of  the  x,  J-plane  in  question. 
One  of  the  points  of  major  interest  in  the  later  discussion  centers 
around  the  question  of  determining  where  and  when  the  character- 
istics cease  to  have  this  property,  and  of  interpreting  the  physical 
meaning  of  such  occurrences. 

The  method  of  finite  differences  used  above  to  determine  the  cha- 
racteristics can  be  interpreted  in  such  a  way  as  to  throw  a  strong  light 
on  the  physical  properties  of  the  solution.  Consider  the  point  10  of 
Fig.  10.2.1  for  example.  We  recall  that  the  approximate  values  ulo 
and  c10  of  u  and  c  at  point  10  were  obtained  through  making  use  of  the 
initial  values  of  u  and  c  at  points  1,  2,  3,  4  on  the  #-axis  only,  and 
furthermore  that  the  values  ulo  and  c10  required  the  use  of  points  con- 
fined solely  to  the  region  within  the  approximate  characteristics  join- 
ing point  10  with  points  1  and  4.  Since  the  finite  difference  scheme 
outlined  above  converges  as  dx  ->  0  to  yield  the  exact  characteristics 
we  are  led  to  make  the  following  important  statement:  the  values  of  u 
and  c  at  any  point  P(x,  t)  within  the  region  of  existence  of  the  solution 
are  determined  solely  by  the  initial  values  prescribed  on  the  segment  of 
the  x-axis  which  is  subtended  by  the  two  characteristics  issuing  from  P. 


Range  of  influence  of  Q 


C,  EL 


Domain  of 
\      determinacy 


J 


Domain  of   dependence    of    P 


Fig.  10.2.2.  Domain  of  dependence  and  range  of  influence 


In  addition,  the  two  characteristics  issuing  from  P  are  also  determined 
solely  by  the  initial  values  on  the  segment  subtended  by  them.  Such 
a  segment  of  the  #-axis  is  often  called  the  domain  of  dependence  of  the 


LONG   WAVES   IN   SHALLOW   WATER  299 

point  P.  Correspondingly  we  may  define  the  range  of  influence  of  a 
point  Q  on  the  #-axis  as  the  region  of  the  x,  £-plane  in  which  the  values 
of  u  and  c  are  influenced  by  the  initial  values  assigned  to  point  Q.  In 
Fig.  10.2.2  we  indicate  these  two  regions.  It  is  also  useful  on  occasion 
to  introduce  the  notion  of  domain  of  determinacy  relative  to  a  given 
domain  of  dependence.  It  is  the  region  in  which  the  motion  is  deter- 
mined solely  by  the  data  over  a  certain  segment  of  the  #-axis.  These 
regions  arc  outlined  by  characteristic  curves,  as  indicated  in  Fig. 
10.2.2,  in  an  easily  understandable  fashion  in  view  of  the  discussion 
above. 

We  are  now  in  a  position  to  understand  why  it  is  appropriate  to 
call  the  quantity  c  the  propagation  or  wave  speed.  To  this  end  we 
suppose  that  a  certain  motion  of  water  exists  at  a  definite  time,  which 
we  take  to  be  t  —  0.  This  means,  of  course,  that  u  and  c  are  known  at 
that  time,  and,  as  we  have  just  seen,  the  motion  would  be  uniquely 
determined  for  t  >  0.  However,  we  raise  the  question:  what  difference 
would  there  be  in  the  subsequent  motion  if  we  created  a  disturbance 
in  some  part  of  the  fluid,  say  over  a  segment  QxQ2  °f  *he  #-axis  (cf. 
Fig.  10.2.3)?  This  amounts  to  asking  for  a  comparison  of  two  solutions 
of  our  equations  which  differ  only  because  of  a  difference  in  the  initial 


0,  Q2 


Fig.  10.2.3.  Propagation  of  disturbances 

conditions  over  the  segment  Q1Q2.  Our  whole  discussion  shows,  that  the 
two  solutions  in  question  would  differ  only  in  the  shaded  region  of 
Fig.  10.2.3,  which  comprises  all  points  of  the  x,  J-plane  influenced  by 
the  data  on  the  segment  QxQ^  and  which  is  bounded  by  characteristics 
C2  and  Cx  issuing  from  the  endpoints  Qt  and  Q2  of  the  segment.  These 
curves,  however,  satisfy  the  differential  equations  dx/dt  =  u  —  c, 
dx/dt  =  u  +  c.  Since  u  represents  the  velocity  of  the  water,  it  is  then 


300  WATER  WAVES 

clear  that  c  represents  the  speed  relative  to  the  flowing  stream  at 
which  the  disturbance  on  the  segment  QXQ2  spreads  over  the  water. 
This  implies  that  the  data  in  our  two  problems  really  differ  at  points 
Qx  and  Q2  and  that  these  differences  persist  along  the  characteristics 
issuing  from  these  points.  Actually,  only  discontinuities  in  derivatives 
at  Ql  and  Q2  (and  not  of  the  functions  themselves)  are  permitted  in  the 
above  theory,  and  it  could  be  shown  that  such  discontinuities  would 
never  smooth  out  entirely  along  the  characteristics  Cl  and  C2.  We  are 
therefore  justified  in  referring  to  the  quantity  c  =  Vg(rj  +  h)  as  the 
(local)  propagation  speed  of  small  disturbances— that  is,  small  in  the 
sense  that  only  discontinuities  in  derivatives  occur  at  the  front  of  a 
disturbance. 

10.3.  The  Notion  of  a  Simple  Wave 

There  is  an  important  class  of  problems  in  which  the  theory  of 
characteristics  as  presented  in  the  preceding  section  becomes  parti- 
cularly simple.  These  are  the  problems  in  which  (1)  the  initial  un- 
disturbed depth  h  of  the  water  is  constant  so  that  the  quantity  m  in 
(10.2.1)  (cf.  also  (10.1.10))  is  zero,  (2)  the  water  extends  from  the 
origin  to  infinity  at  least  in  one  direction,  say  in  the  positive  direction 
of  the  a?-axis,  and  (3)  the  water  is  either  at  rest  or  moves  with  constant 
velocity  and  the  elevation  of  its  free  surface  is  zero  at  the  time  t  =  0. 
In  other  words,  the  water  is  in  a  uniform  state  at  time  t  =  0  such  that 
u  =  u0  =  const,  and  c  =  CQ  =  Vgh  =  const,  at  that  instant.  Our 
discussion  from  here  on  is  modeled  closely  on  the  discussion  given  by 
Courant  and  Friedrichs  [C.9],  Chapter  III. 

We  now  suppose  that  a  disturbance  is  initiated  at  the  origin  x  =  0 
so  that  either  the  particle  velocity  u,  or  the  surface  elevation  r/  (or  the 
wave  velocity  c  =  Vg(7/  +  h))  changes  with  the  time  in  a  prescribed 
manner.*  That  is,  a  disturbance  at  one  point  in  the  water  propagates 
into  water  of  constant  depth  and  uniform  velocity.  Under  these 
circumstances  we  show  that  one  of  the  two  families  of  characteristics 
furnished  by  (10.2.4)  consists  entirely  of  straight  lines  along  each  of  which 
u  and  c  are  constant.  The  corresponding  motion  we  call  a  simple  wave. 

*  One  might  accomplish  this  experimentally  in  a  tank  as  follows:  To  obtain 
a  prescribed  velocity  u  at  one  point  it  would  only  be  necesary  to  place  a  vertical 
plate  in  the  water  extending  from  the  surface  of  the  water  to  the  bottom  of  the 
tank  and  to  move  it  with  the  prescribed  velocity.  To  change  r]  at  one  point  water 
might  be  either  poured  into  the  tank  or  pumped  out  of  it  at  that  point  at  an 
appropriate  rate. 


LONG   WAVES   IN   SHALLOW   WATER 


301 


Our  statement  is  an  immediate  consequence  of  the  following  funda- 
mental fact:  if  the  values  of  u  and  c  on  any  characteristic  curve,  C®  say 
(i.e.  a  solution  curve  of  the  first  of  the  two  ordinary  differential  equations 
(10.2.4)),  are  constant,  then  CJ  is  a  straight  line  and  furthermore  it  is 
embedded  in  a  family  of  straight  line  characteristics  along  each  of  which 
u  and  c  are  constant,  at  least  in  a  region  of  the  x,  f -plane  where  u(x,  t) 
and  c(x,  t)  are  without  singularities  and  which  is  covered  by  the 
two  distinct  families  of  characteristics.  The  proof  is  easily  given.  To 
begin  with,  the  curve  CJ  is  a  straight  line  if  u  and  c  are  constant  along 
it,  since  the  slope  of  the  curve  is  constant  in  that  case  from  (10.2.4). 
Next,  let  Cl  be  another  characteristic  near  to  CJ.  We  consider  any  two 
points  A$  and  BQ  on  CJ  together  with  the  characteristics  of  the  family 
C2  through  AQ  and  BQ  and  suppose  that  the  latter  characteristics 
intersect  Cl  at  points  A  and  B  (cf.  Fig.  10.3.1 ):  To  prove  our  statement 


Fig.  10.3.1.  Region  containing  a  straight  characteristic 

we  need  only  show  that  u(A  )~u(B)  and  c(A  )  =  c(B)  since  then  u  and  c 
would  be  constant  on  Cl  (because  of  the  fact  that  A  and  B  are  any 
arbitrary  points  on  CJ  and  hence  the  slope  of  the  curve  Cr  would  be 
constant,  just  as  was  argued  for  Cj.  We  have  u(A0)  =  u(BQ)  and 
c(AQ)  =  c(BQ)  and  consequently  we  may  write 


(10  81) 


u  A  —  2cA  —  UA    —  2cA  , 
UB  -  2cB  =  UBQ  -  2<?Bo  = 


by  making  use  of  the  second  relation  of  (10.2.5)  (which  holds  along 
the  characteristics  C2)  and  observing  that  m  =  0  since  the  original 


802 


WATER    WAVES 


depth  of  the  water  is  assumed  to  be  constant.  Next  we  make  use  of 
the  first  relation  of  (10.2.5)  for  Cl  to  obtain 

(10.3.2)  UA  +  2cA    =  UB  +  2cB. 
But  from  (10.3.1)  we  have 

(10.3.3)  UA  -  2cA  =  UB  -  2cB, 

and  (10.3.2)  and  (10.3.3)  are  obviously  satisfied  only  if  UA  —  UB 
and  CA  =  CB  .  Our  statement  is  therefore  proved. 

The  problems  formulated  in  the  first  paragraph  of  this  section  are 
at  once  seen  to  have  solutions  (at  least  in  certain  regions  of  the 
x9  /-plane)  of  the  type  we  have  just  defined  as  simple  waves  since 
there  is  a  region  near  the  #-axis  in  the  x,  /-plane  throughout  which  the 
particle  velocity  u  and  wave  speed  c  arc  constant,  and  in  which  there- 
fore the  characteristics  are  two  sets  of  parallel  straight  lines.  The  cir- 
cumstances are  illustrated  in  Fig.  10.3.2  below:  There  is  a  zone  /  along 


Fig.  10.3.2.  A  simple  wave 

the  07-axis  which  might  be  called  the  zone  of  quiet*  inside  which  the 
characteristics  are  obviously  straight  lines  x  ±  c0t  =  const.  (These 
lines  are  not  drawn  in  the  figure).  This  region  is  terminated  on  the 
upper  side  by  an  "initial  characteristic"  x  —  erf  which  divides  the 

*  In  a  "zone  of  quiet"  we  permit  the  particle  velocity  u  to  be  a  non  zero 
constant,  but  the  free  surface  elevation  r\  is  taken  to  be  zero  in  such  a  region. 
In  case  u  —  UQ  =  const.  7^  0  initially,  the  motion  can  be  thought  of  as  observed 
from  a  coordinate  system  moving  with  that  velocity;  thus  there  is  no  real  loss 
of  generality  in  assuming  UQ  =  0,  as  we  frequently  do  in  the  following. 


LONG    WAVES    IN   SHALLOW   WATER  303 

region  of  quiet  from  the  disturbed  region  above  it.  The  physical  inter- 
pretation of  this  is  of  course  that  the  disturbance  initiated  at  the 
time  /  =  0  propagates  into  the  region  of  quiet,  and  the  water  at  any 
point  remains  unaffected  until  sufficient  time  has  elapsed  to  allow 
the  disturbance  to  reach  that  point.  The  exact  nature  of  the  motion 
in  the  disturbed  region  is  determined,  of  course,  by  the  character  of 
the  disturbance  created  at  the  point  x  =  0,  i.e.,  by  appropriate  data 
prescribed  along  the  /-axis.*  One  set  of  characteristics,  i.e.,  the  set 
containing  the  initial  characteristic  C?,  therefore  consists  of  straight 
lines.  (That  the  characteristics  C2  in  the  zone//  are  necessarily  curved 
lines  and  not  straight  lines  can  be  seen  from  the  fact  that  they  would 
otherwise  be  the  continuations  of  the  straight  characteristics  from  the 
zone  /  of  quiet  and  hence  the  zone  //  would  also  be  a  zone  of  quiet,  as 
one  sees  immediately).  Furthermore,  the  set  of  straight  characteristics 
C\  in  zone  //  is  completely  determined  by  appropriate  conditions  pre- 
scribed at  x  —  0  for  all  /,  i.e.,  along  the  /-axis.  What  these  conditions 
should  be  can  be  inferred  from  the  following  discussion.  Consider  any 
straight  characteristic  issuing  from  a  point  /  =  T  on  the  /-axis.  We 
know  that  the  slope  dx/dt  of  this  straight  line  is  given  in  view  of 
(10.2.4),  by 

dx 

(10.3.4)  —  =  u(r)  +  c(r). 

Suppose  now  that  there  is  a  curved  characteristic  C2  going  back 
from  /  =  T  on  the  /-axis  to  the  initial  characteristic  C°  (see  the  dotted 
curve  in  Fig.  10.3.2).  We  have  the  following  relation  from  (10.2.5): 

(10.3.5)  u(r)  -  2c(r)  -  u0  -  2c0, 

in  which  u0  and  CQ  are  the  known  values  of  u  and  c  in  the  zone  of  quiet. 
Hence  the  slope  of  any  of  the  straight  characteristics  issuing  from  the 
/-axis  can  be  given  in  either  of  the  two  forms: 

dx        1  r 
-=;-[3u(T)-uQ]  +  cQ,  or 

(10.3.6) 

(UV 

-  .  8c(r) 

as  one  sees  from  (10.3.4)  and  (10.3.5).  Thus  if  either  u(r)  or  c(r)  is 

*  Our  discussion  in  the  preceding  section  centered  about  the  initial  value 
problem  for  the  case  in  which  the  initial  data  are  prescribed  on  the  #-axis,  but  one 
sees  readily  that  the  same  discussion  would  apply  with  only  slight  modifications 
to  the  present  case,  in  which  what  is  commonly  called  a  boundary  condition  (i.e. 
at  the  boundary  point  x  —  0),  rather  than  an  initial  condition,  is  prescribed. 


304 


WATER  WAVES 


given,  i.e.  if  either  u  or  c  is  prescribed  along  the  J-axis,  then  the  slopes 
of  the  straight  characteristics  Cl  and  with  them  the  characteristics  Cl 
themselves  are  determined.  Since  we  know,  from  (10.3.5),  the  values 
of  both  u  and  c  along  the  /-axis  if  either  one  is  given,  and  since  u  and  c 
are  clearly  constant  along  the  straight  characteristics,  it  follows  that  we 
know  the  values  of  u  and  c  throughout  the  entire  disturbed  region— in 
other  words,  the  motion  is  completely  determined. 

So  far,  we  have  considered  only  the  case  in  which  the  curved  cha- 
racteristics (i.e.,  those  of  the  type  C2)  which  issue  from  the  boundary 
x  =  c0J  of  the  disturbed  region  actually  reach  the  /-axis.  This,  however, 
need  not  be  the  case.  Suppose,  for  example,  that  UQ  is  positive  and 

u0  >  c0  =  Vg/T.  In  this  case  the  slope  dxjdt  of  the  curves  C2  is  positive, 
and  we  cannot  expect  that  they  will  turn  to  the  left,  as  in  Fig.  10.3.2. 
Indeed,  in  such  a  case  one  does  not  expect  that  a  disturbance  will  pro- 
pagate upstream  (that  is,  to  the  left  in  our  case)  since  the  stream  velo- 
city is  greater  than  the  propagation  speed.  In  gas  dynamics  one  would 
say  that  the  flow  is  supersonic,  while  in  hydraulics  the  flow  is  said 
to  be  supercritical.  One  could  also  look  at  the  matter  in  another  way: 
For  not  too  large  values  of  t  the  velocity  u  can  be  expected  to  remain 
supersonic  and  hence  for  such  values  of  t  both  sets  of  characteristics 
issuing  from  the  f-axis  would  go  into  the  right  half  plane  (u  being 
again  taken  positive).  Thus  we  would  have  the  situation  indicated 
in  Fig.  10.3.3,  in  which  a  segment  of  the  /-axis  is  subtended  by  two 
t 


*»<V».c0)t 


Fig.  10.8.3.  The  supercritical  case 


LONG  WAVES   IN  SHALLOW   WATEE  805 

characteristics  drawn  backward  from  P.  In  this  case,  as  in  the  case  of 
the  initial  value  problem  treated  in  the  preceding  section,  we  must 
prescribe  the  values  of  both  u  and  c  along  the  /-axis.  If  we  do  so,  then 
the  solution  is  once  more  determined  through  (10.8.4)  and  the  fact 
that  u  +  2c  is  constant  along  one  set  of  characteristics  and  u  —  2c  is 
constant  along  the  other. 

In  either  of  our  two  cases,  i.e.  of  subcritical  or  supercritical  flow, 
we  see  therefore  that  the  simple  wave  can  be  determined.  One  sees 
also  how  useful  the  formulation  in  terms  of  the  characteristics  can  be 
in  determining  appropriate  subsidiary  conditions  such  as  boundary 
conditions. 

If  we  wish  to  know  the  values  of  u  and  c  for  any  particular  time 
t  =  tQ9  once  the  simple  wave  configuration  is  determined,  we  need 
only  draw  the  line  t  =  J0  and  observe  its  intersections  with  the 
straight  characteristics  since  the  values  of  u  and  c  are  presumably 
known  on  each  one  of  the  latter.  Thus  u  and  c  would  be  known  as 
functions  of  x  for  that  particular  time.  Of  course,  the  surface  elevation 
j\  would  also  be  known  from 

c  =  Vg(h+fi). 

10.4.  Propagation  of  disturbances  into  still  water  of  constant  depth 

In  the  preceding  section  we  have  seen  how  the  method  of  charac- 
teristics leads  to  the  notion  of  a  simple  wave  in  terms  of  which  we  can 
describe  with  surprising  ease  the  propagation  of  a  disturbance  initiat- 
ed at  a  point  into  water  of  constant  depth  moving  with  uniform  speed. 
In  the  present  section  we  consider  in  more  detail  the  character  of  the 
simple  waves  which  occur  in  two  important  special  cases.  We  assume 
always  that  the  pulse  is  initiated  at  x  =  0  and  that  it  then  propagates 
in  the  positive  ^-direction  into  still  water.  Thus  we  are  considering 
cases  in  which  the  flow  is  subcritical  at  the  outset. 

One  of  the  most  striking  and  important  features  of  our  whole  dis- 
cussion is  that  there  is  an  essential  difference  between  the  propagation 
of  a  pulse  which  is  created  by  steadily  decreasing  the  surface  elevation 
rf  at  x  =  0  and  of  a  pulse  which  results  by  steadily  increasing  the  ele- 
vation at  x  =  0.  If  the  pulse  is  created  by  initiating  a  change  in  the 
particle  velocity  u  at  x  =  0  (which  might  be  achieved  simply  by 
moving  a  vertical  barrier  at  x  =  0  with  the  prescribed  particle  velo- 
city) instead  of  by  changing  the  surface  elevation  rj  the  same  typical 
differences  will  result  if  u  is  in  the  first  case  decreased  from  zero 
through  negative  values,  and  in  the  other  case  is  gradually  increased 


306  WATER   WAVES 

so  that  it  becomes  positive  (i.e.  if  the  particles  at  x  =  0  are  given  in 
the  first  case  a  negative  acceleration  and  in  the  second  case  a  positive 
acceleration.)  The  qualitative  difference  between  the  two  cases  from 
the  physical  point  of  view  is  of  course  that  in  the  first  case  it  is  a 
depression  in  the  water  surface  and  in  the  second  case  an  elevation 
above  the  undisturbed  surface—  sometimes  referred  to  later  on  as  a 
hump—  which  propagates  into  still  water. 

If  we  were  to  consider  waves  of  very  small  amplitude  so  that  we  might 
linearize  our  equations  (as  was  done  in  deriving  equation  (10.1.13)) 
there  would  be  no  essential  qualitative  distinction  between  the  motions 
in  the  two  cases;  that  there  is  actually  a  distinction  between  the  two 
is  a  consequence  of  the  nonlinearity  of  the  differential  equations. 

In  the  preceding  section  we  have  seen  that  the  motions  in  either  of 
our  two  cases  can  be  described  in  the  #,  /-plane  by  means  of  a  family  of 
straight  characteristics  which  issue  from  the  /-axis.  In  Figure  10.4.1 
we  show  these  characteristics  together  with  a  curve  indicating  a  set 


of  prescribed  values  for  c  —  Vg(A+^)  =  <*(0  at  oc  =  0,  which  in 
turn  result  from  prescribed  values  of  rj  at  that  point.  We  assume  that 
u  —  UQ  =  0  in  the  zone  of  quiet  /.  Hence  the  slope  docjdt  of  any  straight 
characteristic  issuing  from  a  point  t  ~  r  on  the  /-axis  is  given,  in  ac- 
cordance with  (10.3.6)  by 

dx 

(10.4.1)  _=8c(T)-2cc. 

When  r  is  varied  (10.4.1  )  yields  the  complete  set  of  straight  characteris- 
tics in  the  zone  //.  The  values  oft/  and  c  along  the  same  characteristic 
are  constant  (as  we  have  seen  in  the  preceding  section),  so  that  the 
value  of  u  along  a  characteristic  is  determined,  from  (10.3.5)  by 

(10.4.2)  u(r)  -  2[c(r)  -  c0], 
since  UQ  is  assumed  to  be  zero  and  c(r)  is  given. 

We  are  now  in  a  position  to  note  a  crucial  difference  between  the 
two  cases  described  above.  In  the  first  of  the  two  cases  —  i.e.  that  of  a 
depression  moving  into  still  water—  the  elevation  rj(t)  at  x  —  0  is 
assumed  to  be  a  decreasing  function  of  /  so  that  c(t)  also  decreases 
with  increase  of  /.  It  follows  that  the  slopes  dxjdt  of  the  straight  line 
characteristics  as  given  by  (10.4.1)  decrease  as  /  increases*  so  that  the 
family  of  straight  characteristics  diverge  on  moving  out  from  the 

*  One  should  observe  that  decreasing  values  of  dx/dt  mean  that  the  charac- 
teristics make  increasing  angles  with  the  #-axis,  i.e.  that  they  become  steeper  with 
respect  to  the  horizontal. 


LONG   WAVES   IN    SHALLOW   WATER 


307 


tf-axis.  (This  is  the  case  indicated  in  Fig.  10.4.1).  In  the  second  case, 
however,  the  value  of  77  and  thus  of  c  is  assumed  to  be  an  increasing 
function  of  t  at  x  =  0  so  that  the  straight  characteristics  must  cven- 


c(0,t) 


Fig.  10.4.1.  Propagation  of  pulses  into  still  water 


tually  intersect  — in  fact,  they  will  have  an  envelope  in  general—  and 
this  in  turn  means  that  our  problem  can  riot  be  expected  to  have  a 
continuous  solution  for  values  of  ^r  and  t  beyond  those  for  which 
such  intersections  exist.  In  the  first  case  the  motion  is  continuous 
throughout.  What  happens  in  the  second  case  beyond  the  point 
where  the  solution  is  continuous  can  not  be  discussed  mathematically 
until  we  have  widened  our  basic  theory,  but  in  terms  of  the  physical 
behavior  of  the  water  we  might  expect  the  wave  to  break,  or  to  devel- 
op what  is  called  a  bore,*  some  time  after  the  solution  ceases  to  be 
continuous.  In  later  sections  we  propose  to  discuss  the  question  of  the 
development  of  breakers  and  bores  in  some  detail. 

The  two  cases  discussed  above  are  the  exact  analogues  of  two  cases 
well  known  in  gas  dynamics:  Consider  a  long  tube  filled  with  gas  at 
rest  and  closed  by  a  piston  at  one  section.  If  the  piston  is  moved  away 
from  the  gas  with  increasing  speed  in  such  a  way  as  to  cause  a 
rarefaction  wave  to  move  into  the  quiet  gas,  then  a  continuous  motion 
results.  However,  if  the  piston  is  pushed  with  increasing  speed  into 
the  gas  so  as  to  create  a  compression  wave,  then  such  a  wave  always 

*  In  certain  estuaries  in  various  parts  of  the  world  the  incoming  tides  from  the 
ocean  are  sometimes  observed  to  result  in  the  formation  of  a  nearly  vertical  wall 
of  water,  called  a  bore,  which  advances  more  or  less  unaltered  in  form  over 
quite  large  distances.  What  is  called  a  hydraulic  jump  is  another  phenomenon 
of  the  same  sort.  Such  phenomena  will  be  discussed  in  detail  later  on. 


308 


WATER   WAVES 


develops  eventually  into  a  shock  wave.  That  is,  the  development  of 
a  shock  in  gas  dynamics  is  analogous  to  the  development  of  a  bore 
(and  also  of  a  hydraulic  jump)  in  water. 


10.5.  Propagation  of  depression  waves  into  still  water  of  constant  depth 

In  this  section  we  give  a  detailed  treatment  of  the  first  type  of 
motion  in  which  a  depression  of  the  water  surface  propagates  into 
still  water.  However,  it  is  interesting  and  instructive  to  prescribe  the 
disturbance  in  terms  of  the  velocity  of  the  water  rather  than  in  terms 
of  the  surface  elevation.  We  assume,  in  addition,  that  the  velocity  is 
prescribed  by  giving  the  displacement  x  =  x(t)  of  the  water  particles 
originally  in  the  vertical  plane  at  x  =  0,*  and  this,  as  we  have  remark- 
ed before,  could  be  achieved  experimentally  simply  by  moving  a  ver- 
tical plate  at  the  end  of  a  tank  in  such  a  way  that  its  displacement 
is  x(t).**  Figure  10.5.1  indicates  the  straight  characteristics  which 


t=  const. 


Fig.  10.5.1.  A  depression  wave 

initiate  on  the  "piston  curve"  x  =  x(t).  The  piston  is  assumed  to 
start  from  rest  and  move  in  the  negative  direction  with  increasing 
speed  until  it  reaches  a  certain  speed  w  <  0,  after  which  the  speed 
remains  constant.  That  is,  xt  decreases  monotonically  from  zero  at 
t  =  0  until  it  attains  the  value  w,  after  which  it  stays  constant  at  that 
value.  In  Fig.  10.5.1  this  point  is  marked  B;  clearly  the  piston  curve  is 

*  In  our  theory,  it  should  be  recalled,  the  particles  originally  in  a  vertical 
plane  remain  always  in  a  vertical  plane. 

**  Moving  such  a  plate  at  the  end  of  a  tank  of  course  corresponds  in  gas  dyna- 
mics to  moving  a  piston  in  a  gas-filled  tube. 


LONG   WAVES   IN    SHALLOW   WATER  309 

a  straight  line  from  there  one.  At  any  point  A  on  the  piston  curve  we 
have  UA  =  xt(t),  corresponding  to  the  physical  assumption  that  the 
water  particles  in  contact  with  the  piston  remain  in  contact  with  it 
and  thus  have  the  same  velocity.  If  we  consider  the  curved  character- 
istic drawn  from  A  back  to  the  initial  characteristic  Cj  which  termina- 
tes the  zone  /  of  rest  we  obtain  from  (10.3.5)  the  relation 

(10.5.1)  CA  =  Ki  +  *o. 

since  in  our  case  u0  —  0.  The  slope  of  the  straight  characteristic  at  A 
is  thus  given  by  (cf.  (10.4.1)): 

dx        3 
— 


(10.5.2)  —  =  -UA 


Since  we  have  assumed  that  UA  —  xt(t)  always  decreases  as  t  increases 
until  xt  ~  w  it  follows  from  (10.5.2)  that  dx/dt  also  decreases  as  t  in- 
creases in  this  range  of  values  of  t  so  that  the  characteristics  diverge 
as  they  go  outward  from  the  piston  curve.  Beyond  the  point  B  the 
straight  characteristics  are  parallel  straight  lines,  since  UA  =  w  — 
const,  on  that  part  of  the  piston  curve,  and  the  state  of  the  water  is 
therefore  constant  in  the  zone  marked  ///  in  Fig.  10.5.1.  The  zone  // 
is  thus  a  region  of  non-constant  state  connecting  two  regions  of  differ- 

ent constant  states.  Since  CA  =  ^/g(h-\-r]A),  where  r\A  refers  to  the 
elevation  of  the  water  surface  at  the  piston,  it  follows  from  (10.5.1) 
that  t]A  decreases  in  the  zone  //  as  t  increases,  i.e.  the  water  surface  at 
the  "piston"  moves  downward  as  the  piston  moves  to  the  left,  since  we 
assume*  that  UA  decreases  as  A  moves  out  along  the  piston  curve.  Since 
u  and  c  are  constant  along  any  straight  characteristic  it  is  not  difficult 
to  describe  the  character  of  the  motion  corresponding  to  the  disturbed 
zone  //  at  any  time  t:  Consider  any  straight  line  t  =  const.  Its  inter- 
section with  a  characteristic  yields  the  values  of  u  and  c  at  that  point 
—they  are  the  values  of  u  and  c  which  are  attached  to  that  character- 
istic. Since  the  characteristics  diverge  from  the  piston  curve  one  sees 
that  the  elevation  rj  steadily  increases  upon  moving  from  the  piston  to 
the  right  and  the  particle  velocity  decreases  in  magnitude,  until  the 
initial  characteristic  Cj  is  reached  after  which  the  water  is  undisturbed. 
On  the  other  hand,  if  attention  is  fixed  on  a  definite  point  x  >  0  in  the 
water  and  the  motion  is  observed  as  the  time  increases  it  is  clear—  once 
more  because  the  characteristics  diverge—  that  the  water  remains  un- 
disturbed until  the  time  reaches  the  value  determined  by  x  =  c0£,  after 
which  the  water  surface  falls  steadily  while  the  water  particles  passing 


310 


WATER   WAVES 


that  point  move  more  and  more  rapidly  in  the  negative  ^-direction. 
In  the  foregoing  discussion  of  a  depression  we  have  made  an  assump- 
tion without  saying  so  explicitly,  i.e.  that  the  speed  UA  of  the  piston  is 
such  that  CA  =  \UA  +  CQ  (cf.  (10.5.1))  is  not  negative,  and  this  in 
turn  requires  that 

(10.5.3)  -  UA  ^  2c0. 


Since  —  UA  increases  monotonically  to  the  terminal  value  —  w  it 
follows  that  —  w  must  be  assumed  in  the  above  discussion  to  have  at 


Fig.  10.5.2.  A  limit  case 

most  the  value  2cQ.  The  limit  case  in  which  —  w  just  equals  2c0  is 
interesting.  Since  the  straight  characteristics  have  the  slope  dxfdt  = 
u  +  c  and  since  CA  =  0  from  (10.5.1)  when  UA  =  —  2c0,  it  follows  in 
this  case  that  dx/dt  =  UA  on  the  straight  part  of  the  piston  curve. 
But  this  means  that  the  straight  characteristics  have  all  coalesced 
into  the  piston  curve  itself  in  this  region,  or  in  other  words  that  the 
zone  ///  has  disappeared  in  this  limit  case.  The  circumstances  are 
indicated  in  Fig.  10.5.2.  At  the  front  of  the  wave  for  values  of  x  to  the 
left  of  B  the  elevation  r\A  of  the  water  is  equal  to  —  h  from  CA  = 


+  V!A  )  ==  0>  which  means  that  the  water  surface  just  touches  the 
bottom  at  the  advancing  front  of  the  wave. 

It  is  now  clear  what  would  happen  if  the  terminal  speed  —  w  of  the 
piston  were  greater  than  2c0:  The  zone  //  would  terminate  on  the 
tangent  to  the  piston  curve  drawn  from  the  point  where  the  piston 
speed  —  xt  just  equals  2c0.  The  region  between  this  terminal  charac- 
teristic and  the  remainder  of  the  piston  curve  beyond  it  might  be 
called  the  zone  of  cavitation,  since  no  water  would  exist  for  (x,  t) 


LONG   WAVES    IN    SHALLOW    WATER 


311 


values  in  such  a  region.  In  other  words,  the  piston  eventually  pulls 
itself  completely  free  from  the  water  in  this  case.  Quite  generally  we 
see  that  the  piston  will  lose  contact  with  the  water  (under  the  cir- 
cumstances postulated  in  this  section,  of  course)  if,  and  only  if,  it 
finally  exceeds  the  speed  2c0.  Once  this  happens  it  is  clear  that  the 
piston  has  no  further  effect  on  the  motion  of  the  water.  These  circum- 
stances are  indicated  in  Fig.  10.5.3. 


Covitotion 
Zone 


Fig.  10.5.3.  Case  of  cavitation 

If  the  acceleration  of  the  piston  is  assumed  to  be  infinite  so  that 
its  speed  changes  instantly  from  zero  to  the  constant  terminal  value 
—  w,  the  motion  which  results  can  be  described  very  simply  by  ex- 
plicit formulas.  The  general  situation  in  the  x,  2-plane  is  indicated  in 
Fig.  10.5.4.  This  case  might  be  considered  a  limit  case  of  the  one 
indicated  in  Fig.  10.5.1  which  results  when  the  portion  of  the  piston 
curve  extending  from  the  origin  to  point  B  shrinks  to  a  point.  The 


u=w 


Fig.  10.5.4.  Centered  simple  wave 


812  WATER   WAVES 

consequence  is  that  the  straight  characteristics  in  zone  //  all  pass 
through  the  origin.  The  zone  ///  is  again  one  of  constant  state.  In 
the  zone  //  we  have  obviously  for  the  slopes  of  the  characteristics 

dx       x 

(10.5.4)  Tt=7. 

At  the  same  time  we  have  from  (10.5.2)  dxjdt  —  f  u  +  c0  so  that 

(10.5.5)  ?-  =  ^u+c0. 

It  follows  that  the  zone  //  is  terminated  on  the  upper  side  by  the  line 

(10.5.6)  * 

From  (10.5.5)  and  (10.5.1)  we  can  obtain  the  values  of  u  and  c  within 
zone  Hi 

2  (x          \ 

(10.5.7)  u  =  —  (  —  -  cA     and 

1  1   Ix 

(10.5.8)  C  =  _tt+Co==_^_ 

Since  c  ^  0  we  must  have  —  x/t  ^  2c0  so  that  —  w  must  be  ^  2c0 
from  (10.5.6)  in  conformity  with  a  similar  result  above.  If  w  —  —  2c0, 
the  terminal  characteristic  of  zone  //  is  given,  from  (10.5.6),  by 
x  =  —  2cQt  =  wt  and  this  line  falls  on  the  piston  curve  since  the  slope 
of  the  piston  curve  is  w.  In  this  limit  case,  therefore,  the  zone  /// 
collapses  into  the  piston  curve.  If  the  piston  is  moved  at  still  higher 
speed,  then  cavitation  occurs  as  in  the  cases  discussed  above  since 
c  =  0  at  the  front  of  the  wave,  or  in  other  words,  the  water  surface 
touches  the  bottom. 

From  (10.5.8)  we  can  calculate  the  elevation  rj  of  the  water  surface 
since  c  =  Vg(h  +  77)  ; 

(10.5.9)  >?  +  h  =  +  2c 


In  the  case  of  incipient  cavitation,  i.e.  —  w  =  2c0,  we  have  r\  =  —  h 
at  the  front  of  the  wave.  The  curve  of  the  water  surface  at  any  time  t 
is  a  parabola  from  the  front  of  the  wave  to  the  point  x  =  cQt  (cor- 
responding to  the  characteristic  which  delimits  the  zone  of  quiet), 
after  which  it  is  horizontal.  In  Fig.  10.5.5  the  total  depth  rj  +  h  of 


LONG   WAVES    IN   SHALLOW   WATER 


313 


the  water  is  plotted  against  x  for  a  fixed  time  t.  The  surface  of  the 
water  is  tangent  to  the  bottom  at  the  front  x  —  —  2cQt  of  the  moving 
water.  The  region  in  which  the  water  is  in  motion  extends  from  this 
point  back  to  the  point  x  =  cQt.  From  (10.5.7)  we  can  draw  the  follow- 
ing somewhat  unexpected  conclusion  in  this  case:  Since  t  may  be 
given  arbitrarily  large  values  it  follows  that  the  velocity  u  of  the  water 
at  any  fixed  point  x  tends  to  the  values  —  §c0  as  t  grows  large. 
The  case  of  cavitation  may  have  a  certain  interest  in  practice:  the 
motion  of  the  water  might  be  considered  as  an  approximation  to  the 
flow  which  would  result  from  the  sudden  destruction  of  a  dam  built 
in  a  valley  with  very  steep  sides  and  not  too  great  bottom  slope  (cf. 


Water 


x=-2c0t 


xscol 


Fig.  10.5.5.  Breaking  of  a  dam 


the  paper  of  Re  [R-5])-  If  the  water  behind  the  dam  were  200  feet 
high,  for  example,  our  results  indicate  that  the  front  of  the  wave 
would  move  down  the  valley  at  a  speed  of  about  110  miles  per  hour. 
By  setting  x  =  0  in  (10.5.9)  we  observe  that  the  depth  of  the  water 
at  the  site  of  the  dam  is  always  constant  and  has  the  value  \h, 
i.e.  four-ninths  of  the  original  depth  of  the  water  behind  the  dam. 
The  velocity  of  the  water  at  this  point  is  also  constant  and  has  the 
value  u  =  —  f  CQ  =  —  f  Vgh,  as  we  see  from  (10.5.7).  The  volume 
rate  of  discharge  of  water  at  the  original  location  of  the  dam  is  thus 
constant. 

So  far  we  have  not  considered  the  motion  of  the  individual  water 
particles.  However,  that  is  readily  done  in  all  cases  once  the  velocity 
u(x,  t)  is  known:  We  have  only  to  integrate  the  ordinary  differential 
equation 

dx 


(10.5.10) 


In  zone  //  in  our  present  case  we  have 


314  WATER    WAVES 

dx        2  Ix 


By  setting  £  —  x  +  2c0t  one  finds  readily  that  £  satisfies  the  differen- 
tial equation  d$/dt  =  2£/3t,  from  which  £  ==  At2t3  with  A  an  arbitrary 
constant.  Hence  we  have  for  the  position  x(t)  of  any  particle  in  zone  // 

(10.5.12)  x  =  t{At~11*  —  2c0}. 

In  the  case  of  cavitation  this  formula  holds  for  arbitrarily  large  t  so 
that  we  have  for  large  t  the  asymptotic  expression  for  x: 

(10.5.13)  x~  -  2cQt. 

(This  is  not  in  contradiction  with  our  above  result  that  u  ~  —  f  c0 
for  large  t  and  fixed  x  since  in  that  case  different  particles  pass  the 
point  in  question  at  different  times,  while  (10.5.13)  refers  always  to 
the  same  particle). 

In  the  first  section  of  Chapter  12  this  same  problem  of  the  breaking 
of  a  dam  will  be  treated  by  using  the  exact  nonlinear  theory  in  such 
a  manner  as  to  determine  the  motion  during  its  early  stages  after  the 
dam  has  been  broken—  in  other  words,  at  the  times  when  the  shallow 
water  theory  is  most  likely  to  be  inaccurate. 

10.6.  Discontinuity,  or  shock,  conditions 

The  difference  in  behavior  of  a  depression  which  propagates  into 
still  water  as  compared  with  the  behavior  of  a  hump  has  already  been 
pointed  out:  in  the  first  case  the  motion  is  continuous  throughout,  but 
in  the  second  case  the  motion  can  not  be  continuous  after  a  certain 
time.  The  general  situation  is  indicated  in  Fig.  10.6.1,  which  shows 
the  characteristics  in  the  x,  £-plane  for  the  motion  which  results  when 
a  "piston"  at  the  end  of  a  tank  is  pushed  into  the  water  with  steadily 
increased  speed.  As  before,  the  slope  dxjdt  of  a  straight  characteristic 
issuing  from  the  "piston  curve"  x  =  x(t)  is  given  (cf.  (10.5.2))  by 
dx/dt  =  ^UA  +  CD»  in  which  UA  =  xt(t)  is  the  velocity  of  the  piston. 
Since  UA  is  assumed  to  increase  with  t  it  is  clear  that  the  characteris- 
tics will  cut  each  other.  In  general,  they  have  an  envelope  as  indi- 
cated by  the  heavy  line  in  the  figure.  The  continuous  solutions 
furnished  by  our  theory,  which  have  been  the  only  ones  under  con- 
sideration so  far,  are  thus  valid  in  the  region  of  the  x,  2-plane  between 
-the  initial  characteristic  and  the  piston  curve  up  to  the  curved 
characteristic  (indicated  by  the  curve  segment  ED)  through  the 


LONG    WAVES    IN    SHALLOW    WATER 


315 


"first"  point  E  on  the  envelope  of  the  straight  characteristics,  but 
not  beyond  ED. 

What  happens  "beyond  the  envelope"  can  in  principle  therefore 


Fig.  10.6.1.  Initial  point  of  breaking 

not  be  studied  by  the  theory  presented  up  to  now.  However,  it  seems 
very  likely  that  discontinuous  solutions  may  develop  as  the  time  in- 
creases beyond  the  value  corresponding  to  the  point  E9  which  are 
then  to  be  interpreted  physically  as  motions  involving  the  gradual 
development  of  bores  and  breakers  in  the  water. 


x*<Ut)  £(t)  x»a,(t) 

O  I 

Fig.  10.6.2.  Discontinuity  conditions 

There  is  a  particularly  simple  limit  case  of  the  situation  indicated 
in  Fig.  10.6.2  for  which  a  discontinuous  solution  can  be  found  once 
we  have  obtained  the  discontinuity  conditions  that  result  from  the 


816  WATER   WAVES 

fundamental  laws  of  mechanics.  That  is  the  case  in  which  the  "piston" 
is  accelerated  instantaneously  from  rest  to  a  constant  forward  velocity 
so  that  the  piston  curve  is  a  straight  line  issuing  from  the  origin  in 
the  x,  J-plane.  It  is  the  exact  counterpart  of  the  case  discussed  at  the 
end  of  the  preceding  section  in  which  the  piston  was  withdrawn  from 
the  water  at  a  uniform  speed. 

To  obtain  the  conditions  at  a  discontinuity  we  consider  a  region 
made  up  of  the  water  lying  between  two  vertical  planes  x  —  aQ(t) 
and  x  =  a>i(t)  with  a^  >  a0  and  such  that  these  planes  contain  always 
the  same  particles.  Such  an  assumption  can  be  made,  we  recall  from 
Chapter  2,  since  in  our  theory  the  particles  which  are  in  a  vertical 
plane  at  any  instant  always  remain  in  a  vertical  plane.  Hence  the 
horizontal  particle  velocity  component  u  is  the  same  throughout  any 
vertical  plane.  We  now  suppose  that  there  is  a  finite  discontinuity  in 
the  surface  elevation  rj  at  a  point  x  =  £(t)  within  the  column  of  water 
between  x  =  aQ(t)  and  x  =  a1(^),  as  indicated  in  Fig.  10.6.2. 

The  laws  of  conservation  of  mass  and  of  momentum  as  applied  to 
our  column  of  water  yield  the  relations 

d 
(10.6.1) 

and 

d  f°iW  po  pi 

nnftftl    A  S(r,+h)udx  =  \     p,dy-\     Pldy 

(10.6.2)    ^JaQ(t)  J-n  J-h 


when  the  formula  p  =  gQ(r)  —  y)  for  the  pressure  in  the  water  is 
used.  The  second  relation  states  that  the  change  in  momentum  of  the 
water  column  is  equal  to  the  difference  of  the  resultant  forces  over  the 
end  sections  of  the  column. 

The  integrals  in  these  relations  have  the  form 


/•«h(« 

= 

Ja0(t) 


(x,  t)  dx 

in  which  \p(x,  t)  has  a  discontinuity  at  x  =  g(t).  Differentiation  of  this 
integral  yields  the  relation 

dl       d  f*«  d 

yidx 


(10.6.3) 

-     dx 


LONG    WAVES    IN   SHALLOW   WATER  317 

The  quantities  UQ  and  u±  are  the  velocities  a0(t)  and  a^t)  at  the  ends 
of  the  column,  f  is  the  velocity  of  the  discontinuity,  and  y;(£_,  t)  and 
y(£+,  t)  mean  that  the  limit  values  of  y  to  the  left  and  to  the  right  of 
x  =  f  respectively  are  to  be  taken.  We  wish  to  consider  the  limit  case 
in  which  the  length  of  the  column  tends  to  zero  in  such  a  way  that 
the  discontinuity  remains  inside  the  column.  When  we  do  so  the 
integral  on  the  right-hand  side  of  (10.6.3)  tends  to  zero  and  we  obtain 

dl 

(10.6.4)  lim    —  =  y^  -  Wo 

in  which  vl  and  VQ  are  the  relative  velocities  given  by 
(10.6.5) 


and  if?!  and  y0  refer  to  the  limit  values  of  \p  to  the  right  and  to  the  left 
of  the  discontinuity,  respectively.  The  important  quantities  VQ  and  vl 
are  obviously  the  flow  velocities  relative  to  the  moving  discontinuity. 
Upon  making  use  of  (10.6.4)  and  (10.6.5)  for  the  limit  cases  which 
arise  from  (10.6.1)  and  (10.6.2)  we  obtain  the  following  conditions 

(10.6.6)  gfo  +  hfa  -  Q(rh  +  h)v0  -  0 

and 

(io.6.7)    eOh+AK^-efoo+A^ 

If  we  introduce,  as  in  section  10.1,  the  quantities  £  and  p  (which  are 
the  analogues  of  the  density  and  pressure  in  gas  dynamics)  by  the 
relations  (of.  (10.1.3)  and  (10.1.4)) 

(10.6.8)  Q  =  e(i,  +  h) 
and 

(10.6.9)  p  =  ^(r/+A)i=   Igi, 

we  obtain  in  place  of  (10.6.6)  and  (10.6.7)  the  discontinuity  conditions 

(10.6.10)  QM  =  QOVO, 
and 

(10.6.11)  e&iVi  ~  (?(Wo  =  Po  —  Pi- 

The  last  two  relations  are  identical  in  form  with  the  mechanical  con- 
ditions for  a  shock  wave  in  gas  dynamics  when  the  latter  are  expressed 
in  terms  of  velocity,  density  and  pressure  changes. 


318  WATER   WAVES 

Henceforth  we  shall  often  refer  to  a  discontinuity  satisfying  (10.6. 
10)  and  (10.6.11)  as  a  shock  wave  or  simply  as  a  shock  even  though 
such  an  occurrence  is  better  known  in  fluid  mechanics  as  a  bore,  or 
if  it  is  stationary  as  a  hydraulic  jump. 

Since  u±  —  UQ  —  vl  —  VQ  from  (10.6.5)  it  is  easily  seen  that  the 
shock  conditions  (10.6.10)  and  (10.6.11)  can  be  written  in  the  form 


(10.6.12)  _ 

I  m(^  -  VQ)  =  po  -  plf 

in  which  m  represents  the  mass  flux  across  the  shock  front. 

To  fix  the  motion  on  both  sides  of  the  shock  five  quantities  are 
needed;  i.e.  the  particle  velocities  UQ,  uv  the  elevations  77  0  and  rjl  (or, 
what  is  the  same,  the  "pressures"  p  or  the  "densities"  g  as  given  by 
(10.6.8)  and  (10.6.9)  on  both  sides  of  the  shock),  and  the  velocity  | 
of  the  shock.  Evidently  the  relative  velocities  VQ  and  vl  would  then 
be  determined.  Since  the  five  quantities  satisfy  the  two  relations 
(10.6.12)  we  see  that  in  general  only  three  of  the  five  quantities  could 
be  prescribed  arbitrarily.  Since  the  equations  to  be  satisfied  are  not 
linear  it  is  not  a  priori  clear  whether  solutions  can  be  found  for  two 
of  the  quantities  when  any  other  three  are  arbitrarily  prescribed  or 
whether  such  solutions  would  be  unique.  We  want  to  investigate  this 
question  in  a  number  of  important  special  cases. 

Before  doing  so,  however,  it  is  important  to  consider  the  energy 
balance  across  a  shock.  The  fact  is,  as  we  shall  see  shortly,  that  the 
law  of  conservation  of  energy  does  not  hold  across  a  shock,  but  rather 
the  particles  crossing*  the  shock  must  either  lose  or  gain  in  energy. 
Since  we  do  not  wish  to  postulate  the  existence  of  energy  sources  at 
the  shock  front  capable  of  increasing  the  energy  of  the  water  particles 
as  they  pass  through  it,  we  assume  from  now  on  that  the  water 
particles  do  not  gain  energy  upon  crossing  a  shock  front.  This  will  in 
effect  furnish  us  with  an  inequality  which  in  conjunction  with  the 
two  shock  relations  (10.6.12)  leads  in  all  of  our  cases  to  unique  solu- 
tions of  the  physical  problems.  We  turn,  then,  to  a  consideration  of 
the  energy  balance  across  a  shock,  which  we  can  easily  do  by  following 

*  It  is  important  to  observe  that  the  water  particles  always  do  cross  a  shock 
front:  the  quantity  m  in  (10.6.12),  the  mass  flux  through  the  shock  front,  is 
different  from  zero  if  there  is  an  actual  discontinuity  since  otherwise  vl  =  VQ  =  0, 
tij  =  UQ  =  £,  and  p0  =  pl  and  hence  @0  =  ^  —  in  other  words  the  motion  is 
continuous.  There  is  thus  no  analogue  in  our  theory  of  what  is  called  a  contact 
discontinuity  in  gas  dynamics  in  which  velocity  and  pressure  are  continuous, 
but  the  density  and  temperature  may  be  discontinuous. 


LONG    WAVES    IN    SHALLOW    WATER  319 

the  same  procedure  that  was  used  to  derive  the  shock  relations  (10.6. 
10)  and  (10.6.11  ).  For  the  rate  of  change  dE/dt  of  the  energy  E  in  the 
water  column  of  Fig.  10.6.2  we  have,  as  one  can  readily  verify: 
dE       d 


(10.6.1,3)  <*o«) 


Pi 

J  _h 


l-h* 

and  this  in  turn  yields  in  the  limit  when  a0  -+  av  through  use  of 
(10.6.5),  (10.6.8),  (10.6.9),  and  the  hydrostatic  pressure  law,  the 
relation 

dE 
(10.6.14)  ~  ~ 


for  the  rate  at  which  energy  is  created  or  destroyed  at  the  shock  front. 
If  we  multiply  (10.6.11)  by  |  on  both  sides  and  then  subtract  from 

(10.6.14)  the  result  is  an  equation  which  can  be  written  after  some 
manipulation  and  use  of  (10.6.5)  in  the  form 

dE 

(10.6.15)  —  -=  m  {%(v*  —  vl)  f  2(~pi/ei  —  p0/Q0)} 

dt 

in  which  m  is  the  mass  flux  through  the  shock  front  defined  in 
(10.6.12).  In  this  way  we  express  dE/dt  entirely  in  terms  of  the 
relative  velocities  VQ  and  vl  and  the  change  in  depth.  By  eliminating 
vl  and  v0  through  use  of  v1  =  m/gx  and  v0  =  M/{JO  and  replacing  pl  and 
p0  in  terms  of  QI  and  QO  we  can  express  dE/dt  in  terms  of  QO  and  g^; 
the  result  is  readily  found  to  be  expressible  in  the  simple  form 

dE       mg  (PA  —  pi)3 

(10.6.16)  -7-  —  —        ,  _   — . 

dt         Q       *Qi6Q 

We  sec  therefore  that  energy  is  not  conserved  unless  g0  =  Q19  i.e.  unless 
the  motion  is  continuous.  Since  QQ  —  Qi  =  p(^o  ~~  ^i)  ^  follows  from 
(10.6.16)  that  the  rate  of  change  of  the  energy  of  the  particles  crossing 
the  shock  is  proportional  to  the  cube  of  the  difference  in  the  depth  of 
the  water  on  the  two  sides  of  the  shock,  or  as  we  could  also  put  it  in 
case  rjQ  —  r/l  is  considered  to  be  a  small  quantity:  the  rate  of  change 
of  energy  is  of  third  order  in  the  "jump"  of  elevation  of  the  water 
surface. 

The  statement  that  the  law  of  conservation  of  energy  does  not  hold 
in  the  case  of  a  bore  in  water  must  be  taken  cum  grano  salis.  What  we 
mean  is  of  course  that  the  energy  balance  can  not  be  maintained 


320 


WATER   WAVES 


through  the  sole  action  of  the  mechanical  forces  postulated  in  the 
above  theory.  The  results  of  our  theory  of  the  bore  and  the  hydraulic 
jump  are  therefore  to  be  interpreted  as  an  idealization  of  the  actual 
occurrences  in  which  the  losses  in  mechanical  energy  are  accounted 
for  through  the  production  of  heat  due  to  turbulence  at  the  front  of 
the  shock  (cf.  the  photograph  of  the  bore  in  the  Tsien-Tang  river 
shown  in  Fig.  10.6). 8).  In  compressible  gas  dynamics  the  theory  used 


Fig.  10.6.3.  Bore  in  the  Tsien-Tang  River 

allows  for  the  conversion  of  mechanical  energy  into  heat  so  that  the 
law  of  conservation  of  energy  holds  across  a  shock  in  that  theory. 
The  analogue  of  the  loss  in  mechanical  energy  across  a  shock  in  water 
is  the  increase  in  entropy  across  a  shock  in  gas  dynamics;  furthermore, 
both  of  these  discontinuous  changes  are  of  third  order  in  the  differen- 
ces of  "density"  on  the  two  sides  of  the  shock. 

We  have  tacitly  chosen  as  the  positive  direction  of  the  #-axis,  and 
hence  of  all  velocities,  the  direction  from  the  side  0  toward  the  side  1 
(cf.  Fig.  10.6.2).  Suppose  now  that  the  mass  flux  m  is  assumed  to  be 
positive;  it  follows  from  (10.6.12)  and  the  fact  that  j50  and  QI  are  posi- 
tive that  VQ  and  vl  are  also  positive  and  hence  that  the  water  particles 
cross  the  shock  front  in  the  direction  from  the  side  0  toward  the  side  1. 
Our  condition  that  the  water  particles  can  not  gain  in  energy  on  cross- 
ing the  shock  then  requires,  as  we  see  at  once  from  (10.6.16)  since  w, 


LONG   WAVES    IN    SHALLOW   WATER  321 

g>  £>»  {>o»  anc^  ^i  are  a^  positive,  that  g0  <  gla  In  other  words,  our  energy 
condition  requires  that  the  particles  always  move  across  the  shock  from 
a  region  of  lower  total  depth  to  one  of  higher  total  depth.*  Since  the  mass 
flux  m  is  not  zero  unless  the  flow  is  continuous,  and  hence  there  is  no 
shock,  it  is  possible  to  define  uniquely  the  two  sides  of  the  shock  by  the 
following  useful  convention:  the  front  and  back  sides  of  the  shock  are 
distinguished  by  the  fact  that  the  mass  flux  passes  through  the  shock 
from  front  to  back,  or,  as  one  could  also  put  it,  the  water  crosses  the 
shock  from  the  front  side  toward  the  back  side.  Our  conclusion  based 
on  the  assumed  loss  of  energy  across  the  shock  can  be  interpreted  in 
terms  of  this  convention  as  follows:  the  water  level  is  always  lower  on 
the  front  side  of  the  shock  than  on  the  back  side. 

For  the  further  discussion  of  the  shock  relations  it  is  important  to 
observe  that  all  of  them,  including  the  relation  (10.6.16)  for  the 
energy  loss,  can  be  written  in  such  a  way  as  to  involve  only  the  velocities 
VQ  and  vl  of  the  water  particles  relative  to  the  shock  front  and  not  the  abso- 
lute velocities  UQ  and  ur  It  follows  that  we  may  always  assume  one  of 
the  three  velocities  u0,  ul9 £  to  be  zero  if  we  wish,  with  no  essential  loss 
of  generality,  because  the  laws  of  mechanics  are  in  any  case  invariant 
with  respect  to  axes  moving  with  constant  velocity,  and  adding  the 
same  constant  to  MO,  %  and  £  does  not  affect  the  values  of  v0  and  i^. 

Let  us  assume  then  that  uQ  —  0,  i.e.  that  the  water  is  at  rest  on  one 
side  of  the  shock.  Also,  we  write  the  second  of  the  shock  conditions 
(10.6.12)  in  the  form 

(10.6.17)  ^PQ  =?-  "-?1, 

<?o  —  61 

which  follows  from  mvl  =  QOVOVI  and  mv0  =  pi^t'o  and  (10.6.12). 
From  UQ  =  0  we  have  v^  =  —  f  and  v±  =  ul  —  £  (cf.  (10.6.5))  so  that 
(10.6.17)  takes  the  form 

(10.6.18)  -  fK  -  f )  = :~  (go  +  ei) 

AQ 

upon  making  use  of  p  =  gg2/2?  (cf-  (10.6.9)).  The  first  shock  condition 

now  takes  the  form 

(10.6.19)  giK  _  {)  =  - 

so  that  (10.6.18)  can  be  written 


(10.6.20,  f'= 

This  conclusion  was  first  stated  by  Rayleigh  [R.3J. 


322  WATER   WAVES 

if  u±  is  eliminated,  or  it  may  be  written  in  the  form 


if  g0  is  eliminated.  Thus  (10.6.19)  together  with  either  (10.6.20)  or 
(10.6.21)  are  ways  of  expressing  the  shock  conditions  when  u0  =  0. 

We  are  now  in  a  position  to  discuss  some  important  special  cases. 
Having  fixed  the  value  of  w0,  i.e.  UQ  =  0,  at  most  two  of  the  remaining 
quantities  |,  g0,  QV  and  wx  can  be  prescribed  arbitrarily.  For  our  later 
discussion  it  is  useful  to  single  out  the  following  two  cases:  Case  1.  JDJ. 
and  g0  are  given,  i.e.  the  depth  of  the  water  on  both  sides  of  the  shock 
and  the  velocity  on  one  side  are  given.  Case  2.  ^  and  u±  are  given,  i.e. 
the  velocity  of  the  water  on  both  sides  of  the  shock  and  the  depth  of 
the  water  on  one  side  are  given.  We  proceed  to  discuss  these  cases  in 
detail. 

Case  1.  From  (10.6.20)  we  see  that  |2  is  determined  for  any  arbitrary 
values  (positive,  of  course)  of  £0  and  gl9  i.e.  of  the  water  depths.  Hence 
£  is  determined  by  (10.6.20)  only  within  sign.  Suppose  now  that 
61  -^  (?<)•  ^n  ^is  case  ^e  side  0  is,  as  we  have  seen  above,  the  front 
side  of  the  shock,  and  since  u0  =  0  the  shock  front  must  move  in  the 
direction  from  the  side  1  toward  the  side  0  in  order  that  the  mass  flux 
should  pass  through  the  shock  from  front  to  back. 

Hence  if  it  is  once  decided  whether  the  side  0  is  to  the  left  or  to  the 
right  of  the  side  1  the  sign  of  £  is  uniquely  fixed.  If,  as  in  Fig.  10.6.4, 


€  *- 


0 


1 

-u.  <0 


X 

Fig.  10.0.4.  Bore  advancing  into  still  water 

the  side  0  is  chosen  to  the  left  of  the  side  1,  and  the  ^-direction  is  posi- 
tive to  the  right,  it  follows  that  f  is  negative,  as  indicated.  It  is  useful 
to  introduce  the  depths  A0  and  Ax  of  the  water  on  the  two  sides  of  the 
shock: 

(10.6.22) 


LONG    WAVES    IN   SHALLOW   WATER  328 

and  to  express  (10.6.20)  in  terms  of  these  quantities.  The  result  for  f 
in  our  case  is 


(10.6.28) 


I/"  hi 
«  -  -  K  «  * 


as  one  readily  sees  from  gt  =  ght.  From  (10.6.23)  we  draw  the  im- 
portant conclusion:  Since  h^  >  A0,  the  shock  speed  |  f  |  is  greater  than 
Vgh0  since  h0  <  (Ax  +  A0)/2  <  /&x.  Also,  in  the  case  w0  =  0  we  have 
from  (10.6.19) 


(10.6.24) 


u,  =  |  (l  -  JsJ, 


so  that  the  velocity  of  the  water  behind  the  shock  has  the  same  sign  as 
£  (since  hQ  /hl<  1  )  but  is  less  than  f  numerically. 

Finally,  it  is  very  important  to  consider  the  speed  vl  of  the  shock 
front  relative  to  the  water  particles  behind  it:  from  (10.6.24)  we  have 

(10.6.25)  vl=ul^S=-'T^ 

HI 

and  this  in  turn  can  be  expressed  through  use  of  (10.6.23)  in  the 
form 


^ 

so  that  vl  <  Vghv  In  other  words,  the  speed  of  the  shock  relative  to 
the  water  particles  behind  the  shock  is  less  than  the  ivave  propagation 

speed  Vg/?!  in  the  water  behind  the  shock.  Hence  a  small  disturbance 
created  behind  a  shock  will  eventually  catch  up  with  it.  Although  the 
conclusion  was  drawn  for  the  special  case  u0  =  0  it  holds  quite 
generally  for  the  shock  velocities  relative  to  the  motion  of  the  water 
on  both  sides  of  a  shock,  in  view  of  earlier  remarks  on  the  dependence 
of  the  shock  relations  on  these  relative  velocities. 

The  case  illustrated  by  Fig.  10.6.4  is  that  of  a  shock  advancing  into 
still  water.  The  fact  that  f  is  in  this  case  of  necessity  negative  is  a 
consequence  of  the  assumption  of  an  energy  loss  across  the  shock.  It 
is  worth  while  to  restate  this  conclusion  in  the  negative  sense,  as 
follows:  a  depression  shock  can  not  exist,  i.e.  a  shock  wave  which 
leaves  still  water  at  reduced  depth  behind  it  should  not  be  observed 
in  nature.*  The  observations  bear  out  this  conclusion.  Bores  advancing 

*  In  gas  dynamics  the  analogous  situation  occurs:  only  compression  shocks 
and  not  rarefaction  shocks  can  exist. 


324 


WATER   WAVES 


Rigid        ^ 
Woll  *^ 

-  *• 

u0'0 

h 

0 

h            «  —  ui 

Fig.  1Q.6.5.  Reflection  from  a  rigid  wall 


!     .1 

u,— 

go  » 

1 

Fig.  10.6.6.  Hydraulic  jump 


into  still  water  are  well  known,  but  depression  waves  are  always 
smooth. 

Instead  of  assuming  that  gx  >  g0  (or  that  At  >  h0)  as  in  the  case  of 
Fig.  10.6.4  we  may  assume  ^  <  £0  (or  /&x  <  hQ),  so  that  the  side  1  is 
the  front  side.  In  other  words  the  water  is  at  rest  on  the  baek  side  of 
the  shock  in  this  case.  If  the  front  side  is  taken  on  the  right,  the  situa- 
tion is  as  indicated  in  Fig.  10.6.5.  In  this  case  |  must  be  positive 
and  %  negative  in  order  that  the  mass  flux  should  take  place  from  the 
side  1  to  the  side  0.  The  value  of  t^  is  given  by  (10.6.24)  in  this  case 
also.  The  case  of  Fig.  10.6.5  might  be  realized  in  practice  as  the  result 
of  reflection  of  a  stream  of  water  from  a  rigid  wall  so  that  the  water  in 
contact  with  the  wall  is  brought  to  rest.  We  shall  return  to  this  case 
later. 

In  the  above  two  cases  we  considered  u0  to  be  zero.  However,  we 
know  that  we  may  add  any  constant  velocity  to  the  whole  system 
without  invalidating  the  shock  conditions.  It  is  of  interest  to  consider 
the  motion  which  arises  when  the  velocity  —  £  is  added  to  uQ9  ur  and 
|  in  the  case  shown  in  Fig.  10.6.4.  The  result  is  the  motion  indicated 
by  Fig.  10.6.6  in  which  the  shock  front  is  stationary.  This  case— one 
of  frequent  occurrence  in  nature-— is  commonly  referred  to  as  the 
hydraulic  jump.  From  our  preceding  discussion  we  see  that  the  water 
always  moves  from  the  side  of  lower  elevation  to  the  side  of  higher 
elevation.  The  velocities  UQ  and  u±  are  both  positive,  and  UQ  >  uv 


LONG   WAVES   IN   SHALLOW   WATER 


325 


Also  the  velocity  u0  on  the  incoming  side  is  greater  than  the  wave 
propagation  speed  Vgh0  on  that  side  while  the  velocity  u±  is  less  than 
Vghlt  This  follows  at  once  from  the  known  facts  concerning  the  re- 
lative shock  velocities  and  the  fact  that  u0  and  %  are  the  velocities 
relative  to  the  shock  front  in  this  case.  The  hydraulic  engineers  refer 
to  this  as  a  transition  from  supercritical  to  subcritical  speed. 

Case  2.  We  recall  that  in  this  case  u0  =  0,  u^  and  pL  (or  h^)  are 
assumed  given  and  f  and  h0  are  to  be  determined.  The  value  of  f  is 
to  be  determined  from  (10.6.21).  To  study  this  relation  it  is  conve- 
nient to  set  x  =  —  f  and  y  =  u^  —  £  so  that  (10.6.21)  can  be  replaced 

by 


(10.6.27) 


y  =  k2x/(x2 
y  =  HI  +  x. 


In  Fig.  10.6.7  we  have  indicated  these  two  curves,  whose  intersections 
yield  the  solutions  f  =  —  x  of  (10.6.21).  The  first  equation  is  re- 
presented by  a  curve  with  three  branches  having  two  asymptotes 
%  —  ±  k.  As  one  sees  readily,  there  are  always  three  different  real 
roots  for  —  £  no  matter  what  values  arc  chosen  for  the  positive  quan- 
tity k2  =  g/4/2  and  for  the  velocity  uv  Furthermore,  one  root  £+  = 
x_  is  always  positive,  another  £_  =  —  x+  is  negative,  while  the 
third  |  ~  —  x  lies  between  the  other  two.  However,  the  third  root 
|  =  —  x  must  be  rejected  because  it  is  not  compatible  with  (10.6.19): 
Since  QI  and  QO  are  both  positive  it  follows  that  x  =  — -  f  and  y  = 


Fig.  10.6.7.  Graphical  solution  of  shock  conditions 


326  WATER   WAVES 

M!  —  f  must  have  the  same  sign.  But  the  sign  ofy  —  y  corresponding 
to  x  =  x  is  always  the  negative  of  x  as  one  sees  from  Fig.  10.6.7. 
(If  U-L  =  0  ,then  x  =  y  =  0,  but  there  is  no  shock  discontinuity  in 
this  case.)  The  other  two  roots,  however,  are  such  that  the  signs  of 
—  £  and  ux  — -  f  are  the  same.  In  the  case  2,  therefore,  equation 
(10.6.21)  furnishes  two  different  values  of  f  which  have  opposite 
signs  and  these  values  when  inserted  in  (10.6.19)  furnish  two  values 
of  the  depth  /20.  The  two  cases  are  again  those  illustrated  in  Figs. 
10.6.4  and  10.6.5.  An  appropriate  choice  of  one  of  the  two  roots  must 
be  made  in  accordance  with  the  given  physical  situation,  as  will  be 
illustrated  in  one  of  the  problems  to  be  discussed  in  the  next  section. 
Before  proceeding  to  the  detailed  discussion  of  special  problems 
involving  shocks  it  is  perhaps  worth  while  to  sum  up  briefly  the  main 
facts  derived  in  this  section  concerning  them:  the  five  essential  quan- 
tities defining  a  shock  wave— f,  MO,  u^\  g0,  gx  (or,  what  is  the  same, 
h0  and  h^)— must  satisfy  the  shock  conditions  (10.6.12).  If  it  is  as- 
sumed in  addition  that  the  water  particles  may  lose  energy  on  crossing 
the  shock  but  not  gain  it,  then  it  is  found  that  the  shock  wave  travels 
always  in  such  a  direction  that  the  water  particles  crossing  it  pass  from 
the  side  of  lower  depth  to  the  side  of  higher  depth.  If  hQ  <  hl9  so  that  the 
side  0  is  the  front  side  of  the  shock,  the  speeds  \  v0  \  and  \  v±  \  of  the  water 
relative  to  the  shock  front  satisfy  the  inequalities 

(10.6.28) 

In  hydraulics  it  is  customary  to  say  that  the  velocity  relative  to  the 
shock  is  supercritical  on  the  front  side  (i.e.  greater  than  the  wave 
propagation  speed  corresponding  to  the  water  depth  on  that  side) 
and  subcritical  on  the  back  side  of  the  shock* 

10.7.  Constant    shocks:    bore,    hydraulic  jump,   reflection   from    a 
rigid  wall 

In  the  preceding  section  shock  discontinuities  were  studied  for  the 
purpose  of  obtaining  the  relations  which  must  hold  on  the  two  sides 
of  the  shock,  and  nothing  was  specified  about  the  motion  otherwise 
except  that  the  shock  under  discussion  should  be  the  only  disconti- 

*  In  gas  dynamics  the  analogous  inequalities  lead  to  the  statement  that  the 
flow  velocity  relative  to  the  shock  front  is  supersonic  with  respect  to  the  gas 
on  the  side  of  lower  density  and  subsonic  with  respect  to  the  gas  on  the  other  side. 


LONG   WAVES   IN   SHALLOW   WATER  327 

nuity  in  a  small  portion  of  the  fluid  on  both  sides  of  it.  In  the  present 
and  following  sections  we  wish  to  consider  motions  which  are  conti- 
nuous except  for  the  occurrence  of  a  single  shock.  Furthermore  we 
shall  limit  our  investigations  in  this  section  to  cases  in  which  the  mo- 
tion on  each  of  the  two  sides  of  the  shock  has  constant  velocity  and 
depth.  These  motions,  or  flows,  are  evidently  of  a  very  special  cha- 
racter, but  they  are  easy  to  describe  and  also  of  frequent  occurrence 
in  nature. 

It  is  perhaps  not  without  interest  in  this  connection  to  observe  that 
the  only  steady  and  continuous  wave  motions  (i.e.,  motions  in  which 
the  velocity  u  and  wave  propagation  speed  c  =  Vg(&  +77)  are  in- 
dependent of  the  time )  furnished  by  our  theory  for  the  case  of  constant 
depth  h  are  the  constant  states  u  =  const.,  c  =  const.  This  follows 
from  the  original  dynamical  equations  (10.1.8)  and  (10.1.9).  In  fact, 
when  u  and  c  are  assumed  to  be  functions  of  x  alone  these  equa- 
tions reduce  to 

du  dc 

u  —  +  2c  —  =  0,     and 
ax  ax 

dc  du 

2u~  +    c—  =  0 
ax  ax 

for  the  case  in  which  h  =  const,  (and  so  Hx  —  0).  These  equations  are 
immediately  integrable  to  yield  u2  +  2c2  =  const,  and  uc2  =  const, 
and  these  two  relations  are  simultaneously  satisfied  only  for  constant 
values  of  u  and  c.  On  the  other  hand,  any  constant  values  whatever 
could  be  taken  for  u  and  c.  The  cases  we  discuss  in  this  section  are 
motions  which  result  by  piecing  together  two  such  steady  motions 
(each  with  a  different  constant  value  for  the  depth  and  velocity) 
through  a  shock  which  moves  with  constant  velocity.  In  this  case  the 
motion  as  a  whole  would  be  steady  if  observed  from  a  coordinate 
system  attached  to  the  moving  shock  front.  In  view  of  our  above  dis- 
cussion it  is  clear  that  such  a  motion  with  a  single  shock  discontinuity 
is  the  most  general  progressing  wave  which  propagates  unchanged  in 
form  that  could  be  obtained  from  our  theory.* 

Let  us  consider  now  the  problem  referred  to  at  the  beginning  of  the 
preceding  section:  a  vertical  plate— or  piston,  as  we  have  called  it— at 

*  This  result  should  not  be  taken  to  mean  that  the  so-called  "solitary  wave'* 
does  not  exist.  (By  a  solitary  wave  is  meant  a  continuous  wave  in  the  form  of  a 
single  elevation  which  propagates  unaltered  in  form.)  It  means  only  that  our 
approximate  theory  is  not  accurate  enough  to  furnish  such  a  solitary  wave.  This 
is  a  point  which  will  be  discussed  more  fully  in  section  10.9. 


328  WATER   WAVES 

the  left  end  of  a  tank  full  of  water  at  rest  is  suddenly  pushed  into  the 
water  at  constant  velocity  w.  As  we  could  infer  from  the  discussion 
at  the  beginning  of  the  preceding  section  the  motion  must  be  dis- 
continuous from  the  very  beginning—  or,  as  we  could  also  put  it,  the 
"first"  point  on  the  envelope  of  the  characteristics  would  occur  at 
t  =  0.  Since  the  piston  moves  with  constant  velocity  we  might  expect 
the  resulting  motion  to  be  a  shock  wave  advancing  into  the  still  water 
and  leaving  a  constant  state  behind  such  that  the  water  particles  move 
with  the  piston  velocity  w.  The  circumstances  for  such  an  assumed 
motion  are  indicated  in  Fig.  10.7.1,  which  shows  the  x,  J-plane  together 
with  the  water  surface  at  a  certain  time  tQ.  We  know  that  the  constant 
states  on  each  side  of  the  shock  satisfy  our  differential  equations.  In 
addition,  we  show  that  they  can  always  be  "connected"  through  a 
shock  discontinuity  which  satisfies  the  shock  relations  derived  in  the 
preceding  section.  In  fact,  the  relations  (10.6.18)  and  (10.6.19)  yield 
through  elimination  of  ^  ==  p/^  the  relation 

(10.7.1)  _flw_f) 


for  |  in  terms  of  w  and  the  depth  hQ  in  the  still  water,  when  we  set 
go  =  Qh0.  Equation  (10.7.1)  is  the  same  as  (10.6.21)  except  that  g0 
replaces  glf  and  the  discussion  of  its  roots  |  follows  exactly  the  same 
lines  as  for  (10.6.21):  for  each  A0  >  0  and  any  w  ^  0  the  cubic  equa- 
tion (10.7.1)  has  three  roots  for  f  :  one  negative,  another  positive,  and 
a  third  which  has  a  value  between  these  two.  In  the  present  case  the 
positive  root  for  |  must  be  taken  in  order  to  satisfy  our  energy  con- 
dition (cf.  the  discussion  based  on  (10.6.27)  of  the  preceding  section) 
since  the  side  0  is  the  front  side  of  the  shock.  Once  |+  has  been  calcu- 
lated from  (10.7.1)  we  can  determine  the  depth  of  the  water  h^  behind 
the  shock  from  the  first  shock  condition 

(10.7.2)  h,(w  -£+)=-  AO|+. 

It  is  therefore  clear  that  a  motion  of  the  sort  indicated  in  Fig.  10.7.1 
can  be  determined  in  a  way  which  is  compatible  with  all  of  our  con- 
ditions.* 

A  few  further  remarks  about  the  above  motion  are  of  interest.  In 

*  It  should  be  pointed  out  that  our  discussion  yields  a  discontinuous  solution 
of  the  differential  equations,  but  does  not  prove  that  it  is  the  only  one  which 
might  exist.  However,  it  has  been  shown  by  Goldner  [G.6]  that  our  solution  would 
be  unique  under  rather  general  assumptions  regarding  the  type  of  functions 
admitted  as  possible  solutions. 


LONG   WAVES    IN   SHALLOW   WATER 


829 


wt 


Fig.  10.7.1.  A  bore  with  constant  speed  and  height 


sxs  ft 


Reflected. 
Shock 


(2) 


i 


^_ 


CD 


Inclden 
Shock 


t=t, 


"2=0 


I 


t=t 


*  0 


tlto 
Fig.  10.7.2.  Reflection  of  a  bore  from  a  rigid  wall 


330  WATER   WAVES 

Fig.  10.7.1  we  have  indicated  the  line  x  =  cQt,  c0  =  VghQ>  which 
would  be  the  initial  characteristic  terminating  the  state  of  rest  if  the 
motion  were  continuous,  i.e.  if  the  disturbance  proceeded  into  still 
water  with  the  wave  speed  c0  for  water  of  the  depth  A0.  We  know, 
however,  from  our  discussion  of  the  preceding  section  that  the  shock 
speed  |  is  greater  than  c0,  which  accounts  for  the  position  of  the  shock 
line  x  =  |J  below  the  line  x  =  cQt  in  Fig.  10.7.1.  On  the  other  hand 
we  know  that  the  velocity  w  —  £  of  the  water  particles  behind  the 

shock  relative  to  the  shock  is  less  than  the  wave  speed  c±  =  VgAx  in 
the  water  on  that  side.  It  follows,  therefore,  that  a  new  disturbance 
created  in  the  water  behind  the  shock  should  catch  up  with  it  since 
the  front  of  such  a  disturbance  would  always  move  relative  to  the 
water  particles  with  a  velocity  at  least  equal  to  cr  For  example,  if  the 
piston  were  to  be  decelerated  at  a  certain  moment  a  continuous  de- 
pression wave  would  be  created  at  the  piston  which  would  finally 
catch  up  with  the  shock  front,  and  a  complicated  interaction  process 
would  then  occur. 

The  case  we  have  treated  above  corresponds  to  the  propagation  of 
a  bore  into  still  water.  If  we  were  to  superimpose  a  constant  velocity 
—  |  on  the  water  in  the  motion  illustrated  by  Fig.  10.7.1  the  result 
would  be  the  motion  called  a  hydraulic  jump  in  which  the  shock  front 
is  stationary.  We  need  not  consider  this  case  further. 
.  We  treat  next  the  problem  of  the  reflection  of  a  shock  wave  from 
a  rigid  vertical  wall  by  following  essentially  the  same  procedure  as 
above.  The  circumstances  are  shown  in  Fig.  10.7.2.  We  have  an 
incoming  shock  moving  toward  the  rigid  wall  from  the  left  into  still 
water  of  depth  hQ.  The  shock  is  reflected  from  the  wall  leaving  still 
water  of  depth  h2  behind  it.  Since  the  water  in  contact  with  the  wall 
should  be  at  rest,  such  an  assumed  motion  is  at  least  a  plausible  one. 
We  proceed  to  show  that  the  motion  is  compatible  with  our  shock 
conditions  and  we  calculate  the  height  h2  of  the  reflected  wave. 

We  assume  that  h^  and  uv  =  10,  the  depth  and  the  velocity  of  the 
water  behind  the  shock,  are  known.  The  shock  speed  f  +  is  then  de- 
termined by  taking  the  largest  of  the  three  roots  of  the  cubic  equation 
(10.6.21),  which  we  write  down  again  in  the  form 


Once  |+  has  been  determined,  the  depth  h0  in  front  of  the  shock  is 
fixed  from  the  first  shock  condition,  which  is  in  the  present  case 


LONG    WAVES    IN    SHALLOW    WATER 


381 


(10.7.4)  (W  -  S+fa  =  -  |A- 

To  determine  the  reflected  shock  we  may  once  more  evidently  make 


I          I 


r  •  1.4 


Ol L 


I I I I I I I I 


0  2  4  6  8  10  12  14  16  18 


(a) 


ho 
100 


80 


60 


40 


20 


1      I  I I I I I 


8       •     10  12  14  16 


(b) 

Fig.  l().7.3a,  b.  Reflection  of  a  bore  from  a  rigid  wall 

use  of  (10.7.4),  since  At  and  u^  =  w  remain  the  same  on  one  side  of  the 
shock,  but  we  must  now  choose  the  smallest  of  the  three  roots  of 


382 


WATER   WAVES 


(10.7.3)  as  the  shock  speed  f_  since  the  side  (1)  is  now  obviously  the 
front  side  of  the  shock.  The  depth  h2  of  the  water  behind  the  shock 
after  the  reflection— that  is,  of  the  water  in  contact  with  the  wall 
after  reflection— is  then  obtained  in  the  same  way  as  h0  by  using 

(10.7.4)  with  |_  in  place  of  £+  and  h2  in  place  of  hQ: 

(10.7.4)!  (w  —  £_)&!  =  —  £_/*2. 

By  taking  a  series  of  values  for  w  we  have  determined  the  ratios 
/&2/&J  and  A2/A0  as  functions  of  /?1/A0.  That  is,  the  height  h2  of  the 
reflected  wave  has  been  determined  as  a  function  of  the  ratio  of  the 
depth  AJ  of  the  incoming  wave  to  the  initial  depth  h0  at  the  wall.  The 
results  of  such  a  calculation  are  shown  in  Figs.  10.7.3a  and  10.7. 3b: 
In  Fig.  10.7.4  we  give  a  curve  showing  (h2  —  A0)/^o  as  a  function  of 
(Aj  —  AO)/AO>  that  is,  we  give  a  curve  showing  the  increase  in  depth 
after  reflection  as  a  function  of  the  relative  height  (Ax  —  /*0)/^o  °*  t'ie 
incoming  wave. 


100 


80 


60 


40 


20 


*\> 


1 


I 


1 


•i    o 


2  4  6  8  10  12  14  16 

Fig.  10.7.4.  Height  of  the  reflected  bore 


hrho 


For  A!/AO  near  to  unity,  i.e.  for  (/^  —  hQ)/h0  small,  it  is  not  difficult 
to  show  that 

fin  fin  fl-t     ' lle\ 

(1(\  7  K\  _* -  r*u  9  .       

^AvF.i.tjy  -   £t  . 

f&Q  f?/Q 

From  this  relation  we  may  write  h2  —  h0  ~  2(7^  —  h0)  if  (Aj  —  A0)  is 
small,  i.e.  the  increase  in  the  depth  of  the  water  after  reflection  is 


LONG   WAVES    IN    SHALLOW    WATER 


333 


twice  the  height  of  the  incoming  wave  when  the  latter  is  small.  This 
is  what  one  might  expect  in  analogy  with  the  reflection  of  acoustic 
waves  of  small  amplitude.  However,  if  h^h^  is  not  small,  the  water 
increases  in  depth  after  reflection  by  a  factor  larger  than  2.  For 
instance,  if  AX/A0  is  2,  then  h2  —  h0  ~  3(/&x  —  A0);  while  if  h^h^  is  10, 
then  A2  —  A0^  35(At  —  A0),  as  one  sees  from  the  graph  of  Fig. 
10.7.4.  In  other  words,  the  reflection  of  a  shock  or  bore  from  a  rigid 
wall  results  in  a  considerable  increase  in  height  and  hence  also  in  pres- 
sure against  the  wall  if  the  incoming  wave  is  high.  In  fact,  for  very 
high  waves  the  total  pressure  p  per  unit  width  of  the  wall  could  be 
shown  to  vary  as  the  cube  of  the  depth  ratio  hjh^. 

In  the  upper  curve  of  Fig.  10.7.3a  we  have  drawn  a  curve  for  the 
analogous  problem  in  gas  dynamics,  i.e.  the  reflection  of  a  shock  from 
the  stopped  end  of  a  tube.  In  the  case  of  air  with  an  adiabatic  expo- 
nent y  —  1.4  the  density  ratio  Q2/6i  as  a  function  of  Q^QQ  (in  an 
obvious  notation)  is  plotted  as  a  dotted  curve  in  the  figure.  As  we 
see,  the  density  in  air  on  reflection  is  higher  than  the  corresponding 
quantity,  the  depth,  in  the  analogous  case  in  water.  However,  the 
curve  for  air  ends  at  Ql/QQ  =  6,  since  it  is  not  possible  to  have  a  shock 
wave  in  a  gas  with  y  =  1.4  which  has  a  higher  density  ratio.  In  water 
there  is  no  such  restriction.  The  explanation  for  this  difference  lies 
in  the  fact  that  the  energy  law  is  assumed  to  hold  across  a  shock  in  gas 
dynamics,  but  not  in  our  theory  for  water  waves. 


10.8.  The  breaking  of  a  dam 

At  the  end  of  section  10.5  we  gave  the  solution  to  an  idealized  ver- 
sion of  the  problem  of  determining  the  flow  which  results  from  the  sud- 
den destruction  of  a  dam  if  it  is  assumed  that  the  downstream  side 


Dam 


Fig.  10.8.1.  Breaking  of  a  dam 


384 


WATER    WAVES 


of  the  dam  is  initially  free  of  water.  In  the  present  section  we  consider 
the  more  general  problem  which  arises  when  it  is  assumed  that  there 
is  water  of  constant  depth  on  the  downstream  as  well  as  the  upstream 
side  of  the  dam.  Or,  as  the  situation  could  also  be  described:  a  hori- 
zontal tank  of  constant  cross  section  extending  to  infinity  in  both 
directions  has  a  thin  partition  at  the  section  x  =  0.  For  x  >  0  the 
water  has  the  depth  h0  and  for  x  <  0  the  depth  hv  with  A0  <  hl9  as 
indicated  in  Fig.  10.8.1.  The  water  is  assumed  to  be  at  rest  on  both 
sides  of  the  dam  initially.  At  the  time  t  =  0  the  dam  is  suddenly  des- 
troyed, and  our  problem  is  to  determine  the  subsequent  motion  of 
the  water  for  all  x  and  t. 

The  special  case  h0  =  0— the  cavitation  case— was  treated,  as  we 
have  already  mentioned,  at  the  end  of  section  10.5.  We  found  there 
that  the  discontinuity  for  x  =  0  and  t  —  0  was  instantly  wiped  out 
and  that  the  surface  of  that  portion  of  the  water  in  motion  took  the 
form  of  a  parabola  tangent  to  the  #-axis  (i.e.  to  the  bottom)  at  the 
point  x  =  —  2\/ghlt  =  —  2o1<,  in  which  t  is  the  time  and  cx  the  wave 


x=-c,t 


(I) 


Free  surface  at  t  *  t 


(3) 


(2) 


Fig.  10.8.2.  Breaking  of  a  dam 


LONG   WAVES    IN    SHALLOW    WATER  335 

speed  in  water  of  depth  hv  If  A0  is  different  from  zero  we  might  there- 
fore reasonably  expect  (on  the  basis  of  the  discussion  at  the  beginning 
of  section  10.6)  that  a  shock  wave  would  develop  sooner  or  later  on 
the  downstream  side,  since  the  water  pushing  down  from  above  acts 
somewhat  like  a  piston  being  pushed  downstream  with  an  accelera- 
tion. In  fact,  since  the  water  at  x  =  0  seems  likely  to  acquire  instan- 
taneously a  velocity  different  from  zero  it  is  plausible  that  a  shock 
would  be  created  instantly  on  the  downstream  side.  The  simplest 
assumption  to  make  would  be  that  the  shock  then  moves  downstream 
with  constant  velocity  £  (cf.  Fig.  10.8.2).  If  this  were  so,  the  state  of 
the  water  immediately  behind  the  shock  (i.e.  on  the  upstream  side 
of  it)  would  be  constant  for  all  time,  since  the  velocity  u2  and  depth 
A2  on  the  upstream  side  of  the  shock  would  have  the  constant  values 
determined  from  the  shock  relations  for  the  fixed  values  UQ  =  0  and 
h  —  h0  for  the  velocity  and  depth  on  the  downstream  side  and  the 
assumed  constant  value  |  for  the  shock  velocity.  However,  it  is  clear 
that  the  constant  state  behind  the  shock  could  not  extend  indefinitely 
upstream  since  u2  ^  0  while  the  velocity  of  the  water  far  upstream  is 
zero.  Since  we  undoubtedly  are  dealing  with  a  depression  wave  be- 
hind the  shock  it  seems  plausible  to  expect  that  the  constant  state 
behind  the  shock  changes  eventually  at  some  point  upstream  into  a 
centered  simple  wave  of  the  type  discussed  in  section  10.5.  In  Fig. 
10.8.2  we  indicate  in  an  #,  2-plane  a  motion  which  seems  plausible  as  a 
solution  of  our  problem.  In  the  following  we  shall  show  that  such  a 
motion  can  be  determined  in  a  manner  compatible  with  our  theory 
for  every  value  of  the  ratio  A0//ir 

As  indicated  in  Fig.  10.8.2,  we  consider  four  different  regions  in  the 
fluid  at  any  time  t  =  /0:  the  zone  (0)  is  the  zone  of  quiet  downstream 
which  is  terminated  on  the  upstream  side  by  the  shock  wave,  or  bore; 
the  zone  (2)  is  a  zone  of  constant  state  in  which  the  water,  however,  is 
not  at  rest;  the  zone  (3)  is  a  centered  simple  wave  which  connects  the 
constant  state  (2)  with  the  constant  state  (1 )  of  the  undisturbed  water 
upstream.  We  proceed  to  show  that  such  a  motion  exists  and  to  deter- 
mine it  explicitly  for  all  values  of  the  ratio  h^h^  between  zero  and  one. 

For  this  purpose  it  is  convenient  to  write  the  shock  conditions  for 
the  passage  from  the  state  (0)  to  the  state  (2)  in  the  form 

(10.8.1)  -!(«*•-{)=«<$+*;); 

(10.8.2)  cl(u2  -  f)  =  -  cfe 

which  are  the  same  conditions  as  (10.6.18)  and  (10.6.19;  with 


386  WATER   WAVES 

replaced  by  c\  =  ghi9  i.e.  by  the  square  of  the  wave  propagation  speed 
in  water  of  depth  ht.  By  eliminating  c\  from  (10.8.1  )  by  use  of  (10.8.2) 
and  then  solving  the  resulting  quadratic  for  u2  one  readily  obtains 


(10.8.3)          u2/c0  =  #cc  -  -     1  +  (Vl 

(The  plus  sign  before  the  radical  was  taken  in  order  that  u2  —  f  and 
—  -  f  should  have  the  same  sign.  We  observe  also  that  only  positive 
values  of  |  and  u2  are  in  question  throughout  our  entire  discussion 
since  the  side  of  (0)  is  the  front  side  of  the  shock  and  the  positive 
^-direction  is  taken  to  the  right.)  It  is  also  useful  to  eliminate  u2 
from  (10.8.3)  by  using  (10.8.2);  the  result  is  easily  put  into  the  form 


(10.8.4)  =  (i  (VT+8tfM*  -  1)}*. 
co 

The  relations  (10.8.3)  and  (10.8.4)  yield  the  velocity  u2  and  the  wave 
speed  c2  behind  the  shock  as  functions  of  f  and  the  wave  speed  c0  in 
the  undisturbed  water  on  the  downstream  side  of  the  dam.  We  pro- 
ceed to  connect  the  state  (2)  by  a  centered  simple  wave  (cf.  the  dis- 
cussion in  section  10.5)  with  the  state  (1).  In  the  present  case  the 
straight  characteristics  in  the  zone  (3)  (cf.  Fig.  10.8.2)  are  those  with 
the  slope  u  —  c  (rather  than  the  slope  u  +  c  as  in  section  10.5); 
hence  the  straight  characteristics  which  delimit  the  zone  (3)  are  the 
lines  x  =  —  cj  on  the  left  and  x  —  (u2  —  c2)t  on  the  right.  Along 
each  of  the  curved  characteristics  in  zone  (3)—  one  of  these  is  indicated 
schematically  by  a  dotted  curve  in  Fig.  10.  8.  2  —the  quantity  u  +  2c 
is  a  constant;  it  follows  therefore  that  on  the  one  hand 

(10.8.5)  u  +  2c  =  2cx 
since  u±  =  0,  while  on  the  other  hand 

(10.8.6)  u  +  2c  =  u2  +  2c2 
throughout  the  zone  (3).  The  relation 

(10.8.7)  u2/c0  +  2c2/c0  -  2^/Co 

must  therefore  hold.  Our  statement  that  a  motion  of  the  type  shown 
in  Fig.  10.8.2  exists  for  every  value  of  the  depth  ratio  h^h^—  or,  what 
amounts  to  the  same  thing,  the  ratio  cj/cjj  —  is  equivalent  to  the  state- 
ment that  the  relation  (10.8.7)  furnishes  through  (10.8.3)  and  (10.8.4) 
an  equation  for  f/c0  which  has  a  real  positive  root  for  every  value  of 
C^CQ  larger  than  one.  This  is  actually  the  case.  In  Fig.  10.8.3  we  have 
plotted  curves  for  u2/cQ9  2c2/cQ,  and  u2/cQ  +  2c2/cQ  as  functions  of  f  /c0. 
Once  the  curves  of  Fig.  10.8.3  have  been  obtained,  our  problem  can 


LONG   WAVES    IN    SHALLOW   WATER  337 


be  considered  solved  in  principle:  From  the  given  value  of 
cilco  we  can  determine  £/c0  from  the  graph  (or,  by  solving  (10.8.7)). 
The  values  of  u2/cQ  and  c2/c0  are  then  also  determined,  either  from  the 
graph  or  by  use  of  (10.8.3)  and  (10.8.4).  The  constant  state  in  the 
zone  (2)  would  therefore  be  known.  In  zone  (3)  the  motion  can  now 
be  determined  exactly  as  in  section  10.5;  we  would  have  along  the 
straight  characteristics  in  this  zone  the  relations 

dx        x  , 

-  =  -  =  U-c  =  2c1-3c  =  ^-c1 

from  which 


i 

(10.8.8)  C2  =  j2cl  -l  ,  and 

(10.8.9)  " 

Thus  the  water  surface  in  the  zone  (3)  is  curved  in  the  form  of  a 
parabola  in  all  cases.*  At  the  junctions  with  both  zones  (1)  and  (2) 
the  parabola  does  not  have  a  horizontal  tangent,  so  that  the  slope  of 
the  water  surface  is  discontinuous  at  these  points. 

Some  interesting  conclusions  can  be  drawn  from  (10.8.8)  and 
(10.8.9).  By  comparison  with  Fig.  10.8.2  we  observe  that  the  J-axis, 
i.e.  the  line  x  =  0,  is  a  characteristic  belonging  to  the  zone  (3)  pro- 
vided that  u2  ^  c2  since  the  terminal  characteristic  of  the  zone  (3) 
on  the  right  lies  on  the  J-axis  or  to  the  right  of  it  in  this  case.  If  this 
condition  is  satisfied  we  observe  from  (10.8.8)  and  (10.8.9)—  which 
are  then  valid  on  the  /-axis—  that  c  and  u  are  both  independent  of  t 
at  x  =  0,  which  means  that  the  depth  of  the  water  and  its  velocity  u 
arc  both  independent  of  t  at  this  point,  i.e.  at  the  original  location  of 
the  dam,  and  hence  that  the  volume  of  water  crossing  the  original 
dam  site  per  unit  of  time  (and  unit  of  width)  dQ/dt  =  uh  is  independ- 
ent of  time  although  the  motion  as  a  whole  is  not  a  steady  motion. 
In  fact,  h  —  \hl  and  u  =  f  cx  for  all  time  t  at  this  point.  In 
addition,  u  and  c,  and  thus  also  dQ/dt,  are  not  only  independent  of  t 
as  long  as  u2  ^  c2,  but  also  independent  of  the  undisturbed  depth  A0 
on  the  lower  side  of  the  dam  if  Ax  is  held  fixed.  Of  course,  it  is  clear  that 
AO/A!  must  be  kept  under  a  certain  value  (which  from  section  10.5 
evidently  must  be  less  than  4/9)  or  the  condition  u2  I>  c2  could  not  be 

*  Relations  (10.8.8)  and  (10.8.9)  are  exactly  the  same  as  (10.5.8)  and  (10.5.7) 
except  for  a  change  of  sign  which  arises  from  a  different  choice  of  the  positive 
aj-direction. 


888 


WATER  WAVES 


7, 


I 


I 


V'o 


I 


I 


I 


I 


234567 
Fig.  10.8.3.  Graphical  solution  for  £ 


fulfilled.  In  fact,  the  critical  value  of  the  ratio  /^//^  at  which  u2  •=  c2 
can  be  determined  easily  by  equating  the  right  hand  sides  of  (10.8.3) 
and  (10.8.4)  and  determining  the  value  of  |/c0  for  this  case,  after 

which  c2/c0  =  Vh2/h0  is  known  from  (10.8.4).  Since  c2  ~  f  q  in 
the  critical  case— either  from  the  known  fact  that  we  still  have 
h2  =  |Aj_  in  this  case,  or  from  (10.8.8)  with  x  —  0— we  thus  are 
able  to  compute  the  critical  value  of  cj/cj  =  A1/A0.  A  numerical  cal- 
culation yields  for  the  critical  value  of  the  ratio  h^h^  the  value  7.225, 
or  for  AQ/A!  the  value  .1384.  Thus  if  the  water  depth  on  the  lower  side 
of  the  dam  is  less  than  13.8  percent  of  the  depth  above  the  dam  the 
discharge  rate  on  breaking  the  dam  will  be  independent  of  the  original 
depth  on  the  lower  side  as  well  as  independent  of  the  time.  However, 
if  AQ/AJ  exceeds  the  critical  value  .1384,  the  depth,  velocity,  and  dis- 
charge rate  will  depend  on  h0;  but  they  continue  to  be  independent  of 


LONG    WAVES    IN    SHALLOW   WATER 


330 


the  time  since  the  line  x  =  0  in  the  x>  J-plane  is  under  the  latter  cir- 
cumstances contained  in  the  zone  (2),  which  is  one  of  constant  state. 

The  above  results,  which  at  first  perhaps  seem  strange,  can  be 
made  understandable  rather  easily  from  the  physical  point  of  view, 
as  follows.  If  the  zone  (3)  includes  the  /-axis  (i.e.  if  A0/At  is  below  the 
critical  value)  we  may  apply  (10.8.8)  and  (10.8.9)  for  x  =  0  to  obtain 
at  this  point  c  =  u  —  fcj.  In  other  words,  the  flow  velocity  at  the 
dam  site  is  in  this  case  just  equal  to  the  wave  propagation  speed 
there.  For  x  >  0,  i.e.  downstream  from  the  dam,  we  observe  from 
(10.8.8)  and  (10.8.9)  that  u  is  greater  than  c.  Since  c  is  the  speed  at 
which  the  front  of  a  disturbance  propagates  relative  to  the  moving 
water  we  see  that  changes  in  conditions  below  the  dam  can  have  no 
effect  on  the  flow  above  the  dam  since  the  flow  velocity  at  all  points 
below  the  dam  is  greater  than  the  wave  propagation  speed  at  these 
points  and  hence  disturbances  can  not  travel  upstream.  However, 
once  AQ//^  is  taken  higher  than  the  critical  value,  the  flow  velocity 
at  the  dam  will  be  less  than  the  wave  propagation  speed  at  this  point, 
as  one  can  readily  prove,  and  we  could  no  longer  expect  the  flow  at 
that  point  to  be  independent  of  the  initial  depth  assumed  on  the 
downstream  side. 

The  discharge  rate  dQ/dt  —  hu  per  unit  width  at  the  dam,  i.e.  at 
x  =  0,  is  plotted  in  Fig.  10.8.4  as  a  function  of  the  depth  A0.  In  accord- 
ance with  our  discussion  above  we  observe  that  dQ/dt  remains  con- 


0.3 


0.2 


0.1 


0.2          0.4          0.6          0.8 


1.0 


Fig.  10.8.4.  Discharge  rate  at  the  dam 


stant  at  the  value  dQ/dt  =  .296hlcl  until  h^h^  reaches  the  critical 
value  .138,  after  which  it  decreases  steadily  to  the  value  zero  when 
h0  =  /ir  i.e.  when  the  initial  depth  of  the  water  below  the  dam  is  the 
same  as  that  above  the  dam. 


340 


WATER   WAVES 


Another  feature  of  interest  in  the  present  problem  is  the  height  of 
the  bore,  i.e.  the  quantity  h2  —  A0,  as  a  function  of  the  original  depth 
ratio  h^jhv  When  A0  =  0  we  know  that  there  is  no  bore  and  the  water 
surface  (as  we  found  in  section  10.5)  appears  as  in  Fig.  10.8.5.  The 
water  surface  at  the  front  of  the  wave  on  the  downstream  side  is 
tangent  to  the  bottom  and  moves  with  the  speed  2clB  On  the  other 
hand,  when  h^h^  approaches  the  other  extreme  value,  i.e.  unity,  it  is 
clear  that  the  height  h2  —  h0  of  the  bore  must  again  approach  zero. 
Hence  the  height  of  the  bore  must  attain  a  maximum  for  a  certain 


-Parabola 


x  =  -c,t  x  =  2c,t          x 

Fig.  10.8.5.  Motion  down  the  dry  bed  of  a  stream 

value  of  hQ/hv  In  Fig.  10.8.6  we  give  the  result  of  our  calculations  for 
A2  — -  A0  as  a  function  of  h^h^.  The  curve  rises  very  steeply  to  its 
maximum  A2  —  A0  =  .32^  for  h^h^  =  .176  and  then  falls  to  zero 
again  when  hQ  =  hv  It  is  rather  remarkable  that  the  bore  can  attain 
a  height  which  is  nearly  1/3  as  great  as  the  original  depth  of  the  water 
behind  the  dam. 


0.1 


1 


1 


0          0.2          04          0.6          0.8  IX) 

Fig.  10.8.6.  Maximum  height  of  the  bore 


LONG   WAVES   IN   SHALLOW   WATER 


341 


It  is  instructive  to  describe  the  motion  by  means  of  the  #,  £-plane 
when  AO/AJ  is  near  its  two  limit  values  unity  and  zero.  These  two  cases 
are  schematically  shown  in  Fig.  10.8.7.  When  h0  ~  hv  we  note  that 
the  zone  (3)  is  very  narrow  and  that  the  shock  speed  f  approaches  cl9 
i.e.  the  propagation  speed  of  small  disturbances  in  water  of  depth  hl9 
corresponding  to  the  fact  that  the  height  of  the  shock  wave  tends  to 
zero  as  A0->  h^  (cf.  Fig.  10.8.6).  The  other  limit  situation,  i.e.  h0  c±  0, 
is  more  interesting.  Since  we  tacitly  consider  h±  to  remain  fixed  in  our 
present  discussion,  and  hence  that  h2  is  also  fixed  since  we  are  in  the 
supercritical  case,  it  follows  (for  example  from  (10.6.23)  with  A2  in 
place  of  hi)  that  f ->oo  as  A0->  0.  On  the  other  hand  as  we  see  from 


Free 


(I) 


surface  for  t«tQ 


(3) 


(2) 


(0) 


Fig.  10.8.7.  Limit  cases 

Fig.  10.8.6,  the  height  h2  —  A0  of  the  shock  wave  tends  to  zero  rather 
slowly  as  A0  ->  0.  In  the  limit,  point  P  becomes  the  front  of  the  wave 
in  accordance  with  the  motion  indicated  by  Fig.  10.8.5.  Thus  as 
A0  ->  0  the  shock  wave  becomes  very  small  in  height  but  moves  down- 
stream with  great  speed;  or,  as  we  could  also  say,  in  the  limit  the  water 
in  front  of  the  point  P  is  pinched  out  and  P  is  the  front  of  the  wave. 


342  WATER   WAVES 

10.9.  The  solitary  wave 

It  has  long  been  a  matter  of  observation  that  wave  forms  of  a  per- 
manent type  other  than  the  uniform  flows  with  an  undeformed  free 
surface  occur  in  nature;  for  example,  Scott  Russell  [R.14]  reported 
in  1844  his  observations  on  what  has  since  been  called  the  solitary 
wave,  which  is  a  wave  having  a  symmetrical  form  with  a  single  hump 
and  which  propagates  at  uniform  velocity  without  change  of  form. 
Later  on,  Boussinesq  [B.16]  and  Rayleigh  [R.3]  studied  this  problem 
mathematically  and  found  approximations  for  the  form  and  speed  of 
such  a  solitary  wave.  Korteweg  and  de  Vries  [K.15]  modified  the 
method  of  Rayleigh  in  such  a  way  as  to  obtain  waves  that  arc  periodic 
in  form— called  cnoidal  waves  by  them— and  which  tend  to  the  soli- 
tary wave  found  by  Rayleigh  in  the  limiting  case  of  long  wave  lengths. 
A  systematic  procedure  for  determining  the  velocity  of  the  solitary 
wave  has  been  developed  by  Weinstein  [W.6], 

At  the  beginning  of  section  10.7  we  have  shown  that  the  only  con- 
tinuous waves  furnished  by  the  theory  used  so  far  in  this  chapter 
which  progress  unchanged  in  form  are  of  a  very  special  and  rather 
uninteresting  character,  i.e.,  they  are  the  motions  with  uniform  velo- 
city and  horizontal  free  surface.*  This  would  seem  to  be  in  crass  con- 
tradiction with  our  intention  to  discuss  the  solitary  wave  in  terms  of 
the  shallow  water  theory,  and  it  has  been  regarded  by  some  writers  as  a 
paradox. **The  author's  view  is  that  this  paradox— like  most  others  — 
becomes  not  at  all  paradoxical  when  properly  examined.  What  is 
involved  is  a  matter  of  the  range  of  accuracy  of  a  given  approximate 
theory,  and  also  the  fact  that  a  perturbation  or  iteration  scheme  of 
universal  applicability  does  not  exist:  one  must  always  modify  such 
schemes  in  accordance  with  the  character  of  the  problem.  In  the  pre- 
sent case,  the  salient  fact  is  that  the  theory  used  so  far  in  the  present 
chapter  represents  the  result  of  taking  only  the  lowest  order  terms  in 
the  shallow  water  theory  as  developed  in  section  4  of  Chapter  2,  and 
it  is  necessary  to  carry  out  the  theory  to  include  terms  of  higher  order 

*  If  motions  with  a  discontinuity  are  included  in  the  discussion,  then  the  motion 
of  a  bore  is  the  only  other  possibility  up  to  now  in  this  chapter  with  regard  to 
waves  propagating  unchanged  in  form. 

**  Birkhoff  [B.ll,  p.  23],  is  concerned  more  about  the  fact  that  the  shallow 
water  theory  predicts  that  all  disturbances  eventually  lead  to  a  wave  which 
breaks  when  on  the  other  hand  Struik  [S.29]  has  proved  that  periodic  progressing 
waves  of  finite  amplitude  exist  in  shallow  water.  In  the  next  section  the  problem 
of  the  breaking  of  waves  is  discussed.  Ursell  [U.ll]  casts  doubt  on  the  validity 
of  the  shallow  water  theory  in  general  because  it  supposedly  does  not  give  rise 
to  the  solitary  wave. 


LONG   WAVES    IN    SHALLOW   WATER  348 

if  one  wishes  to  obtain  an  approximation  to  the  solution  of  the  problem 
of  the  solitary  wave.  This  has  been  done  by  Keller  [K.6],  who  finds 
that  the  theory  of  Friedrichs  [F.ll]  presented  in  Chapter  2,  when  car- 
ried out  to  second  order,*  yields  both  the  solitary  wave  and  cnoidal 
waves  of  the  type  found  by  Korteweg  and  de  Vries  [K.15]  (thus  the 
shallow  water  theory  is  capable  of  yielding  periodic  progressing  waves 
of  finite  amplitude).  As  lowest  order  approximation  to  the  solution 
of  the  problem,  Keller  finds  (as  he  must  in  view  of  the  remarks  above), 
that  the  only  possibility  is  the  uniform  flow  with  undeformed  free 
surface,  but  if  the  speed  U  of  the  flow  is  taken  at  the  critical  value 
U  =  \/gk  with  h  the  undisturbed  depth,  then  a  bifurcation  phenome- 
non occurs  (that  is,  among  the  set  of  uniform  flows  of  all  depths  and 
velocities,  the  solitary  wave  occurs  as  a  bifurcation  from  the  special 
flow  with  the  critical  velocity)  and  the  second  order  terms  in  the  de- 
velopment of  Friedrichs  lead  to  solitary  and  cnoidal  waves  with 
speeds  in  the  neighborhood  of  this  value.  To  clinch  the  matter,  it  has 
been  found  by  Friedrichs  and  Hycrs  [F.13]  that  the  existence  of  the 
solitary  wave  can  be  proved  rigorously  by  a  scheme  which  starts  with 
the  solution  of  Keller  as  the  term  of  lowest  order  and  proceeds  by 
iterations  with  respect  to  a  parameter  in  essentially  the  same  manner 
as  in  the  general  shallow  water  theory.**  In  the  following,  we  shall 
derive  the  approximation  to  the  solution  of  the  solitary  Wave  problem 
following  the  method  of  Friedrichs  and  Hyers  rather  than  the  general 
expansion  scheme  which  was  used  by  Keller,  and  we  can  then  state 
the  connection  between  the  two  in  more  detail. 

The  author  thus  regards  the  nonlinear  shallow  water  theory  to  be 
well  founded  and  not  at  all  paradoxical.  Indeed,  the  linear  theory  of 
waves  of  small  amplitude  treated  at  such  length  in  Part  II  of  this 
book  is  in  essentially  the  same  position  as  regards  rigorous  justifica- 
tion as  is  the  shallow  water  theory:  we  have  only  one  or  two  cases  so 
far  in  which  the  linear  theory  of  waves  of  small  amplitude  is  shown  to 
be  the  lowest  order  term  in  a  convergent  development  with  respect  to 
amplitude.  We  refer,  in  particular,  to  the  theory  of  Levi-Civita  [L.7] 
and  Struik  [S.29]  in  which  the  former  shows  the  existence  of  periodic 
progressing  waves  in  water  of  infinite  depth  and  the  latter  the  same 
thing  (and  by  the  same  method)  for  waves  in  water  of  finite  constant 

*  In  order  to  fix  all  terms  of  second  order,  Keller  found  it  necessary  to  make 
use  of  certain  relations  which  result  from  carrying  the  development  of  some  of 
the  equations  up  to  terms  of  third  order. 

**  W.  Littman,  in  a  thesis  to  appear  in  Communs.  Pure  and  AppJ.  Math., 
has  proved  rigorously  in  the  same  way  the  existence  also  of  cnoidal  waves. 


844  WATER   WAVES 

depth.*  This  theory  will  be  developed  in  detail  in  Chapter  12.  It  might 
be  added  that  those  who  find  the  nonlinear  shallow  water  theory 
paradoxical  in  relation  to  the  solitary  wave  phenomenon  should  by 
the  same  type  of  reasoning  also  find  the  linear  theory  paradoxical, 
since  it  too  fails  to  yield  any  approximation  to  the  solitary  wave,  even 
when  carried  out  to  terms  of  arbitrarily  high  order  in  the  amplitude, 
except  the  uniform  flow  with  undisturbed  free  surface.  In  fact,  if  one 
were  to  assume  that  a  development  exists  for  the  solitary  wave  which 
proceeds  in  powers  of  the  amplitude  as  in  the  theory  discussed  in  the 
first  part  of  Chapter  2,  it  is  easily  proved  that  the  terms  of  all  orders 
in  the  amplitude  are  identically  zero.  There  is  no  paradox  here,  how- 
ever; rather,  the  problem  of  the  solitary  wave  is  one  in  which  the 
solution  is  not  analytic  in  the  amplitude  in  the  neighborhood  of  its 
zero  value,  but  rather  has  a  singularity —possibly  of  the  type  of  a 
branch  point— there.  Thus  a  different  kind  of  development  is  needed, 
and,  as  we  have  seen,  one  such  possibility  is  a  development  of  the  type 
of  the  shallow  water  theory  starting  with  a  nonlinear  approximation. 
Another  possibility  has  been  exploited  by  Lavrentieff  [L.4]  in  a 
difficult  paper;  Lavrentieff  proves  the  existence  of  the  solitary  wave 
by  starting  from  the  solutions  of  the  type  found  by  Struik  for  periodic 
waves  of  finite  amplitude  and  then  making  a  passage  to  the  limit  by 
allowing  the  wave  length  to  become  large  and,  presumably,  in  such  a 
way  that  the  parameter  gh/U2  tends  to  unity.  This  procedure  of 
Lavrentieff  thus  also  starts  with  a  nonlinear  first  approximation. 
The  problem  thus  furnishes  another  good  example  of  the  well-known 
fact  that  it  is  not  always  easy  to  guess  how  to  set  up  an  approximation 
scheme  for  solving  nonlinear  boundary  value  problems,  since  the 
solution  may  behave  in  quite  unexpected  ways  for  particular  values  of 
the  parameters.  Hindsight,  however,  can  help  to  make  the  necessity 
for  procedures  like  those  of  Friedrichs  and  Hycrs  and  of  Lavrentieff 
in  the  present  case  more  apparent:  we  have  seen  in  Chapter  7.4  that 
a  steady  flow  with  the  critical  speed  U  =  VgA  is  in  a  certain  sense 
highly  unstable  since  the  slightest  disturbance  would  lead,  in  terms 
of  the  linear  theory  for  waves  of  small  amplitude,  to  a  motion  in  which 
infinite  elevations  of  the  free  surface  would  occur  everywhere;  thus 
the  linear  theory  of  waves  of  small  amplitude  seems  quite  inappro- 
priate as  the  starting  point  for  a  development  which  begins  with  a 

'    *  L.  Nirenberg  [N.2]  has  recently  proved  the  existence  of  steady  waves  of 
finite  amplitude  caused  by  flows  over  obstacles  in  the  bed  of  a  stream. 


LONG    WAVES    IN    SHALLOW   WATER 


345 


uniform  flow  at  the  critical  speed,  and  one  should  consequently  use 
a  basically  nonlinear  treatment  from  the  outset. 

We  turn  now  to  the  discussion  of  the  solution  of  the  solitary  wave 
problem.  The  theory  of  Friedrichs  and  Hyers  begins  with  a  formula- 
tion of  the  general  problem  that  is  the  same  as  that  devised  by  Levi- 
Civita  for  treating  the  problem  of  existence  of  periodic  waves  of  finite 
amplitude,  and  which  was  motivated  by  the  desire  to  reformulate  the 
problem  in  terms  of  the  velocity  potential  <p  and  stream  function  \p  as 
independent  variables  in  order  to  work  in  the  fixed  domain  between 
the  two  stream  lines  \p  =  const,  corresponding  to  the  bottom  and  the 
free  surface  instead  of  in  the  partially  unknown  domain  in  the  physi- 
cal plane.  We  therefore  begin  with  the  general  theory  of  irrotational 
waves  in  water  when  a  free  surface  exists.  The  wave  is  assumed  to 
be  observed  from  a  coordinate  system  which  moves  with  the  same 
velocity  as  the  wave,  and  hence  the  flow  can  be  regarded  as  a  steady 
flow  in  this  coordinate  system. 


Fig.  10.9.1.  The  solitary  wave 
A  complex  velocity  potential  %'(x',  y')  =%'(z'): 

(10.9.1 )  x'  =  ?'  +  *V>     *'  =  x'  +  {y' 

is  sought  in  an  #',  i/'-plane  (cf.  Fig.  10.9.1)  such  that  at  infinity  the 
velocity  is  U  and  the  depth  of  the  water  is  h.  %'  is  of  course  an  analytic 
function  of  z'.  The  real  harmonic  functions  y'  and  \p'  represent  the 
velocity  potential  and  the  stream  function.  The  complex  velocity 
wf  =  d%'/dz'  is  given  by 

(10.9.2)  w'  =  u'  —  iv'9 

in  which  u'  and  v'  are  the  velocity  components.  This  follows  by  virtue 
of  the  Cauchy-Riemann  equations: 

(10.9.3)  <p'x<  =  v>V»         VV  =  ~  V>'*'' 


846  WATER   WAVES 

since  w'  ••=  q>'x>  +  i\p'x>.  It  is  convenient  to  introduce  new  dimensionless 
variables: 

(10.9.4)  z  =  z'/h,        w  =  w'/U,        x  =  V  +  *V  =  X'/(hU)> 
and  a  parameter  y: 

(10.9.5)  y  =  gW2- 

In  terms  of  these  quantities  the  free  surface  corresponds  to  \p  —  1 
if  the  bottom  is  assumed  given  by  y>  =  0,  since  the  total  flow  over  a 
curve  extending  from  the  bottom  to  the  free  surface  is  Uh.  The 
boundary  conditions  are  now  formulated  as  follows: 

(10.9.6)  v  =  —  Jmw  =  0         at  y>  =  0, 

(10.9.7)  \  \w\2  +  yy  =  const.         at  \p  =  1. 

The  second  condition  results  from  Bernoulli's  law  on  taking  the 
pressure  to  be  constant  at  the  free  surface  and  the  density  to  be  unity, 
as  one  sees  from  equation  (1.3.4)  of  Ch.  1.  At  oo  we  have  the  condition 

(10.9.8)  w  ->  1         as  |  x  |  -*  oo. 

We  assume  now  that  the  physical  plane  (i.e.  the  #,  i/-plane)  is  mapped 
by  means  of  %(z)  into  the  99,  y-plane  in  such  a  way  that  the  entire  flow 
is  mapped  in  a  one-to-one  way  on  the  strip  bounded  by  \p  —  0  and 
\p  =  1.*  In  this  case  the  inverse  mapping  function  z(%)  exists,  and  we 
could  regard  the  complex  velocity  w  as  a  function  of  #  defined  in  the 
strip  bounded  by  \p  =  0,  \p  —  1  in  the  ^-plane.  We  then  determine  the 
analytic  function  w(%)  in  that  strip  from  the  boundary  conditions 
(10.9.6),  (10.9.7),  (10.9.8),  after  which  %(z)  can  be  found  by  an  inte- 
gration and  the  free  surface  results  as  the  curve  given  by  \p  =  Jm  # =1. 
It  is  convenient,  however,  again  following  Levi-Civita  to  replace 
the  dependent  variable  w  by  another  (essentially  its  logarithm) 
through  the  equation 

(10.9.9)  w  =  <rl(*+tA). 
It  follows  that 

(10.9.10)  \w\  =  ex,         0  =  argiei, 

and  thus  A  =  log  |io|,  with  \w\  the  magnitude  of  the  velocity  vector, 

*  Our  assumption  that  the  mapping  of  the  flow  on  the  /-plane  is  one-to-one 
can  be  shown  rather  easily  to  follow  from  the  other  assumptions  and  Levi-Civita 
carries  it  out.  The  equivalence  of  the  various  formulations  of  the  problem  is  then 
readily  seen.  In  Chapter  12.2  these  facts  are  proved. 


LONG  WAVES  IN  SHALLOW  WATER  847 

and  0  is  the  inclination  relative  to  the  #-axis  of  the  velocity  vector. 
We  proceed  to  formulate  the  conditions  for  the  determination  of  6 
and  A  in  the  <p,  y  plane.  The  condition  (10.9.6)  becomes,  of  course, 
6  =  0  at  \p  =  0.  To  transform  the  condition  (10.9.7)  we  first  differen- 
tiate with  respect  to  q>  along  the  line  ip  =  1  to  obtain 

d  \w\         dy 

(10.9.11)  |  w  |  —  ^  —  -  +  y-^  =  0         on  w  =  1. 

dcp          f  dq>  r 

Since  x  and  y  are  conjugate  harmonic  functions  of  <p  and  \p  we  may 
write 

doc       dy  (px  u 

dv  ~  dw  ~~  <pl  +  cpl  ~~  j  w  |2 

(10.9.12)  {    7         r       Vx-rVv 

dy          ox         v 

d<p          d\p      \w  |2 

in  accordance  with  well-known  rules  for  calculating  the  derivatives 
of  functions  determined  implicitly,  or  from 

dz         1  1  u  +  iv 

d>X       d%       u  —  iv       u2  +  v2  ' 
dz 

As  a  consequence  we  have  from  (10.9.11): 


or,  since  |  w  \   ~  ex  and  v  =  —  J>w  e~i(P+iK}  —  e*  sinO: 

6^A  ^—  =  --  w~2A  sin  0, 
t/9? 

and  since  ^A/S^  =  —dO/dy  because  A  and  6  are  harmonic  conjugates 
it  follows  finally  that 

dO 

(10.9.13)  —  =  w-3Asin0         atw  =  1. 

oy 

The  boundary  conditions  0  =  0  for  \f  =  0  and  (10.9.13)  at  \p  =  1  are 
Lcvi-Civita's  conditions,  but  the  condition  at  oo  imposed  here  is  re- 
placed in  Levi-Civita's  and  Struik's  work  by  a  periodicity  condition  in 
x,  —and  this  makes  a  great  difference.  Levi-Civita  and  Struik  proceed 
on  the  assumption  that  a  disturbance  of  small  amplitude  is  created 
relative  to  the  uniform  flow  in  which  w  =  const.;  this  is  interpreted  to 


848  WATER  WAVES 

mean  that  6  +  iX  is  a  quantity  which  can  be  developed  in  powers  of 
a  small  parameter  e,  and  the  convergence  of  the  series  for  sufficiently 
small  values  of  e  is  then  proved.  In  Chapter  12.2  we  shall  give  a  proof 
of  the  convergence  of  this  expansion.  (In  lowest  order,  we  note  that 
the  condition  (10.9.13)  leads  for  small  A  and  6  to  the  condition  dO/dy  — 
yd  =  0  at  \f  —  1  —  in  agreement  with  what  we  have  seen  in  Part  II.) 
In  the  case  of  the  solitary  wave  such  a  procedure  will  not  succeed,  as 
was  explained  above,  or  rather  it  would  not  yield  anything  but  a 
uniform  flow.  The  procedure  to  be  adopted  here  consists  in  developing, 
roughly  speaking,  with  respect  to  the  parameter  y  near  y  —  1;  but, 
as  in  the  shallow  water  theory  in  general  in  the  version  presented  in 
section  4  of  Chapter  2,  we  introduce  a  stretching  of  the  horizontal 
coordinate  <p  which  depends  on  y  while  leaving  the  vertical  coordinate 
unaltered  (see  equation  (10.9.19)).  This  stretching  of  only  one  of  the 
coordinates  is  the  characteristic  feature  of  the  shallow  water  theory. 
(The  approximating  functions  are  then  no  longer  harmonic  in  the  new 
independent  variables.)  Specifically,  we  introduce  the  real  parameter 
x  by  means  of  the  equation 

(10.9.14)  e~*»*  =  y  =  gh/U*. 

This  implies  that  gh/U2  <  1,  but  that  seems  reasonable  since  all  of 
the  approximate  theories  for  the  solitary  wave  lead  to  such  an 
inequality.  We  also  introduce  a  new  function  r,  replacing  A,  by  the 
relation 

(10.9.15)  r  =  A  +  x2. 

For  6((f>9  \p)  and  r(y,  y>)  we  then  have  the  boundary  conditions 

(10.9.16)  6  =  0,         \p  =  0 

90 

(10.9.17)  g-  =  e~3x  sin  0,         y>  =  1. 

For  <p  ->  ±  oo  we  have  the  conditions  imposed  by  the  physical  pro- 
blem: 

(10.9.18)  0->0,         r->*2, 

the  latter  resulting  since  A  ->  0  at  oo  from  |  w  \  =  eK  and  |  w  \  ->  1 
at  oo.  As  we  have  already  indicated,  the  development  we  use  requires 
stretching  the  variable  <p  so  that  it  grows  large  relative  to  \p  when  x 
is  small;  this  is  done  in  the  present  case  by  introducing  the  new  in- 
dependent variables 


LONG    WAVES   IN    SHALLOW   WATER  349 

(10.9.19)  y  =  xy,         ijp  =  y. 

The  dependent  variables  6  and  r  are  now  regarded  as  functions  of  y 
and  ip  and  they  are  then  expanded  in  powers  of  x: 

(10  9  20} 

{     ''     ' 

(We  have  omitted  writing  down  a  number  of  terms  which  in  the 
course  of  the  calculation  would  turn  out  to  have  zero  coefficients.) 
Friedrichs  and  Hyers  have  proved  that  the  lowest  order  terms  in 
these  series,  as  obtained  formally  through  the  use  of  the  boundary 
conditions,  are  the  lowest  order  terms  in  a  convergent  iteration  scheme 
using  x  as  small  parameter.  Their  convergence  proof  also  involves  the 
explicit  use  of  the  stretching  process.  However,  the  proof  of  this 
theorem  is  quite  complicated,  and  consequently  we  content  ourselves 
here  with  the  determination  of  the  lowest  order  terms:  we  remark, 
however,  that  higher  order  terms  could  also  be  obtained  explicitly 
from  the  formal  expansion. 

The  series  in  (10.9.20)  are  now  inserted  in  all  of  the  equations  which 
serve  to  determine  6  and  r9  and  relations  for  the  coefficient  functions 
ri(*P>  yO  and  Oiiy*  V)  are  obtained.  The  Cauchy-Riemann  equations 
for  0  and  r  lead  to  the  equations 

(10.9.21)  0^  =  -XT-,          Ty  =  xOj 

in  terms  of  the  variables  <p  and  \p,  and  the  series  (10.9.20)  then  yield 
the  equations 

(10.9.22)  Tl5  -  0,         0<-  =  -  T^  ,         T2^  =  0*  . 

Thus  TJ  =  TI(IJ>)  is  independent  of  y,  and  integration  of  the  remaining 
equations  gives  the  following  results: 


(10.9.23)  T,  =  -  4y«Tj'  + 


The  primes  refer  to  differentiation  with  respect  to  <p.  An  additive 
arbitrary  function  of  (p  in  the  first  of  these  equations  was  taken  to  be 
zero  because  of  the  boundary  condition  6l  =  0  for  \p  =  0. 

Upon  substitution  of  (10.9.20)  into  the  boundary  condition  (10.9. 
17)  we  find 


350  WATER   WAVES 

and  consequently  we  have  the  equations 
(10.9.24) 

The  first  equation  is  automatically  satisfied  because  of  the  first  equa- 
tion of  (10.9.23).  The  second  equation  leads  through  (10.9.23)  to  the 
condition 

(10.9.25)  r["  =  9^, 

for  TV  as  one  readily  verifies.  Once  rl  has  been  determined,  one  sees 
that  0X  is  also  immediately  fixed  by  the  first  equation  in  (10.9.23). 
Boundary  conditions  are  needed  for  the  third  order  nonlinear  differ- 
ential equation  given  by  (10.9.25);  we  assume  these  conditions  to  be 

(o)   -o, 

(10.9.26)  TJ,  (oo)  =  1, 


These  conditions  result  from  our  assumed  physical  situation:  the  first 
is  taken  since  a  symmetrical  form  of  the  wave  about  its  crest  is  ex- 
pected and  hence  0X(0)  =  0,  the  second  arises  from  (10.9.18),  while 
the  third  is  a  reasonable  condition  that  is  taken  in  place  of  what  looks 
like  the  more  natural  condition  ri(oo)  =  0  since  the  latter  condition 
is  automatically  satisfied,  in  view  of  the  first  equation  of  (10.9.23) 
and  0i(oo)  =  0,  and  thus  docs  not  help  in  fixing  r±  uniquely. 
An  integral  of  (10.9.25)  is  readily  found;  it  is: 

Ti   =lri  +  const., 
and  the  boundary  conditions  yield 


From  this  one  obtains,  finally,  the  solution: 

(10.9.27)  T!(<P)  =  1-3  sech2  (3<p/2), 

and  6l  is  then  fixed  by  (10.9.23).  From  these  one  finds  for  the  shape  of 
the  wave—  that  is,  the  value  of  y  corresponding  to  y  =  1  —  ,  and  for 
the  horizontal  component  u  of  the  velocity  the  equations 

(10.9.28)  y  =  1  +  3*2  sech2  —  , 

2 


(10.9.29)  u  =  1  —  3*2sech2-—  . 

2 


LONG   WAVES   IN   SHALLOW   WATER 


351 


Fig.  10.U.2.  A  solitary  wave 


In  calculating  these  quantities, 
higher  order  terms  in  x  have  been 
neglected.  The  expression  for  the 
wave  profile  is  identical  with  those 
found  by  Boussinesq,  Rayleigh, 
and  Keller.  For  the  velocity  u, 
the  two  former  authors  give  u  =  I 
while  Keller  gives  the  same  expres- 
sion as  above  except  that  the  factor 
3#2  is  replaced  by  another  which 
differs  from  it  by  terms  of  order 
tt4  or  higher. 

Thus  a  solitary  wave  of  sym- 
metrical form  has  been  found  with 
an  amplitude  which  increases  with 
its  speed  17.  Careful  experiments 
to  determine  the  wave  profile  and 
speed  of  the  solitary  wave  have 
been  carried  out  by  Daily  and 
Stephan  [D.I],  who  find  the  wave 
profile  and  velocity  to  be  closely 
approximated  by  the  above  formu- 
las with  a  maximum  error  in  the 
latter  of  2.5  %  at  the  highest  ampli- 
tude-depth ratio  tested.  Fig.  10.9.2 
is  a  picture  of  a  solitary  wave  taken 
by  Daily  and  Stephan;  three  frames 
from  a  motion  picture  film  are 
shown. 


10.10.  The  breaking  of  waves  in  shallow  water.  Development  of  bores 

In  sections  10.4  and  10.6  above  it  has  already  been  seen  that  the 
shallow  water  theory,  which  is  mathematically  analogous  to  the 
theory  of  compressible  flows  in  a  gas,  leads  to  a  highly  interesting 
and  significant  result  in  cases  involving  the  propagation  of  disturb- 
ances into  still  water  that  are  the  exact  counterparts  of  the  corre- 
sponding cases  in  gas  dynamics  involving  the  motions  due  to  the  action 
of  a  piston  in  a  tube  filled  with  gas.  These  cases,  which  are  very  easily 


352  WATER   WAVES 

described  in  terms  of  the  concept  of  a  simple  wave  (cf.  sec.  10.3), 
lead,  in  fact  to  the  following  qualitative  results  (cf.  sec.  10.4):  there  is 
a  great  difference  in  the  mode  of  propagation  of  a  depression  wave 
and  of  a  hump  with  an  elevation  above  the  undisturbed  water  line; 
in  the  first  case  the  depression  wave  gradually  smooths  out,  but  in 
the  second  case  the  front  of  the  wave  becomes  progressively  steeper 
until  finally  its  slope  becomes  infinite.  In  the  latter  case,  the  mathe- 
matical theory  ceases  to  be  valid  for  times  larger  than  those  at  which 
the  discontinuity  first  appears,  but  one  expects  in  such  a  case  that  the 
wave  will  continue  to  steepen  in  front  and  will  eventually  break.  This 
is  the  correct  qualitative  explanation,  from  the  point  of  view  of 
hydrodynamical  theory,  for  the  breaking  of  waves  on  shallow  beaches. 
It  was  advanced  by  Jeffreys  in  an  appendix  to  a  book  by  Cornish 
[C.7]  published  in  1934.  Jeffreys  based  his  discussion  on  the  fact  that 
the  propagation  speed  of  a  wave  increases  with  increase  in  the  height 
of  a  wave  above  the  undisturbed  level.  Consequently,  if  a  wave  is 
created  in  such  a  way  as  to  cause  a  rise  in  the  water  surface  it  follows 
that  the  higher  points  on  the  wave  surface  will  propagate  at  higher 
speed  than  the  lower  points  in  front  of  them— in  other  words  there  is 
a  tendency  for  the  higher  portions  of  the  wave  to  overtake  and  to 
crowd  the  lower  portions  in  front  so  that  the  front  of  the  wave  be- 
comes steep  and  eventually  curls  over  and  breaks;  the  same  argument 
indicates  that  a  depression  wave  tends  to  flatten  out  and  become 
smoother  as  it  advances. 

It  is  of  interest  to  recall  how  waves  break  on  a  shallow  beach. 
Figures  10.10.1,  10.10.2,  and  10.10.3  are  photographs*  of  waves  on 
the  California  coast.  Figure  10.10.1  is  a  photograph  from  the  air, 
taken  by  the  Bureau  of  Aeronautics  of  the  U.S.  Navy,  which  shows 
how  the  waves  coming  from  deep  water  are  modified  as  they  move 
toward  shore.  The  waves  are  so  smooth  some  distance  off  shore  that 
they  can  be  seen  only  vaguely  in  the  photograph,  but  as  they  move 
inshore  the  front  of  the  waves  steepens  noticeably  until,  finally, 
breaking  occurs.  Figures  10.10.2  and  10.10.3  are  pictures  of  the  same 
wave,  with  the  picture  of  Figure  10.10.3  taken  at  a  slightly  later  time 
than  the  previous  picture.  The  steepening  and  curling  over  of  the 
wave  are  very  strikingly  shown. 

At  this  point  it  is  useful  to  refer  back  to  the  beginning  of  section 
10.6  and  especially  to  Fig.  10.6.1.  This  figure,  which  is  repeated  here 

*  These  photographs  were  very  kindly  given  to  the  author  by  Dr.  Walter  Munk 
of  the  Sctipps  Institution  of  Oceanography. 


LONG    WAVES    IN    SHALLOW    WATER 


358 


Fig.  10.10.1.  Waves  on  a  beach 

for  the  sake  of  convenience,   indicates  in  terms  of  the  theory  of 
characteristics  what  happens  when  a  wave  of  elevation  is  created  by 


Fig.   10.10.2.  Wave  beginning  to  break 


354 


WATER    WAVES 


Fig.  10.10.3.  Wave  breaking 

pushing  the  moveable  end  of  a  tank  of  water  into  it  so  that  a  disturb- 
ance propagates  into  still  water  of  constant  depth:  the  straight 
characteristics  issuing  from  the  "piston  curve"  AD,  along  each  of 


Fig.  10.10.4.  Initial  point  of  breaking 


LONG    WAVES    IN    SHALLOW    WATER  355 


which  the  velocity  u  and  the  quantity  c  =  Vg(h  +  rj)  are  constant, 
eventually  intersect  at  the  point  E.  The  point  E  is  a  cusp  on  the  enve- 
lope of  the  characteristics,  and  represents  also  the  point  at  which  the 
slope  of  the  wave  surface  first  becomes  infinite.  The  point  E  might 
thus —  somewhat  arbitrarily,  it  is  true— -be  taken  as  defining  the  break- 
ing point  (xb9  tb)  of  the  wave,  since  one  expects  the  wave  to  start 
curling  over  after  this  point  is  reached.  It  is  possible  to  fix  the  values 
of  xb  and  tb  without  difficulty  once  the  surface  elevation  r\  =  77(0,  t)  is 
prescribed  at  x  =  0;  we  carry  out  the  calculation  for  the  interesting 
case  of  a  pulse  in  the  form  of  a  sine  wave: 

(10.10.1)  17(0,  t)  =  A  sin  cot. 

For  t  =  0,  x  >  0  we  assume  the  elevation  r\  of  the  water  to  be  zero 
and  its  velocity  UQ  to  be  constant  (though  not  necessarily  zero,  since 
it  is  of  interest  to  consider  the  effect  of  a  current  on  the  time  and  place 
of  breaking). 

As  we  know,  the  resulting  motion  is  easily  described  in  terms  of  the 
characteristics  in  the  x,  f-plane,  which  arc  straight  lines  emanating 
from  the  J-axis,  as  indicated  in  Figure  10.10.6.  The  values  of  u  and  c 
are  constant  along  each  such  straight  line.  The  slope  dx/dt  of  any 
straight  characteristic  through  the  point  (0,  r)  is  given  by 

dx 

(10.10.2)  —  =  9c  -  2r0  +  MO> 

at 

which  is  the  same  as  (10.3.6).  The  quantity  CQ  has  the  value  c0  =  Vgh, 
while  c  =  Vg(A  +17),  as  always.  On  the  other  hand,  the  slope  of  this 
characteristic  is  clearly  also  given  in  terms  of  a  point  (x,  t)  on  it  by 
xj(t  —  r)  so  that  (10.10.2)  can  be  written  in  the  form 

(10.10.3)  x  =  (t  —  r)[8c(r)  -  2<?0  +  MO] 

in  which  we  have  indicated  explicitly  that  c  depends  only  on  r  since 
it  (as  well  as  all  other  quantities)  is  constant  along  any  straight 
characteristic.  Thus  (10.10.3)  furnishes  the  solution  of  our  problem, 
once  c(r )  is  given,  throughout  a  region  of  the  x,  £-plane  which  is  cover- 
ed by  the  straight  lines  (10.10.3)  without  overlapping.  However,  the 
interesting  cases  for  us  are  just  those  in  which  overlapping  occurs, 
i.e.  those  for  which  the  characteristics  converge  and  eventually  cut 
each  other,  and  this  always  happens  if  an  elevation  is  created  at 
x  —  0.  In  fact,  if  c  is  an  increasing  function  of  r,  then  dx/dt  as  given 
by  (10.10.2)  increases  with  r  and  hence  the  characteristics  for  x  >  0 


856  WATER   WAVES 

must  intersect.  In  this  case,  furthermore,  the  family  of  straight  cha- 
racteristics has  an  envelope  beginning  at  a  point  (xb9  tb),  which  we 
have  defined  to  be  the  point  of  breaking. 

We  proceed  to  determine  the  envelope  of  the  straight  lines  (10.10.3). 
As  is  well  known,  the  envelope  can  be  obtained  as  the  locus  resulting 
from  (10.10.3)  and  the  relation 

(10.10.4)  0  =  -  [8c(r)  -  2c0  +  u0]  +  3(t  -  T)C'(T) 

obtained  from  it  by  differentiation  with  respect  to  T.  For  the  points 
(xC9  tc)  on  the  envelope  we  then  obtain  the  parametric  equations 

MAinKt  [3c(r)  -  2C0  +  utf 

(10.10.5)  *c  =  -  —  -  , 

and 

(10.10.6)  te  =  r 


We  are  interested  mainly  in  the  "first"  point  on  the  envelope,  that 
is,  the  point  (xb9  tb)  for  which  te  has  its  smallest  value  since  we  iden- 
tify this  point  as  the  point  of  breaking.  To  do  so  really  requires  a 
proof  that  the  water  surface  has  infinite  slope  at  this  point.  Such  a 
proof  could  be  easily  given,  but  we  omit  it  here  with  the  observation 
that  an  infinite  slope  is  to  be  expected  since  the  characteristics  which 
intersect  in  the  neighborhood  of  the  first  point  on  the  envelope  all 
carry  different  values  for  c. 

We  have  assumed  that  77(0,  t)  is  given  by  (10.10.1  )  and  consequently 
the  quantity  c(r)  in  (10.10.5)  and  (10.10.6)  is  given  by 


(10.10.7)  c(r)  =  Vg(h  +  A  sin  COT). 

If  we  assume  A  >  0  we  see  that  C'(T)  is  a  positive  decreasing  function 
of  T  for  small  positive  values  of  T.  Since  c(r)  increases  for  small  posi- 
tive values  of  r  it  follows  that  both  xc  and  tc  in  (10.10.5)  and  (10.10.6) 
are  increasing  functions  of  T  near  T  =  0.  A  minimum  value  of  xc  and 
te  must  therefore  occur  for  T  =  0,  so  that  the  breaking  point  is  given 

by 

noin«i  *        2<?o(c°  +  "o)2 

(10.10.8)  x,  = 

and 

••- 

as  one  can  readily  verify.  We  note  that  the  point  (xb,  tb)  lies  on  the 


LONG    WAVES    IN    SHALLOW    WATER  357 

initial  characteristic  x  =  (CQ  +  w0)f,  as  it  should  since  r  =  0  for  this 
characteristic.  From  the  formulas  we  can  draw  a  number  of  interesting 
conclusions.  Since  c0  =  Vgh  we  see  that  breaking  occurs  earlier  in 
shallower  water  for  a  pulse  of  given  amplitude  A  and  frequency  a). 
Breaking  also  occurs  earlier  when  the  amplitude  and  frequency  are 
larger.  It  follows  that  short  waves  will  break  sooner  than  long  waves, 
since  longer  waves  are  correlated  with  lower  frequencies.  Finally  we 
notice  that  early  breaking  of  a  wave  is  favored  by  small  values  for 
uQ,  the  initial  uniform  velocity  of  the  quiet  water.  In  fact,  if  u$  is 
negative,  i.e.  if  the  water  is  flowing  initially  toward  the  point  where 
the  pulse  originates,  the  breaking  can  be  made  to  occur  more  quickly. 
Everyone  has  observed  this  phenomenon  at  the  beach,  where  the  break- 
ing of  an  incoming  wave  is  often  observed  to  be  hastened  by  water 
rushing  down  the  beach  from  the  breaking  of  a  preceding  wave. 

It  is  of  some  importance  to  draw  another  conclusion  from  our  theory 
for  waves  moving  into  water  of  constant  depth:  an  inescapable  con- 
sequence of  our  theory  is  that  the  maxima  and  minima  of  the  surface 
elevation  propagate  into  quiet  water  unchanged  in  magnitude  with  re- 
spect to  both  distance  and  time.  This  follows  immediately  from  the  fact 
that  the  values  of  the  surface  elevation  are  constant  along  the  straight 
characteristics  so  that  if  7?  has  a  relative  maximum  for  x  =  0,  t  =  r, 
say,  then  this  value  of  rj  will  be  a  relative  maximum  all  along  the 
characteristic  which  issues  from  x  —  0,  t  =  r.  The  waves  change 
their  form  and  break,  but  they  do  so  without  changes  in  amplitude. 

In  a  report  of  the  Hydrographic  Office  by  Sverdrup  and  Munk 
[S.36]  some  results  of  observations  of  breakers  on  sloping  beaches  are 
given  in  the  form  of  graphs  showing  the  ratio  of  breaker  height  to 
deep  water  amplitude  and  the  ratio  of  undisturbed  depth  at  the  break- 
ing point  to  the  deep  water  amplitude  as  functions  of  the  "initial 
steepness"  in  deep  water,  the  latter  being  defined  as  the  ratio  of 
amplitude  to  wave  length  in  deep  water.  The  "initial  steepness"  is  thus 
essentially  the  quantity  Ao>  in  our  above  discussion,  and  our  results 
indicate  that  it  is  a  reasonable  parameter  to  choose  for  discussion  of 
breaking  phenomena.  The  graphs  given  in  the  report — reproduced  here 
in  Figures  10.10.5a  and  10.10.5b— show  very  considerable  scattering 
of  the  observational  data,  and  this  is  attributed  in  the  report  to  errors 
in  the  observations,  which  are  apparently  difficult  to  make  with 
accuracy.  On  the  basis  of  our  above  conclusion— that  the  breaking  of 
a  wave  in  water  of  uniform  depth  occurs  no  matter  what  the  amplitude 
of  the  wave  may  be  in  relation  to  the  undisturbed  depth— we  could 


358 


WATER  WAVES 


offer  another  explanation  for  the  scatter  of  the  points  in  Figures 
10.10.5a  and  10.10.5b,  i.e.  that  the  amplitude  ratios  are  relatively 
independent  of  the  initial  steepness.  Of  course,  the  curves  of  Figures 


•    W.H.0.1. 
A     B.E.B  ~  1:6.3 
A    B.E.B  —  1:20.4 
o    B.E.B  —  r.33.3 
©   THEORY 


0.003       0.005 


0.01  0.02       0.03 

INITIAL    STEEPNESS     H' 


0.05 


0.15 


Fig.  10.10.5a.    Ratio  of  breaker  height  to  wave  height  in  deep  water,  //»/ 

assuming  no  refraction 


•f.V 

d   /H1  «  6  66 

'      SLOPE  ' 

1 

•*"*"  b 

0 

0    S.I.O 

3.5 

• 

•    W.H.0.1. 

• 

A    B.E.B    —  1-63 

3-0 

A    B.E  B 

-  1:20.4 

1 

c 
o 

i    B  E  B 

-  1:33.3 

l! 

x°  2.5 

A 

•o 
2.0 

• 

V 

!V 

•o 

^ 

09 

n             0 

II 

o°  ! 

f 

o  6&v 

A   t6°     o 

1.5 
1.0 

- 

0 

I 

0 

1 

1    I 

:*:«*° 

^ 

a 

^ 

i    i 

\ 

__  XHEORE^TICA 

LIMIT 

»      c 

*> 

0.003       0.005               0.01               0.02         0.03          0.05                O.I       0.15 

INITIAL    STEEPNESS     HQ  /LQ 

Fig.  10.10.5b.  Ratio  of  depth  of  water  at  point  of  breaking  to  wave  height  in 
deep  water,  d6/H0',  as  function  of  steepness  in  deep  water,  H07L0,  assuming 
no  refraction 

LONG   WAVES   IN    SHALLOW   WATER 


859 


10.10.5a  and  10.10.5b  refer  to  sloping  beaches  and  hence  to  cases  in 
which  the  wave  amplitudes  increase  as  the  wave  moves  toward  shore; 
but  still  it  would  seem  rather  likely  that  the  amplitude  ratios  would  be 
relatively  independent  of  the  initial  steepness  in  these  cases  also  since 
the  beach  slopes  are  small.  The  detailed  investigation  of  breaking  of 
waves  by  Hamada  [H.2],  which  is  both  theoretical  and  experimental 
in  character,  should  be  consulted  for  still  further  analysis  of  this  and 
other  related  questions.  The  papers  by  Iversen  [1.6]  and  Suquet 
[S.31  ]  also  give  experimental  results  concerning  the  breaking  of  waves. 
We  continue  by  giving  the  results  of  numerical  computations  for 
three  cases  of  propagation  of  sine  pulses  into  still  water  of  constant 
depth.  The  cases  calculated  are  indicated  in  the  following  table: 


Case 


Type  of  pulse 


Case  1  is  a  half-sine  pulse  in  the  form  of  a  positive  elevation,  case  2 
is  a  full  sine  wave  which  starts  with  a  depression  phase,  and  case  3 
consists  of  several  full  sine  waves. 

Figure  10.10.6  shows  the  straight  characteristics  in  the  x,  f-plane 
for  case  1.  (In  all  of  these  cases,  the  quantities  x  and  y  are  now  certain 
dimensionless  quantities,  the  definitions  of  which  are  given  in  [S.19].) 
We  observe  that  the  envelope  begins  on  the  initial  characteristic  in 
this  case,  in  accord  with  earlier  developments  in  this  section.  The 
envelope  has  two  distinct  branches  which  meet  in  a  cusp  at  the 
breaking  point  (xb,  tb).  Figure  10.10.7  gives  the  shape  of  the  wave  for 
two  different  times.  As  we  see,  the  front  of  the  wave  steepens  until  it 
finally  becomes  vertical  for  x  =  tcb,  t  =-  tb,  while  the  back  of  the  wave 
flattens  out.  The  solution  given  by  the  characteristics  in  Figure 


360 


WATER   WAVES 


Region  of 
Constant   State 


Region  of  Constant  State 


Fig.  10.10.6.  Characteristic  diagram  in  the  x,  /-plane 


Fig.  10.10.7.    Wave  height  versus  distance  for  a  half-sine  wave  of  amplitude 
hQ/5  in  water  of  constant  depth  at  two  instants,  where  hn  is  the  height  of  the 

still  water  level 


LONG    WAVES    IN    SHALLOW    WATER 


361 


Fig.    10.1O.8.   Wave  profile  after  breaking 


Fig.   1O.1O.9.  Characteristic  diagram  in  the  a?,  £-plane 


362 


WATER   WAVES 


10.10.6  is  not  valid  for  x  >  x&  t  >  t^  and  we  expect  breaking  to 
ensue.  However,  we  observe  that  the  region  between  the  two  branches 
of  the  envelope  is  quite  narrow,  so  that  the  influence  of  the  developing 


Fig.  10.10.10.  Wave  height  versus  distance  for  a  full  negative  sine  wave  with 
amplitude  /*0/5  in  water  of  constant  depth  at  t  =  3.0,  t  =  5.0,  and  t  =  6.28 

breaker  may  not  seriously  affect  the  motion  of  the  water  behind  it. 
Thus  we  might  feel  justified  in  considering  the  solution  by  characteris- 
tics given  by  Figure  10.10.6  as  being  approximately  valid  for  values  of 


LONG   WAVES   IN   SHALLOW   WATDR 


863 


t  slightly  greater  than  tb.  (This  also  seems  to  the  writer  to  be  intuitive- 
ly rather  plausible  from  the  mechanical  point  of  view.)  Figure  10.10.8 
was  drawn  on  this  basis  for  a  time  considerably  greater  than  tb.  The 
full  portion  of  the  curve  was  obtained  from  the  characteristics  outside 
the  region  between  the  branches  of  the  envelope,  while  the  dotted 
portion— which  is  of  doubtful  validity— was  obtained  by  using  the 
characteristics  between  the  branches  of  the  envelope  in  an  obvious 
manner.  In  this  way  one  is  able  to  approximate  the  early  stages  of  the 
curling  over  of  a  wave. 

Figures  10.10.9,  10.10.10,  and  10.10.11  refer  to  case  2,  in  which  a 
depression  phase  precedes  a  positive  elevation.  In  this  case  the  enve- 
lope of  the  characteristics  does  not  begin,  of  course,  on  the  initial 
characteristic  but  rather  in  the  interior  of  the  simple  wave  region,  as 
indicated  in  Figure  10.10.9.  Figure  10.10.10  shows  three  stages  in  the 
progress  of  the  pulse  into  still  water.  The  steepening  of  the  wave  front 
is  very  marked  by  the  time  the  breaking  point  is  reached —much  more 
marked  than  in  the  preceding  case  for  which  no  depression  phase  oc- 
curs in  front.  Figure  10.10.11  shows  the  shape  of  the  wave  a  short  time 
after  passing  the  braking  point.  This  curve  was  obtained,  as  in  the 
preceding  case,  by  using  the  characteristics  between  the  branches  of 
the  envelope.  Although  this  can  yield  only  a  rough  approximation,  still 


Fig.  10.10.11.  r\  versus  x  at  t  =  7  for  non-sloping  bottom  where  the  pulse  is  an 

entire  negative  sine-wave.  The  dotted  part  of  the  curve  represents  r\  in  the  region 

between  the  branches  of  the  envelope 

one  is  rather  convinced  that  the  wave  really  would  break  very  soon 
after  the  point  we  have  somewhat  arbitrarily  defined  as  the  breaking 
point. 


364  WATER   WAVES 

Figure  10.10.12  shows  the  water  surface  in  case  3  for  a  time  well 
beyond  the  breaking  point. 


Fig.  10.10.12.  Water  profile  after  breaking 

In  gas  dynamics  where  u  and  c  represent  the  velocity  and  sound 
speed  throughout  an  entire  cross  section  of  a  tube  containing  the  gas, 
it  clearly  is  not  possible  to  give  a  physical  interpretation  to  the  region 
between  the  two  branches  of  the  envelope  in  the  cases  analogous  to 
that  shown  in  Figure  10.10.6,  since  the  velocity  and  propagation 
speed  must  of  necessity  be  single-valued  functions  of  x.  However,  in  our 
case  of  water  waves  u  and  c  refer  to  values  on  the  water  surface  so  that 
there  is  no  reason  a  priori  to  reject  solutions  for  u  and  c  which  are  not 
single- valued  in  x.  Thus  we  might  be  tempted  to  think  that  the  dotted 
part  of  the  curve  in  Figure  10.10.8  is  valid  within  the  general  frame- 
work of  our  theory,  but  this  is,  unfortunately,  not  the  case:  our  fun- 
damental differential  equations  are  not  valid  in  the  "overhanging" 
part  of  the  wave,  simply  because  that  part  is  not  resting  on  a  rigid 
bottom.  It  may  be  that  one  could  pursue  the  solutions  beyond  the 
point  where  the  breaking  begins  by  using  the  appropriate  differential 
equations  in  the  overhanging  part  of  the  wave  and  then  piecing  to- 
gether solutions  of  the  two  sets  of  differential  equations  so  that  con- 
tinuity is  preserved,  but  this  would  be  a  problem  of  considerable 
difficulty.  In  this  connection,  however,  it  is  of  interest  to  report  the 
results  of  a  calculation  by  Biesel  [B.10]  for  the  change  of  form  of 
progressing  waves  over  a  beach  of  small  slope.  Not  the  least  interesting 
aspect  of  Biesel's  results  is  the  fact  that  they  are  based  essentially  on 
the  theory  of  waves  of  small  amplitude,  i.e.  on  the  type  of  theory 
which  forms  the  basis  for  the  discussions  in  Part  II  of  this  book. 
However,  in  Part  II  only  the  so-called  Eulerian  representation  was 
used,  in  which  the  dependent  quantities  such  as  velocity,  pressure, 


LONG    WAVES    IN    SHALLOW    WATER  365 

etc.,  are  all  obtained  at  fixed  points  in  space.  As  a  result,  when  lineari- 
zations are  introduced  the  free  surface  elevation  77,  for  example,  is 
a  function  of  x  and  t  and  must  of  necessity  be  single-valued.  Biesel, 
however,  observes  that  one  can  also  use  the  Lagrangian  representa- 
tion* just  about  as  conveniently  as  the  Eulerian  when  a  development 
with  respect  to  amplitude  is  contemplated.  In  this  approach,  all  quan- 
tities are  fixed  in  terms  of  the  initial  positions  of  the  water  particles 
(and  the  time,  of  course).  In  particular,  the  displacements  (£,  77)  of 
the  water  particles  on  the  free  surface  would  be  given  as  functions  of 
a  parameter,  i.e.  £  =  f  (a,  t),  r\  =  r\(a,  t),  and  there  would  be  no  necessity 
a  priori  to  require  that  rj  should  be  a  single-valued  function  of  x. 
Biesel  has  carried  out  this  program  with  the  results  shown  in  Figs. 
10.10.13  to  10.10.16  inclusive.  A  sinusoidal  progressing  wave  in 
deep  water  is  assumed.  The  first  two  figures  refer  to  the  theory  when 
carried  out  only  to  first  order  terms  in  the  displacements  relative  to  the 
rest  position  of  equilibrium.  The  second  figure  is  a  detail  of  the  motion 
in  a  neighborhood  of  the  location  shown  by  the  dotted  circle  in  the 
first  figure.  Fig.  10.10.15  and  Fig.  10.10.16  treat  the  same  problem, 
but  the  solution  is  carried  to  second  order  terms.  In  both  cases  the 
development  of  a  breaker  is  strikingly  shown.  A  comparison  of  the 
results  of  the  first  order  and  second  order  theories  is  of  interest;  the 
main  conclusions  are:  if  second  order  corrections  are  made  the  break- 
ing is  seen  to  occur  earlier  (i.e.  in  deeper  water),  the  height  of  the  wave 
at  breaking  is  much  greater,  and  the  tendency  of  the  wave  to  plunge 
downward  after  curling  over  at  the  top  is  considerably  lessened. 
Actually,  our  shallow  water  theory  cannot  be  expected  to  yield  a 
good  approximation  near  the  breaking  point  since  the  curvature  of 
the  water  surface  is  likely  to  be  large  there.  However,  since  the  motion 
should  be  given  with  good  accuracy  at  points  outside  the  immediate 
vicinity  of  the  breaking  point  it  might  be  possible  to  refine  the  treat- 
ment of  the  breaking  problem  along  the  following  lines:  consider  the 
motion  of  a  fixed  portion  of  the  water  between  a  pair  of  planes  located 
some  distance  in  front  and  in  back  of  the  breaking  point.  The  motion 
of  the  water  particles  outside  the  bounding  planes  can  be  considered  as 
given  by  our  shallow  water  theory.  We  might  then  seek  to  determine 
the  motion  of  the  water  between  these  two  planes  by  making  use  of  a 
refinement  of  the  shallow  water  theory  or  by  reverting  to  the  original 
exact  formulation  of  the  problem  in  terms  of  a  potential  function  with 

*  In  Chapter  12.1  this  representation  is  explained  and  used  to  solve  other 
problems  involving  unsteady  motions. 


366 


WATER   WAVES 


Fig.  10.10.13.  Progression  and  breaking  of  a  wave  on  a  beach  of  1  in  10  slope. 

First-order  theory 


Fig.  10.10.14.  Details  of  breaking  of  wave  shown,  in  Fig.  10.10.18.  First-order 

theory 


LONG    WAVES    IN    SHALLOW    WATER 


367 


Fig.  10.10.15.  Progression  and  breaking  of  a  wave  on  a  beach  of  1  in  10  slope. 

Second-order  theory 


Fig.  10.10.16.  Details  of  breaking  of  wave  shown  in  Fig.  10.10.15.  Second-order 

theory 


368  WATER   WAVES 

the  nonlinear  free  surface  condition  and  determine  it  by  using  finite 
difference  methods  in  a  bounded  region. 

It  is  of  interest  now  to  return  to  the  problem  with  which  we  opened 
the  discussion  of  the  present  section,  i.e.  to  the  problem  of  a  tank  with 
a  moveable  end  which  is  pushed  into  the  water.  As  we  have  seen,  the 
wave  which  arises  will  eventually  break.  Suppose  now  we  assume 
that  the  end  of  the  tank  continues  to  move  into  the  water  with  a  uni- 
form velocity.  The  end  result  after  the  initial  curling  over  and  break- 
ing will  be  the  creation  of  a  steady  progressing  wave  front  which  is 
steep  and  turbulent,  behind  which  the  water  level  is  constant  and  the 


Fig.  10.10.17.  The  bore  in  the  Tsicn  Tang  River 

water  has  everywhere  the  constant  velocity  imparted  to  it  by  the  end 
of  the  tank.  Such  a  steady  progressing  wave  with  a  steep  front  is 
called  a  bore.  It  is  the  exact  analogue  of  a  steady  progressing  shock 
wave  in  a  gas.  In  Figure  10.10.17  we  show  a  photograph,  taken  from 
the  book  by  Thorade  [T.4],  of  the  bore  which  occurs  in  the  Tsien- 
Tang  River  as  a  result  of  the  rising  tide,  which  pushes  the  water  into 
a  narrowing  estuary  at  the  mouth  of  the  river.  The  height  of  this  bore 
apparently  is  as  much  as  20  to  30  feet.  According  to  the  theory  pre- 
sented above,  this  bore  should  have  been  preceded  by  an  unsteady 
phase  during  which  the  smooth  tidal  wave  entering  the  estuary  first 
curled  over  and  broke.  Methods  for  the  treatment  of  problems  in- 
volving the  gradual  development  of  a  bore  in  an  otherwise  smooth 
flow  have  been  worked  out  by  A.  Lax  [L.5]  (see  also  Whitham  [W.I 2] ). 
We  have,  so  far,  used  our  basic  theory— the  nonlinear  shallow 


LONG   WAVES    IN   SHALLOW   WATER  369 

water  theory —to  interpret  the  solutions  of  only  one  type  of  problem, 
i.e.  the  problem  of  the  change  of  form  of  a  pulse  moving  into  still 
water  of  constant  depth.  The  theory,  however,  can  be  used  to  study 
the  propagation  of  a  wave  over  a  beach  with  decreasing  depth  just 
as  well  (cf.  the  author's  paper  [S.19]),  but  the  calculations  are  made 
much  more  difficult  because  of  the  fact  that  no  family  of  straight 
characteristics  exists  unless  the  depth  is  constant.  This  problem,  in 
fact,  brings  to  the  fore  the  difficulties  of  a  computational  nature 
which  occur  in  important  problems  involving  the  propagation  of  flood 
waves  and  other  surges  in  rivers  and  open  channels  in  general.  Such 
problems  will  be  discussed  in  the  next  chapter. 

On  an  actual  beach  on  which  waves  are  breaking,  the  motion  of 
the  water,  of  course,  does  not  consist  in  the  propagation  of  a  single 
pulse  into  still  water,  but  rather  in  the  occurrence  of  an  approximately 
periodic  train  of  waves.  However,  experiments  indicate  that  only  a 
slight  reflection  of  the  wave  motion  from  the  shore  occurs.  The  in- 
coming wave  energy  seems  to  be  destroyed  in  turbulence  due  to  break- 
ing or  to  be  converted  into  the  energy  of  flow  of  the  undertow.  In 
other  words,  each  wave  propagates,  to  a  considerable  degree,  in  a 
manner  unaffected  by  the  waves  which  preceded  it.  Consequently  the 
above  treatment  of  breaking,  in  which  propagation  of  a  wave  into  still 
water  was  assumed,  should  be  at  least  qualitatively  reasonable.  An- 
other objection  to  our  theory  has  already  been  mentioned,  i.e.  that 
large  curvatures  of  the  water  surface  near  the  breaking  point  seem  sure 
to  make  the  results  inaccurate.  Nevertheless,  the  theory  should  be 
valid,  except  near  this  point,  in  many  cases  of  waves  on  sloping  bea- 
ches, since  the  wave  lengths  arc  usually  at  least  10  to  20  or  more  times 
the  depth  of  the  water  in  the  breaker  zone,  hence  the  theory  presented 
above  should  certainly  yield  correct  qualitative  results  and  perhaps 
also  reasonably  accurate  quantitative  results. 

Waves  do  not  by  any  means  always  break  in  the  manner  described 
up  to  this  point.  In  Fig.  10.10.1 8a,  b  we  show  photographs  (given  to 
the  author  by  Dr.  Walter  Munk )  of  waves  breaking  in  a  fashion  con- 
siderably at  variance  with  the  results  of  the  theory  presented  here.  We 
observe  that  the  waves  break,  in  this  instance,  by  curling  over  slightly  at 
the  crest,  but  that  the  wave  remains,  as  a  whole,  symmetrical  in  shape, 
while  the  theory  presented  here  yields  a  marked  steepening  of  the  wave 
front  and  a  very  unsymmetrical  shape  for  the  wave  at  breaking. 

Observation  of  cases  like  that  shown  in  Figure  10.10.18  doubtlessly 
led  to  the  formulation  of  the  theory  of  breaking  due  to  Sverdrup  and 


370 


WATER    WAVES 


(a) 


(b) 


Fig.   l().l().18a,  b.  Waves  breaking  at  crests 

Munk  [S.33];  their  theory  is  based  on  results  taken  from  the  study  of 
the  solitary  wave,  which  has  been  discussed  in  the  preceding  section.* 
The  solitary  wave  is,  by  definition,  a  wave  of  finite  amplitude  con- 

*  An  interesting  mathematical  treatment  of  breaking  phenomena  from  this 
point  of  view  was  given  some  time  ago  by  Kculegan  and  Patterson     K.lllj. 


LONG   WAVES   IN   SHALLOW   WATER  871 

sisting  of  a  single  elevation  of  such  a  shape  that  it  can  propagate  un- 
changed in  form.  At  first  sight,  this  would  seem  to  be  a  rather  curious 
wave  form  to  take  as  a  basis  for  a  discussion  of  the  phenomena  of 
breaking,  since  it  is  precisely  the  change  in  form  resulting  in  breaking 
that  is  in  question.  On  the  other  hand,  the  waves  often  look  as  in 
Figure  10.10.18  and  do  retain,  on  the  whole,  a  symmetrical  shape,* 
with  some  breaking  at  the  crest.  Actually,  the  situation  regarding  the 
two  different  theories  of  breaking  from  the  mathematical  point  of 
view  is  the  following,  as  we  can  infer  from  the  discussion  of  the  pre- 
ceding section:  Both  theories  are  shallow  water  theories.  In  fact,  as 
Keller  [K.6],  and  Friedrichs  and  Hyers  [F.13],  have  shown,  the  theory 
of  the  solitary  wave  can  be  obtained  from  the  approximation  of  next  higher 
order  above  that  used  in  the  present  section,  if  the  assumption  is  made  that 
the  motion  is  a  steady  motion.  In  other  words,  the  theory  used  by 
Sverdrup  and  Munk  is  a  shallow  water  theory  of  higher  order  than 
the  theory  used  in  this  section,  which  furnishes  in  principle  the  con- 
stant state  as  the  only  continuous  wave  which  can  propagate  un- 
altered in  form.  On  the  other  hand,  the  theory  presented  here  makes 


Fig.  10.10.10.  Symmetrical  waves  breaking  at  crests 

it  possible  to  deal  directly  with  the  unsteady  motions,  while  Sverdrup 
and  Munk  are  forced  to  approximate  these  motions  by  a  series  of 
different  steady  motions.  One  could  perhaps  sum  up  the  whole  matter 
by  saying  that  waves  break  in  different  ways  depending  upon  the 
individual  circumstances  (in  particular,  the  depth  of  the  water  com- 
pared with  the  wave  length  is  very  important),  and  the  theory  which 
should  be  used  to  describe  the  phenomena  should  be  chosen  accord- 
ingly. In  fact,  Figures  10.10.17  and  10.10.18  depicting  a  bore  and 

*  Sverdrup  and  Munk,  like  the  author,  assume  that,  when  considering  breaking 
phenomena,  each  wave  in  a  train  can  be  treated  with  reasonable  accuracy  as 
though  it  were  uninfluenced  by  the  presence  of  the  others. 


872 


WATER   WAVES 


waves  breaking  only  at  the  crests  of  otherwise  symmetrical  waves 
perhaps  represent  extremes  in  a  whole  series  of  cases  which  include 
the  breaker  shown  in  Figures  10.10.2  and  10.10.3  as  an  intermediate 
case.  Some  pertinent  observations  on  this  point  have  been  made  by 
Mason  [M.4].  A  theory  has  been  developed  by  Ursell  [U.ll]  which 
differs  from  the  theories  discussed  here  and  which  may  well  be  appro- 
priate in  cases  not  amenable  to  treatment  by  the  shallow  water  theory. 
The  paper  by  Hamada  [H.2]  referred  to  above  should  also  be  men- 
tioned again  in  this  connection.  In  particular,  Fig.  10.10.19,  taken 
from  that  paper,  shows  waves  created  in  a  tank  which  break  by  curling 
at  the  crest  but  still  preserving  a  symmetrical  form.  It  is  interesting 
to  observe  that  the  wave  length  in  this  case  is  almost  the  same  as  the 
depth  of  the  water.  It  is  also  interesting  to  add  that  in  this  case  a 
current  of  air  was  blown  over  the  water  in  the  direction  of  travel  of 
the  waves.  Fig.  10.10.20  shows  a  similar  case,  but  with  somewhat 
greater  wave  length.  The  two  waves  were  both  generated  by  a  wave 


Fig.  10.10.20.  Breaking  induced  by  wind  action 

making  apparatus  at  the  right;  the  only  difference  is  that  a  current 
of  air  was  blown  from  right  to  left  in  the  case  shown  by  the  lower 
photograph.  The  breaking  thus  seems  due  entirely  to  wind  action  in 


LONG    WAVES    IN   SHALLOW   WATER 


878 


this  ease.  Finally,  Fig.  10.10.21  shows  two  stages  in  the  breaking  of  a 
wave  in  shallow  water,  when  marked  dissymmetry  and  the  formation 
of  what  looks  like  a  jet  at  the  summit  of  the  wave  are  seen  to  occur. 
It  is  of  interest,  historically  and  otherwise,  to  refer  once  more  to  the 
case  of  symmetrical  waves  breaking  at  their  crests.  The  wave  crests 
in  such  cases  are  quite  sharp,  as  can  be  seen  in  the  photograph  shown 
in  Fig.  10.10.18.  Stokes  [S.28]  long  ago  gave  an  argument,  based  on 
quite  reasonable  assumptions,  that  steady  progressing  waves  with  an 
angular  crest  of  angle  120°  could  occur;  in  fact,  this  follows  almost  at 
once  from  the  Bernoulli  law  at  the  free  surface  when  the  free  surface  is 
assumed  to  be  a  stream  line  with  an  angular  point.  There  is  another 
fact  pertinent  to  the  present  discussion,  i.e.  that  the  exact  theory  for 
steady  periodic  progressing  waves  of  finite  amplitude,  as  developed 
in  Chapter  12.2,  shows  that  with  increasing  amplitude  the  waves 
flatten  more  and  more  in  the  troughs,  but  sharpen  at  the  crests. 


Fig.  10.10.21.  Breaking  of  a  long  wave  in  shallow  water 

In  fact,  the  terms  of  lowest  order  in  the  development  of  the  free  surface 
amplitude  77  as  given  by  that  theory  can  easily  be  found;  the  result  is 

rf(x)  =  —  a  cos  x  +  a2  cos  2x 
for  a  wave  of  wave  length  2n.  Fig.  10.10.22  shows  the  result  of  super- 


874 


WATER   WAVES 


imposing  the  second-order  term  a2  cos  2x  on  the  wave  —  a  cos  x 
which  would  be  given  by  the  linear  theory;  as  one  sees,  the  effect  is  as 
indicated.  It  would  be  a  most  interesting  achievement  to  show  rigor- 
ously that  the  wave  form  with  a  sharp  crest  of  angle  120°  is  attained 
with  increase  in  amplitude.  An  interesting  approximate  treatment  of 
the  problem  has  been  given  by  Davies  [D.5],  However,  the  problem 
thus  posed  is  not  likely  to  be  easy  to  solve;  certainly  the  method  of 
Levi-Civita  as  developed  in  Ch.  12.2  does  not  yield  such  a  wave  form 
since  it  is  shown  there  that  the  free  surface  is  analytic.  Presumably, 


Fig.  10.10.22.  Sharpening  of  waves  at  the  crest 

any  further  increase  in  amplitude  would  lead  to  breaking  at  the  crests 
—hence  no  solutions  of  the  exact  problem  would  exist  for  amplitudes 
greater  than  a  certain  value. 

10.11.    Gravity  waves  in  the  atmosphere.  Simplified  version  of  the 
problem  of  the  motion  of  cold  and  warm  fronts 

In  practically  all  of  this  book  we  assume  that  the  medium  in  which 
waves  propagate  is  water.  It  is,  however,  a  notable  fact  that  some 
motions  of  the  atmosphere,  such  as  tidal  oscillations  due  to  the  same 
cause  as  the  tides  in  the  oceans,  i.e.  gravitational  effects  of  the  sun 
and  moon,  as  well  as  certain  large  scale  disturbances  in  the  atmosphere 
such  as  wave  disturbances  in  the  prevailing  westerlies  of  the  middle 
latitudes,  and  motions  associated  with  disturbances  on  certain  dis- 
continuity surfaces  called  fronts,  are  all  phenomena  in  which  the  air 
can  be  treated  as  a  gravitating  incompressible  fluid.  In  addition,  one 
of  the  best-founded  laws  in  dynamic  meteorology  is  the  hydrostatic 
pressure  law,  which  states  that  the  pressure  at  any  point  in  the  at- 
mosphere is  very  accurately  given  by  the  static  weight  of  the  column 
pf  air  above  it.  When  we  add  that  the  types  of  motions  enumerated 
above  are  all  such  that  a  typical  wave  length  is  large  compared  with 


LONG   WAVES    IN    SHALLOW   WATER  375 

an  average  thickness  (on  the  basis  of  an  average  density,  that  is)  of 
the  layer  of  air  over  the  earth,  it  becomes  clear  that  these  problems  fall 
into  the  general  class  of  problems  treated  in  the  present  chapter.  Of 
course,  this  means  that  thermodynamic  effects  are  ignored,  and 
with  them  the  ingredients  which  go  to  make  up  the  local  weather, 
but  it  seems  that  these  effects  can  be  regarded  with  a  fair  approxima- 
tion as  small  perturbations  on  the  large  scale  motions  in  question. 

There  are  many  interesting  problems,  including  very  interesting  un- 
solved problems,  in  the  theory  of  tidal  oscillations  in  the  atmosphere. 
These  problems  have  been  treated  at  length  in  the  book  by  Wilkes  [W. 2] ; 
we  shall  not  attempt  to  discuss  them  here.  The  problems  involved  in 
studying  wave  propagation  in  the  prevailing  westerlies  will  also  not 
be  discussed  here,  though  this  interesting  theory,  for  which  papers  by 
Charney  [C.15]  and  Thompson  [T.10]  should  be  consulted,  is  being 
used  as  a  basis  for  forecasting  the  pressure  in  the  atmosphere  by  nu- 
merical means.  In  other  words,  the  dynamical  theory  is  being  used  for 
the  first  time  in  meteorology,  in  conjunction  with  modern  high  speed 
digital  computing  equipment,  to  predict  at  least  one  of  the  elements 
which  enter  into  the  making  of  weather  forecasts. 

In  this  section  we  discuss  only  one  class  of  meteorological  problems, 
i.e.  motions  associated  with  frontal  discontinuities,  or,  rather,  it 
would  be  better  to  say  that  we  discuss  certain  problems  in  fluid 
dynamics  which  are  in  some  sense  at  least  rough  approximations  to  the 
actual  situations  and  from  which  one  might  hope  to  learn  something 
about  the  dynamics  of  frontal  motions.  The  problems  to  be  treated 
here —unlike  the  problems  of  the  type  treated  by  Charney  and 
Thompson  referred  to  above  — are  such  as  to  fit  well  with  the  preceding 
material  in  this  chapter;  it  was  therefore  thought  worthwhile  to  in- 
clude them  in  this  book  in  spite  of  their  somewhat  speculative  charac- 
ter from  the  point  of  view  of  meteorology.  Actually,  the  idea  of  using 
methods  of  the  kind  described  in  this  chapter  for  treating  certain  special 
types  of  motions  in  the  atmosphere  has  been  explored  by  a  number  of 
meteorologists  (cf.  Abdullah  [A.7],  Freeman  [F.10],  Tepper  [T.ll]). 

One  of  the  most  characteristic  features  of  the  motion  of  the  atmos- 
phere in  middle  latitudes  and  also  one  which  is  of  basic  importance 
in  determining  the  weather  there  is  the  motion  of  wave-like  disturb- 
ances which  propagate  on  a  discontinuity  surface  between  a  thin 
wedge-shaped  layer  of  cold  air  on  the  ground  and  an  overlying  layer 
of  warmer  air.  In  addition  to  a  temperature  discontinuity  there  is  also 
in  general  a  discontinuity  in  the  tangential  component  of  the  wind 


376  WATER   WAVES 

velocity  in  the  two  layers.  The  study  of  such  phenomena  was  initiated 
long  ago  by  Bjerknes  and  Solberg  [B.20]  and  has  been  continued 
since  by  many  others.  In  considering  wave  motions  on  discontinuity 
surfaces  it  was  natural  to  begin  by  considering  motions  which  depart 
so  little  from  some  constant  steady  motion  (in  which  the  discontinuity 
surface  remains  fixed  in  space)  that  linearizations  can  be  performed, 
thus  bringing  the  problems  into  the  realm  of  the  classical  linear 
mathematical  physics.  Such  studies  have  led  to  valuable  insights, 
particularly  with  respect  to  the  question  of  stability  of  wave  motions 
in  relation  to  the  wave  length  of  the  perturbations.  (The  problems 
being  linear,  the  motions  in  general  can  be  built  up  as  a  combination, 
roughly  speaking,  of  simple  sine  and  cosine  waves  and  it  is  the 
wave  length  of  such  components  that  is  meant  here,  cf.  Haurwitz 
[H.5,  p.  234].)  One  conjecture  is  that  the  cyclones  of  the  middle  lati- 
tudes are  initiated  because  of  the  occurrence  of  such  unstable  waves 
on  a  discontinuity  surface. 

A  glance  at  a  weather  map,  or,  still  better,  an  examination  of  weath- 
er maps  over  a  period  of  a  few  days,  shows  clearly  that  the  wave 
motions  on  the  discontinuity  surfaces  (which  manifest  themselves  as 
the  so-called  fronts  on  the  ground )  develop  amplitudes  so  rapidly  and 
of  such  a  magnitude  that  a  description  of  the  wave  motions  over  a 
period  of,  say,  a  day  or  two,  by  a  linearization  seems  not  feasible  with 
any  accuracy.  The  object  of  the  present  discussion  is  to  make  a  first 
step  in  the  direction  of  a  nonlinear  theory,  based  on  the  exact  hydro- 
dynamical  equations,  for  the  description  of  these  motions,  that  can  be 
attacked  by  numerical  or  other  methods.  No  claim  is  made  that  the 
problem  is  solved  here  in  any  general  sense.  What  is  done  is  to  start 
with  the  general  hydrodynamical  equations  and  make  a  series  of 
assumptions  regarding  the  flow;  in  this  way  a  sequence  of  three  non- 
linear problems  (we  call  them  Problems  I,  II,  III),  each  one  furnishing 
a  consistent  and  complete  mathematical  problem,  is  formulated. 
One  can  see  then  the  effect  of  each  additional  assumption  in 
simplifying  the  mathematical  problem.  The  first  two  problems  result 
from  a  series  of  assumptions  which  would  probably  be  generally 
accepted  by  meteorologists  as  reasonable,  but  unfortunately  even 
Problem  II  is  still  pretty  much  unmanageable  from  the  point  of  view 
of  numerical  analysis.  Further,  and  more  drastic,  assumptions  lead 
to  a  still  simpler  Problem  III  which  is  formulated  in  terms  of  three 
first  order  partial  differential  equations  in  three  dependent  and  three 
independent  variables  (as  contrasted  with  eight  differential  equations 


LONG   WAVES   IN   SHALLOW   WATER  377 

in  four  independent  variables  in  Problem  I).  The  three  differential 
equations  of  Problem  III  are  probably  capable  of  yielding  reasonably 
accurate  approximations  to  the  frontal  motions  under  consideration, 
but  they  are  still  rather  difficult  to  deal  with,  even  numerically, 
principally  because  they  involve  three  independent  variables*:  such 
equations  are  well  known  to  be  beyond  the  scope  of  even  the  most 
modern  digital  computing  machines  as  a  rule.  Consequently,  still 
further  simplifying  assumptions  are  made  in  order  to  obtain  a  theory 
capable  of  yielding  some  concrete  results  through  calculation. 

At  this  point,  two  different  approaches  to  the  problem  are  proposed. 
One  of  them,  by  Whitham  [W.12],  deals  rather  directly  with  the 
three  differential  equations  of  Problem  III.  Two  of  these  equations 
are  essentially  the  same  as  those  of  the  nonlinear  shallow  water  theory 
treated  in  the  preceding  sections  of  this  chapter.  These  two  equations 
—which  refer  to  motions  in  vertical  planes— -can  therefore  be  inte- 
grated. Afterwards  the  transverse  component  of  the  velocity  is  found 
by  integrating  a  linear  first  order  partial  differential  equation.  In  this 
way  a  quite  reasonable  qualitative  description  of  the  dynamics  of 
frontal  motions  can  be  achieved,  at  least  in  special  cases,  which  is 
in  good  agreement  with  many  of  the  observed  phenomena.  However, 
this  theory  has  a  disadvantage  in  that  it  docs  not  permit  a  complete 
numerieal  integration  because  of  a  peculiar  difficulty  at  cold  fronts. 
(The  difficulty  stems  from  the  fact  that  a  cold  front  corresponds  in 
this  theory  to  what  amounts  to  the  propagation  of  a  bore  down  the 
dry  bed  of  a  stream— a  mathematical  impossibility.  If  one  had  a 
means  of  taking  care  of  turbulence  and  friction  at  the  ground,  it  would 
perhaps  be  possible  to  overcome  this  difficulty.)  Nevertheless,  the 
qualitative  agreement  with  the  observed  phenomena  is  an  indication 
that  the  three  differential  equations  furnishing  the  basic  approximate 
theory  from  which  we  start— i.e.  those  of  our  Problem  III— have  in 
them  the  possibility  of  furnishing  reasonable  results. 

The  author's  method  (ef.  [S.24])  of  treating  the  three  basic  differ- 
ential equations  is  quite  different  from  that  of  Whitham,  but  it  un- 
fortunately involves  a  further  assumption  which  has  the  effect  of 
limiting  the  applicability  of  the  theory.  The  guiding  principle  was  that 

*  The  work  of  Freeman  [F.9,  10]  is  based  on  a  theory  which  could  be  con- 
sidered as  a  special  case  of  Problem  III  in  which  the  Coriolis  terms  due  to  the 
rotation  of  the  earth  are  neglected  and  the  motion  is  assumed  at  the  outset  to 
depend  on  only  one  space  variable  and  the  time.  The  idea  of  deriving  the  theory 
resulting  in  Problem  III  occurred  to  the  author  while  reading  Freeman's  paper 
and,  indeed,  Freeman  indicates  the  desirability  of  generalizing  his  theory. 


378 


WATER   WAVES 


differential  equations  in  only  two  independent  variables  should  be 
found,  but  that  the  number  of  dependent  variables  need  not  be  so 
ruthlessly  limited.  Finally,  it  is  highly  desirable  to  obtain  differential 
equations  of  hyperbolic  type  in  order  that  the  theory  embodied  in  the 
method  of  characteristics  becomes  available  in  formulating  and  solv- 
ing concrete  problems.  These  objectives  can  be  attained  by  making 
quite  a  few  further  simplifying  assumptions  with  respect  to  the  me- 
chanics of  the  situation.  The  result  is  what  might  be  called  Problem 
IV.  The  theory  formulated  in  Problem  IV  is  embodied  in  a  system  of 
four  nonlinear  first  order  partial  differential  equations  of  hyperbolic 
type  in  four  dependent  and  two  independent  variables.  A  numerical 
integration  of  these  equations  is  possible,  but  the  labor  of  integrating 
the  equations  is  so  great  that  only  meagre  results  are  so  far  available. 
Once  Whitham's  theory  and  Problem  IV  have  been  formulated, 
one  is  led  once  more  to  consider  dealing  with  Problem  III  numerically 
in  spite  of  the  fact  that  there  are  three  independent  variables  in  this 
case;  in  Problem  IV,  and  also  in  the  theory  by  Whitham,  for  that 
matter,  the  basic  idea  is  that  variations  in  the  ^/-direction  are  less 
rapid  than  those  in  the  ^-direction,  and  thus  a  finite  difference  scheme 
in  two  space  variables  and  the  time  might  be  possible  under  such 
special  circumstances. 


Worm 


Worm 


Ground 


Fig.  10.11.1.  A  stationary  front 


Cold 
TT7? 


We  proceed  to  the  derivation  of  the  basic  approximate  theory.  To 
begin  with,  a  certain  steady  motion  (called  a  stationary  front)  is  taken 
as  an  initial  state,  and  this  consists  of  a  uniform  flow  of  two  super- 
imposed layers  of  cold  and  warm  air,  as  indicated  in  Figure  10.11.1. 
The  s-axis  is  taken  positive  upward*  and  the  x,  t/-plane  is  a  tangent 

*  Here  we  deviate  from  our  standard  practice  of  taking  the  t/-axis  as  the 
vertical  axis,  in  order  to  conform  to  the  usual  practice  in  dynamic  meteorology. 
This  should  cause  no  confusion,  since  this  section  can  be  read  to  a  large  extent 
independently  of  the  rest  of  the  book. 


LONG    WAVES    IN    SHALLOW    WATER  379 

plane  to  the  earth.  The  rotation  of  the  earth  is  to  be  taken  into 
account  but,  for  the  sake  of  simplicity,  not  its  sphericity —a  common 
practice  in  dynamic  meteorology.  The  coordinate  system  is  assumed 
to  be  rotating  about  the  js-axis  with  a  constant  angular  velocity 
Q  =  co  sin  (p,  with  co  the  angular  velocity  of  the  earth  and  <p  the  lati- 
tude of  the  origin  of  our  coordinate  system.  (The  motivation  for  this  is 
that  the  main  effects  one  cares  about  are  found  if  the  Coriolis  terms 
are  included,  and  that  neglect  of  the  curvature  of  the  earth  has  no 
serious  qualitative  effect.)  As  indicated  in  Figure  10.11.1,  the  cold  air 
lies  in  a  wedge  under  the  warm  air  and  the  discontinuity  surface 
between  the  two  layers  is  inclined  at  angle  a  to  the  horizontal.  The 
term  "front"  is  always  applied  to  the  intersection  of  the  discontinuity 
surface  with  the  ground,  and  in  the  present  case  we  have  therefore  as 
initial  state  a  stationary  front  running  along  the  iT-axis.  The  wind 
velocity  in  the  two  layers  is  parallel  to  the  #-axis  (otherwise  the  dis- 
continuity surface  could  not  be  stationary),  but  it  will  in  general  be 
different  in  magnitude  and  perhaps  even  opposite  in  direction  in  the 
two  layers.  The  situation  shown  in  Figure  10.11.1  is  not  uncommon. 
For  instance,  the  j?-axis  might  be  in  the  eastward  direction,  the  t/-axis 
in  the  northward  direction  and  the  warm  air  would  be  moving  in  the 
direction  of  the  prevailing  westerlies.  The  origin  of  the  cold  air  at  the 
ground  is,  of  course,  the  eold  polar  regions.  We  shall  see  later  that 
sueh  configurations  are  dynamically  eorreet  and  that  the  angle  a 
is  uniquely  determined  (and  quite  small,  of  the  order  of  £°)  once 
the  state  of  the  warm  air  and  eold  air  is  given.  (The  discontinuity 
surface  is  not  horizontal  because  of  the  Coriolis  force  arising  from 
the  rotation  of  the  earth.) 

We  proceed  next  to  describe  what  is  observed  to  happen  in  many 
eases  once  sueh  a  stationary  front  starts  moving.  In  Figure  10.11.2  a 
sequence  of  diagrammatic  sketches  is  given  which  indicate  in  a  general 
way  what  can  happen.  All  of  the  sketches  show  the  intersection  of  the 
moving  discontinuity  surface  (cf.  Figure  10.11.1)  with  the  ground  (the 
x,  jy-plane  with  the  j/-axis  taken  northward,  theo?-axis  taken  eastward). 
The  shaded  area  indicates  the  region  on  the  ground  covered  by  cold 
air,  while  the  unshaded  region  is  covered  at  the  ground  by  warm  air. 
Of  course,  the  cold  air  always  lies  in  a  thin  wedge  under  a  thick  layer 
of  warm  air.  In  Figure  10.11. 2a  the  development  of  a  bulge  in  the 
stationary  front  toward  the  north  is  indicated.*  Such  a  bulge  then 

*  What  agency  serves  to  initiate  and  to  maintain  such  motions  appears  to 
be  a  mystery.  Naturally  such  an  important  matter  has  been  the  subject  of  a  great 

(footnote  continued) 


880 


WATER   WAVES 


frequently  deepens  and  at  the  same  time  propagates  eastward  with  a 
velocity  of  the  order  of  500  miles  per  day.  It  now  becomes  possible  to 
define  the  terms  cold  front  and  warm  front.  As  indicated  in  Figure  10. 
11.26,  the  cold  front  is  that  part  of  the  whole  front  at  which  cold  air  is 
taking  the  place  of  warm  air  at  the  ground,  and  the  warm  front  is  the 


Cold    / 
Front 


Warm 
Front 


(o) 


(b) 


Occluded 
Front 


(c)  (d) 

Fig.  10.11.2.  Stages  in  the  motion  of  a  frontal  disturbance 


portion  of  the  whole  front  where  cold  air  is  retreating  with  warm  air 
taking  its  place  at  the  ground.  Since  such  cold  and  warm  fronts  are 
accompanied  by  winds,  and  by  precipitation  in  various  forms  — in 
fact,  by  all  of  the  ingredients  that  go  to  make  up  what  one  calls 

deal  of  discussion  and  speculation,  but  there  seems  to  be  no  consistent  view  about 
it  among  meteorologists.  In  applying  the  theory  derived  here  no  attempt  is  made 
to  settle  this  question  a  priori:  we  would  simply  take  our  dynamical  model, 
assume  an  initial  condition  which  in  effect  states  that  a  bulge  of  the  kind  just 
described  is  initiated,  and  then  study  the  subsequent  motion  by  integrating  the 
differential  equations  subject  to  appropriate  initial  and  boundary  conditions. 
However,  if  the  approximate  theory  is  really  valid,  such  studies  might  perhaps 
be  used,  or  could  be  modified,  in  such  a  way  as  to  throw  some  light  on  this  im- 
portant and  vexing  question. 


LONG   WAVES    IN   SHALLOW   WATER  381 

weather— it  follows  that  the  weather  at  a  given  locality  in  the  middle 
latitudes  is  largely  conditioned  by  the  passage  of  such  frontal  dis- 
turbances. Cold  fronts  and  warm  fronts  behave  differently  in  many 
ways.  For  example,  the  cold  front  in  general  moves  faster  than  the 
warm  front  and  steepens  relative  to  it,  so  that  an  originally  symme- 
trical disturbance  or  wave  gradually  becomes  distorted  in  the  manner 
indicated  in  Figure  10.11.2c.  This  process  sometimes— though  by  no 
means  always— continues  until  the  greater  portion  of  the  cold  front 
has  overrun  the  warm  front;  an  occluded  front,  as  indicated  in  Figure 
10.11.2d,  is  then  said  to  occur.  The  prime  object  of  what  follows  is  to 
derive  a  theory— or  perhaps  better,  to  invent  a  simplified  dynamical 
model— capable  of  dealing  with  fluid  motions  of  this  type  that  is  not 
on  the  one  hand  so  crude  as  to  fail  to  yield  at  least  roughly  the  observed 
motions,  and  on  the  other  hand  is  not  impossibly  difficult  to  use 
for  the  purpose  of  mathematical  discussion  and  numerical  calculation. 
Since  it  is  desired  that  this  section  should  be  as  much  as  possible 
self  contained,  we  do  not  lean  on  the  basic  theory  developed  earlier 
in  this  book.  Thus  we  begin  with  the  classical  hydrodynamical  equa- 
tions. The  equations  of  motion  in  the  Eulerian  form  arc  taken: 

du  dv 


dv 
(10.11.1) 

dw  dp 

}  "77  ==  ~  a~  +  6FM  ~  OS 
at  OZ 

with  d/dt  (the1  particle  derivative)  defineel  by  the  operator  d/dt  + 
u  dfdx  +  v  d/dy  +  w  d/dz.  In  these  equations  u9  v9  re  are  the  velocity 
components  relative  to  our  rotating  coordinate  system,  p  is  the  pres- 
sure, Q  the  density,  £>/<%)  etc.  the  components  of  the  Coriolis  force 
due  to  the  rotation  of  the  coordinate  system,  anei  pg  is  the  force  of 
gravity  (assumed  to  be  constant).  These  equations  hold  in  both  the 
warm  air  and  the  cold  air,  but  it  is  preferable  to  distinguish  the  eie- 
pendent  quantities  in  the  two  different  layers;  this  is  done  here 
throughout  by  writing  u'9  v',  w'  for  the  velocity  components  in  the 
warm  air  and  similarly  for  the  other  elependent  quantities. 

We  n6w  introduce  an  assumption  which  is  commonly  made  in 
dynamic  meteorology  in  discussing  large-scale  motions  of  the  atmos- 
phere, i.e.  that  the  air  is  incompressible.  In  spite  of  the  fact  that  such 


382  WATEE  WAVES 

an  assumption  rules  out  thermodynamic  processes,  it  does  seem  rather 
reasonable  since  the  pressure  gradients  which  operate  to  create  the 
flows  of  interest  to  us  are  quite  small  and,  what  is  perhaps  the  decisive 
point,  the  propagation  speed  of  the  disturbances  to  be  studied  is  very 
small  compared  with  the  speed  of  sound  in  air  (i.e.  with  disturbances 
governed  by  compressibility  effects).  It  would  be  possible  to  consider 
the  atmosphere,  though  incompressible,  to  be  of  variable  density. 
However,  for  the  purpose  of  obtaining  as  simple  a  dynamical  model  as 
possible  it  seems  reasonable  to  begin  with  an  atmosphere  having  a 
constant  density  in  each  of  the  two  layers.  As  a  consequence  of  these 
assumptions  we  have  the  following  equation  of  continuity: 

(10.11.2)  u*+vy+w,  =  0. 

The  equations  (10.11.1)  and  (10.11.2)  together  with  the  conditions 
of  continuity  of  the  pressure  and  of  the  normal  velocity  components 
on  the  discontinuity  surface,  the  condition  w  —  0  at  the  ground, 
appropriate  initial  conditions,  etc.  doubtlessly  yield  a  mathematical 
problem— call  it  Problem  I— the  solution  of  which  would  furnish  a 
reasonably  good  approximation  to  the  observed  phenomena.  Unfor- 
tunately, such  a  problem  is  still  so  difficult  as  to  be  far  beyond  the 
scope  of  known  methods  of  analysis— including  analysis  by  numerical 
computation.  Thus  still  further  simplifications  arc  in  order. 

One  of  the  best-founded  empirical  laws  in  dynamic  meteorology  is 
the  hydrostatic  pressure  law,  which  states  that  the  pressure  at  any 
point  in  the  atmosphere  is  very  closely  equal  to  the  static  weight  of 
the  column  of  air  vertically  above  it.  This  is  equivalent  to  saying  that 
the  vertical  acceleration  terms  and  the  Coriolis  force  in  the  third 
equation  of  (10.11.1)  can  be  ignored  with  the  result 

(10.11.3)  £  =  -  &. 

This  is  also  the  basis  of  the  long-wave  or  shallow  water  theory  of 
surface  gravity  waves,  as  was  already  mentioned  above.  Since  the 
vertical  component  of  the  acceleration  of  the  particles  is  thus  ignored, 
it  follows  on  purely  kinematical  grounds  that  the  horizontal  compo- 
nents of  the  velocity  will  remain  independent  of  the  vertical  coordinate 
z  for  all  time  if  that  was  the  case  at  the  initial  instant  t  =  0.  Since  we 
do  in  fact  assume  an  initial  motion  with  such  a  property,  it  follows 
that  we  have 


LONG    WAVES    IN    SHALLOW    WATER  883 

(10.11.4)        u  =  u(x,  y,  t),        v  =  v(x,  y,  t),        w  =  0.* 

The  first  two  of  the  equations  of  motion  (10.11.1)  and  the  equation  of 
continuity  (10.11.2)  therefore  reduce  to 


(10.11.5) 


vuy  =  —  —  px  +  F(x) 
vvy  = py  +  F(y) 


ux+vy  =  0, 

where  we  use  subscripts  to  denote  partial  derivatives  and  subscripts 
enclosed  in  parentheses  to  indicate  components  of  a  vector.  The 
Coriolis  acceleration  terms  are  now  given  by 

f  F(x)  =  2co  sin  q>  •  v  =  Av 
(10.11.6)  '*'  \ 

[  F(y}  =  —  2w  sin  (p  •  u  =  —  Aw 

when  use  is  again  made  of  the  fact  that  w  =  0.  (The  latitude  angle  9?, 
as  was  indicated  earlier,  is  assumed  to  be  constant.)  We  observe  once 
more  that  all  of  these  relations  hold  in  both  the  warm  and  cold  layers, 
and  we  distinguish  between  the  two  when  necessary  by  a  prime  on  the 
symbols  for  quantities  in  the  warm  air.  It  is  perhaps  also  worth  men- 
tioning that  the  equations  (10.11.5)  with  F(x}  and  F(y)  defined  by 

(10.11.6)  arc  valid  for  all  orientations  of  the  x,  j/-axes;  thus  it  is  not 
necessary  to  assume  (as  we  did  earlier,  for  example)  that  the  original 
stationary  front  runs  in  the  east-west  direction. 

We  have  not  so  far  made  full  use  of  the  hydrostatic  pressure  law 
(10.11.3).  To  this  end  it  is  useful  to  introduce  the  vertical  height 
h  —  /i(o%  y,  t)  of  the  discontinuity  surface  between  the  warm  and  cold 
layers  and  the  height  h'  =  h'(x,  y,t)  of  the  warm  layer  itself  (see 
Figure  10.11.3).  Assuming  that  the  pressure  p'  is  zero  at  the  top  of 
the  warm  layer  we  find  by  integrating  (10.11.3): 

(10.11.7)  p'Or,  y,  z,  t)  =  e'g(h'  -  z) 

for  the  pressure  at  any  point  in  the  warm  air.  In  the  cold  air  we  have, 
in  similar  fashion: 

(10.11.8)  p(xt  y,  z,  t)  ==  q'g(h'  -  h)  +  Qg(h  -  z) 

*  It  would  be  wrong,  however,  to  infer  that  we  assume  the  vertical  displacements 
to  be  zero.  This  is  a  peculiarity  of  the  shallow  water  theory  in  general  which 
results,  when  a  formal  perturbation  series  is  used,  because  of  the  manner  in 
which  the  independent  variables  are  made  to  depend  on  the  depth  (cf.  Ch.  2  and 
early  parts  of  the  present  chapter). 


384 


WATER    WAVES 


when  the  condition  of  continuity  of  pressure,  p'  =  p  for  z  =  h,  is 
used.  (The  formula  (10.11.8)  is  the  starting  point  of  the  paper  by 
Freeman  [F.10]  which  was  mentioned  earlier.)  Insertion  of  (10.11.8) 


h'(xty,t) 


Worm 


Ps  P> 


h(x,y,t)     Cold 


''•^  X 

Fig.  10.11.3.  Vertical  height  of  the  two  layers 

in  (10.11.5)  and  of  (10.11.7)  in  (10.11.5)'  yields  the  following  six 
equations  for  the  six  quantities  u9  v9  h,  ur,  v\  h': 


(10.11.9) 
(cold  air) 


(10.11.10) 
(warm  air) 


u 


vt 


u 


uu 


4 


u'u 


uv 


wv  =  — 


0 


vuy  = 


vv    = 


=  0. 


gh'x 
gh'y 


These  equations  together  with  the  kinematic  conditions  appropriate 
at  the  surfaces  z  =  h  and  z  —  h',  and  initial  conditions  at  t  =-  0, 
would  again  constitute  a  reasonable  mathematical  problem—  call  it 
Problem  II—  which  could  be  used  to  study  the  dynamics  of  frontal 
motions.  The  Problem  II  is  much  simpler  than  the  Problem  I  formu- 
lated above  in  that  the  number  of  dependent  quantities  is  reduced 
from  eight  to  six  and,  probably  still  more  important,  the  number  of 
independent  variables  is  reduced  from  four  to  three.  These  simplifica- 
tions, it  should  be  noted,  come  about  solely  as  a  consequence  of  assum- 
ing the  hydrostatic  pressure  law,  and  since  meteorologists  have  much 


LONG   WAVES   IN   SHALLOW   WATER  385 

evidence  supporting  the  validity  of  such  an  assumption,  the  Problem 
II  should  then  furnish  a  reasonable  basis  for  discussing  the  problem 
of  frontal  motions.  Unfortunately,  Problem  II  is  just  about  as  in- 
accessible as  Problem  I  from  the  point  of  view  of  mathematical  and 
numerical  analysis.  Consequently,  we  make  still  further  hypotheses 
leading  to  a  simpler  theory. 

As  a  preliminary  to  the  formulation  of  Problem  III  we  write  down 
the  kinematic  free  surface  conditions  at  z  =  h  and  z  =  h'  (the  dyna- 
mical free  surface  conditions,  p  =  0  at  z  =  h'  and  p  =  p'  at  z  =  h, 
have  already  been  used.)  These  conditions  state  simply  that  the 
particle  derivatives  of  the  functions  z  —  h(x,  y,  t)  and  z  —  h'(z,  y,  t) 
vanish,  since  any  particle  on  the  surface  z  —  h  =  0  or  the  surface 
z  —  h'  =  0  remains  on  it.  We  have  therefore  the  conditions 

uhx  +  vkv  +  ht  =  0 

(10.11.11)  u'hr  +  v'hy  +  ht  =  0 

>  *x  +  »'*;  +  tit  -  o, 

in  view  of  the  fact  that  w  vanishes  everywhere.  It  is  convenient  to 
replace  the  third  equations  (the  continuity  equations)  in  the  sets 
(10.11.9)  and  (10.11.10)  by 

(10.11.12)  (uh)x  +  (vh)v  +  ht  =  0,         and 

(10.11.13)  [u'(hf  -  h)]x  +  [v'(h'  -  h)]y  +  (h'  -  h)t  =  0, 

which  are  readily  seen  to  hold  because  of  (10.11.11).  In  fact,  the  last 
two  equations  simply  state  the  continuity  conditions  for  a  vertical 
column  of  air  extending  (in  the  cold  air)  from  the  ground  up  to  z  =  h, 
and  (in  the  warm  air)  from  z  =  h  to  z  =  h'. 

We  now  make  a  really  trenchant  assumption,  i.e.  that  the  motion 
of  the  warm  air  layer  is  not  affected  by  the  motion  of  the  cold  air  layer. 
This  assumption  has  a  rather  reasonable  physical  basis,  as  might  be 
argued  in  the  following  way:  Imagine  the  stationary  front  to  have 
developed  a  bulge  in  the  ^-direction,  say,  as  in  Figure  10.1 1.4a.  The 
warm  air  can  adjust  itself  to  the  new  condition  simply  through  a 
slight  change  in  its  vertical  component,  without  any  need  for  a  change 
in  u'  and  v'9  the  horizontal  components.  This  is  indicated  in  Figure 
10, 11. 46,  which  is  a  vertical  section  of  the  air  taken  along  the  line 
AB  in  Figure  10. 11.40;  in  this  figure  the  cold  layer  is  shown  with  a 
quite  small  height— which  is  what  one  always  assumes.  Since  we 
ignore  changes  in  the  vertical  velocity  components  in  any  case,  it  thus 
seems  reasonable  to  make  our  assumption  of  unaltered  flow  conditions 


386 


WATER   WAVES 


in  the  warm  air.  However,  in  the  cold  air  one  sees  readily —as  indicat- 
ed in  Figure  10.11.4c— that  quite  large  changes  in  the  components 
u,  v  of  the  velocity  in  the  cold  air  may  be  needed  when  a  frontal  dis- 
turbance is  created.  Thus  we  assume  from  now  on  that  u',  v',  h'  have 
for  all  time  the  known  values  they  had  in  the  initial  steady  state  in 


////  B 


7/m/ 


(o) 


Fig.  10.11.4.  Flows  in  warm  and  cold  air  layers 

which  v'  —  0,  u'  —  const.  The  differential  equations  for  our  Problem 
III  can  now  be  written  as  follows: 


(10.11.14) 


ut  +  uux 


=  -  g  f—  h'x  +  il  -  ^-\  hA  + 

=  -  g  P-A;  +  (i  -  ^~\  AJ  - 


(uk)x  +  (vh)v  =  0, 


LONG    WAVES    IN    SHALLOW    WATER 


387 


in  which  h'x  and  h'y  are  known  functions  given  in  terms  of  the  initial 
state  in  the  warm  air.  The  initial  state,  in  which  v'  =  v  =  0,  u  = 
const.,  u1  =  const.,  must  satisfy  the  equations  (10. 11. 9)  and  (10.11.10); 
this  leads  at  once  to  the  conditions 


*;- 


A 

g1 

^ 
g' 


*(e'   ,      \ 

-I  —  u    —  U  I 

g\0  / 


(10.11.15) 


for  the  slopes  of  the  free  surfaces  initially.  The  slope  hy  of  the  dis- 
continuity surface  between  the  two  layers  is  nearly  proportional  to 
the  velocity  difference  u'  —  u  since  g'/p  differs  only  slightly  from 
unity,  and  it  is  made  quite  small  under  the  conditions  normally  en- 
countered because  of  the  factor  A,  which  is  a  fraction  of  the  angular 
velocity  of  the  earth.  The  relation  for  the  slope  hy  of  the  stationary 
discontinuity  surface  is  an  expression  of  the  law  of  Margulcs  in  meteor- 
ology. The  differential  equations  for  Problem  III  can,  finally,  be  ex- 
pressed in  the  form: 


(10.11.16) 
Problem  III 


wt  +  t<ux  +  ruv  +  g  J 1  -  —  | 
vt  +  uvf  +  vvv  +  (>  ( 1 l/t, 

/?,   -f  (U/t),   +   (!'/*)„  =  0. 


by  using  the  formulas  for  h'x  and  h'v  given  in  (10.11.15).  We  note  that 
the  influence  of  the  warm  air  expresses  itself  through  its  density  g' 
and  its  velocity  u'.  The  three  equations  (10.11.16)  undoubtedly  have 
uniquely  determined  solutions  once  the  values  of  u,  v,  and  h  are  given 
at  the  initial  instant  t  =-  0,  together  with  appropriate  boundary  con- 
ditions if  the  domain  in  #,  y  is  not  the  whole  space,  and  such  solutions 
might  reasonably  be  expected  to  furnish  an  approximate  descrip- 
tion of  the  dynamics  of  frontal  motions.*  Unfortunately,  these  equa- 
tions are  still  quite  complicated.  They  could  be  integrated  numerically 

*  These  equations  are  in  fact  quite  similar  to  the  equations  for  two-dimen- 
sional unsteady  motion  of  a  compressible  fluid  with  h  playing  the  role  of  the 
density  of  the  fluid. 


388  WATER   WAVES 

only  with  great  difficulty  even  with  the  aid  of  the  most  modern  high- 
speed digital  computers— -mostly  because  there  are  three  independent 
variables. 

Consequently,  one  casts  about  for  still  other  possibilities,  either  of 
specialization  or  simplification,  which  might  yield  a  manageable 
theory.  One  possibility  of  specialization  has  already  been  mentioned: 
if  one  assumes  no  Coriolis  force  and  also  assumes  that  the  motion  is 
independent  of  the  ^-coordinate,  one  obtains  the  pair  of  equations 


ut  +  uux  +  g    l  -  —  \hx  = 


(10.11.17) 

(Hfc).  =  0 

which  are  identical  with  the  equations  of  the  one-dimensional  shallow 
water  gravity  wave  theory.  These  equations  contain  in  them  the 
possibility  of  the  development  of  discontinuous  motions— called 
bores  in  sec.  10.7— and  this  fact  lies  at  the  basis  of  the  discussions  by 
Freeman  [F.10]  and  Abdullah  [A.7].  In  such  one-dimensional  treat- 
ments, it  is  clear  that  it  is  in  principle  not  possible  to  deal  with  the 
bulges  on  fronts  and  their  deformation  in  time  and  space,  since  such 
problems  depend  essentially  on  both  space  variables  x  and  y.  Another 
possibility  would  be  a  linearization  of  the  differential  equations 
(10.11.16)  based  on  assuming  small  perturbations  of  the  frontal  sur- 
face and  of  the  velocities  from  the  initial  uniform  state.  This  procedure 
might  be  of  some  interest,  since  such  a  formulation  would  take  care  of 
the  boundary  condition  at  the  ground,  while  the  existing  linear  treat- 
ments of  this  problem  do  not.  However,  our  interest  here  is  in  a  non- 
linear treatment  which  permits  of  large  displacements  of  the  fronts. 
One  such  possibility,  devised  by  Whitham  [W.12],  involves  essentially 
the  integration  of  the  first  and  third  equations  for  u  and  h  as  functions 
of  x  and  t,  regarding  y  as  a  parameter,  and  derivatives  with  respect  to 
y  as  negligible  compared  with  derivatives  with  respect  to  x,  and  assum- 
ing initial  values  for  v;  this  is  feasible  by  the  method  of  characteristics. 
Afterwards,  v  would  be  determined  by  integrating  the  second  equa- 
tion considering  u  and  h  as  known,  and  this  can  in  principle  be  done 
because  the  equation  is  a  linear  first  order  equation  under  these  con- 
ditions. As  stated  earlier,  this  procedure  furnishes  qualitative  results 
which  agree  with  observations.  In  addition,  the  discussion  can  be 
carried  through  explicitly  in  certain  cases,  by  making  use  of  solutions 
of  the  type  called  simple  waves,  along  exactly  the  same  lines  as  in 
sec.  10.8  above.  We  turn,  therefore,  to  this  first  of  two  proposed 


LONG   WAVES    IN    SHALLOW   WATER  389 

approximate  treatments  of  Problem  III,  as  embodied  in  equations 
(10.11.16). 

The  basic  fact  from  which  Whitham  starts  is  that  the  slope  a  =  hy 
of  the  discontinuity  surface  is  small  initially,  as  we  have  already  seen 
in  connection  with  the  second  equation  of  (10.11.15),  and  the  fact  that 
A  is  a  fraction  of  the  earth's  angular  velocity,  and  is  expected  to  remain 
in  general  small  throughout  the  motions  considered.  Since  the  Coriolis 
forces  are  of  order  a  also  (since  they  are  proportional  to  A)  it  seems 
clear  that  derivatives  of  all  quantities  with  respect  to  y  will  be  small  of 
a  different  order  from  those  with  respect  to  x\  it  is  assumed  therefore 
that  uy9  hy  and  vy  are  all  small  of  order  a,  but  that  ux  and  hx  are  finite. 
Furthermore  we  can  expect  that  the  main  motion  will  continue  to  be 
a  motion  in  the  ^-direction,  so  that  the  i/-component  v  of  the  velocity 
will  be  small  of  order  a  while  the  ^-component  u  remains  of  course 
finite.  Under  these  circumstances,  the  equations  (10.11.16)  can  be 
replaced  by  simpler  equations  through  neglect  of  all  but  the  lowest 
order  terms  in  a  in  each  equation;  the  result  is  the  set  of  equations 

ht  +  uhx  +  hux  =  0 
«  +  uux  +  khx  =  0 


IQ'         \ 

vt  +  uvx  =  —  khy  +  A  (  —  u'  —  u\ 
with  the  constant  k  defined  by 

(10.11.19)  *  = 

A  considerable  simplification  has  been  achieved  by  this  process,  since 
the  variable  y  enters  into  the  first  two  equations  of  (10.11.18)  only  as 
a  parameter  and  these  two  equations  are  identical  with  the  equations 
of  the  shallow  water  theory  developed  in  the  preceding  sections  of  this 
chapter  if  k  is  identified  with  g  and  h  with  77.  This  means  that  the 
theory  developed  for  these  equations  now  becomes  available  to  dis- 
cuss our  meteorological  problems.  Of  course,  the  solutions  for  h  and 
u  will  depend  on  the  variable  y  through  the  agency  of  initial  and 
boundary  conditions.  Once  u(x9  y,  t)  and  h(x,  y>  t)  have  been  obtained, 
they  can  be  inserted  in  the  third  equation  of  (10.11.18),  which  then 
is  a  first  order  linear  partial  differential  equation  which,  in  principle 
at  least,  can  be  integrated  to  obtain  v  when  arbitrary  initial  conditions 
v  =  v(x9  y,  0)  are  prescribed.  The  procedure  contemplated  can  thus  be 
summed  up  as  follows:  the  motion  is  to  be  studied  first  in  each  vertical 


390  WATER   WAVES 

plane  y  =  constant  by  the  same  methods  as  in  the  shallow  water 
theory  for  two-dimensional  motions  (which  means  gas  dynamics 
methods  for  one-dimensional  unsteady  motions),  to  be  followed  by 
a  determination  of  the  "cross-component"  v  of  the  velocity  through 
integration  of  a  first  order  linear  equation  which  also  contains  the 
variable  y,  but  only  as  a  parameter. 

This  is  in  principle  a  feasible  program,  but  it  presents  problems  too 
complicated  to  be  solved  in  general  without  using  numerical  com- 
putations. On  the  other  hand  we  know  from  the  earlier  parts  of  this 
chapter  that  interesting  special  solutions  of  the  first  two  equations  of 
(10.11.18)  exist  in  the  form  of  what  were  called  simple  waves,  and 
these  solutions  lend  themselves  to  an  easy  discussion  of  a  variety  of 
motions  in  an  explicit  way  through  the  use  of  the  characteristic  form 
of  the  equations.  In  order  to  preserve  the  continuity  of  the  discussion 
it  is  necessary  to  repeat  here  some  of  the  facts  about  the  characteristic 
theory  and  the  theory  of  simple  waves;  for  details,  sees.  10.2  and  10.3 
should  be  consulted. 

By  introducing  the  new  function  c2  —  kh,  replacing  h,  we  obtain 
instead  of  the  first  two  equations  in  (10.11.18)  the  following  equations: 

( 2ct  +  2ucx  +  cux  ==  0 
(10.11.20) 

I  ut  +  uux  +  2ccx  ==  0. 

Thus  the  quantity  c  =  Vkh,  which  has  the  dimensions  of  a  velocity, 
is  the  propagation  speed  of  small  disturbances,  or  wavelets  — in  ana- 
logy with  the  facts  derived  in  sec.  10.2.  These  equations  can  in  turn 
be  written  in  the  form 


which  can  be  interpreted  to  mean  that  the  quantities  u±  2c  are  con- 
stant along  curves  C±  in  the  #,  2-plane  such  that  dx/dt  =  u  ±  c: 

„     dx 
u  -\-  2c  =  const,  along  C+:  —  =  u  +  c 

(10.11.21)     < 

ax 
u  —  2c  =  const,  along  C_:  —  =  u  —  c. 

These  relations  hold  in  general  for  any  solutions  of  (10.11.20).  Under 
special  circumstances  it  may  happen  that  u  — •  2c9  for  example,  has  the 
same  constant  value  on  all  C_  characteristics  in  a  certain  region;  in 


LONG    WAVES    IN    SHALLOW   WATER  391 

that  case  since  u  +  2c  is  constant  along  each  C+  characteristic  it 
follows  that  u  and  c  would  separately  by  constant  along  each  of  the 
C+  characteristics,  which  means  that  these  curves  would  all  be  straight 
lines.  Such  a  region  of  the  flow  (the  term  region  here  being  applied 
with  respect  to  some  portion  of  an  «r,  £-plane)  is  called  a  simple  wave. 
It  is  then  a  very  important  general  fact  that  any  flow  region  adjacent 
to  a  region  in  which  the  flow  is  uniform,  i.e.  in  which  both  c  and  u  are 
everywhere  constant  (in  both  space  and  time,  that  is),  is  a  simple 
wave,  provided  that  u  and  c  are  continuous  in  the  region  in  question. 

It  is  reasonable  to  suppose  that  simple  waves  would  occur  in  cases 
of  interest  to  us  in  our  study  of  the  dynamics  of  frontal  motions, 
simply  because  we  do  actually  begin  with  a  flow  in  which  u  and  h 
(hence  also  c)  are  constant  in  space  and  time,  and  it  seems  reasonable 
to  suppose  that  disturbances  are  initiated,  not  everywhere  in  the  flow 
region,  but  only  in  certain  areas.  In  other  words,  flows  adjacent  to 
uniform  flows  would  occur  in  the  nature  of  things.  Just  how  in  detail 
initial  or  boundary  conditions,  or  both,  should  be  prescribed  in  order 
to  conform  with  what  actually  occurs  in  nature  is,  as  has  already  been 
pointed  out,  something  of  a  mystery;  in  fact  one  of  the  principal 
objects  of  the  ideas  presented  here  could  be  to  make  a  comparison  of 
calculated  motions  under  prescribed  initial  and  boundary  conditions 
with  observed  motions  in  the  hope  of  learning  something  by  inference 
concerning  the  causes  for  the  initiation  and  development  of  frontal 
disturbances  as  seen  in  nature. 

One  fairly  obvious  and  rather  reasonable  assumption  to  begin  with 
might  be  that  u>  v,  and  h  are  prescribed  at  the  time  t  =  0  to  have 
values  over  a  certain  bounded  region  of  the  upper  half  (y  >  0)  of 
the  x,  y-plane  (cf.  Fig.  10.11.1 )  in  such  a  fashion  that  they  differ  from 
the  constant  values  in  the  original  uniform  flow  with  a  stationary 
front.  According  to  the  approximate  theory  based  on  equations  (10.11. 
18),  this  means,  in  particular,  that  in  each  vertical  plane  y  =  t/0— an 
x,  J-plane  — initial  conditions  for  u(x,y09t)  and  h(x,y0,t)  would  be 
prescribed  over  the  entire  #-axis,  but  in  such  a  way  that  u  and  h  are 
constant  with  values  u  =  UQ  >  0,  h  =  A0  ^  0  (hence  c  =  c0  = 

Vkh0)*  everywhere  except  over  a  certain  segment  xl  ^  x  ^  #2,  as 
indicated  in  Fig.  10.11.5.  The  positive  characteristics  C+  are  drawn  in 
full  lines,  the  characteristics  C_  with  dashed  lines  in  this  diagram, 

*  It  should,  however,  always  be  kept  in  mind  in  the  discussion  to  follow 
that  c0,  particularly,  will  usually  have  different  values  in  different  vertical  planes 
y  =  const. 


892 


WATER   WAVES 


which  is  to  be  interpreted  as  follows.  Simple  waves  exist  everywhere 
in  the  x,  J-plane  except  in  the  triangular  region  bounded  by  the  C+ 
characteristic  through  A  and  the  C_  characteristic  through  B  and 
terminating  at  point  C;  in  this  region  the  flow  could  be  determined 


A  B 

Fig.  10.11.5.  Simple  waves  arising  from  initial  conditions 

numerically,  for  example  by  the  method  indicated  in  sec.  10.2  above 
(in  connection  with  Fig.  10.2.1).  The  disturbance  created  over  the 
segment  AB  propagates  both  "upstream"  and  "downstream"  after  a 
certain  time  in  the  form  of  two  simple  waves,  which  cover  the  regions 
bounded  by  the  straight  (and  parallel)  characteristics  issuing  from 
A,  B9  and  C.  In  other  words  the  disturbance  eventually  results  in  two 
distinct  simple  waves,  one  propagating  upstream,  the  other  down- 
stream, and  separated  by  a  uniform  flow  identical  with  the  initial 
state.  In  our  diagram  it  is  tacitly  assumed  that  c  >  \  u  |,  i.e.  that  the 
flow  is  subcritical  in  the  terminology  of  water  waves  (subsonic  in 
gas  dynamics)— otherwise  no  propagation  upstream  could  occur.  We 
have  supposed  u  to  be  positive,  i.e.  that  the  ^-component  of  the  flow 
velocity  in  the  cold  air  layer  has  the  same  direction  as  the  velocity  in 
the  warm  air,  which  in  general  flows  from  the  west,  but  it  can  be  (and 
not  infrequently  is)  in  the  westward  rather  than  the  eastward  direc- 
tion. Since  the  observed  fronts  seem  to  move  almost  invariably  to  the 
eastward,  it  follows,  for  example,  that  it  would  be  the  wave  moving 
upstream  which  would  be  important  in  the  case  of  a  wind  to  the  west- 
ward in  the  cold  layer,  and  a  model  of  the  type  considered  here  — in 
which  the  disturbance  is  prescribed  by  means  of  an  initial  condition 


LONG   WAVES    IN    SHALLOW   WATER  393 

and  the  flow  is  subcritical— implies  that  the  initial  disturbances  are 
always  of  such  a  special  character  that  the  downstream  wave  has  a 
negligible  amplitude.  For  a  wind  to  the  eastward,  the  reverse  would 
be  the  case.  All  of  this  is,  naturally,  of  an  extremely  hypothetical 
character,  but  nevertheless  one  sees  that  certain  important  elements 
pertinent  to  a  discussion  of  possible  motions  are  put  in  evidence. 

The  last  remarks  indicate  that  a  model  based  on  such  an  initial 
disturbance  may  not  be  the  most  appropriate  in  the  majority  of  cases. 
In  fact,  such  a  formulation  of  the  problem  is  open  to  an  objection 
which  is  probably  rather  serious.  The  objection  is  that  such  a  motion 
has  its  origin  in  an  initial  impulse,  and  this  provides  no  mechanism  by 
which  energy  could  be  constantly  fed  into  the  system  to  "drive"  the 
wave.  Of  course,  it  would  be  possible  to  introduce  external  body  forces 
in  various  ways  to  achieve  such  a  purpose,  but  it  is  not  easy  to  see  how 
to  do  that  in  a  rational  way  from  the  point  of  view  of  mechanics. 
Another  way  to  introduce  energy  into  the  system  would  be  to  feed  it 
in  through  a  boundary —in  other  words  formulate  appropriate  bound- 
ary conditions  as  well  as  initial  conditions.  For  the  case  of  fronts 
moving  eastward  across  the  United  States,  a  boundary  condition 
might  be  reasonably  applied  at  some  point  to  the  east  of  the  high 
mountain  system  bordering  the  west  coast  of  the  continent,  since 
these  mountain  ranges  form  a  rather  effective  north-south  barrier 
between  the  motions  at  the  ground  on  its  two  sides.  In  fact,  a  cold 
front  is  not  infrequently  seen  running  nearly  parallel  to  the  moun- 
tains and  to  the  cast  of  them— as  though  cold  air  had  been  deflected 
southward  at  this  barrier.  Hence  a  boundary  condition  applied  at 
some  point  on  the  west  seems  not  entirely  without  reason.  In  any  case, 
we  seek  models  from  which  knowledge  about  the  dynamics  of  fronts 
might  be  obtained,  and  a  model  making  use  of  boundary  conditions 
should  be  studied.  We  suppose,  therefore,  that  a  boundary  condition 
is  applied  at  x  =  0,  and  that  the  initial  condition  for  t  —  0,  x  >  0  is 
that  the  flow  is  undisturbed,  i.e.  u  =  UQ  =  const.,  c  =  c0  =  const.. 
(Again  we  remark  that  we  are  considering  the  motion  in  a  definite 
vertical  plane  y  =  yQ. )  In  this  case  we  would  have  only  a  wave  propa- 
gating eastward— in  effect,  we  replace  the  influence  of  the  air  to  the 
westward  by  an  assumed  boundary  condition.  The  general  situation 
is  indicated  in  Fig.  10.11.6.  There  is  again  a  simple  wave  in  the  region 
of  the  #,  $-plane  above  the  straight  line  x  —  (UQ  +  cQ)t  which  marks 
the  boundary  between  the  undisturbed  flow  and  the  wave  arising 
from  disturbances  created  at  x  =  0.  This  is  exactly  the  situation  which 


394 


WATER   WAVES 


is  treated  at  length  in  sec.  10.3;  in  particular,  an  explicit  solution  of 
the  problem  is  easily  obtained  (cf.  the  discussion  in  sec.  10.4)  for 
arbitrarily  prescribed  disturbances  in  the  values  of  either  of  the  two 
quantities  u  or  c.  Through  various  choices  of  boundary  conditions  it 
is  possible  to  supply  energy  to  the  system  in  a  variety  of  ways. 


Fig.  10.11.6.  Wave  arising  from  conditions  applied  at  a  boundary 

We  proceed  next  to  discuss  qualitatively  a  few  consequences  which 
result  if  it  is  assumed  that  frontal  disturbances  can  be  described  in 
terms  of  simple  waves  in  all  vertical  planes  y  =  yQ  =  const,  at  least 
over  some  ranges  in  the  values  of  the  ^-coordinate.  (We  shall  see  later 
that  simple  waves  are  not  possible  for  all  values  of  y. )  In  this  discussion 
we  do  not  specify  how  the  simple  wave  was  originated  — we  simply 
assume  it  to  exist.  Since  we  consider  only  waves  moving  eastward 
(i.e.  in  the  positive  ^-direction)  it  follows  that  the  straight  character- 
istics are  C+  characteristics,  and  hence  that  u  —  2c  is  constant  (in 
each  plane  y  =•  const.)  throughout  the  wave;  we  have  therefore 

(10.11.22)  u  -  2c  =  A(y), 

with  A(y)  fixed  by  the  values  UQ  and  c0(y)  in  the  undisturbed  flow: 
(10.11.22^  A(y)  =  UQ  -  2c0(y). 

In  addition,  as  explained  before,  we  know  that  u  +  2c  is  a  function  of 
y  alone  on  each  positive  characteristic  dx/dt  =  u  +  c;  hence  u  and  c 
are  individually  functions  of  y  on  each  of  these  characteristics.  There- 
fore, the  characteristic  equation  may  be  integrated  to  yield 

(10.11.23)  x  ==  £  +  (u  +  c)t, 

where  |  is  the  value  of  x  at  t  =  0.  (The  time  t  =  0  should  be  thought 
of  as  corresponding  to  an  arbitrary  instant  at  which  simple  waves 
exist  in  certain  planes  y  =  const. )  Now,  the  values  of  u  and  c  on  the 


LONG    WAVES    IN    SHALLOW   WATER  395 

characteristic  given  by  (10.11.23)  are  exactly  the  same  as  the  values 
(for  the  same  value  of  y)  at  the  point  t  =  0,  x  =  f ;  therefore,  if  we 
suppose,  for  example,  that  c  is  a  given  function  C(#,  y)  at  t  =  0,  the 
value  of  c  in  (10.11.23)  is  C(f,  y)  and  the  value  of  u  is,  from  (10.11.22), 
A(y)  +  2C(f,  t/).  Thus  the  simple  wave  solution  can  be  described  by 
the  equations 

(  c  =  C(£,  y), 
(10.11.24)  •  u  =  A(y)  +2C(f,y), 

,  a?  =  £  +  {^(y)  +  8C(f ,  y)}t. 

(Although  the  arbitrary  function  occurring  in  a  simple  wave  could  be 
specified  in  other  ways,  it  is  convenient  for  our  purposes  to  give  the 
value  of  h,  and  hence  c,  at  t  =  0.) 

We  could  write  down  the  solution  for  the  "cross  component",  or 
north-south  component,  v  of  the  velocity  in  this  case;  by  standard 
methods  (cf.  the  report  of  Whitham  [W.12])  it  can  be  obtained  by 
integrating  the  linear  first  order  partial  differential  equation  which 
occurs  third  in  the  basic  equations  (10.11.18).  To  specify  the  solution 
of  this  equation  uniquely  an  initial  condition  is  needed;  this  might 
reasonably  be  furnished  by  the  values  v  =  v(x,  y)  at  the  time  t  =  0. 
The  result  is  a  rather  complicated  expression  from  which  not  much 
can  be  said  in  a  general  way.  One  of  the  weaknesses  of  the  present 
attack  on  our  problem  through  the  use  of  simple  waves  now  becomes 
apparent:  it  is  necessary  to  know  values  of  v  some  time  subsequent  to 
the  initiation  of  a  disturbance  in  order  to  predict  them  for  the  future. 

It  is  possible,  however,  to  draw  some  interesting  conclusions  from 
the  simple  wave  motions  without  considering  the  north-south  com- 
ponent of  the  velocity.  For  example,  suppose  we  consider  a  motion 
after  a  bulge  to  the  northward  in  an  initially  stationary  front  had 
developed  as  indicated  schematically  in  Fig.  10.11.2.  In  a  plane  y  = 
const,  somewhat  to  the  north  of  the  bulge  we  could  expect  the  top  of 
the  cold  air  layer  (the  discontinuity  surface,  that  is)  as  given  by  h(x,  t) 
to  appear,  for  t  =  0  say,  as  indicated  in  Fig.  10.11.7.  The  main  fea- 
tures of  the  graph  arc  that  there  is  a  depression  in  the  discontinuity 
surface,  but  that  h  >  0  so  that  this  surface  does  not  touch  the  ground. 
(The  latter  possibility  will  be  discussed  later.)  Assuming  that  the 
motion  is  described  as  a  simple  wave,  we  see  from  (10.11.24)  that  the 
value  of  c  =  Vkh  at  the  point  x  =  xl  is  equal  to  the  value  of  c  which 
was  at  the  point  x  =  ^  at  t  =  0,  where  f  1  —  xl  —  (A(y)  +  3C(f1,t/)}/1. 
That  is,  the  value  c  =  C(flf  y)  has  been  displaced  to  the  right  by  an 


896 


WATER    WAVES 


amount  (A(y)  +  BC(^V  y)}tr  Since  this  quantity  is  greater  for  greater 
values  of  C,  the  graph  of  h  becomes  distorted  in  the  manner  shown  in 
Fig.  10.11.7:  the  "negative  region"  (where  hx  <  0)  steepens  whilst  the 
"positive  region"  (where  hx  >  0)  flattens  out.  The  positive  region 
continues  to  smooth  out,  but,  if  the  steepening  of  the  negative  region 


Fig.  10.11.7.  Deformation  of  the  discontinuity  surface 

were  to  continue  indefinitely,  there  would  ultimately  be  more  than 
one  value  of  C  at  the  same  point,  and  the  wave,  as  in  our  discussion 
of  water  waves  (cf.  sec.  10.6  and  10.7),  starts  to  break.  Clearly  the 
latter  event  occurs  when  the  tangent  at  a  point  of  the  curve  in 
Fig.  10.11.7  first  becomes  vertical.  At  this  time,  the  continuous  solu- 
tion breaks  down  (since  c  and  u  would  cease  to  be  single-valued 
functions)  and  a  discontinuous  jump  in  height  and  velocity  must  be 
permitted.  In  terms  of  the  description  of  the  wave  by  means  of  the 
characteristics,  what  happens  is  that  the  straight  line  characteristics 
converge  and  eventually  form  a  region  with  a  fold.  Such  a  disconti- 
nuous "bore"  propagates  faster  than  the  wavelets  ahead  of  it  (the 
paths  of  the  wavelets  in  the  x,  /-plane  are  the  characteristics)  in  a 
manner  analogous  to  the  propagation  of  shock  waves  in  gas  dynamics 
and  bores  in  water. 

In  the  above  paragraph  we  supposed  that  h  and  c  were  different 
from  zero,  and  hence  the  discussion  does  not  apply  to  the  fronts,  which 


LONG    WAVES    IN    SHALLOW   WATER  397 

are  by  definition  the  intersection  of  the  discontinuity  surface  with 
the  ground.  When  c  =  0  there  are  difficulties,  especially  at  cold 
fronts,  but  nevertheless  a  few  pertinent  observations  can  be  made, 
assuming  the  motion  to  be  a  simple  compression  wave  with  u  —  2c 
constant.  When  c  =  0,  it  follows  that  u  =  UQ  —  2c0,  and  since  c0  = 
VkhQ  and  hQ  =  at/0  with  a  the  initial  inclination  of  the  top  of  the  cold 
air  layer,  it  follows  that  u  =  UQ  —  2\/<x.ky  in  this  case.  But  u  then 
measures  the  speed  of  the  front  itself  in  the  ^-direction,  since  a  particle 
once  on  the  front  stays  there;  consequently  for  the  speed  uf  of  the 
front  we  have 

(10.11.25)  uf  =  UQ  —  2\/o%. 

Thus  the  speed  of  the  front  decreases  with  y,  and  on  this  basis  it 
follows  that  a  northward  bulge  would  become  distorted  in  the  fashion 
indicated  by  Fig.  10.11.8,  and  this  coincides  qualitatively  with  obser- 
vations of  actual  fronts. 

Actually,  things  are  not  quite  as  simple  as  this.  If  c  =  0,  it  follows 
from  the  first  equation  of  (10.11.20)  that  ct  +  ucx  =  0  on  such  a 
locus,  and  this  in  turn  means  that  c  =  0  on  the  particle  path  defined 
by  dx/dt  ~-  u.  At  the  same  time  the  C+  and  C_  characteristics  have 
the  same  direction,  since  they  are  given  by  dx/dt  —  u  ±  c.  On  the 
other  hand,  we  have,  again  from  (10.11.20),  ut  +  uux  —  —  khx  and 
we  see  that  the  relation  u  =  const,  along  a  characteristic  for  which 
c  —  0  cannot  be  satisfied  unless  hx  =  0.  In  connection  with  Fig. 
10.11.7  we  have  seen  that  the  rising  portion  toward  the  east  of  a  de- 
pression in  the  discontinuity  surface  tends  to  flatten  out,  while  the 
falling  part  from  the  west  tends  to  steepen  and  break  because  the 
higher  portions  tend  to  move  faster  and  crowd  the  lower  portions. 
Thus  when  /?,  and  hence  r,  tends  to  zero  the  tendency  will  be  for 
breaking  to  occur  at  the  cold  front,  but  not  at  the  warm  front.  The 
slope  of  the  discontinuity  surface  at  the  cold  front  will  then  be  infinite. 
However,  a  bore  in  the  sense  described  above  cannot  occur  since  there 
must  always  be  a  mass  flux  through  a  bore:  the  motion  of  the  cold 
front  is  analogous  to  what  would  happen  if  a  dam  were  broken  and 
water  rushed  down  the  dry  bed  of  a  stream.  Without  considering  in 
some  special  way  what  happens  in  the  turbulent  motion  caused  by  such 
continuous  breaking  at  the  ground,  it  is  not  possible  to  continue  our 
discussion  of  the  motion  of  a  cold  front  along  the  present  lines,  al- 
though such  a  problem  is  susceptible  to  an  approximate  treatment. 
Nevertheless,  this  discussion  has  led  in  a  rational  way  to  a  qualitative 


398 


WATER   WAVES 


explanation  for  the  well-known  fact  that  a  warm  front  does  indeed 
progress  in  a  relatively  smooth  fashion  as  compared  with  the  turbu- 
lence which  is  commonly  observed  at  cold  fronts.  Thus  near  a  cold 
front  the  height  of  the  cold  air  layer  may  be  considerably  greater  than 
in  the  vicinity  of  the  warm  front,  where  h  ~  0;  consequently  the  speed 
of  propagation  of  the  cold  front  could  be  expected  to  be  greater  than 


Cold 


Warm 


t  =  0 


Warm 


t=  t,  >  0 


Cold 


Warm 


tst2>t, 


Fig.  10.11.8.  Deformation  of  a  moving  front 

near  the  warm  front  (as  indicated  by  the  dotted  modifications  of  the 
shape  of  the  cold  front  in  Fig.  10.11.8),  with  the  consequence  that  the 
gap  between  the  two  tends  to  close,  and  this  hints  at  a  possible  ex- 
planation for  the  occlusion  process.  One  might  also  look  at  the  matter 
in  this  way:  Suppose  c  ^  0,  but  is  small  in  the  trough  of  the  wave 
shown  in  Fig.  10.11.7.  If  breaking  once  begins,  it  is  well  known  that 


LONG   WAVES   IN   SHALLOW   WATER  399 

the  resulting  bore  moves  with  a  speed  that  is  greater  than  the  propa- 
gation speed  of  wavelets  in  the  medium  in  front  (to  the  right)  of  it. 
although  slower  than  the  propagation  speed  in  the  medium  behind  it. 
Again  one  sees  that  the  tendency  for  the  wave  on  the  cold  front  side 
to  catch  up  with  the  wave  on  the  warm  front  side  is  to  be  expected  on 
the  basis  of  the  theory  presented  here. 

Finally  we  observe  that  the  velocity  of  the  wave  near  the  undisturbed 
stationary  front  is  UQ,  but  well  to  the  north  it  is  given  roughly  by 
uf  —  UQ  ~~  %  vaky,  which  is  less  than  UQ.  There  is  thus  a  tendency  to 
produce  what  is  called  in  meteorology  a  cyclonic  rotation  around  the 
center  of  the  wave  disturbance. 

To  sum  up,  it  seems  fair  to  say  that  the  approximate  theory  embo- 
died in  equations  (10.11.18),  even  when  applied  to  a  very  special  type 
of  motions  (i.e.  simple  waves  in  each  plane  y  =.  const.),  yields  a 
variety  of  results  which  are  at  least  qualitatively  in  accord  with 
observations  of  actual  fronts  in  the  atmosphere.  Among  the  pheno- 
mena given  correctly  in  a  qualitative  way  are:  the  change  in  shape  of 
a  wave  as  it  progresses  eastward,  the  occurrence  of  a  smooth  wave  at 
a  warm  front  but  a  turbulent  wave  at  a  cold  front,  and  a  tendency  to 
produce  the  type  of  motion  called  a  cyclone. 

It  therefore  seems  reasonable  to  suppose  that  the  differential 
equations  of  our  Problem  III,  which  were  the  starting  point  of  the 
discussion  just  concluded,  contain  in  them  the  possibility  of  dealing 
with  motions  which  have  the  general  characteristics  of  frontal  motions 
in  the  atmosphere,  and  that  numerical  solutions  of  the  equations  of 
Problem  III  might  well  furnish  valuable  insights.  This  is  a  difficult 
task,  as  has  already  been  mentioned.  However,  an  approximate  theory 
different  from  that  of  Whitham  is  possible,  which  has  the  advantage 
that  no  especial  difficulty  arises  at  cold  fronts,  and  which  would  per- 
mit a  numerical  treatment.  This  approximate  theory  might  be  con- 
sidered as  a  new  Problem  IV. 

The  formulation  of  Problem  IV  was  motivated  by  the  following 
considerations.  If  one  looks  at  a  sequence  of  weather  maps  and 
thinks  of  the  wave  motion  in  our  thin  wedge  of  cold  air,  the  re- 
semblance to  the  motion  of  waves  in  water  which  deform  into  brea- 
kers (especially  in  the  case  of  frontal  disturbances  which  develop  into 
occluded  fronts)  is  very  strong.  The  great  difference  is  that  the  wave 
motion  in  water  takes  place  in  the  vertical  plane  while  the  wave  mo- 
tion in  our  thin  layer  of  cold  air  takes  place  essentially  in  the  horizon- 
tal plane.  When  the  hydrostatic  pressure  assumption  is  made  in  the 


400  WATER   WAVES 

case  of  water  waves  the  result  is  a  theory  in  exact  analogy  to  gas 
dynamics,  and  thus  wave  motions  with  an  appropriate  "sound  speed" 
become  possible  even  though  the  fluid  is  incompressible— the  free 
surface  permits  the  introduction  of  the  depth  of  the  water  as  a  de- 
pendent quantity,  this  quantity  plays  the  role  of  the  density  in  gas 
dynamics,  and  thus  a  dynamical  model  in  the  form  of  a  compressible 
fluid  is  obtained.  It  would  seem  therefore  reasonable  to  try  to  invent 
a  similar  theory  for  frontal  motions  in  the  form  of  a  long- wave  theory 
suitable  for  waves  which  move  essentially  in  the  horizontal,  rather 
than  the  vertical,  plane,  and  in  which  the  waves  propagate  essentially 
parallel  to  the  edge  of  the  original  stationary  front,  i.e.  the  #-axis.  In 
this  way  one  might  hope  to  be  rid  of  the  dependence  on  the  variable  y 
at  right  angles  to  the  stationary  front,  thus  reducing  the  independent 
variables  to  two,  x  and  t;  and  if  one  still  could  obtain  a  hyperbolic 
system  of  differential  equations  then  numerical  treatments  by  finite 
differences  would  be  feasible.  This  program  can,  in  fact,  be  carried  out 
in  such  a  way  as  to  yield  a  system  of  four  first  order  nonlinear  differ- 
ential equations  in  two  independent  and  four  dependent  variables 
which  are  of  the  hyperbolic  type. 

Once  having  decided  to  obtain  a  long-wave  theory  for  the  horizontal 
plane,  the  procedure  to  be  followed  can  be  inferred  to  a  large  extent 
from  what  one  does  in  developing  the  same  type  of  theory  for  gravity 
waves  in  water,  as  we  have  seen  in  Chapter  2  and  at  the  beginning  of 
the  present  chapter.  To  begin  with  it  seems  clear  that  the  displace- 
ment r)(x,t)  of  the  front  itself  in  the  ^-direction  should  be  introduced 
as  one  of  the  dependent  quantities— all  the  more  since  this  quantity 
is  anyway  the  most  obvious  one  on  the  weather  maps.  To  have  such  a 
"shallow  water"  theory  in  the  horizontal  plane  requires— unfortuna- 
tely—a  rigid  "bottom"  somewhere  (which  is,  of  course,  vertical  in 
this  case),  and  this  we  simply  postulate,  i.e.  we  assume  that  the 
t/-component  v  of  the  velocity  vanishes  for  all  time  on  a  vertical  plane 
y  =  d  =  const,  parallel  to  the  stationary  front  along  the  #-axis  (see 
Figure  10.11.9).  The  velocity  v(x,  j/,  t)  is  then  assumed  to  vary 
linearly*  in  y,  and  its  value  at  the  front,  y  =  r\(x,  t),  is  called  v  (x,  t). 
The  intersection  of  the  discontinuity  surface  z  =  h(x9 1/,  t)  with  the 
plane  y  =  d  is  a  curve  given  by  z  =  Ti(x,  t),  and  we  assume  that  the 
discontinuity  surface  is  a  ruled  surface  having  straight  line  generators 
running  from  the  front,  y  =  rj(x9 1),  to  the  curve  z  —  Ti(x,  t),  and 

*  The  analogous  statement  holds  also  in  the  long-wave  theory  in  water  (to 
lowest  order  in  the  development  parameter,  that  is). 


LONG   WAVES   IN    SHALLOW   WATER 


401 


parallel  to  the  t/,  2-plane.  Finally,  we  assume  (as  in  the  shallow  water 
theory)  that  u9  the  ^-component  of  the  velocity,  depends  on  x  and  / 


a  z 


Fig.  10.11.9.  Notations  for  Problem  IV 

only:  u  =•  u(x,  t).  The  effect  of  these  assumptions  is  to  yield  the  re- 
lations 

^    y  -*?(*»*) 


(10.11.26) 


(10.11.27) 


V(JT,  y,  t)  = 


d  —  YI(X,  t) 
d-y 


h(x,  t), 


d-f,(x,t)    ~v~'"" 

as  one  readily  sees.  In  addition,  we  assume  that  a  particle  that  is  once 
on  the  front  y  -  r)(x,  t)  =  0  always  remains  on  it,  so  that  the  relation: 

(10.11.28)  v(x,  t)  =  r\t  +  ur\x 

must  hold.  The  four  quantities  u(x9  /),  rj(x,  t),  h(x,  t),  and  v(x,  t)  are 
our  new  dependent  variables.  Differential  equations  for  them  will  be 
obtained  by  integrating  the  basic  equations  (10.11.16)  of  Problem  III 
with  respect  to  y  from  y  =  rj  to  y  =  d —which  can  be  done  since  the 
dependence  of  w,  u,  and  h  on  y  is  now  explicitly  given— and  these  three 
equations  together  with  (10.11.28)  will  yield  the  four  equations  we 
want. 

Before  writing  these  equations  down  it  should  be  said  that  the 
most  trenchant  assumption  made  here  is  the  assumption  concerning 
the  existence  of  the  rigid  boundary  y  =  d.  One  might  think  that  as 


402  WATER   WAVES 

long  as  the  velocity  component  v  dies  out  with  sufficient  rapidity  in 
the  ^-direction  such  an  assumption  would  yield  a  good  approximation, 
but  the  facts  in  the  case  of  water  waves  indicate  this  to  be  not  suffi- 
cient for  the  accuracy  of  the  approximation:  with  water  waves  in  very 
deep  water  the  vertical  component  of  the  velocity  (corresponding  to 
our  v  here)  dies  out  very  rapidly  in  the  depth,  but  it  is  nevertheless 
essential  for  a  good  approximation  to  assume  that  the  ratio  of  the 
depth  down  to  a  rigid  bottom  to  the  wave  length  is  small.  However, 
such  a  rigid  vertical  barrier  to  the  winds  does  exist  in  some  cases  of 
interest  to  us  in  the  form  of  mountain  ranges,  which  are  often  much 
higher  than  the  top  of  the  cold  surface  layer  (i.e.  higher  than  the 
curve  z  =  h(x,  t)  in  Figure  10.11.9).  In  any  case,  severe  though  this 
restriction  is,  it  still  seems  to  the  author  to  be  worth  while  to  study 
the  motions  which  are  compatible  with  it  since  something  about  the 
dynamics  of  frontal  motions  with  large  deformations  may  be  learned 
in  the  process.  In  particular,  one  might  learn  something  about  the 
kind  of  perturbations  that  are  necessary  to  initiate  motions  of  the 
type  observed,  and  under  what  circumstances  the  motions  can  be 
maintained. 

In  carrying  out  the  derivation  of  the  differential  equations  of  our 
theory  according  to  the  plan  outlined  above,  we  calculate  first  a 
number  of  integrals.  The  first  of  these  arise  from  (10.11.26)  and 
(10.11.27): 

r6  h     r*  i  _ 

hdy  =  -  -       (y  ~ 

J  n  o  —rj  J  ^ 


From  these  we  derive  by  differentiations  with  respect  to  x  and  t  an- 
other set  of  relations: 


J 

C8  1 

J    vxdy  =  -vx(d 


=  oM<5  -n)  -*%*> 
n  i  * 

6  i  i_ 

2 
1 

05??«' 


r*  i  i_ 

J  r\  "  " 


LONG    WAVES    IN    SHALLOW    WATER 


403 


(In  deriving  these  relations,  it  is  necessary  to  observe  that  the  lower 
limit  T]  is  a  function  of  x  and  t.)  One  additional  relation  is  needed,  as 
follows: 


r* 

I 


(hu)xdy  ^  — 


1         \      I  -  1  _ 

h  dy  |  =  -  (hxu  +  hux)(6  -»?)--  hur}x. 

T)  vyu    \        %}   ri  1  &  ~ 

We  now  integrate  both  sides  of  the  equations  (10.11.16)  with  re- 
spect to  y  from  77  to  d,  make  use  of  the  above  integrals,  note  that 
u  =  u(x,  t)  is  independent  of  j/,  and  divide  by  d  —  TJ.  The  result  is  the 
equations 

1    _         1     kh  1 

*t  +  MM,  +  -  khx  -  -  i -nx  -  -Av, 


(10.11.29) 


.- 
<5—  77 


jo        _    v2         2kh      ^  I        Q'     \ 


- 
uhx 


u 


h,  -   - 


=  0. 


with  k  a  constant  replacing  the  quantity  g(l  —  Q'/'Q).  These  equations, 
together  with  (10.11.28),  form  a  system  of  four  partial  differential 
equations  for  the  four  functions  u,  77.  £,  and  7L  By  analogy  with  gas 
dynamics  and  the  nonlinear  shallow  water  theory,  it  is  convenient  to 
introduce  a  new  dependent  quantity  c  (which  will  turn  out  to  be  the 
propagation  speed  of  wavelets)  through  the  relation 


(10.11.30) 


--\h. 


The  quantity  c  is  real  if  Q'  is  less  than  Q,  and  this  holds  since  the  warm 
air  is  lighter  than  the  cold  air.  In  terms  of  this  new  quantity  the 
equations  (10.11.28)  and  (10.11.29)  take  the  form 

ut  +  uux  +  2rr,  — TJX  = ••  -  Ai", 


(10.11.31) 


vt  +  uvx  =  — 


6  — 


/          Q'      \ 
2A  1  u       —  u  \  , 


cv 


cux  +  2ucx  = 

d  —  r) 


rjt  +  ur/x  =  i". 
It  is  now  easy  to  write  the  equations  (10.11.31)  in  the  characteristic 


404  WATER   WAVES 

form  simply  by  replacing  the  first  and  third  equations  by  their  sum 
and  by  their  difference.  The  result  is: 


(10  11  32) 


4c2  /          o'      \ 

x  =  --  --  2A  [u  -  —  u'\, 
o  —  ri  \          Q      J 


t  +  urjx  =  v. 

As  one  sees,  the  equations  are  in  characteristic  form:  the  characteristic 
curves  satisfy  the  differential  equations 

dx  dx  dx 

(10.11.33)  —  =  u  +c,     —  =  u  —  c,     —  =  u, 

dt  dt  dt 

and  each  of  the  equations  (10.11.32)  contains  only  derivatives  in  the 
direction  of  one  of  these  curves.  The  characteristic  curves  defined  by 
dxjdt  —  u  are  taken  twice.  Thus  one  sees  that  the  quantity  c  is  indeed 
entitled  to  be  called  a  propagation  speed,  and  small  disturbances  can 
be  expected  to  propagate  with  this  speed  in  both  directions  relative  to 
the  stream  of  velocity  u.  (In  the  theory  by  Whitham,  in  which  the 
motion  in  each  vertical  plane  y  =  const,  is  treated  separately,  the 
propagation  or  sound  speed  of  small  disturbances  is  given  by  Vkh. 
The  sound  speed  in  the  theory  given  here  thus  represents  a  kind  of 
average  with  respect  to  y  of  the  sound  speeds  of  Whitham's  theory. ) 
Since  the  propagation  speed  depends  on  the  height  of  the  disconti- 
nuity surface,  it  is  clear  that  the  possibility  of  motions  leading  to 
breaking  is  inherent  in  this  theory. 

Once  the  dynamical  equations  have  been  formulated  in  character- 
istic form  it  becomes  possible  to  see  rather  easily  what  sort  of  subsi- 
diary initial  and  boundary  conditions  are  reasonable.  In  fact,  there 
are  many  possibilities  in  this  respect.  One  such  possibility  is  the 
following.  At  time  t  —  0  it  is  assumed  that  u  =  const.,  r)  =  0,  h  = 
const,  (as  in  a  stationary  front),  but  that  rjt  =  f(x)  over  a  segment  of 
the  «r-axis.  In  other  words,  it  is  assumed  that  a  transverse  impulse  is 
given  to  the  stationary  front  over  a  portion  of  its  length.  The  sub- 
'sequent  motion  is  uniquely  determined  and  can  be  calculated  nume- 
rically. Another  possibility  is  to  prescribe  a  stationary  front  at  t  =  0 


LONG   WAVES   IN   SHALLOW   WATER  405 

for  x  >  0,  say,  and  then  to  give  the  values  of  all  dependent  quantities* 
at  x  =  0  as  arbitrary  functions  of  the  time;  i.e.  to  prescribe  a  boundary 
condition  which  allows  energy  to  be  introduced  gradually  into  the 
system.  One  might  visualize  this  case  as  one  in  which,  for  example, 
cold  air  is  being  added  or  withdrawn  at  a  particular  point  (x  =  0  in 
the  present  case).  This  again  yields  a  problem  with  a  uniquely  deter- 
mined solution,  and  various  possibilities  are  being  explored  numeri- 
cally. 

It  was  stated  above  that  the  most  objectionable  feature  of  the 
present  theory  is  the  assumption  of  a  fixed  vertical  barrier  back  of  the 
front.  There  is,  however,  a  different  way  of  looking  at  the  problem  as 
a  whole  which  may  mitigate  this  restriction  somewhat.  One  might  try 
to  consider  the  motion  of  the  entire  cap  of  cold  air  that  lies  over  the 
polar  region,  using  polar  coordinates  (0,  y)  (with  0  the  latitude  angle, 
say).  One  might  then  consider  motions  once  more  that  depend 
essentially  only  on  q>  and  t  by  getting  rid  of  the  dependence  on  6 
through  use  of  the  same  type  of  assumptions  (linear  behavior  in  0, 
say)  as  above.  Here  the  North  Pole  itself  would  take  the  place  of  the 
vertical  barrier  (v  —  0!).  The  result  is  again  a  system  of  nonlinear 
equations— this  time  with  variable  coefficients.  Of  course,  it  would 
be  necessary  to  begin  with  a  stationary  flow  in  which  the  motion  takes 
place  along  the  parallels  of  latitude. 

All  in  all,  the  ideas  presented  here  would  seem  to  yield  theories 
flexible  enough  to  permit  a  good  deal  of  freedom  with  regard  to  initial 
and  other  conditions,  so  that  one  might  hope  to  gain  some  insight  into 
the  complicated  dynamics  of  frontal  motions  by  carrying  out  numeri- 
cal solutions  in  well-chosen  special  cases. 

10.12.  Supercritical  steady  flows  in  two  dimensions.  Flow  around  bends. 
Aerodynamic  applications 

The  title  of  this  section  is  a  slight  misnomer,  since  the  flows  in 
question  are  really  three-dimensional  in  nature;  however,  since  we 
consider  them  here  only  in  terms  of  the  shallow  water  theory,  the 
depth  dimension  is  left  out.  Thus  the  velocity  is  characterized  by  the 
two  components  (u,  w)  in  the  horizontal  plane  (the  x,  2-plane);  and 
these  quantities,  together  with  the  depth  h  of  the  water  at  any  point 
constitute  the  quantities  to  be  determined  in  any  given  problem.  By 

*  In  the  numerical  cases  so  far  considered  we  have  had  |  c  \  <  |  u  \  so  that 
even  on  the  /-axis  all  four  dependent  quantities  can  be  prescribed. 


406  WATER   WAVES 

specializing  the  general  equations  (2.4.18),  (2.4.19),  (2.4.20),  of  the 
shallow  water  theory  as  derived  in  Chapter  2  for  the  case  of  a  steady 
flow,  the  differential  equations  relevant  for  this  section  result.  They 
can  also  be  derived  readily  from  first  principles,  as  follows:  Assuming 
that  the  hydrostatic  pressure  law  holds  and  that  the  fluid  starts  from 
rest  (or  any  other  motion  in  which  the  vertical  component  of  the 
velocity  of  the  water  is  zero)  it  follows  that  the  vertical  component  of 
the  velocity  remains  zero  and  that  u  and  w  are  independent  of  the  ver- 
tical coordinate.  The  law  of  continuity  can  thus  be  readily  derived  for 
a  vertical  column;  for  a  steady  flow  it  is 

(10.12.1)  (hu)x  +  (hw)z  =  0. 

We  assume  that  the  flows  we  study  are  irrotational,  and  hence  that 

(10.12.2)  uz  —  wx  =  0. 

The  Bernoulli  law  then  holds  and  can  be  written  in  the  form 

(10.12.3)  (u2  +  w2)  +  2gh  =  const. 

In  these  equations  h  is  the  depth  of  the  water  at  any  point.  By  using 

(10.12.3)  to  express  h  in  terms  of  u  and  w,  and  introducing  the  quan- 
tity c  by  the  relation 

(10.12.4)  c2  =  gh 
we  obtain  the  equation 

(10.12.5)  (c2  -  u2)ux  —  uw(wx  +  uz)  +  (c2  —  w2)wz  =-  0, 

and  this  equation  together  with  (10.12.2),  with  c  defined  in  terms  of 
u  and  w  through  (10.12.4)  and  (10.12.3),  constitute  a  pair  of  first 
order  partial  differential  equations  for  the  determination  of  u(x,  z) 
and  w(x,  z). 

The  theory  of  these  latter  equations  can  be  developed,  as  in  the 
cases  treated  previously  in  this  chapter,  by  using  the  method  of 
characteristics,  provided  that  the  quantity  c  remains  always  less  than 
the  flow  speed  everywhere,  i.e.  provided  that 

(10.12.6)  c2  <  u2  +  w2. 

The  flow  is  then  said  to  be  supercritical.  (In  hydraulics  the  contrast 
subcritical— supercritical  is  commonly  expressed  as  tranquil-shoot- 
ing.) Only  then  do  real  characteristics  exist.  We  shall  not  develop 
this  theory  here,  but  rather  indicate  some  of  the  problems  which  have 
been  treated  by  using  the  theory.  Complete  expositions  of  the  char- 


LONG    WAVES    IN    SHALLOW    WATER 


407 


acteristic  theory  can  be  found  in  the  paper  by  Preiswerk  [P.16],  and 
in  Chapter  IV  of  the  book  by  Courant  and  Friedrichs  [C.9].  The  theory 
is,  of  course,  again  perfectly  analogous  to  the  theory  of  steady  two- 
dimensional  supersonic  flows  in  gas  dynamics. 


Fig.  10.12.1.  Hydraulic  jump 

We  have  already  encountered  an  interesting  example  of  a  flow 
which  is  in  part  supercritical,  in  part  subcritical,  i.e.  the  case  of  a 
hydraulic  jump  in  which  the  character  of  the  flow  changes  on  passage 
through  the  discontinuity.  Figure  10.12.1  is  a  photograph,  taken 
from  the  paper  of  Preiswerk,  of  such  a  hydraulic  jump.  Figure  10.12.2, 
also  taken  from  the  paper  of  Preiswerk,  shows  a  more  complicated 
case  in  which  hydraulic  jumps  occur  at  oblique  angles  to  the  direction 
of  the  flow.  The  picture  shows  a  flow  through  a  sluice  in  a  dam,  with 
conditions  (i.e.  depth  differences  above  and  below  the  dam)  such  that 
supercritical  flow  develops  in  the  sluice,  and  changes  in  level  take  place 
so  abruptly  that  they  might  well  be  treated  as  discontinuities  (as  was 
done  in  earlier  sections  in  the  treatment  of  bores).  The  two  disconti- 
nuities at  the  sides  of  the  sluice  (marked  1  and  2  in  the  figure)  are 
turned  toward  each  other  and  eventually  intersect  to  form  a  still 


408 


WATER   WAVES 


higher  one  (marked  1+2).  Such  oblique  discontinuities  can  be 
treated  mathematically;  the  details  can  be  found  in  the  works  cited 
above. 

Another  interesting  problem  of  the  category  considered  here  is  the 


1+2 


Fig.  10.12.2.  Hydraulic  jumps  at  oblique  angles  to  the  direction  of  the  flow 

problem  of  supercritical  flow  around  a  bend  in  a  stream.  This  type  of 
problem  is  relevant  not  only  for  flows  in  water,  but  also  for  certain 
flows  in  the  atmosphere  (for  which  see  Freeman  [F.9]).  It  is  possible 
in  these  cases  to  have  flows  of  the  type  which  are  mathematically  of 
the  kind  called  simple  waves  in  earlier  sections.  This  means  that  one 
of  the  families  of  characteristics  is  a  set  of  straight  lines  along  each  of 
which  wand  w  (hence  also  h)  are  constant.  Even  the  notion  of  a  cen- 
tered simple  wave  can  be  realized  in  these  cases.  Suppose  that  the 
flow  comes  with  constant  supercritical  velocity  along  a  straight  wall 


LONG    WAVES    IN   SHALLOW   WATER  409 

(cf.  Fig.  10.12.3)  until  a  smooth  bend  begins  at  point  A.  The  straight 
characteristics  are  denoted  by  C+  in  the  figure;  they  form  a  set  of 


Fig.  10.12.3.  Supercritical  flow  around  a  smooth  bend 

parallel  lines  in  the  region  of  constant  flow,  which  then  terminates 
along  the  C+  characteristic  through  the  point  A,  where  the  bend  be- 
gins; beyond  that  characteristic  a  variable  regime  begins.  The  straight 
characteristics  themselves  are  called  Mach  lines;  they  have  physical 
significance  and  would  be  visible  to  the  eye:  the  Mach  lines  are  lines 
along  which  infinitesimal  disturbances  of  a  supercritical  flow  are 
propagated;  in  an  actual  flow  they  would  be  made  visible  because  of 
the  existence  of  small  irregularities  on  the  surface  of  the  wall  of 
the  bend.  If  the  bend  contracts  into  a  sharp  corner,  the  straight 
characteristics,  or  Mach  lines,  which  lie  in  the  region  in  which  the 
flow  is  variable,  all  emanate  from  the  corner,  as  indicated  in  Fig. 
10.12.4;  the  flow  as  a  whole  consists  of  two  different  uniform  flows 


Fig.  10.12.4.  Supercritical  flow  around  a  sharp  corner 

connected  through  a  centered  simple  wave.  If  the  bend  in  the  stream 
is  concave  toward  the  flow,  rather  than  convex  as  in  the  preceding 
two  cases,  the  circumstances  are  quite  different,  since  the  Mach  lines 
would  now  converge,  rather  than  diverge,  in  certain  portions  of  the 
flow,  as  indicated  in  Fig.  10.12.5.  Overlapping  of  the  characteristics 
would  mean  mathematically  that  the  depth  and  velocity  would  be 
multi- valued  at  some  points  in  the  flow;  this  being  physically  impos- 


410  WATER   WAVES 

sible  it  is  to  be  expected  that  something  new  happens  and,  in  fact, 
the  development  of  a  hydraulic  jump  is  to  be  expected.  If  the  bend 


Fig.  10.12.5.  Mach  lines  for  a  supercritical  flow  around  a  concave  bend 

is  a  sharp  angle,  as  in  Fig.  10.12.6,  the  configuration  consisting  of  two 
uniform  flows  parallel  to  the  walls  of  the  bend  and  connected  by  an 
oblique  hydraulic  jump  is  mathematically  possible,  and  it  occurs  in 
practice. 

Having  considered  flows  delimited  on  one  side  only  by  a  wall,  it  is 
natural  to  consider  next  flows  between  two  walls  as  in  a  sluice  or 
channel  of  variable  breadth.  (Such  flows  are  analogous  to  two-dimen- 


Fig.  10.12.6.  Oblique  hydraulic  jump 

sional  steady  flows  through  nozzles  in  gas  dynamics.)  The  possibilities 
here  are  very  numerous,  and  most  of  them  lead  to  cases  not  describ- 
able  solely  in  terms  of  simple  waves.  They  are  of  considerable  import- 
ance in  practice.  For  example,  v.  Karman  [K.2]  was  led  to  the  study 
of  particular  flows  of  this  type  because  of  their  occurrence  in  bends  in 
the  concrete  spillways  designed  to  carry  the  flows  of  the  Los  Angeles 


LONG    WAVES   IN    SHALLOW   WATER 


411 


river  basin  through  the  city  of  Los  Angeles;  the  seasonal  rainfall  is  so 
heavy  and  the  terrain  so  steep  that  supercritical  flows  are  the  rule 
rather  than  the  exception  during  the  rainy  season.  Experiments  for 
sluices  of  special  form  were  carried  out  by  Preiswerk;  Fig.  10.12.7, 


b) 


gemessen    bei    ha=31,1mm 

Fig.  10.12.7.  Laval  nozzle  a)  Mach  lines  b)  contour  lines  of  the  water  surface 

for  example,  shows  the  result  of  an  experiment  in  a  particular  case. 
The  upper  figure  shows  the  Mach  lines,  the  lower  figure  shows  the 
contour  lines  of  the  water  surface  as  given  by  the  theory  as  well  as  by 
experiment;  as  one  sees,  the  agreement  is  quite  good. 

Finally,  we  discuss  briefly  some  applications  of  interest  because  of 
their  connection  with  aerodynamics.  Because  of  the  analogy  of  the 
shallow  water  theory  with  compressible  gas  dynamics,  it  is  of  course 
possible  to  interpret  experiments  with  flows  in  shallow  water  in  terms 
of  the  analogous  flows  in  gases.  Since  it  is  much  cheaper  and  simpler 


412  WATER   WAVES 

to  obtain  supercritical  flows  experimentally  in  water  than  it  is  to 
obtain  supersonic  flows  in  gases,  it  follows  that  "water  table"  experi- 
ments (as  they  are  sometimes  called)  may  have  considerable  import- 
ance for  those  whose  principle  interest  is  in  aerodynamics.  There  is  a 
considerable  literature  devoted  to  this  subject;  we  mention,  for  exam- 
ple, papers  by  Crossley  [C.12],  Einstein  and  Baird  [E.5],  Harleman 


Fig.  10.12.8.  Photogram  of  hydraulic- jump  intersection 

[H.8],  Laitone  [L.I],  Bruman  [B.19].  Figure  10.12.8  is  a  photograph, 
taken  from  the  paper  by  Crossley,  showing  the  interaction  of  two 
hydraulic  jumps;  this  is  a  case  essentially  the  same  as  that  shown  by 
Fig.  10.12.2.  The  ripples  with  short  wave  lengths  constitute  an  effect 
due  to  surface  tension,  and  the  discontinuities  are  smoothed  out  so  that 
a  hydraulic  jump  does  not  really  occur;  the  changes  in  depth  are  quite 
abrupt,  however.  Another  important  case  that  has  been  studied  by 
means  of  the  hydraulic  analogy  is,  as  a  matter  of  course,  the  flow 


LONG    WAVES    IN   SHALLOW   WATER  413 

pattern  which  results  when  a  rigid  body  (simulating  a  projectile  or  an 
airfoil)  is  immersed  in  a  stream.  Figure  10.12.9  shows  a  photograph  of 


Fig.  10.12.9.  Shock  wave  in  front  of  a  projectile 


Fig.  10.12.10.  Flow  pattern  of  a  projectile 

such  a  flow  (taken  from  the  paper  by  Laitone).  The  shock  wave  in 
front  of  the  projectile  is  well  shown.  Figure  10.12.10  is  another  photo- 
graph made  by  Preiswerk;  here,  Mach  lines  are  clearly  visible. 


414 


WATER   WAVES 


10.13.  Linear  shallow  water  theory.  Tides.  Seiches.  Oscillations  in 
harbors.  Floating  breakwaters 

Up  to  now  in  this  chapter  we  have  considered  problems  of  wave 
motion  in  water  sufficiently  shallow  to  permit  of  an  approximation  in 
terms  of  what  we  call  the  shallow  water  theory.  This  theory  is  non- 
linear in  character,  and  consequently  presents  difficulties  which  are 
often  quite  formidable.  By  making  the  assumption  that  the  wave 
amplitudes  in  the  motions  under  study  are  small  in  addition  to  the 
assumption  that  the  water  is  shallow,  it  is  possible  to  obtain  a  theory 
which  is  linear— and  thus  attackable  by  many  known  methods  — 
and  which  is  also  applicable  with  good  approximation  in  a  variety 
of  interesting  physical  situations.  We  begin  by  deriving  the  linear 
shallow  water  theory  under  conditions  sufficiently  general  to  permit 
us  to  discuss  the  cases  indicated  in  the  heading  of  this  section.  (A 
brief  mention  of  the  linear  shallow  water  theory  was  made  in  Chapter 
2  and  in  Chapter  10.1.) 

The  linear  shallow  water  theory  could  of  course  be  derived  by 
appropriate  linearizations  of  the  nonlinear  shallow  water  theory.  It  is, 
however,  more  convenient— and  perhaps  also  interesting  from  the 
standpoint  of  method— to  proceed  by  linearizing  first  the  basic  general 
theory  as  developed  in  Chapter  1,  and  afterwards  making  the  ap- 
proximations arising  from  the  assumption  that  the  water  is  shallow. 
In  other  words,  we  shall  begin  with  the  exact  linear  theory  of  Chapter 
2.1,  and  proceed  to  derive  the  linear  shallow  water  theory  from  it.  One 
of  the  advantages  of  this  procedure  is  that  the  error  terms  involved 
in  the  shallow  water  approximation  can  be  exhibited  explicitly. 

We  suppose  the  water  to  fill  a  region  lying  above  a  fixed  surface 


Fig.  10.13.1.  Linear  shallow  water  theory 

(the  bottom)  y  =  —  h(z,  z)9  and  beneath  a  surface  y  —  Y(x9  z\  t), 
the  motion  of  which  is  for  the  time  being  supposed  known  (cf.  Fig. 


LONG    WAVES    IN    SHALLOW    WATER  415 

10.13.1).  The  t/-axis  is  taken  vertically  upward,  and  the  x,  s-plane  is 
horizontal.  The  upper  surface  of  the  water  given  by  y  =  Y(x,  z;  t) 
will  consist  partly  of  the  free  surface  (to  be  determined,  for  example, 
by  the  condition  that  the  pressure  vanishes  there)  and  partly  of  the 
surfaces  of  immersed  bodies;  it  is,  however,  not  necessary  to  specify 
more  about  this  surface  for  the  present  than  that  it  should  represent 
always  a  small  displacement  fro?n  a  rest  position  of  equilibrium  of  the 
combined  system  consisting  of  water  and  immersed  bodies. 

We  recapitulate  the  equations  of  the  exact  linear  theory  as  derived 
in  Chapter  2.1.  The  velocity  components  are  determined  as  the  deri- 
vatives of  the  velocity  potential  0(x,  y.  z;  t)  which  satisfies  the  Laplace 
equation 

(10.13.1)  0xx+0yv+0zz==Q 

in  the  space  filled  by  the  water.  It  is  legitimate  to  assume  that  all 
boundary  conditions  at  the  upper  surface  of  the  water  are  to  be  satis- 
fied at  the  equilibrium  position;  this  position  is  supposed  given  by 

(10.13.2)  y^fj(x,z). 

(The  bar  over  the  quantity  rj  points  to  the  fact  that  fj  could  also  be 
interpreted  as  the  average  position  of  the  water  in  the  important 
special  case  in  which  the  motion  is  a  simple  harmonic  motion  in  the 
time.)  The  x,  z-planc  is  taken  in  the  undisturbed  position  of  the  free 
surface,  and  this  in  turn  means  that  fj  in  (10.13.2)  has  the  value  zero 
there.  Under  any  immersed  bodies  the  value  of  fj  will  be  fixed  by  the 
static  equilibrium  position  of  the  given  bodies.  Thus  fj  is  in  all  cases 
a  given  function  of  x  and  z;  for  a  floating  rigid  body,  for  example,  it 
would  be  determined  by  hydrostatics. 

The  condition  to  be  satisfied  at  the  upper  surface  is  the  kinematic 
condition : 

(10.13.3)  0XYX  +  0ZYZ  -  0y  +  Yt  =  0, 

which  states  that  a  particle  once  on  the  surface  remains  on  it.  At  the 
bottom  surface,  the  condition  to  be  satisfied  is 

(10.13.3)!      0xhx  +  0zhz  +  0V  =  0        at  y  =  -  h(x,  z). 
Bernoulli's  law  for  determining  the  pressure  at  any  point  in  the  water  is 

(10.13.4)  -  +  0t  +  gy  -  W  =  0. 

Q 

Here  we  have  assumed  that  there  may  be  other  external  forces  beside 


416  WATER   WAVES 

gravity,  and  these  forces  are  assumed  to  be  determined  by  a  work 
function  W(x9  j/,  z;  t)  whose  space  derivatives  furnish  the  force  com- 
ponents; in  this  case  it  is  known  that  the  motion  can  be  irrotational 
and  that  Bernoulli's  law  holds  in  the  above  form  (cf.  the  derivations 
in  Chapter  1  ).  We  now  write  the  equation  of  the  moving  upper  surface 
in  the  form 

(10.13.5)  y  =  Y(x,  z;  t)  =  ij(x9  z)  +  *?(#,  z;  t) 

and  assume  in  accordance  with  our  statement  above  that  r)(x,  z;  t)  re- 
presents a  small  vertical  displacement  from  the  equilibrium  position 
given  by  y  =  q.  Upon  insertion  in  (10.13.3)  and  (10.13.4)  we  find 
after  ignoring  quadratic  terms  in  r/  and  0  and  their  derivatives: 

(10.13.6)  0xfjx  +  0efjz  -  0y  +  r,t  =  0 

at  y  =  fj(x,  z) 
(10.13.7) 


as  boundary  conditions  to  be  satisfied  at  the  equilibrium  position  of 
the  upper  surface  of  the  water.  At  points  corresponding  to  a  free  sur- 
face where  p  =  0  we  would  have,  for  example,  fj  =  0  and  hence 

(10.13.8)  _0y+^==o 


(10.13.9)  ' 

A  special  case  might  be  that  in  which  the  motion  of  a  portion  of  the 
upper  surface  is  prescribed,  i.e.  rj(x,  z;  t)  as  well  as  fj  would  be  pre- 
sumed known;  in  such  a  case  the  condition  (10.13.6)  alone  would 
suffice  as  a  boundary  condition  for  the  harmonic  function  0.  In  some 
of  the  problems  to  be  treated  here,  however,  we  do  not  wish  to  assume 
that  the  motion  of  some  immersed  body,  for  example,  is  known  in 
advance;  rather,  it  is  to  be  determined  by  the  interaction  with  the 
water  which  exerts  a  pressure  p(x,  z;  t)  on  it  in  accordance  with  (10. 
13.7).  Thus  the  exact  formulation  of  our  problems  would  require 
the  determination  of  a  harmonic  function  0(x,  t/,  z;  t)  in  the  space 
between  y  =  —  h(x,  z)  and  y  =  fj  which  satisfies  the  conditions  (10. 
13.6)  and  (10.13.7)  at  the  upper  surface  (in  particular  the  conditions 
(10.13.8)  and  (10.13.9)  on  the  free  surface)  and  (10.13.3)!  at  the 
bottom.  Additional  conditions  where  immersed  bodies  occur  (to  be 
obtained  from  the  appropriate  dynamical  conditions  for  such  bodies) 
would  be  necessary  to  determine  the  pressure  p,  which  provides  the 
"coupling"  between  the  water  on  the  one  hand  and  the  immersed 
bodies  on  the  other.  Finally,  appropriate  initial  conditions  for  the 


LONG   WAVES   IN    SHALLOW   WATER  417 

water  and  the  immersed  bodies  at  the  initial  instant  would  be  needed 
if  one  were  to  study  non-steady  motions,  or—  as  will  be  the  case  here 
—conditions  at  oo  of  the  radiation  type  would  be  needed  if  simple 
harmonic  motions  (that  is,  steady  vibrations  instead  of  transients) 
are  studied.  It  need  hardly  be  said  that  the  difficulties  of  carrying 
out  the  solutions  of  such  problems  are  very  great  indeed  (cf.  Chapter 
9,  for  example)—  so  much  so  that  we  turn  to  an  approximate  theory 
which  is  based  on  the  assumption  that  the  depth  of  the  water  is 
sufficiently  small  and  that  the  immersed  bodies  are  rather  flat.* 

In  the  derivation  of  the  shallow  water  theory  we  start  from  the 
Laplace  equation  (10.13.1)  for  0  and  integrate  it  with  respect  to  y 
from  the  bottom  to  the  equilibrium  position**  of  the  top  surface 
y  =  fj(x,  z)  to  obtain,  after  integration  by  parts: 

(10.13.10)       "_&yvdy  =  0y  -  0y  =  -         (0XX  +  0zz)dy 


Here,  and  in  what  follows,  a  bar  over  the  quantity  0  means  that  it  is 
to  be  evaluated  at  the  equilibrium  position  of  the  upper  surface  of  the 
water,  i.e.  for  y  =  ?/(#,  2),  and  a  bar  under  the  quantity  means  that  it 
is  to  be  evaluated  at  the  bottom  y  =  —  h(x,  z).  From  the  kinematic 
surface  condition  (10.13.6)  and  the  condition  (10.13.3  )t  at  the  bottom, 
we  have  therefore  (due  regard  being  paid  to  the  fact  that  a  bar  should 
now  be  put  over  0  in  (10.13.6)  and  under  0  in  (10.13.3)!): 

(10.13.11)  nt=- 

This  condition—  really  a  continuity  condition—  expresses  the  fact 
that  the  water  is  incompressible.  Consider  next  the  result  of  integrat- 
ing by  parts  the  right  hand  side  of  (10.13.11);  in  particular: 


(10.13.12)  0X  dy  =  fj0x  +  h0x  -         y0xy  dy. 

Since  we  have 

*  In  the  course  of  the  derivation  the  terms  neglected  are  given  explicitly  so 
that  a  precise  statement  about  them  can  be  made. 

**  One  sees  readily  that  carrying  out  the  integration  to  y  ~  7]  rather  than 
to  y  =  rj  +  //  yields  the  same  results  within  terms  of  second  order  in  small 
quantities. 


418  WATER  WAVES 


(10.13.18)  h$xv  dy  =  h$x  -  h&x 

J  —h 

we  may  eliminate  0X  from  (10.13.12)  to  obtain: 
(10.13.14)  0X  dy=(rj  +  h)0x  -         (h  +  y)0xy  dy. 


Indeed,  we  have  quite  generally  for  any  function  F(x,  y,z;t)  the 
formula: 

(10.13.14)!     (*    Fdy=(fj  +  h)F  -  P    (h  +  y)Fy  dy. 

J  —/i  J  —n 

Making  use  of  the  analogous  expression  for  the  integral  of  0Z  we 
obtain  from  (10.13.11)  the  relation 

(10.13.15)  r,t  =  -  [(fj  +  h)0x]x  -  [(fj  +  h)0z]z  +  IX  +  Jz 
in  which 

(10.13.16)  /  =    *    (h  +  y)0xy  dy,     J  =    "    (h  +  y)0zy  dy. 


In  addition,  we  have  from  (10.13.10)  in  combination  with  (10.13.14) 
the  condition: 

(10.13.17)  0y  =  -  (fj  +  h)[0xx  +  0ZZ]   -   hx0x  -  hz0z  +IX+  J29 

as  one  can  readily  verify. 

Up  to  this  point  we  have  made  no  approximations  other  than  those 
arising  from  linearizing.  The  essential  step  in  obtaining  our  approxi- 
mate theory  is  now  taken  in  neglecting  the  terms  Ix  and  Jz.  This  in 
turn  is  justified  if  it  is  assumed  that  certain  second  and  third  deriva- 
tives of  0  are  bounded  when  h  is  small  and  that  77  and  its  first  deriva- 
tives are  small*  of  the  same  order  as  h:  one  sees  that  the  terms  Ix 
and  Jz  in  the  right  hand  sides  of  (10.13.15)  and  (10.13.17)  are  then  of 
order  h2  while  the  remaining  terms  are  of  order  h.  Under  the  free  sur- 
face in  the  case  of  a  simple  harmonic  oscillation  one  can  show  that  this 
approximation  requires  the  depth  to  be  small  in  comparison  with  the 
wave  length. 

Upon  differentiating  the  relation  (10.13.7)  for  the  pressure  at  the 
upper  surface  of  the  water  with  respect  to  t  (again  noting  that  a  bar 
should  be  placed  over  the  term  0t  in  (10.13.7))  and  using  (10.13.15) 
we  find  the  equation 

*  This  means  that  the  theory  developed  here  applies  to  immersed  bodies 
which  are  flat. 


LONG   WAVES    IN    SHALLOW   WATER  419 


(10.13.18)  &tt+L-wt=[(jj+  h)0x]x  + 

after  dropping  the  terms  Ix  and  Jz.  This  is  the  basic  differential  equa- 
tion for  the  function  0(x,  z;  t)  which  holds  everywhere  on  the  upper 
surface  of  the  water.  In  particular,  we  have  at  the  free  surface  where 
p  —  0  and  fj  ~  0  the  equation 

(10.13.19)  (h$x)x  +  (A0,),  -  -$tt  =  -  Wt. 

o  o 

We  recall  that  W(x,  y,  z;  t)  represents  the  work  function  for  any 
external  forces  in  addition  to  gravity  (tide  generating  forces,  for 
example),  so  that  W9  its  value  on  the  free  surface,  would  be  given  by 
W(x,  0,  2;  t).  If,  in  addition,  it  is  assumed  that  h  is  a  constant,  i.e.  that 
the  depth  of  the  water  is  uniform,  and  that  gravity  is  the  only  external 
force,  we  would  have  the  equation 

(10.13.19)!  &xx  +  $„-  —  $it  =  0, 

that  is,  the  linear  wave  equation  in  the  two  space  variables  #,  z  and 
the  time  /.  As  a  consequence,  all  disturbances  propagate  in  such  a 
case  with  the  constant  speed  c  =  Vgh,  as  is  well  known  for  this 
equation. 

If  there  is  an  immersed  object  in  the  water,  the  equation  (10.13.19) 
holds  everywhere  in  the  x,  z-plane  exterior  to  the  curve  C  which 
defines  the  water  line  on  the  immersed  body  in  its  equilibrium  posi- 
tion. The  curve  C  is  supposed  given  by  the  equations 

(10.13.20)  x  =  x(s\         z  =  z(s) 

in  terms  of  a  parameter  s.  We  must  have  boundary,  or  perhaps  it  is 
better  to  say,  transition  conditions  at  the  curve  C  which  connect  the 
solutions  of  (10.13.19)  in  the  exterior  of  C  in  an  appropriate  manner 
with  the  motion  of  the  water  under  the  immersed  body.  Reasonable 
conditions  for  this  purpose  can  be  obtained  from  the  laws  of  conser- 
vation of  mass  and  energy.  In  deriving  these  conditions  we  assume 
W  =  0,  since  we  wish  to  deal  only  with  gravity  as  the  external  force 
when  considering  problems  involving  immersed  bodies.  Consider  an 
element  of  length  ds  of  the  curve  C  representing  the  water  line  of  the 
immersed  body  (cf.  Fig.  10.13.2).  The  expression 


420 


WATER   WAVES 


represents  the  mass  flux  through  a  vertical  strip  having  the  normal  n 


C(x(s),y(s)) 


Fig.  10.18.2.  Boundary  at  water  line  of  an  immersed  body 

and  extending  from  the  bottom  to  the  top  of  the  water.  From  (10.13. 
x  applied  for  F  =  0n  we  have 


(10.13.21  )     e        0n  dy  =  e(ij  +  h)0n  -  Q         (h  +  y)0nv  dy 

^  Q(fj  +  h)0n 

where  the  second  term  is  ignored  because  it  is  of  order  h2.  Thus  it 
would  be  reasonable  to  require  that  (77  +  h)0n  should  be  continuous 
on  C  since  this  is  the  same  as  requiring  that  the  mass  of  the  water  is  con- 
served within  terms  of  the  order  retained  otherwise  in  our  theory.  For 
the  flux  of  energy  across  a  vertical  strip  with  the  normal  n  we  have 

(10.13.22)     (J^  p0n  dy  )  ds  =  (-  Q  J^0«#n  dy  -  gQ  J^  y0n  dy)  ds 

upon  making  use  of  the  Bernoulli  law  (10.13.4)  for  the  pressure  p 
(when  W  =  0).  Once  more  we  may  ignore  the  second  term  in  the 
brackets  since  it  is  of  order  h2.  Upon  applying  (10.13.14)!  with 
F  =  0t0n  and  again  ignoring  a  term  of  order  h2  we  find 


(10.13.23)       j  F    p0n  dy)ds  =  {- 

\  J  —n  i 


0t}  ds. 


Since  we  have  already  required  that  (fj  +  h)0n  should  be  continuous, 
we  see  that  the  additional  requirement,  0t  continuous,  ensures  the 
continuity  of  the  energy  flux. 

As  reasonable  transition  conditions  on  the  curve  C  delimiting  the 
immersed  body  at  its  water  line  we  have  therefore 

(10.13.24)  (fj  +  h)0n,         0t  continuous  on  C. 


LONG    WAVES    IN    SHALLOW   WATER  421 

Of  course,  if  fj  is  continuous  (e.g.  if  the  sides  of  the  body  in  contact 
with  the  water  do  not  extend  vertically  below  the  undisturbed  free 
surface)  it  follows  that  <pn  would  then  be  continuous. 

In  order  to  make  further  progress  it  would  be  necessary  to  specify 
the  properties  of  the  immersed  body.  However,  we  have  succeeded  in 
obtaining  the  equation  (10.13.18)  which  is  generally  valid  and  of  basic 
importance  for  our  theory  together  with  the  transition  conditions 
(10.13.24)  valid  at  the  edge  of  immersed  bodies;  in  particular,  we  have 
the  definitive  equation  for  the  free  surface  itself  in  the  form  of  the 
linear  wave  equation  (10.13.19).  The  idea  behind  this  method  of  ap- 
proximation is  to  get  rid  of  the  depth  variable  by  an  integration  over 
the  depth  so  that  the  problems  then  arc  considered  only  in  the  x,  z- 
plane.  As  a  result,  the  problems  are  no  longer  problems  in  potential 
theory  in  three  space  variables,  but  rather  problems  involving  the  wave 
cq  nation  with  only  two  space  variables,  and  hence  they  are  more  open 
to  attack  by  known  methods.  How  this  comes  about  will  be  seen  in 
special  cases  in  the  following. 

As  a  first  example  of  the  application  of  the  above  theory  we  con- 
sider briefly  the  problem  of  the  tides  in  the  oceans,  with  a  view  to 
indicating  where  this  theory  fits  into  the  theory  of  gravity  waves  in 
general,  but  not  with  the  purpose  of  giving  a  detailed  exposition.  (For 
details,  the  long  Chapter  VIII  in  Lamb  [L.3]  should  be  consulted.) 
To  begin  with,  it  might  seem  incredible  at  first  sight  that  the  shallow 
water  theory  could  possibly  be  accurate  for  the  oceans,  since  depths  of 
five  miles  or  more  occur.  However,  it  is  the  depth  in  relation  to  the 
wave  length  of  the  motions  under  consideration  which  is  relevant. 
The  tides  are  forced  oscillations  caused  by  the  tide-generating  at- 
tractions of  the  moon  and  the  sun,  and  hence  have  the  same  periods 
as  the  motions  of  the  sun  and  moon  relative  to  the  earth.  These  periods 
are  measured  in  hours,  and  consequently  the  tidal  motions  in  the 
water  result  in  waves  having  wave  lengths  of  hundreds  of  miles;* 
the  depth-wave  length  ratio  is  thus  quite  small  and  the  shallow 
water  theory  should  be  amply  accurate  to  describe  the  tides.  This 
means,  in  effect,  that  the  differential  equation  (10.13.19),  or  rather,  its 
analogue  for  the  case  of  water  lying  on  a  rotating  spheroid  (with  Cori- 
olis  terms  put  in  if  a  coordinate  system  rotating  with  the  earth  is 

*  For  example,  in  water  of  depth  10,000  feet  (perhaps  a  fairly  reasonable 
average  value  for  the  depth  of  the  oceans)  a  steady  progressing  wave  having 
a  length  of  10,000  feet  has  a  period  of  only  44.2  sec.  (cf.  Lamb  [L.3],  p.  369). 
Since  the  wave  length  varies  as  the  square  of  the  period,  the  correctness  of  our 
statement  is  obvious. 


422  WATER   WAVES 

used),  would  serve  as  a  basis  for  calculating  tidal  motions.  Of  course, 
the  function  Wt  would  be  defined  in  terms  of  the  gravitational  forces 
due  to  the  attraction  of  the  sun  and  moon.  The  variable  depth  of  the 
water  in  the  oceans  would  come  into  play,  as  well  as  boundary  condi- 
tions at  the  shore  lines  of  the  continents.  Presumably,  Wt  would  be 
analyzed  into  its  harmonic  components  (which  could  be  obtained 
from  astronomical  data),  the  response  to  each  such  harmonic  would 
be  calculated,  and  the  results  superimposed.  Such  a  problem  consti- 
tutes a  linear  vibration  problem  of  classical  type— it  is  essentially  the 
same  as  the  problem  of  transverse  forced  oscillations  of  a  tightly 
stretched  non-uniform  membrane  with  an  irregular  boundary.  If  it 
were  not  for  one  essential  difficulty,  to  be  mentioned  in  a  moment, 
such  a  problem  would  in  all  likelihood  be  solvable  numerically  by 
using  modern  high  speed  computational  equipment.  The  difficulty 
mentioned  was  pointed  out  to  the  author  in  a  conversation  with 
H.  Jeffreys,  and  it  is  that  there  are  difficulties  in  prescribing  an  ap- 
propriate boundary  condition  in  coastal  regions  where  there  is  dissi- 
pation of  energy  in  the  tidal  motions  (in  the  bay  of  Fundy,  for  exam- 
ple, to  take  what  is  probably  an  extreme  case).  At  other  eoastal  re- 
gions the  correct  boundary  condition  would  of  course  often  be  simply 
that  the  component  of  the  velocity  normal  to  the  coast  line  vanishes. 
Of  course,  there  would  also  be  a  difficulty  in  using  a  differential 
equation  like  (10.13.19)  near  any  shores  where  h  =  0,  since  the  dif- 
ferential equation  becomes  singular  at  such  points.  Nevertheless,  a 
computation  of  the  tides  on  a  dynamical  basis  would  seem  to  be  a 
worthwhile  problem— perhaps  it  could  be  managed  in  such  a  way  as 
to  help,  in  conjunction  with  observations  of  the  actual  tides,  in 
providing  information  concerning  the  dissipation  of  energy  in  such 
motions. 

These  remarks  might  be  taken  to  imply  that  the  dynamical  theory 
is  not  at  present  used  to  compute  the  tides.  This  is  not  entirely  correet, 
since  the  tide  tables  for  predicting  the  tides  in  various  parts  of  the 
world  are  based  on  fundamental  consequences  of  the  assumption  that 
the  tides  are  indeed  governed  by  a  differential  equation  of  the  same 
general  type  as  (10.13.19).  The  point  is  that  the  oceans  are  regarded  as 
a  linear  vibrating  system  under  forced  oscillations  due  to  excitation 
from  the  periodic  forces  of  attraction  of  the  sun  and  moon.  It  is 
assumed  that  all  free  vibrations  of  the  oceans  were  long  ago  damped 
out,  and  hence,  as  remarked  above,  that  the  tidal  motions  now  exist- 
ing in  the  oceans  are  a  superposition  of  simple  harmonic  oscillations 


LONG   WAVES    IN    SHALLOW    WATER  423 

having  periods  which  are  very  accurately  known  from  astronomical 
observations.  To  obtain  tide  tables  for  any  given  point  a  superposition 
of  oscillations  of  these  frequencies  is  taken  with  undetermined  ampli- 
tudes and  phases  which  are  then  fixed  by  comparing  them  with  a 
harmonic  analysis  of  actual  tidal  observations  made  at  the  point  in 
question.  The  tide  predictions  are  then  made  by  using  the  result  of 
such  a  calculation  to  prepare  tables  for  future  times.  The  dynamical 
theory  is  thus  used  only  in  a  qualitative  way.  An  interesting  addition- 
al point  might  be  mentioned,  i.e.  that  tides  of  observable  amplitudes 
are  sometimes  measured  which  have  as  frequency  the  sums  or  differ- 
ences (or  also  other  linear  combinations  with  integers)  of  certain  of 
the  astronomical  frequencies,  which  means  from  the  point  of  view  of 
vibration  theory  that  observable  nonlinear  effects  must  be  present. 
Another  type  of  phenomenon  in  nature  which  can  be  treated  by 
the  theory  derived  here  concerns  periodic  motions  of  rather  long 
period,  called  seiches,  which  occur  in  lakes  in  various  parts  of  the 
world.  The  first  observations  of  this  kind  seem  to  have  been  made  by 
Forel  [F.7]  in  the  lake  at  Geneva  in  Switzerland,  in  which  oscillations 
having  a  period  of  the  order  of  an  hour  and  amplitudes  of  up  to  six 
feet  have  been  observed.  In  larger  lakes  still  larger  periods  of  oscilla- 
tion arc  observed— about  fifteen  hours  in  Lake  Erie,  for  example. 
A  rather  destructive  oscillation,  generally  supposed  to  be  of  the  type 
of  a  seiche,  occurred  in  Lake  Michigan  in  June  1054;  a  wave  with  an 
amplitude  of  the  order  of  ten  feet  occurred  and  swept  away  a  number 
of  people  who  were  fishing  from  piers  and  breakwaters.  What  the 
mechanism  is  that  gives  rise  to  seiches  in  lakes  has  been  the  object  of 
considerable  discussion,  but  it  seems  rather  clear  that  the  motions 
represent  free  vibrations  of  the  water  in  a  lake  which  are  excited  by 
external  forces  of  an  impulsive  character,  the  most  likely  type  arising 
from  sudden  differences  in  atmospheric  pressure  over  various  portions 
of  the  water  surface.  Bouasse  [B.I 5,  p.  158]  reports,  however,  that 
the  Lisbon  earthquake  of  1755  caused  oscillations  in  Loch  Lomond 
with  a  period  of  about  5  minutes  and  amplitudes  of  several  feet. 
In  any  case,  the  periods  observed  seem  to  correspond  to  those  calcu- 
lated on  the  basis  of  the  linear  shallow  water  theory,  which  should  be 
quite  accurate  for  the  study  of  seiches  because  of  their  long  periods  and 
small  amplitudes.  It  follows,  therefore,  that  the  differential  equation 
(10.13.18)  is  applicable;  we  suppose  that  Wt  =  0  (since  tidal  forces 
play  no  role  in  this  case),  and  also  set  fj  =  0  since  there  are  no  immersed 
bodies  to  be  considered.  The  differential  equation  for  0(x,  z\  t )  is  thus 


424  WATER  WAVES 


(10.18.25) 

&  e«r 

The  free  natural  vibrations  of  the  lake  are  investigated  by  setting 
pt  =  0  and  0(x,  z;  t)  =  99(01,  *)£*"  in  (10.13.25)  with  the  result 

a2 
(10.13.26)  (hpx)x  +  (%)2  +  -<p  =  0. 


As  boundary  condition  along  the  shore  of  the  lake  we  would  have 
(10.13.27)  (pn  =  0. 

The  problem  thus  posed  is  one  of  the  classical  eigenvalue  problems  of 
mathematical  physics.  Solutions  <p  other  than  the  trivial  solution 
q>  =  0  of  (10.13.26)  under  the  homogeneous  boundary  condition 
(10.13.27)  are  wanted;  such  solutions  exist  only  for  special  values  of 
the  circular  frequency  a,  and  these  values  yield  the  natural  frequencies 
corresponding  to  the  natural  modes  y(x,  z)  which  are  correlated  with 
them.  In  general,  an  infinite  set  of  such  natural  frequencies  occurs. 
For  particular  shapes  and  depths  h—  rectangular  or  circular  lakes  of 
constant  depth,  for  example  —it  is  possible  to  solve  such  problems 
more  or  less  explicitly.  In  practice  however,  lakes  have  such  irregular 
outlines  and  depths  that  the  determination  of  the  natural  frequencies 
and  modes  requires  numerical  computation.  A  reasonable  and  gener- 
ally applicable  method  of  carrying  out  such  computations  is  furnished 
here,  as  in  other  instances  in  this  and  the  subsequent  chapter,  by 
the  method  of  finite  differences.*  In  this  method,  the  derivatives  in 
the  differential  equation  and  boundary  conditions  are  replaced  by 
difference  quotients  defined  by  means  of  the  values  of  the  function 
at  the  discrete  points  of  a  net  in  the  domain  of  the  independent  varia- 
bles. The  resulting  finite  equations  are  then  solved  to  yield  approxi- 
mate values  for  the  unknown  function  at  the  net  points.  The  difference 
approximation  will  be  more  accurate  for  a  closer  spacing  of  the  net 
points.  We  proceed  to  illustrate  the  method  for  the  case  of  a  lake  of 
constant  depth  in  the  form  of  a  square  of  length  /  on  each  side,  with 
a  view  to  comparing  the  result  with  the  exact  solution  which  is  easy 
to  write  down  in  this  case.  The  differential  equation  (10.13.26)  can 
be  written  in  the  form 

*  A  different  method  was  used  by  Chrystal  [C.2]  to  calculate  the  periods  of 
the  free  oscillations  of  Loch  Earn;  he  found  good  agreement  with  the  obser- 
vations for  the  first  six  modes  of  oscillation. 


LONG   WAVES   IN   SHALLOW  WATER 

(10.13.26)!         (pxx  +  <pgz  +  m2<p  =  0,        m2  =  a2/gh 


425 


in  this  case.  A  division  of  the  square  in  a  mesh  with  mesh  width 
8  =  J/7  is  taken,  as  indicated  in  Fig.  10.13.3.  In  numbering  the  net 
points  it  has  been  assumed  that  only  modes  of  oscillation  that  are 
symmetrical  with  respect  to  the  center  lines  parallel  to  the  sides  and  to 
to  the  diagonals  are  sought,  which,  however,  is  not  the  case  for  the 


y 

4 

8     S 
7      9 

> 
0    9 

.      .      . 

3 

2 

§      8 
5      6 

9    8 
7    6 

i 

i 

2      3 

4 

i 

i 

?   . 

X 

°t 

Fig.  10.13.3.  Finite  differences  for  a  seiche 

mode  having  the  lowest  frequency.  The  boundary  condition  q>n  =  0  is 
satisfied  approximately  by  supposing  that  the  solution  is  reflected 
over  the  boundaries  in  such  a  way  as  to  yield  values  which  arc  equal  at 
the  mirror  images  in  the  boundaries,  as  is  also  indicated  in  Fig.  10.13.3. 
The  formulas  used  for  approximating  the  derivatives  are  defined  as 
follows  (cf.  Fig.  10.13.3): 


dz 


ffm.  n+l  -"  ^n?,  w-1 


%9m,n  +  9V  n+l   +  Vm.n-l 


Consequently,  the  differential  equation  (10.13.26)1  is  replaced  at  each 
net  point  (m ,  n)  by  the  difference  equation 

(10.13.28)        -4<pmfn+<pm,n+1+<pm^ 

Such  an  equation  is  written  for  each  of  the  net  points  in  Fig.  10.13.3. 
The  results  for  points  1,6,9,  for  example,  are: 


426  WATER   WAVES 


(10.13.29) 


(<5w)Vi  -  0 

6:   —  4<pe  +  g?3  +  <p7  +  9?8  +  <p5  +  (dw)2^  =  0 
9:   —  4<p9  +  2y8  +  <p7  +  <p10  +  (5m)29?9  =  °- 

These  homogeneous  linear  equations  of  course  have  always  the  solu- 
tion <fi :  =  0,  i  =  1,  2,  .  .  .,  10  unless  their  determinant  vanishes,  and 
this  condition  is  a  tenth  degree  equation  in  the  quantity  (dm)2,  the 
smallest  root  of  which  furnishes  an  approximation  to  the  lowest 
frequency.  The  exact  solution  of  the  differential  equation  (10.13.26)! 
which  satisfies  the  boundary  condition  is,  in  the  present  case,  <p  = 
A  cos  (knx/l)  cos  (jnz/l),  with  k  and  /  any  integers,  provided  that 
m2  =  n2(k2  +  j2)/l2.  A  numerical  comparison  of  the  lowest  value  of 
in  for  the  mode  having  double  symmetry— i.e.  the  value  for  k  =  1, 
j  =  1  — with  the  value  computed  from  the  determinantal  equation 
shows  the  approximate  value  of  m  to  be  too  low  by  6.5  %. 

However,  this  mode  corresponds  to  one  of  the  higher  eigenvalues, 
so  that  the  accuracy  of  the  finite  difference  method  is  rather  good. 
The  error  for  the  lowest  mode  is  very  much  smaller,  but  because  of  the 
lack  of  symmetries  the  amount  of  calculation  needed  to  determine 
the  corresponding  frequency  would  be  much  greater  for  the  present 
case.  If  one  were  to  treat  a  long  narrow  lake,  the  calculation  would 
be  simpler.  It  could  also  be  advantageous  to  employ  the  Rayleigh-Ritz 
method.  In  principle,  similar  calculations  could  be  made  in  more 
complicated  cases  (for  many  examples  of  problems  solved  along  these 
lines  see  the  book  by  Southwell  [S.14]). 

Wave  motions  in  harbors  are  often  of  a  type  suitable  for  discussion 
in  terms  of  the  linear  shallow  water  theory:  they  are  indeed  often  of 
the  type  called  seiches  above.  In  these  cases  oscillations  of  the  water 
in  the  harbor  are  also  commonly  excited  by  the  motion  at  the  harbor 
mouth,  which  in  its  turn  is  due,  of  course,  to  wave  motions  generated 
in  the  open  sea.  An  experimental  and  theoretical  investigation  of  such 
waves  in  a  model  has  been  carried  out  by  McNown  [M.7].  The  model 
was  in  the  form  of  a  circle  3.2  meters  in  diameter  with  vertical  walls. 
The  depth  of  the  water  in  this  idealized  harbor  was  16  cm.  An  opening 
of  angle  n/8  radians  in  the  harbor  wall  permitted  a  connection  with  a 
large  tank  in  which  waves  (simulating  the  open  sea)  were  produced. 
Figures  10.13.4  and  10.13.5  are  photographs  of  the  model  (taken  from 
the  paper  by  McNown),  which  also  show  two  specific  cases  of  symme- 
trical oscillations.  The  free  vibrations  again  are  governed  by  equation 
(10.13.26).  (It  might  be  noted  that  McNown  makes  use  of  the  exact 


LONG   WAVES   IN   SHALLOW   WATER 


427 


linear  theory  rather  than  the  shallow  water  theory.  The  only  difference 
is  that  the  relation  between  cr2  and  m  is  a2  =  gm  tanh  mh,  instead  of 
(T2  =  ghm2,  as  given  above:  the  differential  equation  for  the  velocity 
potential  0(x,  y,  z;  t)  in  the  exact  linear  theory  treated  in  Part  I  is 


Fig.  10.13.4.  and  10.13.5.  Waves  in  a  harbor  model 

written  in  the  form  0  =  A  cosh  m(y  +  h)eiat(p(x,  z),  and  (p(x,  z)  then 

satisfies  V2<p  +  m2<p  =  0.)  Solutions  of  the  differential  equation  are 

sought  in  the  form 

(10.13.30)  <p(r,  0)  =  Jn(mr)  cos  nO 

in  polar  coordinates  (r,  0),  under  the  assumption  that  the  port  is 

closed,  i.e.  that  its  boundary  is  the  whole  circle  r  =  R.  As  is  well 


428 


WATER   WAVES 


known,  <p(r,  6)  is  a  solution  of  (10.13.26 )x  only  if  Jn(mr)  is  a  Bessel 
function  of  order  w,  and  since  it  is  reasonable  to  look  only  for  solutions 
that  are  bounded  we  choose  the  Bessel  functions  of  the  first  kind 
which  are  regular  at  the  origin.  The  boundary  condition  requires  that 

(10.13.31)  -^-  =  0        at  r  =  R 

and  this  in  turn  leads  to  the  condition  dJJdr  =  0  for  r  =  R.  For 
each  n  this  transcendental  equation  has  infinitely  many  roots  m^ , 
each  corresponding  to  a  mode  of  oscillation  with  various  nodal  dia- 
meters and  circles,  and  with  a  definite  frequency  which  is  fixed  by 
a2  =  gm^  (or,  more  accurately,  by  a2  —  gm^  tanh  hm(^  ).  Figure 
10,13.6,  obtained  by  McNown,  shows  a  comparison  of  observed  and 
calculated  amplitudes  for  two  modes  of  oscillation;  the  upper  curve 
is  drawn  for  a  motion  having  no  diametral  nodes  and  two  nodal  circles, 
while  the  lower  is  for  a  motion  having  two  nodal  diameters  and  one 


-2 

-3 
-4 


Y_ 


(kr) 


entrance 


o    Observed 
—  Theoretical 


center 


Fig.  10.13.6.  Comparison  of  results  of  experiment  and  theory  for  resonant  move- 
ments in  a  circular  port 


LONG    WAVES    IN   SHALLOW   WATER 


429 


nodal  circle.  The  motions  were  excited  by  making  waves  in  the  tank, 
and  providing  an  opening  for  communication  with  the  harbor,  as 
noted  above.  The  figures  were  drawn  assuming  that  the  amplitudes 
would  agree  at  the  entrance  to  the  harbor— the  experimental  check 


10.18.7.  of  a 


Fig.  10.13.8.  Model  of  a  harbor  with  breakwater 

thus  applies  only  to  the  shapes  of  the  curves.  As  one  sees,  the  ex- 
perimental and  theoretical  values  are  remarkably  close.  The  ampli- 
tudes used  were  large  enough  so  that  nonlinear  effects  were  observed: 
the  troughs  are  flatter  than  the  crests  by  measurable  amounts.  Of 
course,  having  an  opening  in  the  harbor  wall  violates  the  boundary 
condition  assumed,  but  this  effect  apparently  is  slight:  changing  the 
angle  of  the  opening  at  the  harbor  mouth  had  practically  no  effect  on 


430 


WATER   WAVES 


the  waves  produced,  and,  in  addition,  it  was  found  that  very  little 
wave  energy  radiates  outward  through  the  harbor  entrance. 

Problems  of  harbor  design,  involving  construction  of  breakwaters, 
location  of  docks,  etc.  are  commonly  studied  by  constructing  models. 
Figs.  10.13.7  and  10.13.8  show  two  photographs  of  a  model  of  a  har- 
bor,* the  first  before  a  breakwater  was  constructed,  the  second 
afterward.  As  one  sees,  the  breakwater  has  a  quite  noticeable  effect. 
Fig.  10.13.9.  shows  the  same  model,  with  the  waves  approaching  the 
harbor  mouth  at  a  different  angle,  however;  as  one  sees  the  break- 


Fig.  10.13.9.  Model  of  a  harbor  with  breakwater 

water  seems  to  be  on  the  whole  less  effective  when  the  wave  fronts 
are  less  oblique  to  the  breakwater.  The  diamond-shaped  pattern,  due 
to  reflection,  of  the  waves  on  the  sea  side  of  the  breakwater  is  inter- 


Y////////7* 


/////////A 


Fig.  10.13.10.  Floating  plane  slab 

*  These  photographs  were  given  to  the  author  by  the  Hydrodynamics  Labora- 
tory at  California  Institute  of  Technology. 


LONG    WAVES    IN    SHALLOW   WATER  481 

esting.  Model  studies  are  rather  expensive,  and  consequently  it  might 
well  be  reasonable  to  explore  the  possibilities  of  numerical  solution 
of  the  problems,  perhaps  by  using  appropriate  modifications  of  the 
method  of  finite  differences  outlined  above  for  a  simple  case. 

We  turn  next  to  a  discussion  of  the  effect  of  floating  bodies  on 
waves  in  shallow  water,  on  the  basis  of  the  theory  presented  in  this 
section.  Only  two-dimensional  motions  will  be  considered  (so  that  all 
quantities  are  independent  of  the  variable  z).  The  first  case  to  be 
studied  is  that  of  the  motion  of  a  floating  rigid  body  in  the  form  of  a 
thin  plane  slab  (cf.  Fig.  10.13.10)  in  water  of  uniform  depth.  Such 
problems  have  been  treated  by  F.  John  [J.5].  The  ends  of  the  slab 
arc  at  jc  —  i  a.  In  accordance  with  the  theory  presented  above  we 
must  determine  the  surface  value  3>(tT,  t)  and  the  displacement  r](x9  t) 
of  the  board  from  the  differential  equations  (cf.  (10.13.19),  (10.13.15) 
with  rj  =  0,  and  dropping  /^  and  «/2): 

(10.13.32)  #,,-^0,,,         |*|  >a 

(10.13.33)  rjt  =  -  h&xx,         \  *  \  <  a. 

We  have  dropped  the  bar  over  the  quantity  0.  We  have  also  assumed 
that  fj(x)  for  \  x  \  <  a,  the  rest  position  of  equilibrium  of  the  board,  is 
zero;  this  is  an  approximation  that  is  justified  because  we  assume 
that  the  board  is  so  light  that  it  does  not  sink  appreciably  below  the 
water  surface  when  in  equilibrium.  (This  assumption  is  by  no  means 
necessary  —  it  would  not  be  difficult  to  deal  with  the  problem  if  this 
simplifying  assumption  were  not  made.) 

Since  fj  is  zero,  it  follows  (cf.  (10.13.24))  that  the  transition  con- 
ditions at  the  ends  of  the  board  are 

(10.13.34)  0X,  0t  continuous  at  x  —  ±  a. 

We  are  interested  in  the  problem  of  the  effectiveness  of  the  floating 
board  as  a  barrier  to  a  train  of  waves  coming  from  the  right  (x  = 
+  00).  The  equation  (10.13.32)  has  as  its  general  solution 

0(x, 1)  =  F(x  -  ct)  +  G(x  +  ct),         c  =  VgA 

in  terms  of  two  arbitrary  functions  F  and  G  (as  one  can  readily  verify ) 
which  clearly  represent  a  superposition  of  two  progressing  waves 
moving  to  the  right  and  to  the  left,  respectively,  with  the  speed  Vgh. 
It  is  natural,  in  our  present  problem,  to  expect  that  for  x  >  a  there 
would  exist  in  general  both  an  incoming  and  an  outgoing  wave  be- 
cause of  reflection  from  the  barrier,  while  for  x  <  —  a  we  would  pre- 


432  WATER   WAVES 

scribe  only  a  wave  going  outward  (i.e.  to  the  left).  We  shall  see  that 

these  qualitative  requirements  lead  to  a  unique  solution  of  our  problem. 

We  consider  only  simple  harmonic  waves;  it  is  thus  natural  to  write 

(10.13.35)  &(x,  t)  =  (p(x)eiat,         \  x  \  >  a, 

(10.13.36)  ri(x,  t)  =  v(x)eia\         -  a  <  x  <  a, 

with  the  stipulation  that  the  real  part  is  to  be  taken  at  the  end.  (It  is 
necessary  also  to  permit  <p(x)  and  v(x)  to  be  complex-valued  func- 
tions of  the  real  variable  x.)  The  conditions  (10.13.32)  and  (10.13.33) 
now  become 

d2cp      a2 

(10.13.37)  I  +  =  0.         1*1  >a 


a 

(10.13.38)  Tl  +  T^0'          1*1  <«• 

* 


The  equation   (10.13.37)  has  as  general  solution 

(10.13.39)  y(x)  =  Ae~ikx  +  Beik* 
with  k  given  by 

(10.13.40)  k  =  a/Vgh. 
For  0(x,  t)  we  have  therefore 

(10.13.41)  &(x,  t)  =  Ae-Wx-a»  +  Be'l***"*, 

the  first  term  representing  a  progressing  wave  moving  to  the  right, 
the  second  a  wave  moving  to  the  left.  In  our  problems  we  prescribe 
the  incoming  wave  from  the  right,  and  hence  for  cp(x)  we  write 

(10.13.42)  <p(x)  =  Beikx  +  Re~tkx,         x  >  a, 

in  which  B  is  prescribed,  while  .K—  the  amplitude  of  the  reflected 
wave  (more  precisely,  |  R  \  is  its  amplitude)  —  is  to  be  determined. 
At  the  left  we  write 

(10.13.43)  <p(x)  =  Teikx, 

with  T—  the  amplitude  of  the  transmitted  wave  —to  be  determined. 
To  complete  the  formulation  of  the  problem  it  is  necessary  to  con- 
sider the  dynamics  of  our  floating  rigid  body.  We  shall  treat  two  cases: 
a)  the  board  is  held  rigidly  fixed  in  a  horizontal  position,  b)  the  board 
floats  freely  in  the  water. 

a)  Rigidly  Fixed  Board. 

If  the  board  is  rigidly  fixed  we  have  77(0?,  t)  =  0,  and  hence  (cf. 
(10.13.36))  v(x)  =0.  It  follows  from   (10.13.88)  that  <pxx  vanishes 


LONG   WAVES    IN    SHALLOW   WATER 


433 


identically  under  the  board  and  hence  that  <p(x)  is  a  linear  function: 
(10.13.44)  <p(x)  =  yx  +  d,  —  a  <  x  <  +  a. 

Since  0x(z,  t)  furnishes  the  horizontal  velocity  component  of  the 
water,  it  follows  from  (10.13.44)  and  (10.13.35)  that  the  velocity  under 
the  board  is  given  by  yeiat,  i.e.  it  is  constant  everywhere  under  the 
board  at  each  instant—  a  not  unexpected  result. 

We  now  write  down  the  transition  conditions  at  x  =  ±  «  from 
(10.13.34),  making  use  of  (10.13.35)  and  of  (10.13.42)  at  x  =  +  a 
and  (10.13.43)  at  x  =  -  a;  the  result  is: 


Betka  +  Re~lka  =  ya  +  6 
Bezka  -  Re~ika  =  y/ffc 

Te-ika  =   _  ya  _| 
Te-ika  = 


(10.13.45) 


Once  the  real  number  B— which  fixes  the  amplitude  of  the  incoming 
wave  — has  been  prescribed,  these  four  equations  serve  to  fix  the  con- 
stants R,  T,  y,  and  6  and  hence  the  functions  0(x,  t)  and  rj(x,  t).  The 
pressure  under  the  board  can  then  be  determined  (cf.  the  expression 
(10.13.4)  for  Bernoulli's  law)  from 

(10.13.46)  p(x,  /.)  =  -  Q&t  -  —  Qia(p(x)eiat. 

(Observe  that  the  quantity  y  in  (10.13.4)  is  zero  in  the  present  case.) 
In  terms  of  the  dimensionless  parameter 

(10.13.47)  0  =  2a/A, 

the  ratio  of  the  length  of  the  board  to  the  wave  length  A  on  the  free 
surface,  given  by  (cf.  (10.13.41)) 

(10.13.48)  A  =  2^/fc, 
the  solution  of  (10.13.45)  is 

R  = 


(10.13.49) 


Oni  +  1 


_ 
7 


dni 


a      07i-l 
d  =  Beeni. 
The  reflection  and  transmission  coefficients  are  obtained  at  once: 


434 


WATER   WAVES 


(10.13.50) 


Cr  = 


Qn 


Vi  + 


c*              c* 

0.54 

0.85 

0.30 

0.95 

0.157 

0.986 

They  depend  only  upon  the  ratio  0  =•  2a/A,  as  one  would  expect.  They 
also  satisfy  the  relation  C*  +  Cf  —  1,  as  they  should:  this  is  an  ex- 
pression of  the  fact  that  the  incoming  and  outgoing  energies  are  the 
same.  The  following  table  gives  a  few  specific  values  for  these  co- 
efficients: 


0 

0.5 
1.0 
2.0 


Thus  a  fixed  board  whose  length  is  half  the  incoming  wave  length  has 
the  effect  of  reducing  the  amplitude  behind  it  by  about  50  percent  and 
of  reflecting  about  72  percent  of  the  incoming  energy.  One  should, 
however,  remember  that  the  theory  is  only  for  long  waves  in  shallow 
water,  and,  in  addition,  it  seems  likely  that  the  length  of  the  board 
will  also  play  a  role  in  determining  the  accuracy  of  the  approximation. 
This  question  has  been  investigated  by  Wells  [W.10]  by  deriving  the 
shallow  water  theory  in  such  a  way  as  to  include  all  terms  of  third  order 
in  the  depth  h  and  studying  the  magnitude  of  the  neglected  terms  in 
special  cases;  in  particular,  the  present  case  of  a  floating  rigid  body 
is  investigated.  Wells  finds  that  if  A/A  is  small  and  if  a/A  (the  ratio  of 
the  half-length  of  the  board  to  the  wave  length)  is  not  smaller  than  1, 
the  neglected  higher  order  terms  are  indeed  negligible,  but  if  a/A  is  less 
than  1,  the  higher  order  terms  need  not  be  small.  In  other  words, 
floating  obstacles  ought  to  have  lengths  of  the  order  of  the  wave 
length  of  the  incoming  waves  if  the  shallow  water  theory  to  lowest 
order  in  h  is  expected  to  furnish  a  good  approximation. 

It  is  of  interest  to  study  the  pressure  variation  under  the  board. 
This  is  given  in  the  present  case  (cf.  (10.13.46))  by 


(10.13.51) 


(x,  t)  =  — 


d)e 


dot 


the  real  part  only  to  be  taken.  Thus  the  pressure  varies  linearly  in  x> 
but  it  is  a  different  linear  function  at  different  times  since  y  and  d  are 


LONG    WAVES    IN    SHALLOW   WATER 


435 


complex  constants.  The  result  of  taking  the  real  part  of  the  right  hand 
side  of  (10.13.51)  can  be  readily  put  in  the  form 
(10.13.52)  p(x,  t)  =  Pi(x)  cos  at  —  p2(x)  sin  at 

with 

Pi(x)  =  agBfiifa)  sin  r  +  b2(x)  cos  r) 

p2(x)  =  agB(b2(x)  sin  r  —  b^x)  cos  r) 
and 


(10.13.53) 


(10.13.54) 


-i  -5  0  5       */0       i 

Fig.  10.13.11.  Pressure  variations  for  a  stationary  board.  0  =  1,  p  in  poundsl(ft)2 


436  WATER   WAVES 

We  have  assumed  in  making  these  calculations,  as  stated  above,  that 
B.  which  represents  the  amplitude  of  the  incoming  wave,  is  a  real 
number. 

In  Fig.  10.13.11  the  results  of  computations  for  the  pressure  distri- 
bution for  time  intervals  of  1/4  cycle  over  the  full  period  are  given  for 
a  special  numerical  case  in  which  the  parameter  6  has  the  value  0  =  1, 
i.e.  the  length  of  the  board  is  the  same  as  the  wave  length.  One  ob- 
serves that  the  pressure  variation  is  greater  at  the  right  end  than  at 
the  left,  which  is  not  surprising  since  the  board  has  a  damping  effect  on 
the  waves.  One  observes  also  that  the  pressure  is  sometimes  less  than 
atmospheric  (i.e.  it  is  negative  at  times,  while  p  —  0  is  the  assumed 
pressure  at  the  free  surface). 

b)  Freely  Floating  Board. 

In  Fig.  10.13.12  the  notation  for  the  present  case  is  indicated: 
u(t),  v(t)  represent  the  coordinates  of  the  center  of  gravity  of  the 
board  in  the  displaced  position,  and  co(t)  the  angular  displacement. 
As  before,  we  consider  only  simple  harmonic  oscillations  and  thus 
take  u9  v,  and  CD  in  the  form 

(10.13.55)  u  =  xeiQ\         v  =  yeiat,         CD  =  we™*, 


Fig.  10.13.12.  A  freely  floating  board 

in  which  x,  y,  and  w  arc  constants  representing  the  complex  ampli- 
tudes of  these  components  of  the  oscillation.  For  rj(x,  t)  we  have, 
therefore 

(10.13.56)  ri(x9 1)  =  [y  +  (x  -  x)w]eiat 

=  (y  +  xw)eiat 

when  terms  of  first  order  in  x,  y,  and  w  only  are  considered.  (The  hori- 
zontal component  of  the  oscillation  is  thus  seen  to  yield  only  a  second 
order  effect.)  The  relation  (10.13.56)  now  yields  (cf.  (10.13.38)): 

(10.13.57)  <pxx=-^(y  + 


LONG  WAVES   IN  SHALLOW  WATEE 


487 


in  which  (p  is  the  complex  amplitude  of  the  velocity  potential  0(x,  t)  = 
(p(x)eiat.  Hence  <p  is  the  following  cubic  polynomial: 

(10.13.58)  <p(x)  =  -  ! 

Since  the  pressure  is  given  by  p  =  —  g0f  —  gg^  we  have  in  the 
present  case 

(10.13.59)  p(x)  =  [-  icre0>(0)  -  Qg(y  +  xw)]eiat. 

The  transition  conditions  (10.13.34)  at  x  =  ±  fl  now  lead,  in  the 
same  way  as  above  from  (10.13.42)  and  (10.13.43),  to  the  equations 

Beika  +  Re~n 


(10.13.60) 


Te 


tffei 

~tka 


i  —    o                \             4* 
= (  —  ya\ V. 

*A\  2        9  J       kY 


These  four  equations  are  not  sufficient  to  determine  the  six  constants 
ff,  T,  w,  y<  y,  and  6.  We  must  make  use  of  the  dynamical  equations  of 
motion  of  the  floating  rigid  body  for  this  purpose.  We  have  the 
equations  of  motion: 

(10.13.61)  F  -  Alv,        and        L  =  /a 

at  our  disposal.  In  the  first  equation  F  and  M  are  the  total  vertical 

force  on  the  board  and  its  mass,  per  unit  width,  and  v  is  the  vertical 
acceleration  of  its  center  of  gravity,  /  the  moment  of  inertia,  L  the 
torque,  and  a  the  angular  acceleration.  These  dynamical  conditions 
then  yield  the  following  relations: 

(10.13.62)  p  dx  =  Mv,          px  dx  =  lib, 

J-o  J-a 

and  these  in  turn  lead  to  the  equations 


0.13.63) 


1" 

J_a 


--!««. 


488  WATER   WAVES 

In  the  first  equation  we  have  ignored  the  weight  of  the  board,  since 
it  is  balanced  by  the  hydrostatic  pressure.  The  equations  (10.13.60) 
and  (10.13.63)  now  determine  all  of  the  unknown  complex  amplitu- 
des. 

We  omit  the  details  of  the  calculations,  which  can  be  found  in  the 
paper  by  Fleishman  [F.5],  In  Fig.  10.13.13  the  results  of  calculations 
for  the  pressure  distribution  in  a  numerical  case  are  given.  The  para- 
meters were  chosen  as  follows: 

0  =  1,  h  =  1  ft,         B  =  1  ft*/sec,  a  =  4  ft, 

M  =  18.72  pounds/ft,        a  =  4.46  rad/sec,        A  =  8  ft. 

It  might  be  added  that  the  value  chosen  for  M  is  such  that  the  struc- 
ture sinks  down  0.0375  feet  when  in  equilibrium. 

A  few  observations  should  be  made.  First  of  all,  we  note  that  in  both 
cases  the  pressure  variation  at  the  right  end  (x  =  +  a),  where  the 
incoming  wave  is  incident,  is  greater  than  at  the  left  end.  This  is  to 
be  expected,  since  the  barrier  exercises  a  damping  effect  on  the  wave 
going  under  it.  The  pressure  distribution  in  the  case  of  the  floating 
board  is  quadratic  in  #,  in  contrast  with  the  case  of  the  fixed  board  in 
which  the  distribution  of  pressure  was  linear  in  x.  Next,  we  note  that 
the  pressure  variation  near  the  right  end  of  the  stationary  board  is 
greater  than  at  the  same  end  of  the  floating  one;  this  too  might  be 
expected  since  the  fixed  board  receives  the  full  impact  of  the  incident 
wave,  while  the  floating  one  yields  somewhat.  Finally,  we  see  that  at 
the  left  end  the  opposite  effect  occurs:  there  the  pressure  variation 
under  the  stationary  barrier  is  less  than  that  under  the  floating  barrier. 
This  is  not  surprising  either,  since  the  fixed  board  should  damp  the 
wave  more  successfully  than  the  movable  board. 

Finally,  we  take  up  the  case  of  a  floating  elastic  beam  (cf.  [F.5] ). 
The  beam  is  assumed  to  extend  from  x  ~  —  ltox  =  Q  and,  as  in  the 
above  cases,  to  be  in  simple  harmonic  motion  due  to  an  incoming 
wave  from  x  =  +  oo.  The  basic  relations  for  <£(#,  t)  on  the  free  sur- 
face, and  for  rf(x,  t)  under  the  beam  are  the  same  as  before: 

(10.13.64)  0XX  =  -1  0tt9         x  >  0,         x  <  -  I, 

gh 

(10.13.65)  ijt  =  -  h0xx,         -  /  <  x  <  0. 

We  assume  once  more  that  the  beam  sinks  very  little  below  the  water 
surface  when  in  equilibrium  (i.e.  very  little  in  relation  to  the  depth  of 


LONG   WAVES    IN    SHALLOW   WATER 


439 


p- 

200 

0 

-200 


•500 


<T 1    =  IT/4 


— i — 
-5 


x/o 


-500 


P 
400 

200 


•200 


•5 


.5 


x/a 


P 
400 

200 

0 

•200 


-1-5  0  5       *'°         J 

Fig.  10.13.13.  Pressure  variations  for  floating  board.   6  =  1,  p  in  pounds  I  (ft)* 


440  WATER   WAVES 

the  water),  so  that  the  coefficient  of  <&xx  in  (10.13.65)  can  be  taken  as 
h  rather  than  (h  +r))  (cf.  (10.13.15)),  and  also  the  transition  condi- 
tions at  the  ends  of  the  beam  are 

(10.13.66)  0X  and  <Pt  continuous  at  x  =  0,     x  =  —  /. 
After  writing 

(10.13.67)  <P(x,  t)  =  (p(x)eiat9         17(0,  /)  =  V(X)CM 
we  find,  as  before: 

a2 

(10.13.68)  <pxx  H  --  <p  =  0,         x  >  0,         x  <  ~  I 

gh 

(10.13.69)  <pxx  +  ^  v  =  0,         -  I  <  x  <  0. 

h 

The  conditions  at   oo  have  the  effect  that  (cf.  (10.13.41)  et  seq.): 

(10.13.70)  <p(x)  =  Beik*  +  Ife-"*         x  >  0, 

(10.13.71)  <p(x)  -  2V**          x  <  -  Z, 

with  A:  =  cr/Vgfe-  All  of  this  is  the  same  as  for  the  previous  cases.  We 
turn  now  to  the  conditions  which  result  from  the  assumption  that  the 
floating  body  is  a  beam. 

The  differential  equation  governing  small  transverse  oscillations 
of  a  beam  is 

(10.13.72)  EIr,xxxx  +  mr,tt  =  p, 

in  which  E  is  the  modulus  of  elasticity,  /  the  moment  of  inertia  of  a 
cross  section  of  unit  breadth  (or,  perhaps  better,  El  is  the  bending 
stiffness  factor),  m  the  mass  per  unit  area,  and  p  is  the  pressure.  We 
ignore  the  weight  of  the  beam  and  at  the  same  time  disregard  the 
contribution  of  the  hydrostatic  pressure  term  in  p  corresponding  to 
the  equilibrium  position  of  the  beam—  i.e.  the  pressure  here  is  that  due 
entirely  to  the  dynamics  of  the  situation.  Thus 

(10.13.73)  p  =  ~  Q&t 


Insertion  of  this  relation  in  (10.13.72)  and  use  of  (10.13.69)  leads  at 
once  to  the  differential  equation  for  <p(cc): 

(10.18.T4) 


El       dx*       Elh 


LONG   WAVES    INT   SHALLOW   WATER  441 

that  is  valid  under  the  beam.  The  case  of  greatest  importance  for  us— 
that  of  a  floating  beam  used  as  a  breakwater—  leads  obviously  to  the 
boundary  conditions  for  the  beam  which  correspond  to  free  ends,  i.e. 
to  the  conditions  that  the  shear  and  bending  moments  should  vanish 
at  the  ends  of  the  beam.  These  conditions  in  turn  mean  that  rjxx  and 
r]xxx  should  vanish  at  the  ends  of  the  beam,  and  from  (10.13.67)  and 
(10.13.69)  we  thus  have  for  q>  the  boundary  conditions 


(10.13.75)         -J?  =  -      =  0         at  x  =  0,     x  =  -  I. 
ax*       ax5 

The  transition  conditions  (10.13.66)  require  that  <p  and  q>x  be  conti- 
nuous at  x  =  0,  x  =  —  I,  and  this,  in  view  of  (10.18.70)  and  (10.13. 
71),  requires  that 


(10  13  76)          f 

*     '     '     '         \<p(-  1)  =  Ter*»,     <pm(-l)  =  ikTe~ikl. 

We  remark  once  more  that  the  constant  B  is  assumed  to  be  real,  but 
that  R  and  T  will  in  general  be  complex  constants,  and  that  the  real 
parts  of  0  and  rj  as  given  by  (10.13.67)  are  to  be  taken  at  the  end. 

In  order  to  solve  our  problem  we  must  solve  the  differential  equa- 
tion (10.13.74)  subject  to  the  conditions  (10.13.75)  and  (10.13.76). 
A  count  of  the  relations  available  to  determine  the  solution  should  be 
made:  The  general  solution  of  (10.13.74)  contains  six  arbitrary  con- 
stants, and  we  wish  to  determine  the  constants  R  and  T  (the  am- 
plitudes of  the  reflected  and  transmitted  waves)  occurring  in  (10.13. 
76)  once  the  constant  B  (the  amplitude  of  the  incoming  wave)  has  been 
fixed.  In  all  there  are  thus  eight  constants  to  be  found,  and  we  have  in 
(10.18.75)  and  (10.13.76)  eight  relations  to  determine  them.  Once 
these  constants  have  been  determined,  the  reflection  and  transmission 
coefficients  are  known,  and  the  deflection  of  the  beam  can  be  found 
from  (10.13.69).  The  maximum  bending  stresses  in  the  beam  can  then 
be  calculated  from  the  usual  formula:  s  =  Me//,  with  M  =  Elrfxx 
and  c  the  distance  from  the  neutral  axis  to  the  extreme  outer  fibres  of 
the  beam. 

In  principle,  therefore,  the  solution  of  the  problem  is  straightfor- 
ward. However,  the  carrying  out  of  the  details  in  the  case  of  the  beam 
of  finite  length  is  very  tedious,  involving  as  it  does  a  system  of  eight 
linear  equations  for  eight  unknowns  with  complex  coefficients.  In 
addition,  one  must  determine  the  roots  of  a  sixth  degree  algebraic 
equation  in  order  to  find  the  general  solution  of  (10.13.74).  These 


442  WATER   WAVES 

roots  are  in  general  complex  numbers  and  they  involve  the  essential 
parameters  of  the  mechanical  system.  Thus  it  is  clear  that  a  dis- 
cussion of  the  behavior  of  the  system  in  general  terms  with  respect  to 
arbitrary  values  of  the  parameters  of  the  system  is  not  feasible,  and 
one  must  turn  rather  to  concrete  cases  in  which  most  of  the  parameters 
have  been  given  specific  numerical  values.  The  results  of  some  calcula- 
tions of  this  kind,  for  a  case  proposed  as  a  practical  possibility,  will  be 
given  a  little  later. 

The  case  of  a  semi-infinite  beam—  i.e.  a  beam  extending  from  x  =  0 
to  x  =  —  oo  —  is  simpler  to  deal  with  in  that  the  conditions  in  the 
second  line  of  (10.13.76)  fall  away,  and  the  conditions  (10.13.75)  at 
x  —  —  oo  can  be  replaced  by  the  requirement  that  q>  be  bounded  at 
x  —  —  oo.  The  number  of  constants  to  be  fixed  then  reduces  to  four 
instead  of  eight,  but  the  determination  of  the  deflection  of  the  beam 
still  remains  a  formidable  problem;  we  shall  consider  this  case  as  well 
as  the  case  of  a  beam  of  finite  length. 

We  begin  the  program  indicated  with  a  discussion  of  the  general 
solution  of  the  differential  equation  (10.13.74).  Since  it  is  a  linear 
differential  equation  with  constant  coefficients  we  proceed  in  the 
standard  fashion  by  setting  q>  =  £*,  inserting  in  (10.13.74),  to  find 
for  K  the  equation 

(10'.13.77)  x«  +  ax2  +  b  --=  0 

with 


This  is  a  cubic  equation  in  x2  =  /?,  which  happens  to  be  in  the  standard 
form  to  which  the  Cardan  formula  for  the  roots  of  a  cubic  applies 
directly.  For  the  roots  /^  of  this  equation  one  has  therefore 


(10.13.79) 


=  u  +  v 


s2v 


H  CT 
with  u  and  v  defined  by 

/       b        /b2       as\i\*  /       b        /b2      a3\i\i 

(10.18.80)    u=l-  —  +  (—  +  —I   |  ,     0  =  1 _+_ )) 

\       2T\4       27//'  \       2        \4       27/  / 

and  e  the  following  cube  root  of  unity: 


LONG    WAVES    IN    SHALLOW    WATER  443 

The  constant  a  is  positive,  since  or,  the  frequency  of  the  incoming 
wave,  is  so  small  in  the  cases  of  interest  in  practice  that  qg  is  much 
larger  than  ma2.  The  constant  b  is  obviously  positive.  Consequently 
the  root  ^  is  real  and  negative  since  \u\  <\v\  and  v  is  negative. 
Thus  the  roots  »x  =  +  /?J'2,  x2  =  —  £J/2  are  pure  imaginary.  The 
quantities  f}2  and  /?3  are  complex  conjugates,  and  their  square  roots 
yield  two  pairs  of  complex  conjugates 


For  /92  and  /?3  we  have 

(10.13.82)  &  -  -  -  (it  +  v)  +  i       -  (u  -  v), 

2  2 

(10.13.83)  /?,  =  -  1  (u  +  »)  -  i  ~  (u  -  v). 

&  £ 

Thus  f}2  and  ^3  both  have  positive  real  parts.  We  suppose  the  roots 
x&  #4,  H5,  x6  to  be  numbered  to  that  x3  and  x5  are  taken  to  have  posi- 
tive real  parts,  while  x4  and  x6  have  negative  real  parts.  The  general 
solution  of  (10.13.74)  thus  is 

(10.13.84)  <p(x)  -  a^i*  +  a<£**x  +  atf**x  +  a4^4*  +  a^x  +  a^x. 

In  the  case  of  a  beam  covering  the  whole  surface  of  the  water,  i.e. 
extending  from  —  oo  to  +  oo,  the  condition  that  <p  be  bounded  at 
cr  = :  ^  oo  would  require  that  «3  =  «4  =  «5  =  «6  =  0  since  the  ex- 
ponentials in  the  corresponding  terms  have  non-vanishing  real  parts. 
The  remaining  terms  yield  progressing  waves  traveling  in  opposite 
directions;  their  wave  lengths  are  given  by  A  =  2n/\  K±  \  =  2n/\  x2  I 
and  thus  by 

(10.13.85)  A  =  2n/V\  u  +  v  |, 

with  u  +  v  defined  by  (10.13.80).  The  wave  length  and  frequency  are 
thus  connected  by  a  rather  complicated  relation,  and,  unlike  the  case 
of  waves  in  shallow  water  with  no  immersed  bodies  or  constraints  on 
the  free  surface,  the  wave  length  is  not  independent  of  the  frequency 
and  the  wave  phenomena  are  subject  to  dispersion. 

In  the  case  of  a  beam  extending  from  the  origin  to  —  oo  while  the 
water  surface  is  free  for  x  >  0,  the  boundedness  conditions  for  <p  at 
—  oo  requires  that  we  take  a4  =  ae  =  0  since  x4  and  x6  have  negative 
real  parts  and  consequently  «*«*  and  e*«*  would  yield  exponentially 
unbounded  contributions  to  <p  at  x  =  —  oo.  We  know  that  «j  and  x2 


444 


WATER    WAVES 


are  pure  imaginary  with  opposite  signs,  with  x2,  say,  negative  imagin- 
ary. Since  no  progressing  wave  is  assumed  to  come  from  the  left,  we 
must  then  take  a2  =  0.  Thus  the  term  a^x  yields  the  transmitted 
wave  and  the  terms  involving  a3  and  a5  yield  disturbances  which  die 
out  exponentially  at  oo.  The  conditions  (10.13.70)  and  (10.13.71 )  at 
x  =  0  now  yield  the  following  four  linear  equations: 


(10.13.86) 


=  0 


a     +  " 


-  B  +  R 

x6at~ik(B-R) 


for  the  constants  av  a3,  #5,  R.  For  the  amplitude  R  of  the  reflected 
wave  one  finds 


0 

(10.13.87) 

-  1       -  1      -    1      -f  1  , 

K!  *3  *5  /* 

1  3  5 

Even  in  this  relatively  simple  case  of  the  semi-infinite  beam  the  re- 
flection coefficient  is  so  complicated  a  function  of  the  parameters 
(even  though  it  is  algebraic  in  them)  that  it  seems  not  worthwhile  to 
write  it  down  explicitly.  The  results  of  numerical  calculations  based 
on  (10.13.87)  will  be  given  shortly. 

In  the  case  of  the  beam  of  finite  length  extending  from  <r  —  / 
to  x  =  0  the  eight  conditions  given  by  (10.13.75)  and  (10.13.76)  must 
be  satisfied  by  the  solution  (10.13.84)  of  the  differential  equation 
(10.13.74),  and  these  conditions  serve  to  determine  the  six  constants 
of  integration  and  the  amplitudes  R  and  T  of  the  reflected  and  trans- 
mitted waves.  The  problem  thus  posed  is  quite  straightforward  but 
extremely  tedious  as  it  involves  solving  eight  linear  eq  nations  for 
eight  complex  constants.  For  details  reference  is  again  made  to  the 
work  of  Wells  [F.5]. 

This  case  of  a  floating  beam  was  suggested  to  the  author  by  J.  H. 


LONG    WAVES    IN    SHALLOW    WATER  445 

Carr  of  the  Hydraulics  Structures  Laboratory  at  the  California  Insti- 
tute of  Technology  as  one  having  practical  possibilities;  at  his  sug- 
gestion calculations  in  specific  numerical  cases  were  carried  out 
in  order  to  determine  the  effectiveness  of  such  a  breakwater.  The 
reason  for  considering  such  a  structure  for  a  breakwater  as  a 
means  of  creating  relatively  calm  water  between  it  and  the  shore 
is  the  following:  a  structure  which  floats  on  the  surface  without  sink- 
ing far  into  the  water  need  not  be  subjected  to  large  horizontal  forces 
and  hence  would  not  necessarily  require  a  massive  anchorage.  How- 
ever, in  order  to  be  effective  as  a  reflector  of  waves  such  a  floating 
structure  would  probably  have  to  be  built  with  a  fairly  large  dimen- 
sion in  the  direction  of  travel  of  the  incoming  waves.  As  a  consequence 
of  the  length  of  the  structure,  it  would  be  bent  like  a  beam  under  the 
action  of  the  waves  and  hence  could  not  in  general  be  treated  with 
accuracy  as  a  rigid  body  in  determining  its  effectiveness  as  a  barrier. 
This  brings  with  it  the  possibility  that  the  structure  might  be  bent 
so  much  that  the  stresses  set  up  would  be  a  limiting  feature  in  the 
design.  The  specifications  (as  suggested  by  Carr)  for  a  beam  having 
a  width  of  one  foot  (parallel  to  the  wave  crest,  that  is)  were: 

Weight:  85  pounds •/ 'ft2 

Moment  of  inertia  (of  area  of  cross-section):  0.2  /24 

Modulus  of  elasticity:  437  X  107  pounds  I  ft2. 

The  depth  of  the  water  is  taken  as  40  feet.  Simple  harmonic  progres- 
sing waves  having  periods  of  8  and  of  15  sees,  were  to  be  considered, 
and  these  correspond  to  wave  lengths  of  287  and  539  feet,  and  to  cir- 
cular frequencies  a  of  785  x  10~3  and  418  x  10~3  cycles  per  second, 
respectively.  The  problem  is  to  determine  the  reflecting  power  of  the 
beam  under  these  circumstances  when  the  length  of  the  beam  is  varied. 
In  other  words,  we  assume  a  wave  train  to  come  from  the  right  hand 
side  of  the  beam  and  that  it  is  partly  transmitted  under  the  beam  to 
the  left  hand  side  and  partly  reflected  back  to  the  right  hand  side. 
The  ratio  R/B  of  the  amplitude  R  of  the  reflected  wave  and  the  am- 
plitude B  of  the  incoming  wave  is  a  measure  of  the  effectiveness  of 
the  beam  as  a  breakwater. 

Before  discussing  the  case  of  beams  of  finite  length  it  is  interesting 
and  worthwhile  to  consider  semi-infinite  beams  first.  Since  the  calcu- 
lations are  easier  than  for  beams  of  finite  length  it  was  found  possible 
to  consider  a  larger  range  of  values  of  the  parameters  than  was  given 


446  WATER   WAVES 

above.  The  results  are  summarized  in  the  following  tables  (taken 
from   [F.5]): 

TABLE  A 


X  (ft)  a  l—\  W (pounds)  I  (ft)  E^^-}  h  (ft)  R/B 
\secj \     ft2     J 

539            0.418            85  0.20  437  X  107  40  0.14 

287            0.785            85  0.20  437  X  107  40  0.19 

225             1.0                 85  0.20  437  X  107  40  0.23 

150             1.5                 85  0.20  437  X  107  40  0.32 

113            2.0                 85  0.20  437  X  107  40  0.43 


In  Table  A  the  beam  design  data  are  as  given  above.  At  the  two  speci- 
fied circular  frequencies  of  0.418  and  0.785  one  sees  that  the  floating 
beam  is  quite  ineffective  as  a  breakwater  since  the  reflected  wave  has 
an  amplitude  of  less  than  1/5  of  the  amplitude  of  the  incoming  wave, 
even  for  the  higher  frequency  (and  hence  shorter  wave  length),  which 
means  that  less  than  4  %  of  the  incoming  energy  is  reflected  back.  At 
higher  frequencies,  and  hence  smaller  wave  lengths,  the  breakwater 
is  more  effective,  as  one  would  expect.  However  the  approximate 
theory  used  to  calculate  the  reflection  coefficient  R/B  can  be  expected 
to  be  accurate  only  if  the  ratio  Xjh  of  wave  length  to  depth  is  suffi- 
ciently large,  and  even  for  the  case  A  —  287  ft.  (a  =-  .785)  the  re- 
flection coefficient  of  value  0.19  may  be  in  error  by  perhaps  10  %  or 
more  since  A/A  is  only  about  7,  and  the  errors  for  the  shorter  wave 
lengths  would  be  greater.  Calculations  for  still  other  values  of  the 
parameters  are  shown  in  Table  B.  The  only  change  as  compared  with 

TABLE  B 


W  I  E  h  RIB 


539         0.418         384         0.20         437  X  107         40         0.51 
287         0.785         384         0.20         437  X  107         40         0.75 

the  first  two  rows  of  Table  A  is  that  the  weight  per  foot  of  the  beam 
has  been  increased  by  a  factor  of  more  than  4  from  a  value  of  85 
pounds] ft*  to  a  value  of  384  pounds] 'ft2.  The  result  is  a  decided  increase 
in  the  effectiveness  of  the  breakwater,  especially  at  the  shorter  wave 
length,  since  more  than  half  (i.e.  (.75)2)  of  the  incoming  energy  would 
be  reflected  back.  However,  this  beneficial  effect  is  coupled  with  a 
decided  disadvantage,  since  quadrupling  the  weight  of  the  beam 


LONG    WAVES    IN   SHALLOW    WATER  447 

would  cause  it  to  sink  deeper  in  the  water  in  like  proportion  and  hence 
might  make  heavy  anchorages  necessary.  Table  C  is  the  same  as  the 


TABLE  C 


A 

(7 

W 

/ 

E 

h 

R/B 

539 

.418 

85 

2.0 

437  X  107 

40 

.26 

287 

.785 

85 

2.0 

437  X  107 

40 

.32 

00 

1 

first  two  rows  of  Table  A  except  that  the  bending  stiffness  has  been 
increased  by  a  factor  of  10  by  increasing  the  moment  of  inertia  of  the 
beam  cross-section  from  0.2  //4  to  2.0  //4.  Such  an  increase  in  stiffness 
results  in  a  noticeable  increase  in  the  effectiveness  of  the  breakwater, 
but  by  far  not  as  great  an  increase  as  is  achieved  by  multiplying  the 
weight  by  a  factor  of  four.  If  the  stiffness  were  to  be  made  infinite 
(i.e.  if  the  beam  were  made  rigid)  the  reflection  coefficient  could  be 
made  unity,  and  no  wave  motion  would  be  transmitted.  This  is 
evidently  true  for  a  semi-infinite  beam,  but  would  not  be  true  for  a 
rigid  body  of  finite  length. 


TABLE  D 

A 

a 

W 

7 

E 

h 

RIB 

539 

.418 

85  ~ 

(T 

~Q 

~  40 

.  ooT 

287 

.785 

85 

0 

0 

40 

.007 

In  Table  D  the  difference  as  compared  with  Table  A  is  that  the 
beam  stiffness  is  taken  to  be  zero.  This  means  that  the  surface  of  the 
water  is  assumed  to  be  covered  by  a  distribution  of  inert  particles 
weighing  85  pounds  per  foot.  (Such  cases  have  been  studied  by  Gold- 
stein and  Keller  [G.I].)  As  we  observe,  there  is  practically  no  reflec- 
tion and  this  is  perhaps  not  surprising  since  the  mass  distribution  per 
unit  length  has  such  a  value  that  the  beam  sinks  down  into  the  water 
only  slightly. 

One  might  perhaps  summarize  the  above  results  as  follows:  A  very 
long  beam  can  be  effective  as  a  floating  breakwater  if  it  is  stiff  enough. 
However,  a  reasonable  value  for  the  stiffness  (the  value  0.2  given 
above)  leads  to  an  ineffective  breakwater  unless  the  weight  of  the 
beam  per  square  foot  is  a  fairly  large  multiple  (say  8  or  10)  of  the 
weight  of  water. 

In  practice  it  seems  unlikely  that  beams  long  enough  to  be  considered 


448  WATER    WAVES 

semi-infinite  would  be  practicable  as  breakwaters.  (The  term  "long 
enough"  might  be  interpreted  to  mean  a  sufficiently  large  multiple  of 
the  wave  length,  but  since  the  wave  lengths  are  of  the  order  of  200 
feet  or  more  the  correctness  of  this  statement  seems  obvious. )  It  there- 
fore is  necessary  to  investigate  the  effectiveness  of  beams  of  finite 
length.  Such  an  investigation  requires  extremely  tedious  calculations 
—so  much  so  that  only  a  certain  number  of  numerical  cases  have  been 
treated.  These  are  summarized  in  the  following  tables. 

a  =  .785,     A  =  287  a  =  .418,     A  =  539 


I  (ft)  R/B  I  (ft)  R/B 


17.5  0  145.9  .17 

49.2  .93  196.9  .53 

72.9  0  291.8  .13 

98.5  .75  443.0  .90 

145.9  .10  583.6  .74 

196.9  0  656.2  .62 

291.8  .33  874.9  .07 

450.4  .32  875.4  .08 

583.6  .12  948.3  .54 

656.3  .13  oo  .14 

875.4  .32 
oo  .19 


In  these  tables  the  parameters  have  values  the  same  as  in  the  first 
two  rows  of  Table  A,  except  that  now  lengths  other  than  infinite 
length  are  considered.  The  most  noticeable  feature  of  the  results  given 
in  the  tables  is  their  irregularity  and  the  fact  that  at  certain  lengths 
—even  certain  rather  short  lengths— the  beam  proposed  by  Carr 
seems  to  be  quite  effective.  For  example,  when  the  wave  length  is 
287  ft.  a  beam  less  than  50  ft.  long  reflects  more  than  80  %  of  the 
incoming  energy.  A  beam  of  length  443  ft.  is  also  equally  effective  at 
the  longer  wave  length  of  539  ft.* 

*  It  might  not  be  amiss  to  consider  the  physical  reason  why  it  is  possible  that 
a  beam  of  finite  length  could  be  more  effective  as  a  breakwater  than  a  beam  of 
infinite  length.  Such  a  phenomenon  comes  about,  of  course,  through  multiple 
reflections  that  take  place  at  the  ends  of  the  beam.  Apparently  the  phases  some- 
times arrange  themselves  in  the  course  of  these  complicated  interactions  in  such 
a  way  as  to  yield  a  small  amplitude  for  the  transmitted  wave.  That  such  a  process 
might  well  be  sensitive  to  small  changes  in  the  parameters,  as  is  noted  in  the 
discussion,  cannot  be  wondered  at. 


LONG    WAVES   IN    SHALLOW   WATER  449 

However,  the  maximum  effectiveness  of  any  such  breakwater 
occurs  for  a  specific  wave  length  within  a  certain  range  of  wave 
lengths;  thus  the  reflection  of  a  given  percentage  of  the  incoming  wave 
energy  would  involve  changing  the  length  (or  some  other  parameter)  of 
the  structure  in  accordance  with  changes  in  the  wave  length  of  the 
incoming  waves.  Also,  the  reflection  coefficient  seems  to  be  rather 
sensitive  to  changes  in  the  parameters,  particularly  for  the  shorter 
structures  (a  relatively  slight  change  in  length  from  an  optimum  value, 
or  a  slight  change  in  frequency,  leads  to  a  sharp  decrease  in  the  re- 
flection coefficient).  It  is  also  probable —as  was  indicated  earlier  on  the 
basis  of  calculations  by  Wells  [W.10]  —  that  the  shallow  water  approx- 
imation used  here  as  a  basis  for  the  theory  is  not  sufficiently  accurate 
for  a  floating  beam  whose  length  is  too  much  less  than  the  wave  length. 
Nevertheless,  it  does  seem  possible  to  design  floating  breakwaters  of 
reasonable  length  which  would  be  effective  at  a  given  wave  length. 
Perhaps  it  is  not  too  far-fetched  to  imagine  that  sections  could  be 
added  to  or  taken  away  from  the  breakwater  in  accordance  with 
changing  conditions. 

Another  consequence  of  the  theory— which  is  also  obvious  on 
general  grounds— is  that  there  is  always  the  chance  of  creating  a 
large  standing  wave  between  the  shore  and  the  breakwater  because  of 
reflection  from  the  shore,  unless  the  waves  break  at  the  shore;  this 
effect  is  perhaps  not  important  if  the  main  interest  is  in  breakwaters 
off  beaches  of  not  too  large  slope,  since  breaking  at  the  shore  line  then 
always  occurs.  (The  theory  developed  here  could  be  extended  to  cases 
in  which  the  shore  reflects  all  of  the  incoming  energy,  it  might  be 
noted.)  In  principle,  the  calculation  of  the  deflection  curve  of  the 
structure,  and  hence  also  of  the  bending  stresses  in  it,  as  given  by  the 
theory  is  straightforward,  but  it  is  very  tedious;  consequently  only 
the  reflection  coefficients  have  been  calculated. 


CHAPTER  11 


Mathematical  Hydraulics 


In  this  chapter  the  problems  to  be  treated  are,  from  the  mathema- 
tical point  of  view,  much  like  the  problems  of  the  preceding  chapter, 
but  the  emphasis  is  on  problems  of  rather  concrete  practical  signifi- 
cance. Aside  from  this,  the  essential  difference  is  that  external  forces 
other  than  gravity,  such  as  friction,  for  example,  play  a  major  role  in 
the  phenomena.  Problems  of  various  types  concerning  flows  and 
wave  motions  in  open  channels  form  the  contents  of  the  chapter.  The 
basic  differential  equations  suitable  for  dealing  with  such  flows  under 
rather  general  circumstances  are  first  derived.  This  is  followed  by  a 
study  of  steady  motions  in  uniform  channels,  and  of  progressing  waves 
of  uniform  shape,  including  roll  waves  in  inclined  channels.  Flood 
waves  in  rivers  are  next  taken  up,  including  a  discussion  of  numerical 
methods  appropriate  in  such  cases;  the  results  of  such  calculations 
toy  a  flood  wave  in  a  simplified  model  of  the  Ohio  River  and  for  a 
model  of  its  junction  with  the  Mississippi  are  given.  This  discussion 
follows  rather  closely  the  two  reports  made  to  the  Corps  of  Engineers 
of  the  U.S.  Army  by  Stoker  [S.23]  and  by  Isaacson,  Stoker,  and 
Trocsch  [1.4],  These  methods  of  dealing  with  flood  waves  have  been 
applied,  with  good  results,  to  a  400-mile  stretch  of  the  Ohio  as  it 
actually  is  for  the  case  of  the  big  flood  of  1945,  and  also  to  a  flood 
through  the  junction  of  the  Ohio  and  the  Mississippi;  these  results 
will  be  discussed  toward  the  end  of  this  chapter. 

There  is  an  extensive  literature  devoted  to  the  subject  of  flow  in 
open  channels.  We  mention  here  only  a  few  items  more  or  less  directly 
connected  with  the  material  of  this  chapter:  the  famous  Essai  of 
Boussinesq  [B.17],  the  books  of  Bakhmeteff  [B.3]  and  Rouse  [R.10, 
11]  (in  particular,  the  article  by  Gilcrest  in  [R.ll]),  the  Enzyklopadie 
article  of  Forchheimer  [F.6]  and  the  booklet  by  Thomas  [T.2]. 

451 


452 


WATER   WAVES 


11.1.  Differential  equations  of  flow  in  open  channels 

It  has  already  been  stated  that  the  basic  mathematical  theory  to 
be  used  in  this  chapter  does  not  differ  essentially  from  the  theory 
derived  in  the  preceding  chapter.  However,  there  are  additional 
complications  due  to  the  existence  of  significant  forces  beside  gravity, 
and  we  wish  to  permit  the  occurrence  of  variable  cross-sections  in  the 
channels.  Consequently  the  theory  is  derived  here  again,  and  a  some- 
what different  notation  from  that  used  in  previous  chapters  is  em- 
ployed both  for  the  sake  of  convenience  and  also  to  conform  somewhat 
with  notations  used  in  the  engineering  literature. 

The  theory  is  one-dimensional,  i.e.  the  actual  flow  in  the  channel  is 
assumed  to  be  well  approximated  by  a  flow  with  uniform  velocity  over 
each  cross-section,  and  the  free  surface  is  taken  to  be  a  level  line  in 
each  cross-section.  The  channel  is  assumed  also  to  be  straight  enough 
so  that  its  course  can  be  thought  of  as  developed  into  a  straight  line 
without  causing  serious  errors  in  the  flow.  The  flow  velocity  is  denoted 
by  v,  the  depth  of  the  stream  (commonly  called  the  stage  in  the 
engineering  literature)  by  y,  and  these  quantities  are  functions  of  the 


Fig.  11.1.1.  River  cross-section  and  profile 

distance  x  down  the  stream  and  of  the  time  t  (cf.  Fig.  11.1.1).  The 
vertical  coordinates  of  the  bottom  and  of  the  free  surface  of  the  stream, 
as  measured  from  the  horizontal  axis  x9  are  denoted  by  z(x)  and 


MATHEMATICAL   HYDRAULICS  453 

h(x,  t),  with  z  positive  downward,  h  positive  upward;  thus  y  =•  h  +  z. 
The  slope  of  the  bed  is  therefore  counted  positive  in  the  positive 
^-direction,  i.e.  downward.  The  breadth  of  the  free  surface  at  any 
section  of  the  stream  is  denoted  by  B. 

The  differential  equations  governing  the  flow  are  expressions  of  the 
laws  of  conservation  of  mass  and  momentum.  In  deriving  them  the 
following  assumptions,  in  addition  to  those  mentioned  above,  are 
made  *:  1)  the  pressure  in  the  water  obeys  the  hydrostatic  pressure 
law,  2 )  the  slope  of  the  bed  of  the  river  is  small,  3)  the  effects  of  friction 
and  turbulence  can  be  accounted  for  through  the  introduction  of  a 
resistance  force  depending  on  the  square  of  the  velocity  v  and  also,  in 
a  certain  way  to  be  specified,  on  the  depth  y. 

We  first  derive  the  equation  of  continuity  from  the  fact  that  the 
mass  gAAx  included  in  a  layer  of  water  of  density  p,  thickness  Ax, 
and  cross-section  area  A,  changes  in  its  flow  along  the  stream  only 
through  a  possible  inflow  along  the  banks  of  the  stream,  say  at  the 
rate  qq  per  unit  length  along  the  river.  The  total  flow  out  of  the  ele- 
ment of  volume  A  Ax  is  given  by  the  net  contributions  Q(Av)xAx  from 
the  flow  through  the  vertical  faces  plus  the  contribution  @BhtAx  due 
to  the  rise  of  the  free  surface,  with  B  the  width  of  the  channel;  since 
Bht  represents  the  area  change  At  it  follows  that  the  sum  [(Av)x  + 
At}Ax  equals  the  volume  influx  qAx  over  the  sides  of  the  channel, 
with  q  the  influx  per  unit  length  of  channel.  The  subscripts  x  and  t 
refer,  of  course,  to  partial  derivatives  with  respect  to  these  variables. 
Tfye  equation  of  continuity  therefore  has  the  form 

(11.1.1)  (Av)x+At  =  q. 

It  should  be  observed  that  A  —  A(y(x,  t)9  x)  is  in  the  nature  of 
things  a  given  function  of  y  and  x,  although  y(x,  t)  is  an  unknown 
function  to  be  determined;  in  addition,  q  =  q(x,  t)  depends  in  general 
on  both  x  and  /  in  a  way  that  is  supposed  given.  In  the  important 
special  case  of  a  rectangular  channel  of  constant  breadth  B,  so  that 
A  —  By,  the  equation  of  continuity  takes  the  form 

(11.1.2)  vxy  +  vyx  +  yt  =  q/B. 

The  equation  of  motion  is  next  derived  for  the  same  slice  of  mass 
m  =  qAAx  by  equating  the  rate  of  change  of  momentum  d(mv)jdt 

*  These  assumptions  are  not  the  minimum  number  necessary:  for  example, 
assumption  1 )  has  as  a  consequence  the  independence  of  the  velocity  on  the  vertical 
coordinate  if  that  were  true  at  any  one  instant  (cf.  the  remarks  on  this  point 
in  Ch.  2  and  Ch.  10). 


454  WATER   WAVES 

to  the  net  force  on  the  element.  We  write  the  equation  of  motion  for 
the  horizontal  direction: 

(11.1.8)      Q  —  (AvAx)  =  HAx—FfAx  cos  9?  +QgA  Ax  sin  q>. 
dt 

In  this  equation  H  represents  the  unbalanced  horizontal  pressure 
force  at  the  surface  of  the  element.  The  angle  9?  is  the  slope  angle  of  the 
bed  of  the  channel,  reckoned  positive  downward.  The  quantity  Ff  re- 
presents the  friction  force  along  the  sides  and  bottom  of  the  channel, 
and  the  term  QgAAx  sin  <p  represents  the  effect  of  gravity  in  accelerat- 
ing the  slice  down-hill  as  manifested  through  the  normal  reaction  of 
the  stream  bed.  Since  9?  was  assumed  small  we  may  replace  sin  (p  by 
the  slope  S  —  dz/dx  and  cos  99  by  1.  In  the  frictional  resistance  term 
we  set 


This  is  an  empirical  formula  called  Manning's  formula.  The  resistance 
is  thus  proportional  to  the  square  of  the  velocity  and  is  opposite  to  its 
direction;  in  addition,  the  friction  is  inversely  proportional  to  the 
4/3-power  of  the  hydraulic  radius  R,  defined  as  the  ratio  of  the  cross- 
section  area  A  to  the  wetted  perimeter  (thus  R  =  Byj(B  -\-  2y)  for  a 
rectangular  channel  and  R  =  y  for  a  very  wide  rectangular  channel), 
and  inversely  proportional  to  y,  a  roughness  coefficient. 

We  calculate  next  the  momentum  change  Qd(AvAx)/dt.  In  doing  so, 
we  observe  that  the  symbol  d/dt  must  be  interpreted  as  the  particle 
derivative  (cf.  Chapter  1.1  and  equation  (1.1.8))  d/dt  +  vd/dx  since 
Newton's  law  must  be  applied  in  following  a  given  mass  particle  along 
its  path  x  =  x(t).  However,  the  law  of  continuity  (11.1.1  )  derived  above 
is  clearly  equivalent  to  writing  d(AAx)ldt  —  qAx,  with  d/dt  again  in- 
terpreted as  the  particle  derivative.  Since 

—  (AvAx)=v  —  (AAx)+AAx  — 
dt  dt  dt 

it  follows  that 

—  (AvAx)—AAx(vvx+vt)-}-qvAx. 
dt 

Finally,  the  net  contribution  HAx  of  the  pressure  forces  over  the 
surface  of  the  slice  is  calculated  as  follows:  The  total  pressure  over  a 

vertical  face  of  the  slab  is  given  by  f  v  Qg[y(x,  t)  —  £]b(x,  f  )  d£  from 
the  hydrostatic  pressure  law  (cf.  Fig.  11.1.1);  while  the  component 


MATHEMATICAL   HYDRAULICS  455 

in  the  ^-direction  of  the  total  pressure  over  the  part  of  the  slice  in 
contact  with  the  banks  of  the  river  is  given  by 

try  \ 

j  I    6S[y  "~  S]bx(x9  f )  d£    Ax,  we  have  for  HAx  the  following  equation: 

(11.1.5) 

/o 

my 


-t-f  f 

WO 


[1 

=  --  I    @gyxb(%,  %)d!;~  ~ogAyx. 
Jo 

In  this  calculation  the  integrals  involving  bx  cancel  out,  and  we  have 
used  the  fact  that  yx  is  independent  of  f . 

Adding  all  of  the  various  contributions  we  have 


(11.1.6)  vt+wx+  -    v=Sg-S,g-gyx 

scL 

upon  defining  what  is  called  the  friction  slope  Sf  by  the  formula 
(11-1.7)  S, 


with  Ff  defined  by  (11.1.4).  It  should  perhaps  be  mentioned  that  the 
term  qv/A  on  the  left  hand  side  of  (11.1.6)  arises  because  of  the 
tacit  assumption  that  flows  enter  the  main  stream  from  tributaries 
or  by  flow  over  the  banks  at  zero  velocity  in  the  direction  of  the  main 
stream;  if  such  flows  were  assumed  to  enter  with  the  velocity  of  the 
main  stream,  the  term  would  not  be  present—  it  is,  in  any  case,  a 
term  which  is  quite  small.  If  we  introduce  A  =  A(y(x,  t),  x)  in 
(11.1.1)  the  result  is 

(11.1.8)  Ayyxv  +  Axv  +  Avx  +  Ayyt  —  q. 

The  two  differential  equations  (11.1.6)  and  (11.1.8),  which  serve 
to  determine  the  two  unknown  functions,  the  depth  y(x,  t)  and  the 
velocity  v(x,  t),  are  the  basic  equations  for  the  study  of  flood  waves  in 
rivers  and  flows  in  open  channels  generally.  For  any  given  river  or 
channel  it  is  thus  necessary  to  have  data  available  for  determining 
the  cross-section  area  A  and  the  quantities  y  and  R  in  the  resistance 
term  Ff  as  functions  of  x  and  j/,  and  of  the  slope  S  of  its  bed  as  a 
function  of  x  in  order  to  have  the  coefficients  in  the  differential  equa- 
tions (11.1.6)  and  (11.1.8)  defined.  Three  of  these  quantities  are 
purely  geometrical  in  character  and  could  in  principle  be  determined 


456  WATER   WAVES 

from  an  accurate  contour  map  of  the  river  valley,  but  the  determina- 
tion of  the  roughness  coefficient  y  of  course  requires  measurements  of 
actual  flows  for  its  determination. 

11.2.  Steady  flows.  A  junction  problem 

We  define  a  steady  flow  in  the  usual  fashion  to  be  one  for  which  the 
velocity  v  and  depth  y  are  independent  of  the  time,  that  is,  vt—  yt  —  0. 
In  this  section  channels  of  constant  rectangular  cross-section  and 
constant  slope  will  be  considered  for  the  most  part.  It  follows  from  the 
equation  of  continuity  (cf.  (11.1.2)): 

Vt  +  vyx  +  yvx  =  0, 
that  for  steady  flow 

(11.2.1)  (vy)x  =  0  whence  vy  —  D     (Da  constant), 

when  no  flow  into  the  channel  from  its  sides  occurs  (i.e.  q  =  0  in 
(11.1.2)).  Similarly,  the  equation  of  motion  (cf.  (11.1.6)) 

^  +  vvx  +  gyx  +  g(S,  -S)  =  0 
yields 

(11.2.2)  vvx  +  gy,  +  g(St  -S)  =  Q. 

It  follows  from  equation  (11.2.1)  that 

D  A  D 

v  =  _     and  vx  =  --  -  yx, 

y  y2 

so  that  equation  (11.2.2)  becomes 


Here  the  hydraulic  radius  is  given  by  R  =  y/(l  +  2y/B)  because  the 
channel  is  assumed  to  be  rectangular  in  cross-section. 

For  a  channel  with  given  physical  parameters  such  as  cross-section, 
resistance  coefficient,  etc.  the  steady  flows  would  provide  what  are 
called  backwater  curves.  In  general,  one  could  in  principle  always 
find  steady  solutions  y  =  y(x)  and  v  =  v(x)  for  a  non-uniform  chan- 
nel. The  explicit  determination  of  the  stage  y  and  discharge  rate  BD 
as  functions  of  x  would  be  possible  by  numerical  integration  of  the 
pair  of  first  order  ordinary  differential  equations  arising  from  (11.1.6) 
and  (11.1.8)  when  time  derivatives  are  assumed  to  vanish. 


MATHEMATICAL   HYDRAULICS  457 

We  note  that  equation  (11.2.3)  has  the  simple  solution  y  =  constant 
for  y  satisfying 

This  means  that  we  can  find  a  flow  of  uniform  depth  and  velocity 
having  a  constant  discharge  rate  BD  (B  is,  as  in  the  preceding  section, 
the  width  of  the  channel).  Conversely,  by  fixing  the  depth  y  we  can 
find  the  discharge  from  (11.2.4)  appropriate  to  the  corresponding 
uniform  flow.  Physically  this  means  that  the  flow  velocity  is  chosen 
so  that  the  resistance  due  to  turbulence  and  friction  and  the  effect  of 
gravity  down  the  slope  of  the  stream  just  balance  each  other.  We  re- 
mark that  if  (11.2.4)  is  satisfied  at  any  point  where  the  coefficient 
g  —  D2lyz  of  yx  in  (11.2.3)  does  not  vanish,  then  y  =  constant  is  the 
only  solution  of  (11.2.3)  because  of  the  fact  that  the  solution  is  then 
uniquely  determined  by  giving  the  value  of  y  at  any  point  x.  We  note 
that  g  —  D2/y3  =  0  corresponds  to  v  =  VliJ/>  *-e-  to  a  fl°w  at  critical 
speed  (a  term  to  be  discussed  in  the  next  section),  since  D  =  vy. 
Furthermore,  the  differential  equation  (11.2.3)  can  be  integrated  to 
yield  x  as  a  function  of  y: 

(11.2.5)        a?  = 


when  x  —  0  for  y  —  yQ. 

We  proceed  to  make  use  of  (11.2.5)  in  order  to  study  a  problem 
involving  a  steady  flow  at  the  junction  of  two  rivers  each  having  a 
rectangular  channel.  Later  on,  the  same  problem  will  be  treated  but 
for  an  unsteady  motion  resulting  from  a  flood  wave  traveling  down 
one  of  the  branches,  and  such  that  the  steady  flow  to  be  treated  here 
is  expected  to  result  as  a  limit  state  after  a  long  time.  The  numerical 
data  arc  chosen  here  for  the  problem  in  such  a  way  as  to  correspond 
roughly  with  the  actual  data  for  the  junction  of  the  Ohio  River  with 
the  Mississippi  River.  Thus  the  Ohio  is  assumed  to  have  a  rectangular 
channel  1000  feet  in  width  and  a  constant  slope  of  .5  feet/mile.  In 
Manning's  formula  for  the  resistance  the  constant  y  is  assumed  given 
by  y  =.  (1.49/n)2  in  terms  of  Manning's  roughness  coefficient  n,  and 
n  is  given  the  value  0.03.  The  upstream  branch  of  the  Mississippi  was 
taken  the  same  in  all  respects  as  the  Ohio,  but  the  downstream  branch 
is  assumed  to  have  twice  the  breadth,  i.e.  2000  feet,  and  its  slope  to 


458  WATER   WAVES 

have  a  slightly  smaller  value,  i.e.  0.49  feet/mile  instead  of  0.5  feet/mile. 
With  these  values  of  the  parameters,  a  flow  having  the  same  uniform 
depth  of  20  feet  in  all  three  branches  is  possible— the  choice  of  the 


Lower 
Mississippi 


Fig.  11.2.1.  Junction  of  Ohio  and  Mississippi  Rivers 

value  0.49  feet/mile  for  the  slope  of  the  downstream  branch  of  the 
Mississippi  River  was  in  fact  made  in  order  to  ensure  this.  Later  on 
we  intend  to  calculate  the  progress  of  a  flood  which  originates  at  a 
moment  when  the  flow  is  such  a  uniform  flow  of  depth  20  feet.  The 
flood  wave  will  be  supposed  to  initiate  at  a  point  50  miles  up  the  Ohio 
from  the  junction  and  to  be  such  that  the  Ohio  rises  rapidly  at  that 
point  from  the  initial  depth  of  20  feet  to  a  depth  of  40  feet  in  4  hours. 
A  wave  then  moves  down  the  Ohio  to  the  junction  and  creates  waves 
which  travel  both  upstream  and  downstream  in  the  Mississippi  as  well 
as  a  reflected  wave  which  travels  back  up  the  Ohio.  After  a  long  time 
we  would  expect  a  steady  state  to  develop  in  which  the  depth  at  the 
point  50  miles  up  the  Ohio  is  40  feet,  while  the  depth  far  upstream  in 
the  Mississippi  would  be  the  original  value,  i.e.  20  feet  (since  we  would 
not  expect  a  retardation  of  the  flow  far  upstream  because  of  an  inflow  at 
the  junction).  Downstream  in  the  Mississippi  we  expect  a  change  in  the 
flow  extending  to  infinity.  It  is  the  steady  flow  with  these  latter  charac- 
teristics that  we  wish  to  calculate  in  the  present  section.  ( See  Fig.  11.2.1) 
We  remark  first  of  all  that  the  stream  velocities  in  all  of  the  three 
branches  will  always  be  subcritical— in  fact,  they  are  of  the  order  of 
a  few  miles  per  hour  while  the  critical  velocities  \/gt/  are  of  the  order 
of  15  to  25  miles  per  hour.  It  follows  that  the  quantity  g  —  D2/t/3  in  the 
integrand  of  the  basic  formula  (11.2.5)  for  the  river  profiles  (i.e.  the 
curve  of  the  free  surface)  is  always  positive.  The  integrand  I(y)  in 


MATHEMATICAL   HYDRAULICS 


459 


that  formula  has  the  general  form  indicated  by  Fig.  11.2.2  in  the  case 
of  flows  at  subcritical  velocities.  The  vertical  asymptote  corresponds 
to  the  value  of  y  for  which  a  steady  flow  of  constant  depth  exists 


Fig.  11.2.2.  The  integrand  iu  the  wave  profile  formula 

(cf.  (11.2.4)),  since  the  square  bracket  (the  denominator  in  the  inte- 
grand )  vanishes  for  this  value.  It  follows  that  x  can  become  positive 
infinite  for  finite  values  of  y  only  if  y  takes  on  somewhere  this  value; 
but  in  that  case  we  have  seen  that  the  whole  flow  is  then  one  with 
constant  depth  everywhere.  Consequently  the  downstream  side  of  the 
Mississippi  carries  a  flow  of  constant  speed  and  depth,  though  the 
values  of  these  quantities  are  not  known  in  advance.  However,  in  the 
upstream  branch  of  the  Mississippi  the  flow  need  not  be  constant,  and 
of  course  we  do  not  expect  it  to  be  constant  in  the  Ohio:  in  these 
branches  x  must  be  taken  to  be  decreasing  on  going  upstream  and 
consequently  the  negative  portion  of  I(y)  indicated  in  Fig.  11.2.2 
comes  into  use  since  we  may,  and  do,  set  x  =  0  at  the  junction. 
For  the  sake  of  convenience  we  use  subscripts  1,  2,  and  3  to  refer  to 
all  quantities  in  the  Ohio,  the  upstream  branch  of  the  Mississippi,  and 
the  downstream  branch  of  the  Mississippi  respectively.  The  conditions 
to  be  satisfied  at  the  junction  are  chosen  to  be 

(11.2.6)  *  =  y,  =  y,  =  y, 

(11.2.7)  D1+D2 


460  WATER    WAVES 

The  first  condition  simply  requires  the  water  level  to  have  the  same 
value  yj  (which  is,  however,  not  known  in  advance)  in  all  three  bran- 
ches, while  the  second  states,  upon  taking  account  of  the  first  condition, 
that  the  combined  discharge  of  the  two  tributaries  makes  up  the  total 
discharge  in  the  main  stream.  The  quantity  Z)2,  the  discharge  in  the 
upper  Mississippi,  is  known  since  the  flow  far  upstream  in  this  branch 
is  supposed  known— i.e.  it  is  a  uniform  flow  of  depth  20  feet. 

By  using  (11.2.7)  in  (11.2.4)  as  applied  to  the  lower  branch  of  the 
Mississippi  (in  which  the  flow  is  known  to  be  constant)  we  have 


Next,  we  write  equation  (11.2.5)  for  the  50-mile  stretch  of  the  Ohio 
which  ends  at  the  point  where  the  depth  in  that  branch  was  prescribed 
to  be  40  feet  (and  which  was  the  point  of  initiation  of  a  flood  wave); 
the  result  is 

(11.2.9)  50-  \    I(y,Dl,Bl)dy 


in  which  it  is  indicated  that  D  and  B  (as  well  as  all  other  parameters) 
are  to  be  evaluated  for  the  Ohio;  the  quantity  y  has  the  value  40/5280 
in  miles.  Equations  (11.2.8)  and  (11.2.9)  are  two  equations  containing 
j/;-  and  D!  as  unknowns,  since  Z)2  is  known.  They  were  solved  by  an 
iterative  process,  i.e.  by  taking  for  Dl  an  estimated  value,  determining 
a  value  for  yj  from  (11.2.9),  reinserting  this  value  in  (11.2.8)  to  deter- 
mine a  new  value  for  Z)1,  etc.  The  results  obtained  by  such  a  calcula- 
tion are  as  follows: 

y\  =  2/2  =  2/3  =  Vt  =  81-2  feet 
vl  =  4.83  miles/hour,    v2  =  1.53  miles/hour,    v3  ~  3.18  miles/hour. 

The  profiles  of  the  river  surface  can  now  be  computed  from  (11.2.5); 
the  results  are  given  in  Fig.  11.2.3. 

The  solution  of  the  mathematical  problem  has  the  features  we 
would  expect  in  the  physical  problem.  The  flow  velocity  and  stage  are 
increased  at  the  junction,  even  quite  noticeably,  by  the  influx  from 
the  Ohio.  Upstream  in  the  Mississippi  the  stage  decreases  rather 
rapidly  on  going  away  from  the  junction,  and  very  little  backwater 
effect  is  noticeable  at  distances  greater  than  50  miles  from  the  junc- 
tion. This  illustrates  a  fact  of  general  importance,  i.e.  that  backwater 


MATHEMATICAL   HYDRAULICS 


461 


effects  in  long  rivers  arising  from  even  fairly  large  discharges  of  tri- 
butaries into  the  main  stream  do  not  persist  very  far  upstream, 
but  such  an  influx  has  an  influence  on  the  flow  far  downstream. 
For  unsteady  motions  this  general  observation  also  holds,  and  is  in 
fact  one  of  the  basic  assumptions  used  by  hydraulics  engineers  in 


Ohio 


Upstream 
Mississippi 


y  feet 


40 


Downstream 
Mississippi 


4- 


,  -50  Junction  50  100        miles 

Fig.  11.2.3.  Steady  flow  profile  in  a  model  of  the  Ohio  and  Mississippi  Rivers 

computing  the  passage  of  flood  waves  down  rivers  (a  process  called 
flood  routing  by  them).  Later  on,  in  sec.  6  of  this  chapter,  we  shall 
deal  with  the  unsteady  motion  described  above  in  our  model  of  the 
Ohio-Mississippi  system,  and  we  will  see  that  the  unsteady  motion 
approaches  the  steady  motion  found  here  as  the  time  increases. 

11.3.  Progressing  waves  of  fixed  shape.  Roll  waves 

In  addition  to  the  uniform  steady  flows  treated  above  there  also 
exist  a  variety  of  possible  flows  in  uniform  channels  in  the  form  of 
progressing  waves  moving  downstream  at  constant  speed  without 
change  in  shape.  Such  waves  arc  expressed  mathematically  by  depths 
y(x,  t)  and  velocities  v(x9t)  in  the  form 
(11.8.1)  y(x,  t)  =  y(x  —  Ut)<  v(x,  t)  =  v(x  —  Ut),  U  =  const. 


462  WATEE   WAVES 

The  constant  U  is  of  course  the  propagation  speed  of  the  wave  as 
viewed  from  a  fixed  coordinate  system;  if  viewed  from  a  coordinate 
system  moving  downstream  with  constant  velocity  U  the  wave  profile 
would  appear  fixed,  and  the  flow  would  appear  to  be  a  steady  flow 
relative  to  the  moving  system.  It  is  convenient  to  introduce  the  va- 
riable £  by  setting 

(11.3.2)  C  =  x  -  Ut 

so  that  y  and  v  are  functions  of  £  only.  In  this  case  the  equations  of 
continuity  and  motion  given  by  (11.1.6)  and  (11.1.8)  become,  for  a 
rectangular  channel  of  fixed  breadth  and  slope: 


m  oo^  f    (*>  -  U)y:  +  yvc  =  0, 

(11.3.3)  |       ^  ^        ^ 

with  «S  the  slope  of  the  channel  and  Sf  defined,  as  before,  by 

(11.3.4)  Sf= 


The  first  equation  of  (11.3.3)  can  be  integrated  to  yield 
(11.3.5)  (v  —  U)y  =  D  =  const. 

as  one  readily  verifies,  and  the  second  equation  then  takes  the  form 


The  first  order  differential  equation  (11.3.6)  has  a  great  variety  of 
solutions,  which  have  been  studied  extensively,  for  example  by  Tho- 
mas [T.I],  but  most  of  them  are  not  very  interesting  from  the  physical 
point  of  view.  However,  one  type  of  solution  of  (11.3.6)  is  particularly 
interesting  from  the  point  of  view  of  the  applications,  and  we  there- 
fore proceed  to  discuss  it  briefly.  The  solution  in  question  furnishes 
the  so-called  monoclinal  rising  flood  wave  in  a  uniform  channel  (see 
the  article  by  Gilcrest  in  the  book  of  Rouse  [R.ll,  p.  644]).  This,  as 
its  name  suggests,  is  a  progressing  wave  the  profile  of  which  tends  to 
different  constant  values  (and  the  flow  velocity  v  also  to  different  con- 
stant values)  downstream  and  upstream,  with  the  lower  depth  down- 
stream, connected  by  a  steadily  falling  portion,  as  indicated  schema- 
tically in  Fig.  11.3.1.  In  this  wave  the  propagation  speed  U  is  larger 


MATHEMATICAL   HYDRAULICS 


463 


than  the  flow  velocity  v.  It  is  always  a  possible  type  of  solution  of 
(11 .3.6)  if  the  speed  of  propagation  of  the  wave  relative  to  the  flow  is 


Fig.  11.3.1.  Monoclinal  rising  flood  wave 

subcritical,  i.e.  if  (£7— v)2  is  less  than  gy,  in  which  case  the  coefficient 
of  the  first  derivative  term  in  (11.3.6)  is  seen  to  be  positive.  This  can 
be  verified  along  the  following  lines.  The  differential  equation  can  be 
solved  explicitly  for  f  as  a  function  of  y\ 


(11.3.7) 


-f 

Jv* 


with  the  integrand  I(y)  defined  in  the  obvious  manner;  here  y*  is  the 
value  of  y  corresponding  to  £  —  0.  The  function  I(y)  has  the  general 


Fig.   11.8.2.  The  integrand  I(y)  for  a  monoclinal  wave 


464  WATER   WAVES 

form  shown  in  Fig.  11.3.2  if  the  propagation  speed  U  and  the  constant 
D  in  (11.8.5)  are  chosen  properly.  The  main  point  is  that  the  curve  has 
two  vertical  asymptotes  at  y  —  yQ  and  y  =  yl  between  which  I(y) 
is  negative.  By  choosing  y*  between  y0  and  yl  we  can  hope  that 
f  -*  +  oo  as  y  -*  t/0,  while  £  ->  —  oo  as  y  ->  yx:  all  that  is  necessary 
is  that  I(y)  becomes  infinite  at  j/0  and  yl  of  sufficiently  high  order. 
This  is,  in  fact,  the  case;  we  can  select  values  of  D  and  U  in  such  a 
way  that  I(y)  becomes  infinite  at  t/0  and  yl  through  setting  the  quan- 
tity Sj  —  S  in  (11.3.6)  equal  to  zero,  i.e.  by  choosing  D  and  U  such  that 


(11.3.8) 


For  given  positive  values  of  z/0  and  yl  these  are  a  pair  of  linear  equa- 
tions (after  taking  a  square  root)  which  determine  U  and  D  uniquely. 
An  elementary  discussion  of  the  possible  solutions  of  these  equations 
shows  that  U  must  be  positive  and  D  negative,  and  this  means,  as  wo 
see  from  (11.3.5),  that  U  is  larger  than  v9  i.e.  the  propagation  speed  of 
the  wave  is  greater  than  the  flow  speed. 

By  taking  the  numerical  data  for  the  model  of  the  Ohio  given  in  the 
preceding  section  and  assuming  the  depth  far  upstream  to  be  40  feet, 
far  downstream  20  feet,  it  was  found  that  the  corresponding  mono- 
clinal  flood  wave  in  the  Ohio  would  propagate  with  a  speed  of  5 
miles/hour.  The  shape  of  the  wave  will  be  given  later  in  sec.  6  of  this 
chapter,  where  it  will  be  compared  with  an  unsteady  wave  obtained  by 
gradually  raising  the  level  in  the  Ohio  at  one  point  from  20  feet  to 
40  feet,  then  holding  the  level  fixed  there  at  the  latter  value,  and  cal- 
culating the  downstream  motion  which  results.  We  shall  see  that  the 
motion  tends  to  the  monoclinal  flood  wave  obtained  in  the  manner 
just  now  described.  Thus  the  unsteady  wave  tends  to  move  eventually 
at  a  speed  of  about  5  miles/hour,  while  on  the  other  hand,  as  we  know 
from  Chapter  10  (and  will  discuss  again  later  on  in  this  chapter),  the 
propagation  speed  of  small  disturbances  relative  to  the  stream  is  <\/gy 
and  hence  is  considerably  larger  in  the  present  case,  i.e.  of  the  order 
of  15  to  25  miles/hour.  This  important  and  interesting  point  will  be 
discussed  in  sec.  6  below. 


MATHEMATICAL   HYDRAULICS 


465 


Fig.  11.8.8.  Roll  waves,  looking  down  stream  (The  Grunnbach,  Switzerland) 


466  WATER   WAVES 

We  turn  next  to  another  type  of  progressing  waves  in  a  uniform 
channel  which  can  be  described  with  the  aid  of  the  differential  equa- 
tion (11.3.6),  i.e.  the  type  of  wave  called  a  roll  wave.  A  famous  exam- 
ple of  such  waves  is  shown  in  Fig.  11.3.3,  which  is  a  photograph  taken 
from  a  book  of  Cornish  [C.7],  and  printed  here  by  the  courtesy  of  the 
Cambridge  University  Press.  As  one  sees,  these  waves  consist  of  a 
series  of  bores  (cf.  Chapter  10.7)  separated  by  stretches  of  smooth 
flow.  The  sketch  of  Fig.  11.3.4  indicates  this  more  specifically.  Such 


Fig.  11.3.4.  Roll  waves 

waves  frequently  occur  in  sufficiently  steep  channels  as,  for  example, 
spill-ways  in  dams  or  in  open  channels  such  as  that  of  Fig.  11.3.3. 
Roll-waves  sometimes  manifest  themselves  in  quite  unwanted  places, 
as  for  example  in  the  Los  Angeles  River  in  California.  The  run-off  from 
the  steep  drainage  area  of  this  river  is  carried  through  the  city  of  Los 
Angeles  by  a  concrete  spill-  way;  in  the  brief  rainy  season  a  large 
amount  of  water  is  carried  off  at  high  velocity.  It  sometimes  happens 
that  roll  waves  occur  with  amplitudes  high  enough  to  cause  spilling 
over  the  banks,  though  a  uniform  flow  carrying  the  same  total  amount 
of  water  would  be  confined  to  the  banks.  The  phenomenon  of  roll 
waves  thus  has  some  interest  from  a  practical  as  well  as  from  a  theore- 
tical point  of  view;  we  proceed  to  give  a  brief  treatment  of  it  in  the 
remainder  of  this  section  following  the  paper  of  Dressier  [D.12].  In 
doing  so,  we  follow  Dressier  in  taking  what  is  called  the  Ch^zy  for- 
mula for  the  resistance  rather  than  Manning's  formula,  as  has  been 
done  up  to  now.  The  Ch£zy  formula  gives  the  quantity  Sf  the  following 
definition: 


in  which  r2  is  a  "roughness  coefficient"  and  R  is,  as  before,  the  hy- 
draulic radius.  For  a  very  broad  rectangular  channel,  the  only  case 
we  consider,  R  =  y.  Under  these  circumstances  the  differential 
equation  (11.3.6)  takes  the  form 


MATHEMATICAL    HYDRAULICS  467 

sS_^(Uy+D)\Uy+D\ 

0  o  " 


dy 


Z)2 

-    _ 

q 


as  can  be  readily  seen. 

It  is  natural  to  inquire  first  of  all  whether  (11.3.10)  admits  of  solu- 
tions which  are  continuous  periodic  functions  of  £  since  this  is  the 
general  type  of  motion  we  seek.  There  are,  however,  no  such  periodic 
and  continuous  solutions  (cf.  the  previously  cited  paper  of  Thomas 
[T.I])  of  the  equations;  in  fact,  since  the  right  hand  side  of  (11.3.10) 
can  be  expressed  as  the  quotient  of  cubic  polynomials  in  y  the 
types  of  functions  which  arise  on  integrating  it  are  linear  combina- 
tions of  the  powers,  the  logarithm,  and  the  arc  tangent  function  and 
one  hardly  expects  to  find  periodic  functions  on  inverting  solutions 
C(j/)  of  this  type.  This  fact,  together  with  observations  of  roll  waves  of 
the  kind  shown  in  Fig.  11.3.3,  leads  one  to  wonder  whether  there  might 
not  be  discontinuous  periodic  solutions  of  (11.3.10)  with  discontinui- 
ties in  the  form  of  bores,  which  should  be  fitted  in  so  that  the  discon- 
tinuity or  shock  conditions  described  in  sec.  6  of  the  preceding  chap- 
ter *  are  satisfied.  This  Dressier  shows  to  be  the  case;  he  also  gives  a 
complete  quantitative  analysis  of  the  various  possibilities. 

The  starting  point  of  the  investigation  is  the  observation,  due  to 
Thomas  [T.I],  that  only  quite  special  types  of  solutions  of  (11.3.10) 
come  in  question  once  the  roll  wave  problem  has  been  formulated  in 
terms  of  a  periodic  distribution  of  bores.  In  fact,  we  know  from  Chap- 
ter 10  that  the  flow  relative  to  a  bore  must  be  subcritical  behind  a 
bore  but  supercritical  in  front  of  it;  consequently  there  must  be  an  in- 
termediate point  of  depth  j/0,  say,  (cf.  Fig.  11.3.4)  where  the  smooth 
flow  has  the  critical  speed,  i.e.  where 

(11.3.11)  K-  tf)2  =  g2/0> 

since  C7,  the  speed  of  the  bore,  coincides  with  the  propagation  speed 
of  the  wave.  At  such  a  point  the  denominator  on  the  right  hand  side 
of  (11.3.10)  vanishes,  since  D  =  (v  —  U)y9  and  hence  dy/d£  would  be 
infinite  there—  contrary  to  the  observations—  unless  the  numerator 
of  the  right  hand  side  also  vanishes  at  that  point.  The  right  hand  side 
can  now  be  written  as  a  quotient  of  cubic  polynomials,  and  we  know 

*  The  shock  conditions  were  derived  in  Chapter  10  under  the  assumption 
that  no  resistances  were  present.  As  one  would  expect,  the  resistance  terms  play 
no  role  in  shock  conditions,  as  Dressier  [D.12]  verifies  in  his  paper. 


468  WATER   WAVES 

that  numerator  and  denominator  have  yQ  as  a  common  root;  it  follows 
that  a  factor  y  —  yQ  can  be  cancelled  and  the  differential  equation 
then  can  be  put  in  the  form 


after  a  little  algebraic  manipulation.  Since  the  denominator  on  the 

right  hand  side  is  positive  and  since  we  seek  solutions  for  which 

dy/d£  is  everywhere  (cf.  Fig.  11.3.4)  positive,  it  follows  in  particular 

that  the  quadratic  in  the  numerator  must  be  positive  for  y  =  yQ.  This 

leads  to  the  following  necessary  condition  for  the  formation  of  roll 

waves 

(11.3.13)  4r2  <  5, 

obtained  by  using  (11.3.11  )  and  other  relations.  A  practically  identical 
inequality  was  derived  by  Thomas  on  the  basis  of  the  same  type  of 
reasoning.  The  inequality  states  that  the  channel  roughness,  which  is 
larger  or  smaller  with  r2,  must  not  be  too  great  in  relation  to  the  steep- 
ness of  the  channel,  and  this  corroborates  observations  by  Rouse 
[R.10]  that  roll  waves  can  be  prevented  by  making  a  channel  suffi- 
ciently rough.  Dressier  also  shows  in  his  paper  that  it  is  important  for 
the  formation  of  roll  waves  that  the  friction  force  for  the  same  rough- 
ness coefficient  and  velocity  should  increase  when  the  depth  decreases; 
he  finds,  in  fact,  that  roll  waves  would  not  occur  if  the  hydraulic  ra- 
dius R  in  the  Ch£zy  formula  (11.3.9)  were  to  be  assumed  independent 
of  the  depth  y. 

Dressier  goes  on  in  his  paper  to  show  that  smooth  solutions  of 
(11.3.12)  can  be  pieced  together  through  bores  in  such  a  way  that  the 
conditions  referring  to  continuity  of  mass  and  momentum  across  the 
discontinuity  are  satisfied  as  well  as  the  inequality  requiring  a  loss 
rather  than  a  gain  in  energy.  For  the  details  of  the  calculations  and  a 
quantitative  analysis  in  terms  of  the  parameters,  the  paper  of  Dressier 
should  be  consulted,  but  a  few  of  the  results  might  be  mentioned  here. 
Once  the  values  of  the  slope  S  and  the  roughness  coefficient  r2  are 
prescribed  by  the  physical  situation,  and  the  wave  propagation  speed 
U  is  arbitrarily  given,  there  exists  a  one-parameter  family  of  possible 
roll-waves.  As  parameter  the  wave  length  A,  i.e.  the  distance  between 
two  successive  bores,  can  be  chosen;  if  this  parameter  is  also  fixed, 
the  roll  wave  solution  is  uniquely  determined.  A  specific  solution 


MATHEMATICAL   HYDRAULICS  469 

would  also  be  fixed  if  the  time  period  of  the  oscillation  were  to  be 
fixed  together  with  one  other  parameter— the  average  discharge  rate, 
say.  Perhaps  it  is  in  this  fashion  that  the  roll  waves  are  definitely 
fixed  in  some  cases  — for  example,  the  roll  waves  down  the  spill- way 
of  a  dam  are  perhaps  fixed  by  the  period  of  surface  waves  in  the  dam 
at  the  crest  of  the  spill-way.  Schonfeld  [S.4a]  discusses  the  problem 
from  the  point  of  view  of  stability  and  arrives  at  the  conclusion  that 
only  one  of  the  solutions  obtained  by  Dressier  would  be  stable,  and 
hence  it  would  be  the  one  likely  to  be  observed. 

11.4.  Unsteady  flows  in  open  channels.  The  method  of  characteristics 

In  treating  unsteady  flows  it  becomes  necessary  to  integrate  the 
nonlinear  partial  differential  equations  (11.1.1)  and  (11.1.6)  for  pre- 
scribed initial  and  boundary  conditions.  It  has  already  been  mentioned 
that  such  problems  fall  into  the  same  category  as  the  problems  treated 
in  the  preceding  chapter,  since  they  are  of  hyperbolic  type  in  two 
independent  variables  and  thus  amenable  to  solution  by  the  method 
of  characteristics.  It  is  true  that  the  equations  (11.1.1)  and  (11.1.6) 
are  more  complicated  than  those  of  Chapter  10  because  of  the  occur- 
rence of  the  variable  coefficient  A  and  of  the  resistance  term,  so  that 
solutions  of  the  type  called  simple  waves  (cf.  Ch.  10.3)  do  not  exist  for 
these  equations.  Nevertheless  the  theory  of  characteristics  is  still 
available  and  leads  to  a  variety  of  valuable  and  pertinent  observa- 
tions regarding  the  integration  theory  of  equations  (11.1.1)  and  (11.1. 
6)  which  are  very  important.  The  essential  facts  have  already  been 
stated  in  Chapter  10.2,  but  we  repeat  them  briefly  here  for  the  sake 
of  preserving  the  continuity  of  the  discussion.  Our  emphasis  in  this 
chapter  is  on  numerical  solutions,  which  can  be  obtained  by  operating 
with  the  characteristic  form  of  the  differential  equations,  but  since  we 
shall  not  actually  use  the  characteristic  form  for  such  purposes  we 
shall  base  the  discussion  immediately  following  on  a  special  case,  al- 
though the  results  and  observations  are  applicable  in  the  most  general 
case.  The  special  case  in  question  is  that  of  a  river  of  constant  rectan- 
gular section  and  uniform  slope,  with  no  flow  into  the  river  from  the 
banks  (i.e.  q  =  0  in  (11.1.2)  and  (11.1.6)).  In  this  case  the  differential 
equations  can  be  written  as  follows: 

(11.4.1)  vxy  +vyx  +  yt  =  0, 

(11.4.2)  vt  +  vvx  +  gyx  +  E  =  0. 


470  WATER   WAVES 

We  have  introduced  the  symbol  E  for  the  external  forces  per  unit 
mass: 

(11.4.3)  E  =  -  gS  +  gSf9         S  =  const. 

The  term  E  differs  from  the  others  in  that  it  contains  no  derivatives 
of  y  or  v. 

The  theory  of  characteristics  for  these  equations  can  be  approached 
very  directly  *  in  the  present  special  case  by  introducing  a  new 
quantity  c  to  replace  y,  as  follows: 

(11.4.4)  c  -  ^gy. 

This  quantity  has  great  physical  significance,  since  it  represents— as 
we  have  seen  in  Chapter  10— the  propagation  speed  of  small  disturb- 
ances in  the  river.  From  (11.4.4)  we  obtain  at  once  the  relations 

(11-4.5)  2ccx  =  gyX9         2cct  =  gyt, 

and  the  differential  equations  (11.4.1)  and  (11.4.2)  take  the  form 

2cc  x  +  vt  +  vvx  +  E  =  0, 


(11.4.6) 

'   2w?a  +  2ct  =  0. 

These  equations  are  next  added,  then  subtracted,  to  obtain  the  follow- 
ing equivalent  pair  of  equations: 

a 

—   — I —  — —   > 

dx       dt  j 
(11.4.7) 


We  observe  that  the  derivatives  in  these  equations  now  have  the  form 
of  directional  derivatives— indeed,  to  achieve  that  was  the  purpose  of 
the  transformation— so  that  c  and  v  in  the  first  equation,  for  example, 
are  both  subject  to  the  operator  (c  +  v)d/dx  +  d/dt,  which  means 
that  these  functions  are  differentiated  along  curves  in  the  a%  £-plane 
which  satisfy  the  differential  equation  dx/dt  =  c  +  v.  In  similar 
fashion,  the  functions  c  and  v  in  the  second  equation  are  both  subject 
to  differentiation  along  curves  satisfying  the  differential  equation 
dx/dt  =  —  c  +  fl- 
it is  entirely  feasible  to  develop  the  integration  theory  of  equations 
(11.4.7)  quite  generally  on  the  basis  of  these  observations  (as  is  done, 
for  example,  in  Courant-Friedrichs  [C.9,  Ch.  2]),  but  it  is  simpler,  and 
leads  to  the  same  general  results,  to  describe  it  for  the  special  case  in 

*  For  a  treatment  which  shows  quite  generally  how  to  arrive  at  the  for- 
mulation of  the  characteristic  equations,  see  Courant-Friedrichs   [C.9,  Ch.  2]. 


MATHEMATICAL   HYDRAULICS  471 

which  the  resistance  force  Ff  is  neglected  so  that  the  quantity  E  in 
(11.4.7)  is  a  constant  (see  (11.4.3)).  In  this  case  the  equations  (11.4.7) 
can  be  written  in  the  form 


(11.4.7  )t 


as  one  can  readily  verify.  But  the  interpretation  of  the  operations  de- 
fined in  (11.4.7)!  has  just  been  mentioned:  the  relations  state  that  the 
functions  (v  ±  2c  +  Et)  are  constant  for  a  pbint  moving  through  the 
fluid  with  the  velocity  (v  ±  c),  or,  as  we  may  also  put  it,  for  a  point 
whose  motion  in  the  x9  /-plane  is  characterized  by  the  ordinary  dif- 
ferential equations  dxjdt  =  v  ±  c.  That  is,  we  have  the  following 
situation  in  the  tr,  /-plane:  There  are  two  sets  of  curves,  Cl  and  C2, 
called  characteristics,  which  are  the  solution  curves  of  the  ordinary 
differential  equations 

dx  _ 

—  =t;-4-c,  and 

dt 
(11.4.8) 

"c-    dX -v     t 

L/2-       — "  —  ' 

and  we  have  the  relations 

'•v-\-2c-}-Et=kl= const,  along  a  curve  Cx  and 


(11.4.9) 

1  v  —  2c+Et =k2=^ const,  along  a  curve  C2. 

Of  course  the  constants  k^  and  k2  will  be  different  on  different  curves 
in  general.  It  should  also  be  observed  that  the  two  families  of  charac- 
teristics determined  by  (11.4.8)  arc  really  distinct  because  of  the  fact 
that  c  —  \/gy  ^  0  since  we  suppose  that  y  >  0,  i.e.  that  the  water 
surface  never  touches  the  bottom. 

By  reversing  the  above  procedure  it  can  be  seen  rather  easily  that  the 
system  of  relations  (11.4.8)  and  (11.4.9)  is  completely  equivalent  to 
the  system  of  equations  (11.4.6)  for  the  case  of  constant  bottom  slope 
and  zero  resistance,  so  that  a  solution  of  either  system  yields  a  solution 
of  the  other.  In  fact,  if  we  set  /(#,  /)  =  v  +  2c  +  Et  and  observe  that 
f(x,  t)  =  k±  =  const,  along  any  curve  x  —  x(t)  for  which  dxjdt  = 
v  +  c  it  follows  that  along  such  curves 


(11.4.10) 

dt 


472  WATER   WAVES 

In  the  same  way  the  function  g(x,  t)  =  v  —  2c  +  Et  satisfies  relation 
(11.4.11)  ft  +  (v  -  c)gx  =  0 

along  the  curves  for  which  dx/dt  =  v  —  c.  Thus  wherever  the  curve 
families  Cl  and  C2  cover  the  as,  J-plane  in  such  a  way  as  to  furnish  a 
curvilinear  coordinate  system  the  relations  (11.4.10)  and  (11.4.11) 
hold.  If  now  equations  (11.4.10)  and  (11.4.11)  are  added  and  the 
definitions  of  f(x9 1)  and  g(x,  t)  are  recalled  it  is  readily  seen  that  the 
first  of  equations  (11.4.6)  results.  By  subtracting  (11.4.11)  from 
(11.4.10)  the  second  of  equations  (11.4.6)  is  obtained.  In  other  words, 
any  functions  v  and  c  which  satisfy  the  relations  (11.4.8)  and  (11.4.9) 
will  also  satisfy  (11.4.6)  and  the  two  systems  of  equations  are  there- 
fore now  seen  to  be  completely  equivalent. 

As  we  would  expect  on  physical  grounds,  a  solution  of  the  original 
dynamical  equations  (11.4.6)  could  be  shown  to  be  uniquely  deter- 
mined when  appropriate  initial  conditions  (for  t  =  0,  say)  and  boun- 
dary conditions  are  prescribed;  it  follows  that  any  solutions  of  (11.4. 
8)  and  (11.4.9)  are  also  uniquely  determined  when  such  conditions 
are  prescribed  since  we  know  that  the  two  systems  of  equations  are 
equivalent. 

At  first  sight  one  might  be  inclined  to  regard  the  relations  (11.4.8) 
and  (11.4.9)  as  more  complicated  than  the  original  pair  of  partial 
differential  equations,  particularly  since  the  right  hand  sides  of  (11.4.8) 
are  not  known  and  hence  the  characteristic  curves  are  also  not  known. 
Nevertheless,  the  formulation  in  terms  of  the  characteristics  is  quite 
useful  in  studying  properties  of  the  solutions  and  also  in  studying 
questions  referring  to  the  appropriateness  of  various  boundary  and 
initial  conditions.  In  Chapter  10.2  a  detailed  discussion  along  these 
lines  is  given;  we  shall  not  repeat  it  here,  but  will  summarize  the  con- 
clusions. The  description  of  the  properties  of  the  solution  is  given  in 
the  x9  J-plane,  as  indicated  in  Fig.  11.4.1.  In  the  first  place,  the  values 
of  v  and  c  at  any  point  P(x9  t)  within  the  region  of  existence  of  the  solu- 
tion are  determined  solely  by  the  initial  values  prescribed  on  the  segment 
of  the  x-axis  which  is  subtended  by  the  two  characteristics  issuing  from  P. 
In  addition,  the  two  characteristics  issuing  from  P  are  themselves  also 
determined  solely  by  the  initial  values  on  the  segment  subtended  by 
them.  Such  a  segment  of  the  ff-axis  is  often  called  the  domain  of  de- 
pendence of  the  point  P.  Correspondingly  we  may  define  the  range  of 
influence  of  a  point  Q  on  the  #-axis  as  the  region  of  the  x,  £-plane  in 
which  the  values  of  v  and  c  are  influenced  by  the  initial  values  assigned 


MATHEMATICAL   HYDRAULICS 


473 


to  point  Q,  i.e.,  it  is  the  region  between  the  two  characteristics  issuing 
from  Q.  In  Fig.  11.4.1  we  indicate  these  two  regions. 


t 


Range  of  influence  of   Q 


Domain  of 
determmacy 


J 


Domain  of   dependence    of    P 


Fig.  11.4.1.  The  role  of  the  characteristics 

We  are  now  in  a  position  to  understand  the  role  of  the  charac- 
teristics as  curves  along  which  discontinuities  in  the  first  and  higher 
derivatives  of  the  initial  data  are  propagated,  since  it  is  reasonable  to 
expect  (and  could  be  proved)  that  those  points  P  whose  domains  of 
dependence  do  not  contain  such  discontinuities  are  points  at  which 
the  solutions  v  and  c  also  have  continuous  derivatives.  On  the  other 
hand,  it  could  be  shown  that  a  discontinuity  in  the  initial  data  at  a 
certain  point  docs  not  in  general  die  out  along  the  characteristic 
issuing  from  that  point.  Such  a  discontinuity  (or  disturbance  in  the 
water)  therefore  spreads  in  both  directions  over  the  surface  of  the 
water  with  the  speed  v  +  c  in  one  direction  and  v  —  c  in  the  other  in 
view  of  the  interpretation  given  to  the  characteristics  through  the 
relations  (11 .4.7  )r  Since  v  is  the  velocity  of  the  water  particles  we  see 
that  c  represents  quite  generally  the  speed  at  which  a  discontinuity 
in  a  derivative  of  the  initial  data  propagates  relative  to  the  moving 
water.  We  are  therefore  justified  in  referring  to  the  quantity  c  =  ^/gy 
as  the  wave  speed  or  propagation  speed. 

We  considered  above  a  problem  in  which  only  initial  conditions, 
and  no  boundary  conditions,  were  prescribed.  In  the  problems  we 
consider  later,  however,  such  boundary  conditions  will  occur  in  the 
form  of  conditions  prescribed  at  a  certain  fixed  point  of  the  river  in 
terms  of  the  time:  for  example,  the  height,  or  stage,  of  the  river  might 
be  given  at  a  certain  station  as  a  function  of  the  time.  In  other  words, 


474  WATER   WAVES 

conditions  would  be  prescribed  not  only  along  the  #-axis  of  our  x,  t- 
plane,  but  also  along  the  £-axis  (in  general  only  for  t  >  0)  for  a  certain 
fixed  value  of  x.  The  method  of  finite  differences  used  in  Chapter  10.2 
to  discuss  the  initial  value  problem,  with  the  general  result  given  above, 
can  be  modified  in  an  obvious  way  to  deal  with  cases  in  which  bound- 
ary conditions  are  also  imposed.  In  doing  so,  it  would  also  become 
clear  just  what  kind  of  boundary  conditions  could  and  should  be  im- 
posed. For  example,  in  the  great  majority  of  rivers— in  fact,  for  all 
in  which  the  flow  is  subcritical,  i.e.  such  that  v  is  everywhere  less  than 
the  wave  speed  \/^jy—it  is  possible  to  prescribe  only  one  condition 
along  the  J-axis,  which  might  be  either  the  velocity  v  or  the  depth  y, 
in  contrast  with  the  necessity  to  impose  two  conditions  along  the  ay- 
axis.  This  fact  would  become  obvious  on  setting  up  the  finite  differ- 
ence scheme,  and  examples  of  it  will  be  seen  later  on. 

Finally,  it  should  be  stated  that  the  role  of  the  characteristics,  and 
also  the  method  of  finite  differences  applied  to  them  could  be  used 
with  reference  to  the  general  case  of  the  characteristic  equations  as 
embodied  in  the  equations  (11.4.7)  and  (11.4.8)  in  essentially  the 
same  way  as  was  sketched  out  above  for  the  system  comprised  of 
(11.4.8)  and  (11.4.9)  which  referred  to  a  special  case.  In  particular, 
the  role  of  the  characteristics  as  curves  along  which  small  disturbances 
propagate,  and  their  role  in  determining  the  domain  of  dependence, 
range  of  influence,  etc.  remain  the  same. 

11.5.  Numerical  methods  for  calculating  solutions  of  the  differential 
equations  for  flow  in  open  channels 

It  has  already  been  stated  that  while  the  formulation  of  our  pro- 
blems by  the  method  of  characteristics  is  most  valuable  for  studying 
many  questions  concerned  with  general  properties  of  the  solutions 
of  the  differential  equations,  it  is  in  most  cases  not  the  best  formula- 
tion to  use  for  the  purpose  of  calculating  the  solutions  numerically. 
That  is  not  to  say  that  the  device  of  replacing  derivatives  by  differ- 
ence quotients  should  be  given  up,  but  rather  that  this  device  should 
be  used  in  a  different  manner.  The  basic  idea  is  to  operate  with  finite 
differences  by  using  a  fixed  rectangular  net  in  the  x9  /-plane,  in  con- 
trast with  the  method  outlined  in  Chapter  10.2,  in  which  the  net  of 
points  in  the  x9  J-plane  at  which  the  solution  is  to  be  approximated  is 
determined  only  gradually  in  the  course  of  the  computation.  In  the 
latter  procedure  it  is  thus  necessary  to  calculate  not  only  the  values 


MATHEMATICAL    HYDRAULICS 


475 


of  the  unknown  functions  v  and  c,  but  also  the  values  of  the  coordinates 
x,  t  of  the  net  points  themselves,  whereas  a  procedure  making  use  of  a 
fixed  net  would  require  the  calculation  of  v  and  c  only,  and  it  would 
also  have  the  advantage  of  furnishing  these  values  at  a  convenient  set 
of  points. 

However,  the  question  of  the  convergence  of  the  approximate 
solution  to  the  exact  solution  when  the  mesh  width  of  a  rectangular 
net  is  made  to  approach  zero  is  more  delicate  than  it  is  when  the  meth- 
od of  characteristics  is  used.  For  example,  it  is  not  correct,  in  general, 
to  choose  a  net  in  which  the  ratio  of  the  mesh  width  At  along  the  /-axis 
and  the  mesh  width  Ax  along  the  «r-axis  is  kept  constant  independent 
of  the  solution:  such  a  procedure  would  not  in  general  yield  approxi- 
mations converging  to  the  solution  of  the  differential  equation  pro- 
blem. The  reason  for  this  can  be  understood  with  reference  to  one  of 
the  basic  facts  about  the  solution  of  the  differential  equations  which 
was  brought  out  in  the  discussion  of  the  preceding  section.  The  basic 
fact  in  question  is  the  existence  of  what  was  called  there  the  domain 
of  dependence  of  the  solution.  For  example,  suppose  the  solution  were 
to  be  approximated  at  the  points  of  the  net  of  Fig.  11. 5. la  by  advanc- 
ing from  one  row  parallel  to  the  tT-axis  to  the  next  row  a  distance  At 
from  it.  In  addition,  suppose  this  were  to  be  done  by  determining  the 
approximate  values  of  v  and  c  at  any  point  such  as  P  (cf.  Fig.  11.5.1b) 


P\C2 


I 


2  x 

0  b 

11.5.1.  Approximation  by  using  a  rectangular  net 

by  using  the  values  of  these  quantities  at  the  nearest  three  points 
0,  1, 2  in  the  next  line  below,  replacing  derivatives  in  the  two  different- 
ial equations  by  difference  quotients,  and  then  solving  the  resulting 
algebraic  equations  for  the  two  unknowns  vp  and  cp.  It  seems  reason- 
able to  suppose  that  such  a  scheme  would  be  appropriate  only  if  P 
were  in  the  triangular  region  bounded  by  the  characteristics  drawn 
from  points  0  and  2  to  form  the  region  within  which  the  solution  is  de- 


476  WATER   WAVES 

termined  solely  by  the  data  given  on  the  segment  0—2:  otherwise  it 
seems  clear  that  the  initial  values  at  additional  points  on  the  #-axis 
ought  to  be  utilized  since  our  basic  theory  tells  us  that  the  initial  data 
at  some  of  them  would  indeed  influence  the  solution  at  point  P.  On  the 
other  hand,  the  characteristic  curves  themselves  depend  upon  the 
values  of  the  unknown  functions  v  and  c— their  slopes,  in  fact,  are 
given  (cf.  (11.4.8))  by  dxjdt  —  v  ±  c  and  thus  the  interval  At  must  be 
chosen  small  enough  in  relation  to  a  fixed  choice  of  the  interval  Ax 
so  that  the  points  such  as  P  will  fall  within  the  appropriate  domains  of 
determinacy  relative  to  the  points  used  in  calculating  the  solution  at 
P.  In  other  words,  the  theory  of  characteristics,  even  if  it  is  not  used 
directly,  comes  into  play  in  deciding  the  relative  values  of  At  and  Ax 
which  will  insure  convergence  (for  rigorous  treatments  of  these 
questions  see  the  papers  by  Courant,  Isaacson,  and  Rees  [C.ll],  and 
by  Keller  and  Lax  [K.4]). 

We  shall  introduce  two  different  schemes  employing  the  method  of 
finite  differences  in  a  fixed  rectangular  net  of  the  x,  J-plane.  The  first 
of  these  makes  use  of  the  differential  equations  in  the  form  given  by 
(11.4.7),  and  we  no  longer  suppose  that  the  function  E  is  restricted  in 
any  way.  (It  might  be  noted  that  the  slopes  of  the  characteristics  as 
given  by  (11.4.8)  are  determined  by  the  quantities  v  dr  c,  no  matter 
how  the  function  E  is  defined,  and  in  fact  also  for  the  most  general 
case  of  a  river  having  a  variable  cross  section  A9  etc.,  and  hence  we  are 
in  a  position  to  determine  appropriate  lengths  for  the  ^-intervals,  in 
accord  with  the  above  discussion,  in  the  most  general  case.  This  is 
also  a  good  reason  for  working  with  the  quantity  c  in  place  of  y.) 
At  the  same  time,  the  calculation  is  based  on  assuming  that  the  ap- 


Fig.  11.5.2.  A  rectangular  net 

proximate  values  of  c  and  v  have  been  calculated  at  the  net  points 
L,  M,  R  (cf.  Fig.  11.5.2)  and  that  the  differential  equations  are  to  be 


MATHEMATICAL   HYDRAULICS  477 

used  to  advance  the  approximate  solution  to  the  point  P.  The  differ- 
ential equations  to  be  solved  are  thus 

(11.5.1)  2{(c  +  v)cx  +  ct}  +  {(c  +  v)vx  +vt}+E  =  0, 

(11.5.2)  -  2{(-  c  +  v)cx  +  ct}  +  {(-  c  +  v)vx  +  vt}  +  E  =  0, 

and  the  characteristic  directions  are  determined  by  dx/dt  =  v  ±  c. 
The  characteristic  with  slope  v  +  c  we  call  the  forward  characteristic, 
and  that  with  slope  v  —  c  the  backward  characteristic.  We  shall  re- 
place the  derivatives  in  the  equations  by  difference  quotients  which 
approximate  the  values  of  the  derivatives  at  the  point  M.  In  order  to 
advance  the  values  of  v  and  c  from  the  points  L,  M9  R  to  the  point  P 
by  using  (11.5.1)  and  (11.5.2)  it  is  natural  to  replace  the  time  deriva- 
tives vt  and  ct  by  the  following  difference  quotients 

niK«x  y,        Vp  —  vM  cp  —  CM 

(11.5.3)  „,  =  -__,      ci  =  —zr 

in  both  equations.  However,  in  order  to  insure  the  convergence  of 
the  approximations  to  the  exact  solution  when  Ax  ->•  0  and  At  ->  0 
(see  Courant,  Isaacson,  and  Rees  [C.ll]  for  a  proof  of  this  fact)  it  is 
necessary  to  replace  the  derivatives  vx  and  cx  by  difference  quotients 
which  are  defined  differently  for  (11.5.1)  than  for  (11.5.2),  as  follows: 

(11.5.4)  vx  =  V^L^9      Cx  =  ^_I_^       in  (11.5.1), 

Ax  Ax 

(11.5.5)  „.  =  *-*=!*  ,      c.  =  C-*^-**       in  (11.5.2). 

Ax  Ax 

The  reason  for  this  procedure  is,  at  bottom,  that  (11.5.1  )  is  an  equation 
associated  with  the  forward  characteristic,  while  (11.5.2)  is  associated 
with  the  backward  characteristic.  The  coefficients  of  the  derivatives 
and  the  function  E  arc,  of  course,  to  be  evaluated  at  the  point  M.  The 
difference  equations  replacing  (11.5.1)  and  (11.5.2)  are  thus  given  by 

«n.,e,    ,,.„ 


+    <„„  +„„)  !*-p  +  "J^!L  )+£(„„,  cu)  =  0, 
Ax  At       } 


(- 

,     =  o. 


478  WATEE    WAVES 

We  observe  that  the  two  unknowns,  vp  and  cp,  occur  linearly  in  these 
equations;  hence  they  are  easily  found  by  solving  the  equations.  The 
result  is 

(11.5.8)        VP^VM+  — 


(11.5.9)         Cp-cM+|— 


In  accordance  with  the  remarks  made  above,  we  must  also  require  that 
the  ratio  of  At  to  Ax  be  taken  small  enough  so  that  P  lies  within  the 
triangle  formed  by  drawing  lines  from  L  and  R  in  the  directions  of  the 
forward  and  backward  characteristics  respectively,  i.e.  lines  with  the 
slopes  VL  +  CL  at  L  and  VR  —  CR  at  jK:  a  condition  that  is  well-de- 
termined since  the  values  of  v  and  c  are  presumably  known  at  L  and  R. 

One  can  now  see  in  general  terms  how  the  initial  value  problem 
starting  at  t  =  0  can  be  solved  approximately:  One  starts  with  a  net 
along  the  #-axis  with  spacing  Ax.  Since  c  and  v  arc  known  at  all  of 
these  points,  the  values  of  c  and  v  can  be  advanced  through  use  of 
(11.5.8)  and  (11.5.9)  to  a  parallel  row  of  points  on  a  line  distant  At 
along  the  2-axis  from  the  a?-axis.  However,  the  mesh  width  At  must 
be  chosen  small  enough  so  that  the  convergence  condition  is  satisfied 
at  all  net  points  where  new  values  of  v  and  c  are  computed. 

We  can  now  see  also  how  to  take  care  of  boundary  conditions,  i.e. 
of  conditions  imposed  at  a  fixed  point  (say  at  the  origin,  x  =  0)  as 
given  functions  of  the  time.  For  example,  the  depth  y  (corresponding 
to  the  stage  of  the  river)  or  the  velocity  v  (which  together  with  the 
cross-section  area  A  fixes  the  rate  of  discharge)  might  be  given  in 
terms  of  the  time.  Initial  conditions  downstream  from  this  point  (i.e. 
for  x  >  0)  might  also  be  prescribed.  Suppose,  for  example,  that  the 
stage  of  the  river  is  prescribed  at  x  -=  0,  i.e.  that  j/(0,  t)  is  known,  and 
that  the  calculation  had  already  progressed  so  far  as  to  yield  values  of 
v  and  c  at  net  points  along  a  certain  line  parallel  to  the  #-axis  and 
containing  the  points  L,  M,  JB,  as  indicated  in  Fig.  11.5.3.  It  is  clear 
that  the  determination  of  the  values  of  v  and  c  at  point  P  can  be  ob- 
tained from  their  values  at  L,  M,  R  by  using  (11.5.8)  and  (11.5.9), 
as  in  the  above  discussion  of  the  initial  value  problem,  and  similarly 


MATHEMATICAL  HYDRAULICS  479 

for  points  Pv  P2,  etc.  However,  the  value  of  v  at  Q  must  be  deter- 
mined in  a  different  manner;  for  this  purpose  we  use  the  equation 
(11.5.7)  with  the  subscript  Q  replacing  P,  L  replacing  Af,  and  M 
replacing  R.  Since  VM,  CM,  VR,  CR  are  supposed  known,  and  CQ  is  also 


t ' 


-P 
M 


Fig.  11.5.3.  Satisfying  boundary  conditions 

known  since  the  values  of  y  arc  proscribed  on  the  /-axis,  it  follows  that 
equation  (11.5.7)  contains  VQ  as  the  only  unknown;  in  fact  it  is  given 
by  the  equation 

(11.5.10)      vQ^vL+At\  —  (cL 

The  reason  for  using  (11.5.7)  instead  of  (11.5.0)  is,  of  course,  that  the 
points  M  and  Q  are  associated  with  the  backward  characteristic,  and 
hence  (11.5.2)  should  be  used  to  approximate  the  ^-derivatives  at 
..  point  L  (where  the  differential  equations  are  replaced  by  difference 
equations).  It  is  quite  clear  that  the  same  general  procedure  could  be 
used  to  calculate  CQ  if  the  values  of  v  had  been  assumed  given  along 
the  /-axis.  If,  on  the  other  hand,  we  had  a  boundary  condition  on  the 
right  of  our  domain  instead  of  on  the  left,  as  above,  we  could  make  use 
of  (11.5.6)  for  the  forward  characteristic  as  a  basis  for  obtaining  the 
formula  for  advancing  the  solution  along  the  /-axis. 

The  above  discussion  would  seem  to  imply  that  under  all  circum- 
stances only  one  boundary  condition  could  be  imposed— that  is,  that 
either  v  or  c  could  be  prescribed  at  a  fixed  point  on  the  river,  but  not 
both— since  prescribing  one  of  these  quantities  leads  to  a  unique  de- 
termination of  the  other.  This  is,  indeed,  true  in  any  ordinary  river, 
but  not  necessarily  in  all  cases.  In  fact,  we  made  a  tacit  assumption 
in  the  above  discussion,  i.e.  that  of  the  two  characteristics  issuing 
from  any  point  of  the  /-axis  only  the  forward  characteristic  goes  into 


480 


WATER   WAVES 


the  region  x  >  0  to  the  right  of  the  J-axis,  and  this  in  turn  implies  that 
v  +  c  and  v  — -  c,  which  fix  the  slopes  of  the  characteristics,  are  op- 
posite in  sign.  The  physical  interpretation  of  this  is  that  the  value  of 
v  (which  is  positive  here)  must  be  less  than  c  =  ^gy,  i.e.  that  the 
flow  must  be  what  is  called  tranquil,  or  subcritical.*  Otherwise,  as 
we  see  from  Fig.  11.5.4,  we  should  expect  to  determine  the  values  of 
v  and  c  at  points  close  to  and  to  the  right  of  the  J-axis,  say  at  K9  by 


•M 


Fig.  11.5.4.  A  case  of  super-critical  flow 

utilizing  values  for  both  v  and  c  along  the  segment  LQ,  its  domain  of 
dependence.  The  scheme  outlined  above  would  therefore  have  to  be 
modified  in  a  proper  way  under  such  circumstances.  One  sees,  how- 
ever, how  useful  the  theory  based  on  the  characteristics  can  be  even 
though  no  direct  use  of  it  is  made  in  the  numerical  calculations  (aside 
from  decisions  regarding  the  maximum  permissible  size  of  the  ^-inter- 
val). 

The  procedure  sketched  out  above,  while  it  is  recommended  for  use 


.P 

M 


Fig.  11.5.5.  A  staggered  net 
In  gas  dynamics  the  flow  in  an  analogous  case  would  be  called  subsonic. 


MATHEMATICAL   HYDRAULICS 


481 


when  a  boundary  condition  is  to  be  satisfied,  is  not  always  the  best 
one  to  use  for  advancing  the  solution  to  such  points  as  P,  Pl9  P2,  .  .  . 
in  Fig.  11.5.3.  For  such  "interior  points"  a  staggered  rectangular  net, 
as  indicated  in  Fig.  11.5.5,  and  a  difference  equation  scheme  based  on 
the  original  differential  equations  (11.4.6)  may  be  preferable  (cf. 
Keller  and  Lax  [K.4]  for  a  discussion  of  this  scheme).  The  equations 
(11.4.6)  were 


(11.5.11) 


cv 


2vc 


2ct  =  0. 


The  values  VM  and  CM  at  the  mid-point  M  (which  is,  however,  not  a 
net  point)  of  the  segment  LR  are  defined  by  the  averages: 


(11.5.12) 


after  which  the  derivatives  at  M  are  approximated  in  a  quite  natural 
way  by  the  difference  quotients 


(11.5.13) 


Ax     ' 


Ax    ' 


u  M 


LM 


At 


Upon  substitution  of  these  quantities  into  (11.5.11),  evaluation  of  the 
coefficients  c,  u,  and  E  at  point  M9  and  subsequent  solution  of  the 
two  equations  for  vp  and  cp,  the  result  is 


(11.5.14) 


7- 
Ax 

— 
Ax 


As  we  see  on  comparison  with  (11.5.8)  and  (11.5. 9\  these  equations 
are  simpler  than  the  earlier  ones.  The  criterion  for  convergence  re- 
mains the  same  as  before,  i.e.  that  P  should  lie  within  a  triangle  formed 
by  the  segment  LR  and  the  two  characteristics  issuing  from  its 
ends. 


482  WATER   WAVES 

11.6.  Flood  prediction  in  rivers.  Floods  in  models  of  the  Ohio  River 
and  its  junction  with  the  Mississippi  River 

The  theory  developed  in  the  preceding  sections  can  be  used  to  make 
predictions  of  floods  in  rivers  on  the  basis  of  the  observed,  or  estimat- 
ed, flow  into  the  river  from  its  tributaries  and  from  the  local  run-off, 
together  with  the  state  of  the  river  at  some  initial  instant.  Hydraulics 
engineers  have  developed  a  procedure,  called  flood-routing,  to  accom- 
plish the  same  purpose.  The  flood-routing  procedure  can  be  deduced 
as  an  approximation  in  some  sense  to  the  solution  of  the  basic  differ- 
ential equations  for  flow  in  open  channels  (cf.  the  article  by  B.  R. 
Gilcrest  in  the  book  by  Rouse  [R.ll] ),  but  it  makes  no  direct  use  of  the 
differential  equations.  However,  the  flood-routing  procedure  in  ques- 
tion seems  not  to  give  entirely  satisfactory  results  in  cases  other  than 
that  of  determining  the  progress  of  a  flood  down  a  long  river  —  for 
example,  the  problem  of  what  happens  at  a  junction,  such  as  that  of 
the  Ohio  and  Mississippi  Rivers,  or  the  problem  of  calculating  the 
transient  effects  resulting  from  regulation  at  a  dam,  such  as  the 
Kentucky  dam  at  the  mouth  of  the  Tennessee  River,  seem  to  be  diffi- 
cult to  treat  by  methods  that  are  modifications  of  the  more  or  less 
standard  flood-routing  procedures.  Even  for  a  long  river  like  the 
Ohio,  the  usual  procedure  fails  occasionally  to  yield  the  observed  river 
stages  at  some  places.  On  the  other  hand,  the  basic  differential  equa- 
tions for  flow  in  open  channels  are  in  principle  applicable  in  all  cases 
and  can  be  used  to  solve  the  problems  once  the  appropriate  data  de- 
scribing the  physical  characteristics  of  the  river  and  the  appropriate 
initial  and  boundary  conditions  are  known. 

The  idea  of  using  the  differential  equations  directly  as  a  means  of 
treating  problems  of  flow  in  open  channels  is  not  at  all  new.  In  fact, 
it  goes  back  to  Massau  [M.5]  as  long  ago  as  1889.  Since  then  the  idea 
has  been  taken  up  by  many  others  (mostly  in  ignorance  of  the  work 
of  Massau)— for  example,  by  Preiswerk  [P.16],  von  Karman  [K.2], 
Thomas  [T.2],  and  Stoker  [S.19].  Thomas,  in  particular,  attacked  the 
flood-routing  problem  in  his  noteworthy  and  pioneering  paper  and 
outlined  a  numerical  procedure  for  its  solution  based  on  the  idea  of  us- 
ing the  method  of  finite  differences.  However,  his  method  is  very  la- 
borious to  apply  and  would  also  not  necessarily  furnish  a  good 
approximation  to  the  desired  solution  even  if  a  large  number  of 
divisions  of  the  river  into  sections  were  to  be  taken.  In  general,  the 
amount  of  numerical  work  to  be  done  in  a  direct  integration  of  the 


MATHEMATICAL    HYDRAULICS  483 

differential  equations  looked  too  formidable  for  practical  purposes 
until  rather  recently. 

During  and  since  the  late  war  new  developments  have  taken  place 
which  make  the  idea  of  tackling  flood  prediction  and  other  similar 
problems  by  numerical  solution  of  the  relevant  differential  equations 
quite  tempting.  There  have  been,  in  fact,  developments  in  two  differ- 
ent directions,  both  motivated  by  the  desire  to  solve  difficult  problems 
in  compressible  gas  dynamics:  1)  development  of  appropriate  nu- 
merical procedures— for  the  most  part  methods  using  finite  differences 
—  for  solving  the  differential  equations,  and  2)  development  of  com- 
puting machines  of  widely  varying  characteristics  suitable  for  carry- 
ing out  the  numerical  calculations.  As  we  have  seen,  the  differential 
equations  for  flood  control  problems  are  of  the  same  type  as  those  for 
compressible  gas  dynamics,  and  consequently  the  experience  and  cal- 
culating equipment  developed  for  solving  problems  in  gas  dynamics 
can  be  used,  or  suitably  modified,  for  solving  flood  control  problems. 

In  carrying  out  such  a  study  of  an  actual  river  it  is  necessary  to 
make  use  of  a  considerable  bulk  of  observational  data— cross-sections 
and  slopes  of  the  channels,  measurements  of  river  depths  and  dis- 
charges as  functions  of  time  and  distance  down  the  river,  drainage 
areas,  observed  flows  from  tributaries,  etc.— in  order  to  obtain  the 
information  necessary  to  fix  the  coefficients  of  the  differential  equa- 
tions and  to  fix  the  initial  and  boundary  conditions.  This  is  a  task 
with  many  complexities.  For  the  purposes  of  this  book  it  is  more 
reasonable  to  carry  out  numerical  solutions  for  problems  which  are 
simplified  versions  of  actual  problems.  The  present  section  has  as  its 
purpose  the  presentation  of  the  solutions  in  a  few  such  special  cases, 
together  with  an  analysis  of  their  bearing  on  the  concrete  problems 
for  actual  rivers.  In  any  case,  the  general  methods  for  an  actual  river 
would  be  the  same— there  would  simply  be  greater  numerical  compli- 
cations. 

The  simplified  models  chosen  correspond  in  a  rough  general  way 
(a)  to  two  types  of  flow  for  the  Ohio  River  and  (b)  to  the  Ohio  and 
Mississippi  Rivers  at  their  junction.  Rivers  of  constant  slope,  with 
rectangular  cross-sections  having  a  uniform  breadth,  and  with  con- 
stant roughness  coefficients  are  assumed.  In  this  way  differential 
equations  with  constant  coefficients  result.  The  values  of  these  quan- 
tities are,  however,  taken  to  correspond  in  order  of  magnitude  with 
those  for  the  actual  rivers.  In  the  model  of  the  Ohio,  for  example,  the 
slope  of  the  channel  was  assumed  to  be  0.5  ft/mile,  the  quantity  n 


484  WATER   WAVES 

(the  roughness  coefficient  in  Manning's  formula)  was  given  the  value 
0.03,  and  the  breadth  of  the  river  was  taken  as  1000  feet.  It  is  assumed 
that  a  steady  uniform  flow  with  a  depth  of  20  ft  existed  at  the  initial 
instant  t  =  0,  and  that  for  t  >  0  the  depth  of  the  water  was  increased 
at  a  uniform  rate  at  the  point  x  =  0  from  20  ft  to  40  ft  within  4 
hours  and  was  then  held  fixed  at  the  latter  value.  (These  depths  are 
the  same  as  for  the  problem  of  a  steady  progressing  wave  treated  in 
sec.  11.2  above.)  The  problem  is  to  determine  the  flow  downstream, 
i.e.  the  depth  y  and  the  flow  velocity  v  as  functions  of  x  (for  x  >  0) 
and  t. 

The  methods  used  to  obtain  the  solution  of  this  problem  of  a  flood  in 
a  model  of  the  Ohio  River,  together  with  a  discussion  of  the  results, 
will  be  given  in  detail  later  on  in  this  section.  Before  doing  so,  a  few 
general  remarks  and  observations  about  them  should  be  made  at  this 
point.  In  the  first  place,  it  was  found  possible  to  carry  out  the  solution 
numerically  by  hand  computation  over  a  considerable  range  of  dis- 
tances and  times  (values  at  900  net  points  in  the  #,  2-plane  were  de- 
termined by  finite  differences),  and  this  in  itself  shows  that  the 
problems  are  well  within  the  capacity  of  modern  calculating  equip- 
ment. It  might  be  added  that  the  special  case  chosen  for  a  flood  in  the 
Ohio  was  one  in  which  the  rate  of  rise  at  the  starting  point  upstream 
was  extremely  high  (5  feet  per  hour,  in  comparison  with  the  rate  of 
rise  during  the  flood  of  1945— one  of  the  biggest  ever  recorded  in  the 
Ohio— which  was  never  larger  than  0.7  feet  per  hour  at  Wheeling, 
West  Virginia),  so  that  a  rather  severe  test  of  the  finite  difference 
method  was  made  in  view  of  the  rapid  changes  of  the  basic  quantities 
in  space  and  time.  The  decisive  point  in  estimating  the  magnitude  of 
the  computational  work  in  using  finite  differences  is  the  number  of 
net  points  needed;  for  a  river  such  as  the  Ohio  it  is  indicated  that  an 
interval  Ax  of  the  order  of  10  miles  along  the  river  and  an  interval  At 
of  the  order  of  0.3  hours  in  time  in  a  rectangular  net  in  the  x,  J-plane 
will  yield  results  that  are  sufficiently  accurate.  (Of  course,  a  problem 
for  the  Ohio  in  its  actual  state  involves  empirical  coefficients  in  the 
differential  equations  and  other  empirical  data,  which  must  be  coded 
for  calculating  machines,  but  this  would  have  no  great  effect  on  these 
estimates  for  Ax  and  might  under  extreme  flood  conditions  reduce  At 
by  a  factor  of  1/2.) 

As  we  know  from  sec.  11.3  above,  there  is  a  case  in  which  an  exact 
solution  of  the  differential  equations  is  known,  i.e.  the  case  of  a 
steady  progressing  wave  with  two  different  depths  at  great  distances 


MATHEMATICAL    HYDRAULICS  485 

upstream  and  downstream.  The  exact  solution  obtained  in  sec.  11.3 
for  the  case  of  a  wave  of  depth  20  ft  far  downstream  and  40  ft  far 
upstream  was  taken  as  furnishing  the  initial  conditions  at  t  —  0  for 
a  wave  motion  in  the  river.  With  the  initial  conditions  prescribed  in 
this  way  the  finite  difference  method  was  used  to  determine  the  mo- 
tion at  later  times;  of  course  the  calculation,if  accurate,  should  fur- 
nish a  wave  profile  and  velocity  distribution  which  is  the  same  at 
time  t  as  at  the  initial  instant  t  =  0  except  that  all  quantities  are  dis- 
placed downstream  a  distance  Ut,  with  U  the  speed  of  the  steady 
progressing  wave.  In  this  way  an  opportunity  arises  to  compare  the 
approximate  solution  with  an  exact  solution.  In  the  present  case  the 
phase  velocity  U  is  approximately  5  mph.  Interval  sizes  of  Ax  =  5 
miles  in  a  "staggered"  finite  difference  scheme  (cf.  equations  (11.5.14)) 
with  At  =  .08  hr  were  taken  and  a  numerical  solution  was  worked 
out.  We  report  the  results  here.  After  12  hours,  the  calculated  values 
for  the  stage  y  agreed  to  within  .5  per  cent  with  the  exact  values. 
The  discharge  and  the  velocity  deviated  by  less  than  .8  per  cent 
from  the  exact  values. 

One  of  the  valuable  insights  gained  from  working  out  the  solution 
of  the  flood  problem  in  a  model  of  the  Ohio  was  an  insight  into  the 
relation  between  the  methods  used  by  engineers— for  example,  by 
the  engineers  of  the  Ohio  River  Division  of  the  Corps  of  Engineers  in 
Cincinnati  — for  predicting  flood  stages,  and  the  methods  explained 
here,  which  make  use  of  the  basic  differential  equations.  At  first  sight 
the  two  methods  seem  to  have  very  little  in  common,  though  both,  in 
'  the  last  analysis,  must  be  based  on  the  laws  of  conservation  of  mass 
and  momentum;  indeed,  in  one  important  respect  they  even  seem  to 
be  somewhat  contradictory.  The  methods  used  in  engineering  prac- 
tice (which  make  no  direct  use  of  our  differential  equations)  tacitly 
assume  that  a  flood  wave  in  a  long  river  such  as  the  Ohio  propagates 
only  in  the  downstream  direction,  while  the  basic  theory  of  the  dif- 
ferential equations  we  use  tells  us  that  a  disturbance  at  any  point  in 
a  river  flowing  at  subcritical  speed  (the  normal  case  in  general  and 
always  the  case  for  such  a  river  as  the  Ohio)  will  propagate  as  a  wave 
traveling  upstream  as  well  as  downstream.  Not  only  that,  the  speed 
of  propagation  of  small  disturbances  relative  to  the  flowing  stream,  as 
defined  by  the  differential  equations,  is  \/gy  for  small  disturbances 
and  this  is  a  good  deal  larger  (by  a  factor  of  about  4  in  our  model  of 
the  Ohio)  than  the  propagation  speed  used  by  the  engineers  for  their 
flood  wave  traveling  downstream.  There  is,  however,  no  real  dis- 


486  WATER  WAVES 

crepancy.  The  method  used  by  the  engineers  can  be  interpreted  as  a 
method  which  yields  solutions  of  the  differential  equations,  with  cer- 
tain terms  neglected,  that  are  good  approximations  (though  not  under 
all  circumstances,  it  seems)  to  the  actual  solutions  in  some  cases, 
among  them  that  of  flood  waves  in  a  river  such  as  the  Ohio.  The 
neglect  of  terms  in  the  differential  equations  in  this  approximate 
theory  is  so  drastic  as  to  make  the  theory  of  characteristics,  from 
which  the  properties  of  the  solutions  of  the  differential  equations  were 
derived  here,  no  longer  available.  The  numerical  solution  presented 
here  of  the  differential  equations  for  a  flood  wave  in  a  model  of  the 
Ohio  yields,  as  we  have  said,  a  wave  the  front  of  which  travels  down- 
stream at  the  speed  \/^y;  but  the  amplitude  of  this  forerunner  is 
quite  small,*  while  the  portion  of  the  wave  with  an  amplitude  in  the 
range  of  practical  interest  is  found  by  this  method  to  travel  with 
essentially  the  same  speed  as  would  be  determined  by  the  engineers' 
approximate  method.  What  seems  to  happen  is  the  following:  small 
forerunners  of  a  disturbance  travel  with  the  speed  \/gy  relative  to  the 
flowing  stream,  but  the  resistance  forces  act  in  such  a  way  as  to  de- 
crease the  speed  of  the  main  portion  of  the  disturbance  far  below  the 
values  given  by  i/gy,  i.e.  to  a  value  corresponding  closely  to  the  speed 
of  a  steady  progressing  wave  that  travels  unchanged  in  form.  (One 
could  also  interpret  the  engineering  method  as  one  based  on  the  as- 
sumption that  the  waves  encountered  in  practice  differ  but  little  from 
steady  progressing  waves).  As  we  shall  see  a  little  later,  our  unsteady 
flow  tends  to  the  configuration  of  a  steady  progressing  wave  of  depth 
40  ft  upstream  and  20  ft  downstream. 

This  analysis  of  the  relation  between  the  methods  discussed  here 
and  those  commonly  used  in  engineering  practice  indicated  why  it 
may  be  that  the  latter  methods,  while  they  furnish  good  results  in 
many  important  cases,  fail  to  mirror  the  observed  occurrences  in  other 
cases.  For  example,  the  problem  of  what  happens  at  a  junction  of  two 
major  streams,  and  various  problems  arising  in  connection  with  the 
operation  of  such  a  dam  as  the  Kentucky  Dam  in  the  Tennessee  River 
seem  to  be  cases  in  which  the  engineering  methods  do  not  furnish 
accurate  results.  These  would  seem  to  be  eases  in  which  the  motions 
of  interest  depart  too  much  from  those  of  steady  progressing  waves, 
and  cases  in  which  the  propagation  of  waves  upstream  is  as  vital  as  the 
propagation  downstream.  Thus  at  a  major  junction  it  is  clear  that 

In  an  appendix  to  this  chapter  an  exact  statement  on  this  point  is  made. 


MATHEMATICAL   HYDRAULICS  487 

considerable  effects  on  the  upstream  side  of  a  main  stream  are  to  be 
expected  when  a  large  flow  from  a  tributary  occurs.  In  the  same  way, 
a  dam  in  a  stream  (or  any  obstruction,  or  change  in  cross-section,  etc.) 
causes  reflection  of  waves  upstream,  and  neglect  of  such  reflections 
might  well  cause  serious  errors  on  some  occasions. 

The  above  general  description  of  what  happens  when  a  flood  wave 
starts  down  a  long  stream— in  particular,  that  it  has  a  lengthy  front 
portion  which  travels  fast,  but  has  a  small  amplitude,  while  the  main 
part  of  the  disturbance  moves  much  more  slowly— has  an  important 
bearing  on  the  question  of  the  proper  approach  to  the  numerical  solu- 
tion by  the  method  of  finite  differences.  It  is,  as  we  shall  see  shortly, 
necessary  to  calculate— or  else  estimate  in  some  way— the  motion  up 
to  the  front  of  the  disturbance  in  order  to  be  in  a  position  to  calculate 
it  at  the  places  and  times  where  the  disturbances  are  large  enough  to 
be  of  practical  interest.  This  means  that  a  large  number  of  net  points 
in  the  finite  difference  mesh  in  the  #,  J-plane  lie  in  regions  where  the 
solution  is  not  of  much  practical  interest.  Since  the  fixing  of  the  solu- 
tion in  these  regions  costs  as  much  effort  as  for  the  regions  of  greater 
interest,  the  differential  equation  method  is  at  a  certain  disadvantage 
by  comparison  with  the  conventional  method  in  such  a  case.  However, 
it  is  possible  in  simple  cases  to  determine  analytically  the  character 
of  the  front  of  the  wave  and  thus  estimate  accurately  the  places  and 
times  at  which  the  wave  amplitude  is  so  small  as  to  be  negligible; 
these  regions  can  then  be  regarded  as  belonging  to  the  regions  of  the 
x,  f-planc  where  the  flow  is  undisturbed,  with  a  corresponding  re- 
duction in  the  number  of  net  points  at  which  the  solutions  must  be 
calculated.  A  method  which  can  be  used  for  this  purpose  has  been 
derived  by  G.  Whitham  and  A.  Troesch,  and  a  description  of  it  is 
given  in  an  appendix  to  this  chapter.  If  a  modern  high  speed  digital 
computer  were  to  be  used  to  carry  out  the  numerical  work,  however, 
it  would  not  matter  very  much  whether  the  extra  net  points  in  the 
front  portion  of  the  wave  were  to  be  included  or  not:  many  such 
machines  have  ample  capacity  to  carry  out  the  necessary  calculations. 

We  proceed  to  give  a  description  of  the  calculations  made  for  our 
model  of  the  Ohio,  including  a  discussion  of  various  difficulties  which 
occurred  for  the  flood  wave  problem  near  the  front  of  the  disturbance, 
and  particularly  at  the  beginning  of  the  wave  motion  (i.e.  near  x  =  0, 
/  =  0),  and  an  enumeration  of  the  features  of  the  calculation  which 
must  play  a  similar  role  in  the  more  complicated  cases  presented  by 
rivers  in  their  actual  state.  This  will  be  followed  by  a  description  of 


488  WATER   WAVES 

the  method  used  and  the  calculations  made  for  a  problem  simulating 
a  flood  coming  down  the  Ohio  and  its  effect  on  passing  into  the 
Mississippi.  This  problem  and  its  solution  give  rise  to  further  general 
observations  which  will  be  made  later  on. 
The  differential  equations  to  be  solved  are 


(11.6.1) 

2ct  +  2vcx  +  cvx  =  0, 

with  v(x9t)  the  velocity,  and  c  =  Vgy  the  propagation  speed  of 
small  disturbances.  The  assumption  of  a  uniform  cross-section  and 
the  assumption  that  no  flow  over  the  banks  occurs  (i.e.  q  =  0  in  the 
basic  differential  equations  (11.1.1)  and  (11.1.6))  have  already  been 
used.  The  quantity  E  is  given  by 

E  =  -  gS  +  gSf, 

with  S  the  slope  of  the  river  bed  and  Sf,  the  friction  slope,  given  by 
Manning's  formula 


Here  we  assume  the  channel  to  be  rectangular  with  breadth  B. 

The  numerical  data  for  the  problem  of  a  flood  in  a  model  of  the 
Ohio  River  are  as  follows.  For  the  slope  S  a  value  of  0.5  ft/mi  was 
chosen,  and  B  is  given  the  value  1000  ft.  For  y  a  value  of  2500  was 
taken  (in  foot-sec  units),  corresponding  to  a  value  of  Manning's 
constant  n  (in  the  formula  y  —  (1.49/n)2)  of  0.03.  The  special  pro- 
blem considered  was  then  the  following:  At  time  t  =  0,  a  steady  flow 
of  depth  20  ft  is  assumed.  At  the  "headwaters"  of  the  river,  corres- 
ponding to  x  =  0,  we  impose  a  linear  increase  of  depth  with  time  which 
brings  the  level  to  40  ft  in  4  hours.  For  subsequent  times  the  level  of 
40  ft  at  x  =  0  is  maintained.  The  initial  velocity  of  the  water  cor- 
responding to  a  uniform  flow  of  depth  yQ  =  20  ft  is  calculated  from 
Sf  =  S  to  be 

i>0  =  2.38  mph; 

the^propagation  speed  of  small  disturbances  corresponding  to  the 
depth  of  20  ft  is 


MATHEMATICAL   HYDRAULICS 


489 


C    = 


=  17.3  mph. 


The  problem  then  is  to  determine  the  solution  of  (11.6.1)  for  v(x,  t), 
c(x,  t)  for  all  later  times  t  ^  0  along  the  river  x  ^  0.  Figures  11.6.1 
and  11.6.2  present  the  result  of  the  computation  in  the  form  of  stage 
and  discharge  curves  plotted  as  functions  of  distance  along  the  river 
at  various  times. 

In  order  to  indicate  how  the  solution  was  calculated  it  is  conven- 
ient to  refer  to  diagrams  in  the  (x,  t)  plane  given  by  Figs.  11.6.3 
and  11.6.4.  According  to  the  basic  theory,  we  know  that  for  x  ^ 
fao  +  co)t  =  19-7*,  called  region  O  in  Fig.  11.6.3,  the  solution  is  given 


10- 


Legend 

t-  time  m  hours  after  start  of    flood 
y  -  stage  in  feet 
x  -  distance   along  Ohio  m  miles 

^sTJG 


-•<£;| ;:%  ^-A«%3;e$  A^^  ^^      x 

20  4O  60  8O  100  I2O 

Fig.   11.6.1.  Stage  profiles  for  a  flood  in  the  Ohio  River 


by  the  unchanged  initial  data,  v(x9  t)  =  00,  c(x,  t)  =  C0  (since  the 
forerunner  of  the  disturbance  travels  at  the  speed  w0  +  CQ  =  19.7  mph). 

Experiments  were  made  with  various  interval  sizes  and  finite 
difference  schemes  in  order  to  try  to  determine  the  most  efficient  way 
to  calculate  the  progress  of  the  flood.  We  proceed  to  describe  the 
various  schemes  tried  and  the  regions  in  which  they  were  used  on  the 
basis  of  Figs.  11.6.3  and  11.6.4. 

Region   I,    0  ^  x  ^  19.7J,    0  ^  t  ^  .4.    Quite   small   intervals   of 


490 


WATER   WAVES 


300. 


260. 


220. 


Legend 

t  *  time  in  hours  after  start  of  flood 
Q»  discharge  in   1000  c.fs. 
x  3  distance  along  Ohio  in  miles 


t«a> 


0  20  40  60  80  100  120  140  160 

Fig.  11.6.2.  Discharge  records  for  a  flood  in  the  Ohio  River 


1.25 


Fig.  11.6.3.  Regions  in  which  various  computational  methods  were  tried 

Ax  =  l  mile  and  At  =  .048  hours  were  required  owing  to  the  sudden 
increase  of  depth  at  x  =  0,  t  =  0.  The  finite  difference  formulas  given 
above  in  equations  (11.5.8),  (11.5.9)  were  used. 

In    Region    II,    0  ^  x  ^  19.7*,    .4  ^  t  <^  .7,    with    Ax  =  1  mile, 


MATHEMATICAL   HYDRAULICS 


491 


t 
2.0- 


1.5  J 


1.0  J 


Legend 

1  s  time  in  hours 
x  a  distance  in  miles 


°  10  20  30 

Fig.  11.6.4.  Net  points  used  in  the  finite  difference  schemes 

At  —  .024  hr,  the  "staggered"  scheme  was  used.  The  formulas  for  this 
scheme  have  been  given  above  in  equations  (11.5.14).  In  order  to 
calculate  0(0,  2),  the  velocity  at  the  upstream  boundary  of  the  river, 
the  formula  associated  with  the  backward  characteristic,  namely 
equation  (11.5.10),  has  to  be  used  twice  in  succession:  for  the  triangles 
FBM  and  MRP  (cf.  Figs.  11.5.3  and  11.6.5).  The  values  CB  and  VB 
arc  simply  determined  by  linear  interpolation  from  the  values  at  the 
points  F  and  G. 


492  WATER   WAVES 


Region  III,  0  ^  x  ^  5,  .7  ^  t^  1.25,  with  Ax  =  1  mile,  z^  =  .024  hr. 
The  same  procedure  was  used  as  in  Region  II. 

Region  IV,  5  ^  x  ^  19.7J,  .7  ^  *  ^  1.25,  with  Ax  =  2  miles, 
At  =  .048  hr.  The  values  at  the  boundary  between  Regions  III  and 


.G  .H 


x 
Fig.  11.6.5.  Net  point  arrangement  used  at  boundary  in  "staggered"  scheme 

IV  were  obtained  by  linear  interpolation  from  the  neighboring  values. 

Other  quantities  were  computed  by  the  " staggered"  scheme  as  in 

Regions  II  and  III. 

Region  V,  0  ^  x  ^  Ut,   1.25  ^  t  <,  10,  Ax  =  5  miles,  J*  =  .17  hr. 

U  represents  a  variable  speed  which  marks  the  downstream  end  of 

what  might  be  called  the  observable  disturbance  (U  &  10  mph). 

That  is,  by  using  an  expansion  scheme  (see  the  appendix  to  this 

chapter)  we  obtain  the  solution  in 

Region  VI,   defined  by  Ut  ^  as  ^  19.72,   back  of  the  forerunner 

of  the  disturbance,     in   which   the   flow   is   essentially  undisturbed 

for  all  practical  purposes.  The  expansion  valid  near  the  front  of  the 

wave  and  referred  to  above  was  used  to  calculate  the  various  quantities 

in  Region  VI,  and  a  staggered  scheme  was  used  to  compute  the  values 

in  Region  V. 

A  number  of  conclusions  reached  on  the  basis  of  the  experience 

gained  from  these  calculations  of  a  flood  in  a  model  of  the  Ohio  River 

can  be  summarized  as  follows: 

(a)  The  rate  of  rise  of  the  flood —5  feet  per  hour— is  extreme,  and 
such  a  case  exaggerates  the  way  in  which  errors  in  the  finite 
difference  methods  are  propagated.  For  example,  slight  inaccu- 
racies at  the  head,  x  ~  0,  were  found  to  develop  upon  increasing 
the  size  of  the  Ax  interval.  In  spite  of  the  exceptionally  high  rate 
of  rise  of  the  flood,  the  fluctuations  created  by  using  finite  dif- 
ference methods  were  damped  out  rather  strongly  (in  about 
8  —  10  time  steps).  It  is  possible  to  control  these  inaccuracies 


MATHEMATICAL   HYDRAULICS  493 

simply  by  using  small  interval  sizes.  The  process  by  which  the 
small  errors  of  the  finite  difference  scheme  are  caused  to  die  out 
may  be  described  as  follows:  A  value  of  v  which  is  too  large 
produces  a  correspondingly  larger  friction  force  which  slows  down 
the  motion  and  produces  at  a  later  time  a  smaller  velocity.  The 
lower  velocity  in  a  similar  way  then  operates  through  the  resistance 
to  create  a  larger  velocity  and  the  process  repeats  in  an  oscillatory 
fashion  with  a  steady  decrease  in  the  amplitude  of  variation. 

(b)  The  accuracy  of  our  computation  (as  a  function  of  the  interval 
size)  was  checked  by  repeating  the  calculation  for  two  different 
interval  sizes  over  the  same  region  in  space  and  time. 

(c)  A  linearized  theory  of  wave  propagation,  obtained  by  assuming 
a  small  perturbation  about  the  uniform  flow  with  20  ft  depth,  is 
easily  obtained,  and  the  problem  was  solved  using  such  a  theory. 
However,  it  does  not  give  an  accurate  description  of  the  solution 
of  our  problem.  It  was  found  that  the  stage  was  predicted  too  low 
by  the  linear  theory  by  as  much  as  2  feet  after  only  2  hours— a 
very  large  error. 

(d)  It  would  be  convenient  to  be  in  possession  of  a  safe  estimate  for 
the  maximum  value  of  the  particle  velocity,  in  order  to  select  an 
appropriate  safe  value  for  the  time  interval  At,  since  we  must 
have  At  5g  Ax/(v  -)-  c)  in  order  to  make  sure  that  the  finite  dif- 
ference scheme  converges.  The  calculations  in  our  special  case 
indicate  that  this  may  not  be  easy  to  obtain  in  a  theoretical  way, 
since  the  maximum  velocity  at  x  —  0,  for  example,  greatly  ex- 
reeds  its  asymptotic  value,   as  indicated  in  Fig.   11.6.6.   In  a 


5.4- 


24 


v(0,t) 
mph 


velocity   for 
40ft   steady  flow 


t 


4  hours 

Fig.  11.6.6.  Water  velocity  obtained  at  "head"  of  river 

computation  for  an  actual  river,  however,  no  real  difficulty  is 
likely  to  result,  since  c  is  in  general  much  larger  than  v  and  is 
determined  by  the  depth  alone. 


494 


WATER  WAVES 


(e) 


As  was  already  indicated  above,  the  curves  of  constant  stage 
turn  out  to  have  slopes  which  are  closer  to  5  mph  (the  speed  with 
which  a  steady  progressing  flow,  40  ft  upstream  and  20  ft  down- 
stream, moves)  than  they  are  to  the  19.7  mph  speed  of  pro- 
pagation of  small  disturbances.  This  is  shown  by  Fig.  11.6.7. 


10. 


5. 


t  hours 


region  of 
practically 
undisturbed 
flow 


Smph-slope        I97mph-slope 


50 


100 


x  miles 


Fig.  11.6.7.  Curves  of  constant  stage — comparison  with  first  characteristic  and 
steady  progressing  flow  velocity 

The  region  of  practically  undisturbed  flow  (determined  by  an 
expansion  about  the  "first"  characteristic  x  =  19.7J,  for  which 
see  the  appendix  to  this  chapter)  is  shown  above.  In  an  actual 
river,  we  would  of  course  expect  the  local  runoff  discharges  and  the 
non-uniform  flow  conditions  to  eliminate  largely  the  region  of 
practically  undisturbed  flow.  For  this  reason  it  is  not  feasible 
to  use  analytic  expansion  schemes  as  a  means  of  avoiding 
computational  labor. 

We  turn  next  to  our  model  of  the  junction  of  the  Ohio  and  Missis- 
sippi Rivers  and  the  problem  of  what  happens  when  a  flood  wave 
comes  down  the  Ohio  and  passes  through  the  junction.*  The  physical 
data  chosen  are  the  same  as  were  used  above  in  sec.  11.2  in  discussing 
the  problem  of  a  steady  flow  at  a  junction. 

We  suppose  the  upstream  side  of  the  Mississippi  to  be  identical 
with  the  Ohio  River— i.e.  that  it  has  a  rectangular  cross-section 
1000  ft  wide,  a  slope  of  .5  ft/mile,  and  that  Manning's  constant  n  has 
the  value  .03.  The  downstream  Mississippi  is  also  taken  to  be  rectan- 
gular, but  twice  as  wide,  i.e.  2000  ft  in  width,  Manning's  constant  is 
again  assumed  to  have  the  value  .03,  but  the  slope  of  this  branch  is 
given  the  value  .49  ft/mile.  This  modification  of  the  slope  was  made 

*  The  analogous  problem  in  gas  dynamics  would  be  concerned  with  the  pro- 
pagation of  a  wave  at  the  junction  of  two  pipes  containing  a  compressible  gas. 


MATHEMATICAL    HYDRAULICS 


495 


in  order  to  make  possible  an  initial  solution  corresponding  to  a  uni- 
form flow  of  20  ft  depth  in  all  three  branches.  (Such  a  change  is 
necessary  in  order  to  overcome  the  decrease  in  wetted  perimeter 
which  occurs  on  going  downstream  through  the  junction.)  Figure 
11.6.8  shows  a  schematic  plan  of  the  junction.  The  concrete  problem 
to  be  solved  is  formulated  as  follows.  A  flood  is  initiated  in  the  Ohio 


L3J 

Downstream 
Mississippi 


Fig.  11.6.8.  Schematic  plan  of  junction 

at  a  point  50  miles  above  the  junction  by  prescribing  a  rise  in  depth  of 
the  stream  at  that  point  from  20  ft  to  40  ft  in  4  hours  — in  other 
words,  the  same  initial  and  boundary  conditions  were  assumed  as  for 
the  case  of  the  flood  in  the  Ohio  treated  in  detail  above.  After  about 
2.5  hours  the  forerunner,  or  front,  of  the  wave  in  the  Ohio  caused  by 
the  disturbance  50  miles  upstream  reaches  the  junction;  up  to  this 
instant  nothing  will  have  happened  to  disturb  the  Mississippi,  and 
the  numerical  calculations  made  above  for  the  Ohio  remain  valid 
during  the  first  2.5  hours.  Once  the  disturbance  created  in  the  Ohio 
reaches  the  junction,  it  will  cause  disturbances  which  travel  both 
upstream  and  downstream  in  the  Mississippi,  and  of  course  also  a 
reflected  wave  will  start  backward  up  the  Ohio.  The  finite  difference 
calculations  therefore  were  begun  in  all  three  branches  from  the 
moment  that  the  junction  was  reached  by  the  forerunner  of  the  Ohio 
flood,  and  the  solution  was  calculated  for  a  period  of  10  hours. 

We  proceed  to  describe  the  method  of  determining  the  numerical 
solution.  Let  r(1),  c(1),  i>(2),  c(2),  0(8),  c(3)  represent  the  velocity  v  and  the 
propagation  speed  c  for  the  Ohio,  upstream  Mississippi,  and  down- 


496 


WATER   WAVES 


stream  Mississippi,  respectively.  A  "staggered"  scheme  was  used  with 
intervals  Ax  =  5  miles  and  At  —  .17  hr  as  indicated  in  Fig.  11.6.9. 
The  junction  point  is  denoted  by  x  =  0,  the  region  of  the  Ohio  and 


x         -P      x 

•  L      *M      .R 
.K      XA      .F       XB      .G 


Ohio   [II 


Upstream  Mississippi    C2J         Junction 


Downstream  Mississippi    [31 


Fig.  11.6.9.  Junction  net  point  scheme 

the  upstream  Mississippi  are  represented  by  x  ^  0,  while  the  down- 
stream Mississippi  is  described  for  x  S>  0.  The  time  t  =  2.5  hrs,  as 
explained  above,  corresponds  to  the  instant  that  the  forerunner  of 
the  flood  reaches  the  junction. 

The  values  of  the  quantities  v  and  c  at  the  junction  were  determined 
as  follows:  Assume  that  the  values  of  v  and  c  have  been  obtained  at  all 
net  points  for  times  preceding  that  of  the  boundary  net  point  P,  which 
represents  a  point  at  the  junction.  We  use  at  this  point  the  relations 


since  c  =  \/gy  and  the  water  level  is  the  same  in  the  three  branches  at 
the  junction.  In  addition,  we  have 


since  what  flows  into  the  junction  from  the  upstream  side  of  the  Mis- 
sissippi and  from  the  Ohio  must  flow  out  of  the  junction  into  the  down- 
stream branch  of  the  Mississippi.  If  the  values  of  v  and  c  were  known 
at  the  point  M  in  Fig.  11.6.9  in  the  respective  branches  of  the  rivers, 
we  could  find  the  values  at  P  from  equation  (11.5.6)  for  the  Ohio  and 
the  upstream  side  of  the  Mississippi,  and  equation  (11.5.7)  for  the 


MATHEMATICAL    HYDRAULICS  497 

downstream  side  of  the  Mississippi.  We  rewrite  the  equations  for 
convenience,  as  follows: 


CP(l)   •=  CP(2)   ="  <V(3)>  (with    C  = 

+  IJP(2)  =  2^P(3)»  (since  y(l)  =  y(2)  =  j/(3)), 


nifi9\ 

(11.6.2), 


and 


,  •>./,.         ...        \  VL(,)\\ 

+  {  -----  --  +  (Of(J)+PMU))l   —  -       ---  I  j 


A 

(          At  Ax 

l) 


I        """27  ----  M(3)"A/(3)     —  -  ^  -  j 

The  above  system  of  six  linear  equations  determines  uniquely  the 
values  u(1),  c(1),  u(2),  c(2),  z;{3),  c(3)  at  P  in  terms  of  their  values  at  the 
preceding  points  L,  M  and  R.  The  equations  can  be  solved  explicitly. 
The  values  of  the  relevant  quantities  at  M  are  determined  in  the  same 
way  from  the  preceding  values  at  A,  F  and  B.  The  values  at  A  and  B 
arc  determined  by  interpolation  between  the  neighboring  points 
(K,  F)  and  (F,  G)  respectively  (sec  Fig.  11.6.9).  Of  course,  it  is  ne- 
cessary to  treat  the  motions  in  each  of  the  branches  away  from  the 
junction  by  the  same  methods  as  were  described  for  the  problem  of 
the  Ohio  treated  above,  and  this  is  feasible  once  the  values  of  v  and  c 
have  been  obtained  at  the  junction. 

The  results  of  the  calculations  are  shown  in  Fig.  11.6.10,  which 
furnishes  the  river  profiles,  i.e.  the  depths  as  functions  of  the  location 
in  each  of  the  three  branches,  for  times  t  =  (K  2.5,  4,  and  10  hours 
after  the  beginning  of  the  flood  50  miles  up  the  Ohio.  The  curves  for 
t  =  oo  are  those  for  the  steady  flow  which  was  calculated  above  in 
sec*.  11.2  (cf.  Fig.  11.2.3).  The  calculations  indicate  that  the  unsteady 
flow  does  tend  to  the  steady  flow  as  the  time  increases.  Another  no- 
ticeable effect  is  the  backwater  effect  in  the  upper  branch  of  the 
Mississippi.  For  example,  the  stage  is  increased  by  about  2  feet  at  a 
point  in  the  Mississippi  20  miles  above  the  junction  and  7.5  hours  after 
the  flood  wave  from  the  Ohio  first  reaches  the  junction. 


498 


WATER   WAVES 


It  might  be  mentioned  that  the  forerunners  of  the  flood  in  all  three 
branches  were  computed  by  using  the  expansion  scheme  which  is 
explained  in  the  appendix  to  this  chapter. 


x  =  distonce  in  miles 

measured  from  junction 

y  =  stage  measured  in  feet 

t=  time  in  hours  after  start 
of   flood 


40' 


////     i  i i  i  i i i     1 1  i  i i i  i  i  i  i  1 1  i  i 


10' 


Fig.  11.6.10  River  profiles  for  the  junction 


11.7.  Numerical  prediction  of  an  actual  flood  in  the  Ohio,  and  at  its 
junction  with  the  Mississippi.  Comparison  of  the  predicted  with 
the  observed  floods 

The  methods  for  numerical  analysis  of  flood  wave  problems  in 
rivers  developed  above  and  applied  to  simplified  models  of  the  Ohio 
and  its  junction  with  the  Mississippi  have  been  used  to  predict  the 
progress  of  a  flood  in  the  Ohio  as  it  actually  is,  and  likewise  to  predict 
the  progress  of  a  flood  coming  from  the  Ohio  and  passing  through  the 
junction  with  the  Mississippi.  The  data  for  the  flood  in  the  Ohio  were 
taken  for  the  case  of  the  big  flood  of  1945,  and  predictions  were  made 
numerically  for  periods  up  to  sixteen  days  for  the  400-mile  long 
stretch  of  the  Ohio  extending  from  Wheeling,  West  Virginia,  to 
Cincinnati,  Ohio.  For  the  flood  through  the  junction,  the  data  for 
the  1947  flood  were  used,  and  predictions  were  made  in  all  three 
branches  for  distances  of  roughly  40  miles  from  the  junction  along 


MATHEMATICAL   HYDRAULICS  499 

each  branch.  In  each  case  the  state  of  the  river,  or  river  system,  was 
taken  from  the  observed  flood  at  a  certain  time  t  =  0;  for  subsequent 
times  the  inflows  from  tributaries  and  the  local  run-off  in  the  main 
river  valley  were  taken  from  the  actual  records,  and  then  the  differ- 
ential equations  were  integrated  numerically  with  the  use  of  the 
UNIVAC  digital  computer  in  order  to  obtain  the  river  stages  and  dis- 
charges at  future  times.  The  flood  predictions  made  in  this  way  were 
then  compared  with  the  actual  records  of  the  flood. 

A  comparison  of  observed  with  calculated  flood  stages  will  be  given 
later  on;  however,  it  can  be  said  in  general  that  there  is  no  doubt  that 
this  method  of  dealing  with  flood  waves  in  rivers  is  entirely  feasible 
since  it  gives  accurate  results  without  the  necessity  for  unduly  large 
amounts  of  expensive  computing  time  on  a  machine  such  as  the 
UNIVAC.  For  example,  a  prediction  for  six  days  in  the  400-mile 
stretch  of  the  Ohio  requires  less  than  three  hours  of  machine  time. 
This  amount  of  calculating  time  — which  is  anyway  not  unreasonably 
large  — could  almost  certainly  be  materially  reduced  by  modifying 
appropriately  the  basic  methods;  so  far.  no  attention  has  been  given 
to  this  aspect  of  the  problem,  since  it  was  thought  most  important 
first  of  all  to  find  out  whether  the  basic  idea  of  predicting  floods  by 
integrating  the  complete  differential  equations  is  sound.  The  fact  that 
such  problems  can  be  solved  successfully  in  this  way  is,  of  course,  a 
matter  of  considerable  practical  importance  from  various  points  of 
view.  For  example,  this  method  of  dealing  with  flood  problems  in 
rivers  is  far  less  expensive  than  it  is  to  build  models  of  a  long  river  or 
a  river  system,  and  it  appears  to  be  accurate.  Actually,  the  two 
methods— empirically  by  a  model,  or  by  calculation  from  the  theory 
-  -are  in  the  present  case  basically  similar,  since  the  models  are  really 
huge  and  expensive  calculating  machines  of  the  type  called  analogue 
computers,  and  the  processes  used  in  both  methods  are  at  bottom  the 
same,  even  in  details.  An  amplification  of  these  remarks  will  be  made 
later  on. 

It  would  require  an  inordinate  amount  of  space  in  this  book  to  deal 
in  detail  with  the  methods  used  to  convert  the  empirical  data  for  a 
river  into  a  form  suitable  for  computations  of  the  type  under  discussion 
here,  and  with  the  details  of  coding  for  the  calculating  machine;  for 
this,  reference  is  made  to  a  report  [1.4].  Instead,  only  a  brief  outline 
of  the  procedures  used  will  be  given  here. 

In  the  first  place,  it  is  necessary  to  have  records  of  past  floods  with 
stages  up  to  the  maximum  of  any  to  be  predicted.  It  would  be  ideal 


500  WATER   WAVES 

to  have  records  of  flood  stages  and  discharges  (or,  what  comes  to  the 
same  thing,  of  average  velocities  over  a  cross-section)  at  points  closely 
spaced  along  the  river— at  ten  mile  intervals,  say.  Unfortunately, 
measurements  of  this  kind  are  available  only  at  much  wider  inter- 
vals *  — of  the  order  of  50  to  80  miles  or  more— even  in  the  Ohio 
River,  for  which  the  data  are  more  extensive  than  for  most  rivers  in 
the  United  States.  From  such  records,  it  is  possible  to  obtain  the  co- 
efficient of  the  all-important  resistance  term  in  the  differential  equa- 
tion expressing  the  law  of  conservation  of  momentum.  This  coefficient 
depends  on  both  the  location  of  the  point  along  the  river  and  the 
stage.  The  other  essential  quantity,  the  cross-section  area,  also  as  a 
function  of  location  along  the  river  and  of  stage,  could  in  principle 
be  determined  from  contour  maps  of  the  river  valley;  this  is,  in  fact, 
the  method  used  in  building  models,  and  it  could  have  been  used  in 
setting  up  the  problem  for  numerical  calculation  in  the  manner  under 
discussion  here.  If  that  had  been  done,  the  results  obtained  would 
probably  have  been  more  accurate;  however,  such  a  procedure  is 
extremely  laborious  and  time  consuming,  and  since  the  other  equally 
important  empirical  element,  i.e.  the  resistance  coefficient,  is  known 
only  as  an  average  over  each  of  the  reaches  (this  applies  equally  to 
the  models  of  a  river),  it  seems  reasonable  to  make  use  of  an  average 
cross-section  area  over  each  reach  also.  Such  an  average  cross-section 
area  was  obtained  by  analyzing  data  from  past  floods  in  such  a  way 
as  to  determine  the  water  storage  volumes  in  each  reach,  and  from  them 
an  average  cross-section  area  as  a  function  of  the  river  stages  was 
calculated.  In  this  way  the  coefficients  of  the  differential  equations  are 
obtained  as  numerically  tabulated  functions  of  x  and  y.  (It  might 
perhaps  be  reasonable  to  remark  at  this  point  that  the  carrying  out 
of  this  program  is  a  fairly  heavy  task,  which  requires  close  cooperation 
with  the  engineers  who  are  familiar  with  the  data  and  who  understand 
also  what  is  needed  in  order  to  operate  with  the  differential  equations). 
In  Fig.  11.7.1  a  diagrammatic  sketch  of  the  Ohio  River  between 
Wheeling  and  Cincinnati  is  shown,  together  with  the  reaches  and 
observation  stations  at  their  ends.  What  we  now  have  are  resistance 
coefficients  and  cross-section  areas  that  represent  averages  over  any 
given  reach.  However,  the  reaches  are  too  long  to  serve  as  intervals 
for  the  method  of  finite  differences— which  is  basic  for  the  numerical 
integration  of  the  differential  equations.  Rather,  an  interval  between 

*  Each  such  interval  is  called  a  reach  by  those  who  work  practically  with  river 
regulation  problems. 


MATHEMATICAL   HYDRAULICS  501 


net  points  (in  the  staggered  scheme  described  in  the  preceding  section) 
of  10  miles  was  taken  in  order  to  obtain  a  sufficiently  accurate  approx- 
imation to  the  exact  solution  of  the  problem.  A  time  interval  of 


Cincinnati 


Huntmgton 
Fig.  11.7.1.  Reaches  in  the  Ohio 

9  minutes  was  used.  Actually,  calculations  were  first  made  using  a 
5-mile  interval  along  the  river,  but  it  was  found  on  doubling  the  inter- 
val to  10  miles  that  no  appreciable  loss  in  accuracy  resulted. 

To  begin  with,  flood  predictions  for  the  1945  flood  were  made,  start- 
ing at  a  time  when  the  river  was  low  and  the  flow  was  practically  a 
steady  flow.  Calculations  were  first  made  for  a  36  hour  period  during 
which  the  flood  was  rising;  as  stated  earlier,  these  were  made  using 
the  measured  inflows  from  tributaries,  and  the  estimated  run-off 
in  the  main  valley.  Upon  comparison  with  the  actual  records,  it  was 
found  that  the  predicted  flood  stages  were  systematically  higher  than 
the  observed  flood  stages,  and  that  the  discrepancy  increased  steadily 
with  increase  in  the  time.  It  seemed  reasonable  to  suppose  that  the 
error  was  probably  due  to  an  error  in  the  resistance  coefficient.  Con- 
sequently a  series  of  calculations  was  made  on  the  UNIVAC  in  which 
this  coefficient  was  varied  in  different  ways;  from  these  results,  cor- 
rected coefficients  were  estimated  for  each  one  of  the  reaches.  Actually 


502  WATER   WAVES 

this  was  done  rather  roughly,  with  no  attempt  to  make  corrections 
that  would  require  a  modification  in  the  shape  of  these  curves  in  their 
dependence  on  the  stage.  The  new  coefficients,  thus  corrected  on  the 
basis  of  36-hour  predictions  (and  thus  for  flood  stages  far  under  the 
maximum),  were  then  used  to  make  predictions  for  various  6-day 
periods,  as  well  as  some  16-day  periods,  with  quite  good  results,  on 
the  whole. 

It  might  be  said  at  this  point  that  making  such  a  correction  of  the 
resistance  oceff icient  on  the  basis  of  a  comparison  with  an  actual  flood 
corresponds  exactly  to  what  is  done  in  making  model  studies.  There, 
it  is  always  necessary  to  make  a  number  of  verification  runs  after  the 
model  is  built  in  order  to  compare  the  observed  floods  in  the  model 
with  actual  floods.  In  doing  so,  the  first  run  is  normally  made  without 
making  any  effort  to  have  the  resistance  correct  — in  fact,  the  rough- 
ness of  the  concrete  of  the  model  furnishes  the  only  resistance  at  the 
start.  Of  course  it  is  then  observed  that  the  flood  stages  arc  too  low 
because  the  water  runs  off  too  fast.  Brass  knobs  are  then  screwed 
into  the  bed  of  the  model,  and  wire  screen  is  placed  at  some  parts  of 
the  model,  until  it  is  found  that  the  flood  stages  given  by  the  model 
agree  with  the  observations.  This  is,  in  effect,  what  was  done  in 
making  numerical  calculations.  In  other  words,  the  resistance  cannot 
be  scaled  properly  in  a  model,  but  must  be  taken  care  of  in  an  empi- 
rical way.  The  model  is  thus  not  a  true  model,  but,  as  was  stated  earlier, 
it  is  rather  a  calculating  machine  of  the  class  called  analogue  com- 
puters. It  is,  however,  a  very  expensive  calculating  machine  which  can, 
in  addition,  solve  only  one  very  restricted  problem.  A  model  of  two 
fair  sized  rivers,  for  example,  consisting  of  two  branches  perhaps  200 
miles  in  length  upwards  from  their  junction,  together  with  a  short 
portion  below  the  junctions,  could  cost  more  than  a  UNIVAC. 

It  has  already  been  stated  that  average  cross-section  areas  for  the 
individual  reaches  were  used  in  making  the  numerical  computations, 
while  in  the  model  the  cross-sections  arc  obtained  from  the  contour 
maps.  In  operating  numerically  it  is  possible  to  change  the  local  cross- 
section  areas  without  any  difficulty,  and  this  might  be  necessary  at 
certain  places  along  the  river. 

Some  idea  of  the  results  of  the  calculations  for  the  1945  flood  in  the 
Ohio  is  given  by  Fig.  11.7.2.  The  graph  shows  the  river  stage  at  Po- 
meroy  as  a  function  of  the  time.  At  the  other  stations  the  results  were 
on  the  whole  more  accurate.  The  graph  marked  "computation  with 
original  data",  and  which  covers  a  36  hour  period,  was  computed  on 


MATHEMATICAL    HYDRAULICS 


503 


the  basis  of  the  resistance  coefficients  as  estimated  from  the  basic 
flow  data  for  the  river.  As  one  sees,  these  coefficients  resulted  in  much 
too  high  stages,  and  corrections  to  them  were  made  along  the  river 


Computed  hydrogroph 
(resistance  adjusted) 


554 


12  0  12 

Feb  27  Feb  28 


Fig.  1 1 .7.2.  Comparison  of  calculated  with  observed  stages  at  Pomeroy  for  the 
1945  flood  in  the  Ohio  River 

on  the  basis  of  the  results  of  this  computation.  Afterwards,  flood 
predictions  wore  made  for  periods  up  to  16  days  without  further 


Thebes 


Metropolis 


MISSISSIPPI 


Hickmon 


32m. 


Fig.  11.7.3.  The  junction  of  the  Ohio  and  the  Mississippi 


504 


WATER   WAVES 


correction  of  these  coefficients.  The  graph  indicates  results  for  a  6  day 
period  during  which  the  flood  was  rising.  Evidently,  the  calculated 
and  observed  stages  agree  very  well. 


300 


296 


292- 


288 


Stage  at  Hickman 


Jan  15  18          21  24         27          30 


304- 


300 


Observed  stages 
Computed  hydrograph 

Stage  at  Cairo 


Jan    15  18          21  24          27          30 

Fig.  11.7.4.  Calculated  and  observed  stages  at  Cairo  and  Hickman 

In  Fig.  11.7.3  a  diagrammatic  sketch  of  the  junction  of  the  Ohio 
and  the  Mississippi  is  shown  indicating  the  portions  of  these  rivers 
which  entered  into  the  calculation  of  a  flood  coming  down  the  Ohio 
and  passing  through  the  junction.  The  flood  in  question  was  that  of 


MATHEMATICAL   HYDRAULICS  505 

1947.  It  was  assumed  that  the  stages  at  Metropolis  in  the  Ohio  (about 
40  miles  above  Cairo)  and  at  Thebes  in  the  upper  Mississippi  (also 
about  40  miles  above  Cairo)  were  given  as  a  function  of  the  time.  At 
Hickman  in  the  lower  Mississippi  (about  40  miles  below  Cairo)  the 
stage-discharge  relation  at  this  point,  as  known  from  observations, 
was  used  as  a  boundary  condition.  The  results  of  a  calculation  for  a 
16  day  period  are  shown  in  Fig.  11.7.4,  which  gives  the  stages  at 
Cairo,  and  at  the  terminating  point  in  the  lower  Mississippi,  i.e.  at 
Hickman.  As  one  sees,  the  accuracy  of  the  prediction  is  very  high, 
the  error  never  exceeding  0.6  foot.  It  might  be  mentioned  that  a 
prediction  for  6  days  requires  about  one  hour  of  calculating  time 
on  the  UNIVAC,  so  that  the  calculating  time  for  the  16  day  period 
was  under  8  hours,  which  seems  reasonable.  This  problem  of  rout- 
ing a  flood  through  a  junction  is,  as  has  been  mentioned  before, 
one  which  has  not  been  dealt  with  successfully  by  the  engineering 
methods  used  for  flood  routing  in  long  rivers.* 


Appendix  to  Chapter  11 

Expansion  in  the  neighborhood  of  the  first  characteristic 

It  has  been  mentioned  already  that  whereas  the  forerunner  of  a 
disturbance  initiated  at  a  certain  point  in  a  river  at  a  moment  when 
the  flow  is  uniform  travels  downstream  with  the  speed  v  +  \/gy,  the 
main  part  of  the  flood  wave  travels  more  slowly  (cf.  Deymte  [D.9]), 
depending  strongly  on  the  resistance  of  the  river  bed.  An  investigation 
of  the  motion  near  the  head  of  the  wave,  i.e.  near  the  first  characteris- 
tic (cf.  the  first  part  of  sec.  11.6)  with  the  equation  x  =  (VQ  +  <?0)<, 
shows  immediately  why  the  main  part  of  the  disturbance  will  in 
general  fall  behind  the  forerunners  of  the  wave. 

The  motion  is  investigated  in  this  Appendix  by  means  of  an  ex- 
pansion in  terms  of  a  parameter  that  has  been  devised  by  G.  Whitham 
and  A.  Troesch  and  carried  out  to  terms  of  the  two  first  orders  for  the 
model  of  the  Ohio  River,  and  to  the  lowest  order  in  the  much  more 

*  Added  in  proof:  In  the  meantime,  calculations  have  been  completed  ( see  [I.4a] ) 
for  the  case  of  floods  through  the  Kentucky  Reservoir  at  the  mouth  of  the 
Tennessee  River.  The  calculated  and  observed  stages  differed  only  by  inches  for 
a  flood  period  of  three  weeks  over  the  186  miles  of  the  resevoir. 


506 


WATER   WAVES 


complicated  case  of  the  junction  problem.  The  results  obtained  make 
it  possible  to  improve  the  accuracy  of  the  solution  near  the  first 
characteristic  which  separates  the  region  of  undisturbed  flow  from 
that  of  the  flood  wave.  It  turns  out  that  the  finite  difference  scheme 
yields  river  depths  which  are  too  large,  as  indicated  by  Fig.  11. A.I. 


vprofile   computed    by 
finite  differences 


wove 
front 


'  Region  of 
I  undisturbed 
'  uniform   flow 


Fig.  11.  A.I.  Error  introduced  by  finite  difference  scheme  in  neighborhood  of 
first  characteristic  of  a  rapidly  rising  flood  wave 

In  order  to  expand  the  solution  in  the  neighborhood  of  the  wave 
front,  we  introduce  new  coordinates  f  and  r  as  follows: 

£  =  x  and  r  =  (v0  +  c0)t  —  x 

such  that  the  £-axis  (i.e.  r  =  0)  coincides  with  the  first  characteristic. 
Near  the  front  of  the  wave  T  will  be  small,  and  the  expansion  will  be 
carried  out  by  developing  v  and  c  in  powers  of  r.  The  basic  system  of 
equations  is  restated  for  convenience: 

%ccx  +vt+  vvx  -  gS  +  gSf  =  0, 
cvx  +  2vcx  +  2ct  =  0. 


Upon  substitution  of  the  new  variables  |  and  r  we  find 

cT)  +  v(vt-vr)  +  (v0  +  c0)vr-  gS  +  gS,  -  0, 
0  +  c(0£-0T)  +  2(»0  +  cQ)cr  =  0. 


(11.A.1) 


where  the  friction  slope  Sf  for  a  rectangular  channel  of  width  B  is  given 

by 

4/3 


MATHEMATICAL  HYDRAULICS 


507 


We  expand  v  and  c  as  power  series  in  r  with  coefficients  that  are  func- 
tions of  £  as  follows: 


v  = 


This  expansion  is  to  be  used  for  r  >  0  only,  since  for  r  <  0  we  are  in 
the  undisturbed  region  and  all  the  functions  ^(1),  u2(£),  .  .  .,  cx(f  ), 
c2(£),  •  •  •  vanish  identically.  If  we  insert  the  series  for  v  and  c  into 
equations  (ll.A.l)  and  collect  terms  of  the  same  order  in  T,  we  get 
ordinary  differential  equations  for  ^(f  ),  cx(f  ),  .  .  ..  The  equations 
resulting  from  the  terms  of  zero  order  in  r  yield  vl  =  2Cj  .  The 
first  order  terms  become,  after  thus  eliminating  vl  . 


(11.A.2) 


dc± 


1 

*>o 


By  adding  these  two  equations  and  removing  the  common  factor  4, 
we  find  the  differential  equation  for  ^(1)  is: 

1         2  1 


-  0. 


.wove  front 


=  0 


Although  the  solution  of  this  differential  equation  for  ^(f )  is  easily 
obtained,  the  result  expressed  in  general  terms  is  complicated,  and  it 
is  preferable  to  give  it  only  for  the  case  of  the  model  of  the  Ohio  River 


Fig.  11.A.2.  Behavior  near  the  front  of  a  wave 


508 


WATER    WAVES 


using  the  parameter  values  introduced  above.  In  this  case  we  find: 
Cl  =  (1.05  +  8.06  e0'146*)-1,  with  ^  and  f  in  miles  and  hours.  This  re- 
sult has  the  following  physical  meaning:  The  angle  a  of  the  profile 
measured  between  the  wave  front  and  the  undisturbed  water  surface 
dies  out  exponentially:  oc~  1/(1  +  aebx),  with  a  and  b  constants  de- 
pending on  the  river  and  the  boundary  condition  at  x  =  0.  Theore- 
tically, a  could  also  increase  exponentially  downstream  so  that  a  bore 
would  eventually  develop,  but  only  if  the  increase  in  level  at  x  =  0  is 
extremely  fast;  in  our  example  no  bore  will  develop  unless  the  water 
rises  at  the  extremely  rapid  rate  of  at  least  1  ft  per  minute. 
Unfortunately,  the  evaluation  of  c2(£),  which  yields  the  curvature  of 


t  ' 
hours 

10 


=  !9.7t 


50 


100        150          x, 

miles 


Fig.  11.A.3.  Region  of  practically  undisturbed  flow 

the  profile  at  the  wave  front,  is  already  very  cumbersome.  The  curva- 
ture is  found  to  decrease  for  large  x  like  xc~bx,  b  being  a  positive  con- 
stant. With  the  two  highest  order  terms  in  the  expansion  known,  it  is 
possible  to  estimate  the  region  adjacent  to  the  first  characteristic 
where  the  flow  is  practically  undisturbed.  It  is  remarkable  how  far 
behind  the  forerunner  the  first  measurable  disturbance  travels  (see 
Fig.  11.A.3). 

In  a  similar  way,  an  expansion  as  a  power  series  in  T  has  been  carried 
out  for  the  problem  of  the  junction  of  the  Ohio  and  Mississippi,  as 
described  in  earlier  sections.  Here  even  the  lowest  order  term  was 
obtained  only  after  a  complicated  computation,  since  it  was  necessary 
to  work  simultaneously  in  three  different  x,  J-planes,  with  boundary 
conditions  at  the  junction.  The  differential  equations  for  cl  are,  in  all 
three  branches,  of  the  same  type  as  for  the  Ohio,  and  their  solution 
for  the  junction  problem  with  the  parameters  of  section  11.6 


MATHEMATICAL   HYDRAULICS  509 


are  cx  =  .00084  c-145*  for  the  upstream  branch  of  the  Mississippi, 
and  c1  =  .00084  <r~-229*  for  the  downstream  branch  of  the  Mississippi, 
cl  and  f  both  being  given  in  miles  and  hours.  This  means  that  the 
angle  a  also  dies  out  exponentially  in  the  Mississippi,  a  little  faster 
downstream  than  upstream,  as  might  have  been  expected,  since  the 
oncoming  water  in  the  upstream  branch  has  the  affect  of  making  the 
wave  front  steeper. 

In  the  problem  of  the  idealized  Ohio  River  and  of  the  idealized 
problem  of  its  junction  with  the  Mississippi  River  the  expansions 
were  carried  out  numerically  in  full  detail  and  were  used  to  avoid 
computation  by  finite  differences  in  a  region  of  practically  undisturb- 
ed flow.* 


This  would  become  more  and  more  important  if  the  flow  were  to  be  com- 
puted beyond  10  hours. 


PART  IV 


CHAPTER  12 


Problems  in  which  Free  Surface  Conditions  are  Satisfied 
Exactly.  The  Breaking  of  a  Dam.  Levi-Civita's  Theory 

This  concluding  chapter  constitutes  Part  IV  of  the  book.  In  Part  I 
the  basic  general  theory  and  the  two  principal  approximate  theories 
were  derived.  Part  II  deals  with  problems  treated  by  means  of  the 
linearized  theory  arising  from  the  assumption  that  the  motion  is  a 
small  deviation  from  a  state  of  rest  or  from  a  uniform  flow.  Part  III 
is  concerned  with  the  approximate  nonlinear  theory  which  arises 
when  the  depth  of  the  water  is  small,  but  the  amplitude  of  the  waves 
need  not  be  small.  Finally,  in  this  chapter  we  deal  with  a  few  problems 
in  which  no  assumptions  other  than  those  involved  in  the  basic  general 
theory  are  made.  In  particular,  the  nonlinear  free  surface  conditions 
arc  satisfied  exactly. 

The  first  type  of  problem  considered  in  this  chapter  belongs  in  the 
category  of  problems  concerned  with  motions  in  their  early  stages 
after  initial  impulses  have  been  applied.  A  typical  example  is  the 
motion  of  the  water  in  a  dam  when  the  dam  is  suddenly  broken.  This 
problem  will  be  treated  along  lines  worked  out  by  Pohle  [P.ll], 
[P.12],  Similar  problems  involving  the  collapse  of  a  column  of  liquid 
in  the  form  of  a  circular  half-cylinder  or  of  a  hemisphere  resting  on  a 
rigid  bottom  have  been  treated  by  Penney  and  Thornhill  [P.2]  by 
a  method  different  from  that  used  by  Pohle. 

The  second  section  of  the  chapter  deals  with  the  theory  of  steady 
progressing  waves  of  finite  amplitude.  The  existence  of  exact  solutions 
of  this  type  is  proved,  following  in  the  main  the  theory  worked  out 
by  Levi-Civita  [L.7]. 

12.1.  Motion  of  water  due  to  breaking  of  a  dam,  and  related  problems 

With  the  exception  of  the  present  section  we  employ  throughout 
this  book  the  so-called  Euler  representation  in  which  the  velocity  and 
pressure  fields  are  determined  as  functions  of  the  space  variables  and 

513 


514  WATER   WAVES 

the  time.  In  this  section  it  is  convenient  to  make  use  of  what  is  com- 
monly called  the  Lagrange  representation,  in  which  the  displacements 
of  the  individual  fluid  particles  are  determined  with  respect  to  the 
time  and  to  parameters  which  serve  to  identify  the  particles.  Usually 
the  parameters  used  to  specify  individual  particles  are  the  initial 
positions  of  the  particles,  and  we  shall  conform  here  to  that  practice. 
Only  a  two-dimensional  problem  will  be  treated  in  detail  here;  con- 
sequently we  choose  the  quantities  a,  6,  and  t  as  independent  variables, 
with  a  and  b  representing  Cartesian  coordinates  of  the  initial  positions 
of  the  particles  at  the  time  t  —  0.  The  displacements  of  the  particles 
are  denoted  by  X(a,  b;  t)  and  Y(a,  b;  t),  and  the  pressure  by  p(a,  b;  t). 
The  equations  of  motion  are 

Xtt  =  ---  Px 

e> 

Ytt  =  -  -  PY  -    S 
Q 

in  accord  with  Newton's  second  law.  We  assume  gravity  to  be  the 
only  external  force.  These  equations  are  somewhat  peculiar  because 
of  the  fact  that  derivatives  of  the  pressure  p  with  respect  to  the  de- 
pendent variables  X  and  Y  occur.  To  eliminate  them  we  multiply  by 
Xa  and  Fa,  respectively,  and  add,  then  also  by  Xb,  Yb<  and  add; 
the  result  is 


XttXa  +  (Yti 

XttXb  +  (Ytt  +  g)Yb 


(12.1.1)  e 


Q 

and  these  are  the  equations  of  motion  in  the  Lagrangian  form.  These 
equations  are  not  often  used  because  the  nonlinearities  occur  in  an 
awkward  way;  however,  they  have  the  great  advantage  that  a  solu- 
tion is  to  be  found  in  a  fixed  domain  of  the  a,  6-plane  even  though 
a  free  surface  exists.  For  an  incompressible  fluid—  the  only  case 
considered  here—  the  continuity  condition  is  expressed  by  requiring 
that  the  Jacobian  of  X  and  Y  with  respect  to  a  and  b  should  remain 
unchanged  during  the  flow  (since  an  area  element  composed  always 
of  the  same  particles  has  this  property);  but  since  X  —  a  and  Y  =-  b 
initially,  it  follows  that 

(12.1.2)  XaYb  -  XbYa  =  I 


THE   BREAKING   OF   A    DAM  515 

is  the  condition  of  continuity.  If  the  pressure  p  is  eliminated  from 
(12.1.1)  by  differentiation  the  result  is 

(12.1.3)  (XaXbt  +  YaY,t)t  =.  (XbXat  +  YbYat)t. 
Integration  with  respect  to  /  leads  to 

(12.1.4)  (XaXtt  +  YaYbl)  -  (XbXaf  +  Y6Yat)  -  f(a,  b) 

with  /  an  arbitrary  function.  It  can  easily  be  shown  by  a  calculation 
using  the  Eulcrian  representation  that  the  left  hand  side  of  this  equa- 
tion represents  the  vorticity;  consequently  the  equation  is  a  verifi- 
cation of  the  law  of  conservation  of  vorticity.  If  the  fluid  starts  from 
rest,  or  from  any  other  state  with  vanishing  vorticity,  the  function 
j(a,  b)  would  be  zero. 

The  method  used  by  Pohlc  [P.ll],  [P.12]  to  solve  the  equations 
(12.1.1)  and  (12.1.2)— which  furnish  the  necessary  three  equations  for 
the  three  functions  X,  F,  and  p  consists  in  assuming  that  solutions 
exist  in  the  form  of  power  series  developments  in  the  time,  with  co- 
efficients which  depend  on  a  and  b: 


(12.1.5) 


In  these  expansions  we  observe  that  the  terms  of  order  zero  in  X  and  Y 
are  a  and  b  —  in  accordance  with  the  basic  assumption  that  these 
quantities  fix  the  initial  positions  of  the  particles.  It  should  also  be 
noted  that  X(l)  and  F(1)  are  the  components  of  the  initial  velocity, 
and  X(2)  and  F(2)  similarly  for  the  acceleration;  in  general,  we  would 
therefore  expect  that  X(l)  and  F(1)  would  be  prescribed  in  advance  as 
part  of  the  initial  conditions.  Of  course,  boundary  conditions  imposed 
on  .Y,  F,  and  p  would  lead  to  boundary  conditions  for  the  coefficient 
functions  in  the  series  developments.  The  convergence  of  the  series 
for  the  cases  discussed  below  has  not  been  studied,  but  it  seems  likely 
that  the  scries  would  converge  at  least  for  sufficiently  small  values  of 
the  time.  The  convergence  of  developments  of  this  kind  in  some  simp- 
ler problems  in  hydrodynamics  has  been  proved  by  Lichtenstein 
[L.12]. 

The  series  (12.1.5)  are  inserted  first  in  equation  (12.1.2)  and  the 
coefficient  of  each  power  of  /  is  equated  to  zero  with  the  following 
result  for  the  first  two  terms: 


X(a9  b;  t)  =--  a  +  X™(a.  b)  • 1  +  X^(a,  b)  • 1*  +  . . ., 
Y(a,  ft;  0  -  b  +  Y™(a.  b)-t  +  F<2>(0,  b)  •  *2  +  . . ., 
p(a.  ft;  0  -  p(0)(fl,  b)  +  p(l)(a.  b)  •  t  +  p<2)(0,  b)  •  t2  + 


516  WATER   WAVES 


F<2>  =  - 

We  observe  that  X(l}  and  F(1)  are  subject  to  the  above  relation  and 
hence  cannot  both  be  prescribed  arbitrarily;  however,  if  the  fluid 
starts  from  rest  so  that  X(l}  =  F(1)  —  0,  the  condition  is  automatic- 
ally satisfied.  The  equation  for  X(2)  and  F(2)  is  linear  in  these  quanti- 
ties, but  nonlinear  in  X  (1)  and  F(1).  This  would  be  the  situation  in 
general:  X(n)  and  F(n)  would  satisfy  an  equation  of  the  form 

jqw  +  y<«)  =  F(X<»9  y(1),  *<2),  F<2>,  .  .  .,  JC<»-i>,  F<"-»), 

with  F  a  nonlinear  function  in  X(i\  Y(l\  i  =  1,  2,  .  .  .,  n  —  1.  In 
the  following  we  shall  consider  only  motions  starting  from  rest.  Con- 
sequently, we  have  X(l)  =  F(1)  =  0,  and  equation  (12.1.4)  holds 
with  /  =  0;  a  substitution  of  the  series  in  powers  of  t  in  equation 
(12.1.4)  yields  (for  the  lowest  order  term): 

(12.1.7)  X?>  -  Ff  -  0. 

The  higher  order  coefficients  satisfy  an  equation  of  the  form 
X(n)  _  y(n)  =  G(JC<«,  F<2>,  .  .  .,  Xi*-u9  F<w~1>),  with  G  a  nonlinear 
function  of  X(i\  F(t),  i  =  2,  3,  .  .  .,  n  -  1.  Thus  we  observe  that  X™ 
and  F(2)  satisfy  the  Cauchy-Riemann  equations  and  arc  therefore 
conjugate  harmonic  functions  of  a  and  b.  The  higher  order  coefficients 
would  satisfy  Poisson's  equation  with  a  right  hand  side  a  known 
function  fixed  by  the  coefficient  functions  of  lower  order.  Thus  the 
coefficients  in  the  series  for  X  and  F  can  be  determined  step-wise  by 
solving  a  sequence  of  Poisson  equations.  Once  the  functions  X(t)  and 
Y(i)  have  been  determined,  the  coefficients  in  the  series  for  the 
pressure  p  can  also  be  determined  successively  by  solving  a  sequence 
of  Poisson  equations.  To  this  end  we  of  course  make  use  of  equations 
(12.1.1);  the  result  for  p(o)(a,b)  is 

(12.1.8)  p(*  +  pfl  =  -  2Q(XP  +  Ff  )  =  0, 

from  (12.1.6)  and  JC(1)  =  F^>  =  0.  Thus  p(0}(a,  b)  is  a  harmonic 
function.  For  p(n)(a,  b)  one  would  find  a  Poisson  equation  with  a  right 
hand  side  determined  by  X(i)  and  F(t)  for  i  =  2,  3,  .  .  .,  n  +  2. 

It  would  be  possible  to  consider  boundary  conditions  in  a  general 
way,  but  such  a  procedure  would  not  be  very  useful  because  of  its 
complexity.  Instead,  we  proceed  to  formulate  boundary  conditions 
for  the  special  problem  of  breaking  of  a  dam,  which  is  in  any  case 
typical  for  the  type  of  problems  for  which  the  present  procedure  is 


THE    BREAKING    OF   A    DAM 


517 


recommended.  We  assume  therefore  that  the  region  occupied  initially 
by  the  water  (or  rather,  a  vertical  plane  section  of  that  region)  is  the 
half-strip  0  ^  a  <  oo,  0  ^  b  ^  h,  as  indicated  in  Fig.  12.1.1.  The 


b=  h 


b=0  o 

Fig.  12.1.1.  The  breaking  of  a  dam 

damris  of  course  located  at  a  =  0.  Since  we  assume  that  the  water 
is  initially  at  rest  when  filling  the  half-strip  we  have  the  conditions 

(12.1.9)  X(a.b;0)  =  a,         Y(a,b;0)  =  b, 
and 

(12.1.10)  Xt(a,  b;  0)  =  0.         Yt(a,  b;  0)  =  0. 

When  the  dam  is  broken,  the  pressure  along  it  will  be  changed 
suddenly  from  hydrostatic  pressure  to  zero;  it  will  of  course  be  pre- 
scribed to  be  zero  on  the  free  surface.  This  leads  to  the  following 
boundary  conditions  for  the  pressure: 


(12.1.11) 


(  p(a,  h;  t)  =  0, 


\ 


=  0, 


0 
0 


a  <  GO, 
b  <  h, 


t  >  0, 


Finally  the  boundary  condition  at  the  bottom  b  =  0  results  from  the 
assumption  that  the  water  particles  originally  at  the  bottom  remain 
in  contact  with  it;  as  a  result  we  have  the  boundary  condition 


0 


a  <  oo,     t  >  0. 


(12.1.12)  Y(a,  0;*)=0. 

The  conditions  (12.1.9)  are  automatically  satisfied  because  of  the 
form  (cf.  (12.1.5))  chosen  for  the  scries  expansion.  The  conditions 
(12.1.10)  are  satisfied  by  taking  X(l)(a,  b)  =  F(1)(a,  b)  =  0. 

In  order  to  determine  the  functions  X(2)(a,  b)  and  F(2)(a,  6),  it  is 
necessary  to  obtain  boundary  conditions  in  addition  to  the  differential 
equations  given  for  them  by  (12.1.6)  and  (12.1.7).  Such  boundary 
conditions  can  be  obtained  by  using  (12.1.11)  and  (12.1.12)  in  con- 
junction with  (12.1.1)  and  the  power  series  developments.  Thus  from 


518  WATER   WAVES 

(12.1.12)  we  find  F<2)(0,  0)  ±=  0  for  0  ^  a  <  oo  (indeed,  F<">(a,  0) 
would  be  zero  for  all  n).  Insertion  of  the  series  (12.1.5)  and  use  of  the 
boundary  conditions  for  6  =  h  yields 

(12.1.13)  XW(a,  h)  =  0, 

upon  using  the  first  of  the  equations  in  (12.1.1).  The  second  equation 
of  (12.1.1)  leads  to  the  condition 

(12.1.14)  F<2>(0,&)=:  -1 

2 

We  know  that  Z(z)  ==  F<2)  +  iX™  is  an  analytic  function  of  the 
complex  variable  z  =  a  +  ib  in  the  half-strip,  and  we  now  have 
prescribed  values  for  either  its  real  or  its  imaginary  part  on  each  of  the 
three  sides  of  the  strip;  it  follows  that  the  function  Z  can  be  deter- 
mined by  standard  methods— for  example  by  mapping  conformally 
on  a  halfplane.  In  fact,  the  solution  can  be  given  in  closed  form,  as 
follows:  Since  X™(a,  h)  =  0,  we  see  that  X^(a9  h)  =  0,  and  hence 
that  F£2)(a,  h)  =  Q  since  X(2)  and  F(2)  arc  harmonic  conjugates. 
Therefore  the  harmonic  function  F(2)(a,  b)  can  be  continued  over  the 
line  b  =  h  by  reflection  into  a  strip  of  width  2/i,  as  indicated  in  Fig. 
12.1.2;  the  boundary  values  for  F(2)  are  also  shown.  Thus  a  complete- 

Y(2)-b  b:2h 


b=0 


Fig.  12.1.2.  Boundary  value  problem  for  F(2)(a,  b) 

ly  formulated  boundary  value  problem  for  F(2)(0,  b)  in  a  half-strip 
has  been  derived.  To  solve  this  problem  we  map  the  half-strip  on  the 
upper  half  of  a  w-plane  by  means  of  the  function  w  =  cosh  (nz/2h) 
—either  by  inspection  or  by  using  the  Schwarz-Christoffel  mapping 
formula— and  observe  that  the  vertices  z  =  0  and  z  =  2ih  of  the  half- 
strip  map  into  the  points  w  =  ±  1  of  the  wj-plane,  as  indicated  in 
Fig.  12.1.3.  The  appropriate  boundary  values  for  Y(2}(w)  on  the  real 
axis  of  the  w-plane  are  indicated.  The  solution  for  Y(2)(w)  under 


THE    BREAKING    OF   A    DAM 


519 


w-  plane 


-H 


Fig.  12.1.3.  Mapping  on  the  w-plane 

these  conditions  is  well  known;  it  is  the  function  Y(2)(P)  ~ 
—  (g/2n)(62  —  0J,  with  0X  and  02  the  angles  marked  in  Fig.  12.1.3. 
The  analytic  function  of  which  this  is  the  real  part  is  well  known;  it  is 


y<2> 


+  1 


as  can  in  any  case  be  easily  verified.  Transferring  back  to  the  z-plane 
we  have 


u 

cosh  —  —  1 
2h 


Z(z)  =  Y(2)  +  iX™  =  -        log 


, 
cosh 


and  upon  separation  into  real  and  imaginary  parts  we  have  finally: 

cos2  —  +  sinh2  — 


(12.1.15) 


X<*>(a9b)  =  -  -*- 


sn 


sinh2  — 


F<2>(a,  6)  =  - 


n 


.    nb 

sin  — 

2h 


.  ,  na 
smh  — 


One  checks  easily  that  the  boundary  conditions  X(2>(a,  h)  —  0, 
F<2>(a,  0)  =  0  are  satisfied,  and  that  F(2)(0,  6)  =  —  g/2.  The  initial 
pressure  distribution  p(0)(a,  6)  can  be  calculated,  now  that  X(2)(a.  b) 
is  known,  by  using  the  first  equation  of  (12.1.1),  which  yields 


520  WATER   WAVES 


(12.1  16)  pj»  - 

In  the  present  case  there  are  advantages  in  working  first  with  the 
pressure  p(a,  b;  t)  and  determining  the  coefficient  of  the  series  for  it 
directly  by  solving  appropriate  boundary  value  problems;  afterwards 
the  coefficients  of  the  series  for  X  and  Y  are  easily  found.  The  main 
reason  for  basing  the  calculation  on  the  pressure  in  the  first  instance 
is  that  the  boundary  conditions  at  b  =  h  and  a  =  0  are  very  simple, 
i.e.  p  =  0  and  hence  p<*>  =  0  for  all  indices  i.  The  boundary  conditions 
at  the  bottom  6  =  0  involve  the  displacements  Y.  For  instance,  one 
finds  readily  in  the  same  general  way  as  above  that  p^  =  —  gg, 
pW  =.  0,  and  pj,2)  _  __  QgY^  as  boundary  conditions  at  6  =  0. 
Since  p<0)  is  harmonic,  it  is  found  at  once  without  reference  to  dis- 
placements—an interesting  fact  in  itself.  Once  p(o)  is  found,  X(2)  and 
F(2)  can  be  calculated  without  integrations  (cf.  (12.1.16),  for  example). 
Since  p^  =  0  for  b  =  0,  and  p(1)  is  also  harmonic,  it  follows  that 
p(1)(a,  b)  =  0.  Since  F(2)  is  now  known,it  follows  that  a  complete  set 
of  boundary  conditions  for  p(2)(a,  b)  is  known,  and  p(2)(a,  6)  is  then 
determined  by  solving  the  differential  equation 


(12.1.17) 

rd(x(z)  F(2M 
=  e  ^ST-'-M—  -  V2  {(Jf(2))2  +  (F(2))2 

L     d(a,  b) 


L9     /  I      nCL  n^\ 

h*  I  cosh  —  —  cos  —  I 
\  h  hj 

whose  right  hand  side  is  obtained  after  a  certain  amount  of  mani- 
pulation. This  process  can  be  continued.  One  would  find  next  that 
XW  =  F<3)  =  o,  and  that  X(^  and  F(4)  can  be  found  once  p(2)  is 
known.  However,  the  boundary  condition  at  the  bottom,  and  the  right 
hand  sides  in  the  Poisson  equations  for  the  functions  p(i)(a,b)  become 
more  and  more  complicated. 

The  initial  pressure  p(o)(a,  b)  can  be  discussed  more  easily  on  the 
basis  of  a  Fourier  series  representation  than  from  the  solution  in 
closed  form  obtainable  from  (12.1.16);  this  representation  is 

(12.1.18)     p<°>(a,  6) 


., 

=  oglh  —  b)  --  —-  V  -  e  ---  zh  —  cos 
^          } 


THE    BREAKING   OF   A   DAM 


521 


We  note  that  the  first  term  represents  the  hydrostatic  pressure,  and 
that  the  deviation  from  hydrostatic  pressure  dies  out  exponentially 
as  a  ->  oo  and  also  as  h  ->  0,  i.e.  on  going  far  away  from  the  dam  and 
also  on  considering  the  water  behind  the  dam  to  be  shallow  (or,  better, 
considering  a/h  to  be  large).  This  is  at  least  some  slight  evidence  of 
the  validity  of  the  shallow  water  theory  used  in  Chapter  10  to  discuss 
this  same  problem  of  the  breaking  of  a  dam—  at  least  at  points  not 
too  close  to  the  site  of  the  dam. 

The  shape  of  the  free  surface  of  the  water  can  be  obtained  for  small 
times  from  the  equations 


(12.1.19) 


X  =  a 


=  b 


evaluated  for  a  —  0  (for  the  particles  at  the  face  of  the  dam)  and  for 
b  —  h  on  the  upper  free  surface.  The  results  of  such  a  calculation  for 
the  specific  case  of  a  dam  200  feet  high  are  shown  in  Fig.  12.1.4. 


x 
miles 


50  40  30  20  10  100 

Fig.  12.1.4.  Free  water  surface  after  the  breaking  of  a  dam 


One  of  the  peculiarities  of  the  solution  is  a  singularity  at  the  origin 
a  =  0,  6  —  0  which  is  brought  about  by  the  discontinuity  in  the 
pressure  there.  In  fact,  X(2}  has  a  logarithmic  singularity  for  a  =  0, 
b  =  0,  as  one  sees  from  (12.1.15)  and  X  is  negative  infinite  for  all 
t  ^  0.  This,  of  course,  indicates  that  the  approximation  is  not  good 
at  this  point;  in  fact,  there  would  be  turbulence  and  continuous 
breaking  at  the  front  of  the  wave  anyway  so  that  any  solution 
ignoring  these  factors  would  be  unrealistic  for  that  part  of  the  flow. 
In  the  thesis  by  Pohle  [P.ll],  the  solution  of  the  problem  of  the 
collapse  of  a  liquid  half-cylinder  and  of  a  hemisphere  on  a  rigid  plane 
are  treated  by  essentially  the  same  method  as  has  been  explained  for 


522  WATER   WAVES 

the  problem  of  the  breaking  of  a  dam.  These  problems  have  also  been 
treated  by  Penney  and  Thornhill  [P.2],  who  also  use  power  series  in 
the  time  but  work  with  the  Eulerian  rather  than  the  Lagrangian  re- 
presentation, which  leads  to  what  seem  to  the  author  to  be  more  com- 
plicated calculations  than  are  needed  when  the  Lagrangian  represen- 
tation is  used. 

12.2.  The  existence  of  periodic  waves  of  finite  amplitude 

In  this  section  a  proof,  in  detail,  of  the  existence  of  two-dimensional 
periodic  progressing  waves  of  finite  amplitude  in  water  of  infinite 
depth  will  be  given.  This  problem  was  first  solved  by  Nekrassov 
[N.I,  la]  and  later  independently  by  Levi-Civita  [L.7];  Struik 
[S.29]  extended  the  proof  of  Levi-Civita  to  the  same  problem  for 
water  of  finite  constant  depth.  A  generalization  of  the  same  theory 
to  liquids  of  variable  density  has  been  given  by  Dubreuil-Jacotin 
[D.I 5,  15a],  Lichtenstein  [L.ll]  has  given  a  different  method  of 
solution  based  on  E.  Schmidt's  theory  of  nonlinear  integral  equations. 
Davies  [D.5]  has  considered  the  problem  from  still  a  different  point  of 
view.  Gerber  [G.5]  has  recently  derived  theorems  on  steady  flows  in 
water  of  variable  depth  by  making  use  of  the  Schauder-Leray  theory. 

We  shall  start  from  the  formulation  of  the  problem  given  by  Levi- 
Civita  (and  already  derived  in  10.9  above),  but,  instead  of  proving 
directly,  as  he  does,  the  convergence  of  a  power  series  in  the  ampli- 
tude to  the  solution  of  the  problem,  an  iteration  procedure  devised 
by  W.  Littman  and  L.  Nirenberg  will  be  used  to  establish  the  existence 
of  the  solution.  The  two  procedures  are  not,  however,  essentially 
different. 

It  is  convenient  to  break  up  this  rather  long  section  into  sub-sec- 
tions as  a  means  of  focusing  attention  on  separate  phases  of  the 
existence  proof. 

12.2a.  Formulation  of  the  problem 

As  in  sec.  10.9,  the  problem  of  treating  a  progressing  wave  which 
moves  unchanged  in  form  and  with  constant  velocity  is  reduced  to  a 
problem  of  steady  flow  by  observing  the  motion  from  a  coordinate 
system  which  moves  with  the  wave.  A  complex  velocity  potential  (see 
sec.  10.9  for  details)  %(z)  is  therefore  to  be  found  in  the  #,  t/-plane 
(cf.  Fig.  12.2.1): 


LEVI-C1  VITA'S  THEORY 


523 


Fig.  12.2.1.  Periodic  waves  of  finite  amplitude 

(12.2.1)  x  =  <p  +  i\p  =  jf(z),         z  =  x  +  iy. 

The  velocity  at  y  =  —  oo  should  be  [7.  The  real  harmonic  functions 
<p(x,y)  and  y(x,y)  represent  the  velocity  potential  and  the  stream 
function.  The  complex  velocity  w  is  given  by 


(12.2/2) 


w  =  —  —  u  —  iv 
dz 


with  u,  v  the  velocity  components.  This  follows  at  once  from  the 
Cauchy-Riemann  equations: 

(12.2.3)  (px  =  yy  =  w,         <pv  =  -  yx  =  ^ 

since  w  —  993.  -f  iy^. 

We  proceed  to  formulate  the  boundary  conditions  at  the  free  sur- 
face. The  kinematic  free  surface  condition  can  be  expressed  easily  be- 
cause the  free  surface  is  a  stream  line,  and  we  may  choose  y(x>  y)  =  0 
along  it.  The  dynamic  condition  expressed  in  Bernoulli's  law  is  given 

by 

(12.2.4)  |  |  w  |2  +  gy  =  const.         at  y  =  0, 

as  one  can  readily  verify.  The  problem  of  satisfying  this  nonlinear 
condition  is  of  course  the  source  of  the  difficulties  in  deriving  an 
existence  proof.  At  oo  the  boundary  condition  is 

(12.2.5)  w  ->  U  uniformly  as  y  ->  —  oo, 

and  w  is  in  addition  supposed  to  be  nowhere  zero  and  to  be  uniformly 
bounded.  We  seek  waves  which  are  periodic  in  the  ^-coordinate  and 
thus  we  require  ^  to  satisfy  the  condition 

(12.2.6)  X(z  +  h)  -  x(z)  =  0, 
with  h  a  real  constant. 


524  WATER   WAVES 

Following  Levi-Civita,  we  assume  that  the  region  of  flow  in  the 
s-plane  is  mapped  into  the  99,  ^-plane  by  means  of  %(z).  The  free  sur- 
face in  the  physical  plane  corresponds  to  the  real  axis  \p  —  0  of  the 
^-plane,  and  we  assume  that  the  entire  region  of  the  flow  in  the  2-plane 
is  mapped  in  a  one-to-one  way  on  the  lower  half  of  the  ^-plane.  (We 
shall  prove  shortly  that  a  function  %(z)  satisfying  the  conditions  given 
above  would  have  this  property.)  In  this  case  the  inverse  mapping 
z(%)  exists,  and  we  may  regard  the  complex  velocity  w(z)  as  an  ana- 
lytic function  of  #  defined  in  the  lower  half  of  the  ^-plane.  In  this 
way  we  are  enabled  to  work  with  a  domain  in  the  <p,  y-plane  that  is 
fixed  in  advance  instead  of  with  an  unknown  domain  of  the  x,  j/-planc. 
Levi-Civita  goes  a  step  further  by  introducing  a  new  dependent  varia- 
ble co,  replacing  w,\  by  the  relation 

(12.2.7)  w  =  Ue~l°><         co  =  6  +  ir; 

so  that  (jo  is  an  analytic  function  of  cp  -(-  iy.  Consequently  we  have 
(cf.  (12.2.2)) 

(12.2.8)  I  w  |  =  Uer,         0  =  argw. 

Thus  r  —  log  (\w\/U),  while  0  is  the  inclination  of  the  velocity  vector. 
In  the  same  way  as  in  sec.  10.9  (cf.  the  equations  following  (10.9.11)) 
the  boundary  condition  (12.2.4)  can  be  put  in  the  form 

(12.2.9)  0V  =  AV3T  sin  0,         for  y>  =  0, 
with  A'  defined  by 

(12.2.10)  A'  =  g/f/3. 

Our  problem  now  is  to  determine  an  analytic  function  a>(#)  — 
0(<p9  ip)  +  if(<p,  ty)  in  the  lower  halfplane  y  <  0  and  a  constant  A'  in 
(12.2.9)  such  that  a)  o>  is  analytic  for  \p  <  0,  continuous  for  \p  ^  0, 
b)  Ov  is  continuous  for  y  ^  0  and  the  nonlinear  boundary  condition 
(12.2.9)  is  satisfied,  c)  co  has  the  period  Uh  in  99,  d)  co(%)  ->  0  as 
yj  ->  —  oo,  c)  |  a)(%)  |  fg  £.  The  last  two  conditions  are  motivated  by 
the  conditions  imposed  on  w  at  oo  and  the  condition  w  3=  0:  the  con- 
dition d)  from  (12.2.5)  and  (12.2.7),  while  the  condition  e)  is  imposed 
in  order  to  ensure  that  w  is  uniformly  bounded  away  from  both  zero 
and  infinity.  As  we  shall  see,  the  condition  c)  leads  to  the  periodicity 
condition  (12.2.6)  on  #. 

We  proceed  to  show  briefly  (again  following  Levi-Civita)  that  a 
solution  of  the  problem  we  have  formulated  for  co  would  lead  through 
(12.2.7)  to  a  function  w(%)  and  then  to  a  function  %(z)  through  the 


LEVI-CI  VITA'S  THEORY  525 

differential  equation  d%(z)/dz  =  w(%)  which  satisfies  all  of  the  condi- 
tions formulated  above.  The  essential  items  requiring  verification 
are  the  periodicity  condition  and  the  one-to-one  character  of  the 
mapping  z(%)  defined  by 


over  the  half  plane  ^  <  0. 

We  proceed  to  investigate  the  second  property.  From  (12.2.7)  we 
have 


l/w(%)  —  —  e~r  (cos  0  +  i  sin  0). 


and  hence  that 


fM  =  i«-* 

W    t/ 


Since  \  co  \  ^  ^,  it  follows  that  w  is  bounded  away  from  zero,  so  that 

rxci«, 
the  integral        —  converges.   Since  both  |  r  |  ^  J  and  |  0  |  ^  J,  it 

Jo  » 
follows  that  &e(I/w)  is  positive  (we  assume  (7  to  be  positive)  and 

bounded  away  from  0  and  oo.  We  have  ^  +  iyv  =  i/w9  oc^  +  iy^  =  l/w9 
so  that  yv  =  ^(l/w)and  x9  —  &te(\\w}\  since  <#e(I/w)  >  0  it  follows 
therefore  that  y  is  a  strictly  monotonic  increasing  function  of  y,  and 
x  similarly  in  (p.  Consequently  the  mapping  z(%)  is  one-to-one,  and 
in  addition  y  ->  —  oo  when  y  ->  —  oo,  while  #  ->  i  oo  when 
9^  ->  i  °°  since  &e(I/w)  is  positive  and  bounded  away  from  zero;  thus 
the  flow  is  mapped  onto  the  entire  halfplane  \p  <  0. 


& 


dz 


We  consider  the  periodicity  condition  next.  We  have  — 

since  a)  has  the  period  Uh  by  assumption.  This  implies  that  z(%  +  Uh) 
—  %(%)  =  const.  This  constant  is  easily  seen  to  have  the  value  h  by 
letting  ip  -»  oo  in  the  formula 

dx 


I 

J 


=  z(x  +  Uh)  - 


since  w  ->  U  uniformly  when  y>  -»•  —  oo.  Consequently  we  have 

*(*  +  UA)  -  x(X)  =  h, 
y(X  +  Uh)  -  y(x)  =  0. 


526  WATER   WAVES 

We  know  from  (12.2.8)  and  |  0  |  ^  J  that  the  stream  lines  \p  =  const, 
have  no  vertical  tangents,  hence  they  can  be  represented  in  the  form 
y  :=-  y(v),  and  the  last  two  equations  show  that  they  are  periodic  in 
x  of  period  h.  The  problem  of  determining  co(%)  subject  to  the  condi- 
tions a)— e)  is  therefore  equivalent  to  the  problem  formulated  for  %(z). 


1  2.2b.  Outline  of  the  procedure  to  be  followed  in  proving  the  existence 
of  the  function 


The  proof  of  the  existence  of  the  analytic  function  co(%)  which 
solves  our  problem  will  be  carried  out  as  follows.  First  of  all,  we  ob- 
serve that  the  problem  has  always  the  solution  co(%)  =0,  correspond- 
ing to  the  uniform  flow  w  =  U  with  undisturbed  free  surface.  We 
shall  begin  by  assuming  that  a  solution  a>(%)  ^  0  exists,  and  will  then 
proceed,  through  the  use  of  the  properties  assumed  for  co,  to  derive  a 
functional  equation  for  the  values  £0(99,  0)  of  co  on  the  boundary  ip  —  0, 
—  oo  <  <p  <  oo.  It  will  then  be  shown  that  the  functional  equation 
has  a  solution  co(<p9  0)  ^=  0  in  the  form  of  a  complex-  valued  continuous 
function  &>(<p),  and  this  function  will  be  used  to  determine  an  analytic 
function  co(<p,  y})  in  —  oo  <  y>  <  0,  —  oo  <  (p  <  oo,  with  co(<p)  as 
boundary  values,  which  is  then  shown  to  satisfy  all  of  the  conditions 
a)-e). 

It  will  occasion  no  surprise  to  remark  at  this  point  that  the  solution 
we  obtain  will  give  a  motion  in  a  neighborhood  of  the  uniform  flow 
with  horizontal  free  surface,  i.e.  with  an  amplitude  in  a  neighborhood 
of  the  zero  amplitude.  Also,  it  should  be  remarked  that  the  problem 
in  perturbation  theory  which  thus  arises  involves  a  bifurcation  pheno- 
menon, since  the  desired  solution  of  the  nonlinear  problem,  once  the 
wave  length  is  fixed,  requires  that  the  perturbations  take  place  in  the 
neighborhood  of  a  definite  value  of  the  velocity  U.  In  other  words,  the 
desired  solution  bifurcates  from  a  definite  one  of  the  infinitely  many 
possible  flows  with  uniform  velocity  which  are  exact  solutions  of  the 
nonlinear  problem. 

The  decisive  relation  in  the  process  just  outlined  is  the  nonlinear 
boundary  condition  (12.2.9).  It  is  convenient  to  introduce  at  this  point 
some  notations  which  refer  to  it,  to  recast  it  in  a  different  and  more 
convenient  form,  and  also  to  derive  a  number  of  consequences  which 
flow  out  of  it.  At  the  same  time,  some  factors  which  motivate  all  that 
follows  will  be  put  in  evidence. 


LEVI-CIVITA'S  THEORY  527 

Since  we  wish  to  concentrate  attention  on  the  boundary  values 
co  (<p,  0)  of  a>,  it  is  useful  to  introduce  the  notation 

(12.2.11)  eo(p,  0)  =  £>(<p)  =  6(<p)  +  ii(<p), 
and  then  to  introduce  the  operator  f[co]  defined  by 

(12.2.12)  /[£]  -  )i(e~*~  sin  0  -  0)  +  ee~^  sin  0  =  F(<p) 
with  e  defined  by 

(12.2.13)  e  -  V  -  L 

The  constant  A  will  be  given  an  arbitrary  but  fixed  value;  the  quantity 
27T/A  will  then  be  the  period  of  the  function  co(%).  The  constant  e,  and 
with  it  A'  through  (12.2.13),  will  be  fixed  by  the  solution  a)(%)  in  a 
manner  to  be  indicated  below,  and  the  propagation  speed  U  is  then 
determined  by  the  formula  (12.2.10).  As  can  be  seen  at  once,  the 
boundary  condition  (12.2.9)  now  takes  the  form 

(12.2.14)  0V  -  A0  -  F(<p),         y  =  0. 

The  reasons  for  writing  the  free  surface  condition  in  the  form 
(12.2.14)  are  as  follows:  As  remarked  above,  we  seek  a  motion  in  the 
neighborhood  of  a  uniform  flow,  so  that  co  as  defined  by  (12.2.7)  should 
be  small  in  some  sense.  It  would  seem  reasonable  to  set  up  an  iteration 
procedure  which  starts  with  that  solution  (o^y,  y)  of  the  problem 
which  results  when  F(<p),  which  contains  the  nonlinear  terms  in  the 
free  surface  condition,  vanishes  identically.  Afterwards  the  successive 
approximations  will  be  inserted  in  F(cp)  to  obtain  a  sequence  of  linear 
problems  whose  solutions  cok  converge  to  the  desired  solution  of  our 
problem. 

The  problem  of  determining  co^y,  ^>),  when  F  =  0,  is  exactly  the 
problem  posed  by  the  linear  theory  which  was  discussed  at  length  in 
Chapter  3;  in  fact,  if  F(q>)  vanishes,  we  know  from  the  discussion  in 
Chapter  3  that  the  only  bounded  conjugate  harmonic  functions 
0i (^»  V>)»  ri(V>  V0>  °ther  than  01  —  TI  =  0,  in  the  lower  half  plane 
ip  <  0  which  satisfy  the  homogeneous  free  surface  condition 
QIV  —  A0!  =  0  are  the  functions 

01(^,  y>)  =  a^  sin  A<p, 
ri(<P)  V)  —  aie^  cos  ^P» 

once  (p  is  taken  to  be  zero  at  a  crest  or  trough  of  the  wave.  Thus  the 
boundedness  condition  at  oo  and  the  homogeneous  free  surface  con- 
dition lead  automatically  to  waves  which  are  sines  or  cosines  of  <p. 


528  WATER   WAVES 

The  ' 'amplitude"  at  is,  of  course,  arbitrary  on  account  of  the  homo- 
geneity of  the  problem.  The  corresponding  function  co1(^)  is  then 


We  suppose,  naturally,  that  the  "amplitude"  |  ax  |  is  small  and  hence 
that  |  ct^  |  is  also  small  of  the  same  order.  The  basic  parameter  in  the 
iteration  procedure  will  be  the  quantity  av  and  the  procedure  will 
be  so  arranged  that  the  quantity  s  in  (12.2.13)  as  well  as  the  iterates 
a)k  will  be  of  order  |  ox  |.  It  is  then  easily  seen  that  F((p)  will  always 
be  of  order  |  ax  |2,  which  indicates  that  such  a  scheme  of  iteration  is 
reasonable.  We  shall  show  that  it  does  indeed  lead  to  a  sequence  COA 
which  converges  to  the  desired  solution  co  for  all  sufficiently  small 
values  of  |  %  |,  and  that  the  solution  co  fixes  a  value  of  e,  and  hence  of 
A'  since  A  is  once  for  all  fixed,  in  a  manner  to  be  explained  in  a  moment. 
It  might  be  mentioned  that  it  is  not  difficult  to  verify  that  the 
corresponding  motion  furnished  in  the  physical  plane  by  Xi(z)  would, 
up  to  terms  of  first  order  in  |  ax  |,  be  given  by 


and  this  coincides  with  what  was  found  in  Chapter  3  by  a  more  direct 
procedure. 

Iterations,  as  we  have  indicated,  are  to  be  performed,  starting  with 
the  solution  a)^  =  iale*(v~i(p]  of  the  linearized  problem,  with  ax  regarded 
as  a  small  parameter.  This  is  then  inserted  in  the  right  hand  side  of 
(12.2.14);  a  bounded  harmonic  function  02(g9,  y)  in  the  half  plane 
ip  <  0  is  then  determined  through  this  nonhomogeneous  boundary 
condition  and  the  corresponding  analytic  function  co2(%)  —  ^2  +  7'T2 
with  it.  In  order  to  solve  the  boundary  problem  for  02  (and  through  it 
o>2),  however,  it  is  necessary  to  dispose  of  the  parameter  e  in  (12.2.13) 
appropriately.  This  comes  about  because,  as  we  have  just  seen,  the 
homogeneous  linear  boundary  value  problem  for  01  has  a  non-trivial 
solution,  6l  —  a^  sin  h<p,  and  hence  an  orthogonality  condition  on 
F(<p)  is  needed  which  will  ensure  the  existence  of  the  solution  of  the 
nonhomogeneous  problem  for  02.  This  condition  is  well  known  to  be 

that  the  integral  F(<p)  sin  hydy  should  vanish.  It  turns  out  that 

the  value  of  e  so  determined  really  is  of  the  same  order  as  av  Con- 
tinuing the  iterations  in  this  fashion,  the  result  is  a  sequence  of 
functions  eon(#),  and  a  sequence  of  corresponding  values  en  of  s  sueh 


LE VI -CI VITA'S  THEORY  529 

that  |  a)n  |  and  |  en  \  are  all  of  order  |  ax  |.  It  is  to  be  shown  that  both 
sequences  converge  to  yield  a  function  CD  (/)  and  a  number  e  which  solve 
the  problem,  the  quantity  A'  in  (12.2.10)  being  fixed  by  e  =  lim  en  and 
the  arbitrarily  chosen  value  of  A  through  (12.2.13). 

We  observe  that  this  whole  procedure  is  in  marked  contrast  with 
the  method  of  solution  of  the  problem  of  the  solitary  wave  given 
by  Friedrichs  and  Hyers  [F.13]  and  explained  in  sec.  10.9;  in  the  latter 
case  the  iteration  procedure  is  quite  different  and  it  is  carried  out 
with  respect  to  a  parameter  which  has  an  entirely  different  signific- 
ance from  the  parameter  ax  which  is  used  here. 

The  procedure  outlined  here  also  differs  from  the  procedure  followed 
by  Liechtenstein  [L.I  1 J  in  solving  the  same  problem.  Lichtenstcin  applies 
E.  Schmidt's  bifurcation  theory  to  an  appropriate  nonlinear  integral 
equation  (essentially  the  counterpart  of  the  functional  equation  to  be 
used  here).  In  this  procedure,  the  basic  idea  is  to  modify  what  cor- 
responds to  the  function  F(q>)  in  (12.2.14)  in  such  a  way  that  the 
modified  problem  (which  is  arranged  to  contain  one  or  more  parame- 
ters) can  always  be  solved.  Afterwards,  conditions  are  written  down 
to  ensure  that  the  modified  problem  is  identical  with  the  original 
problem;  these  conditions  arc  called  the  bifurcation  conditions.  Such 
a  process  could  have  been  used  here  in  conjunction  with  an  iteration 
scheme,  as  a  substitute  for  the  process  of  fixing  the  parameters  en  at 
t-cich  stage  of  the  iteration  procedure  in  the  manner  indicated  above. 

Basically  the  method  of  solution  of  our  nonlinear  problem  just  out- 
lined requires  the  solution  of  a  sequence  of  linear  problems.  We  turn 
next,  therefore,  in  sec.  12.2c  to  the  derivation  of  the  solution  of  these 
linear  problems,  and  afterwards,  in  sec.  12.2d  we  shall  prove  that  an 
appropriate  sequence  of  solutions  of  the  linear  problems  converges  to 
the  desired  solution  of  the  nonlinear  problem. 

12.2c.  The  solution  of  a  class  of  linear  problems 

The  linear  problems  we  have  in  mind  to  solve,  in  accordance  with 
the  above  discussion,  are  problems  for  co(<p,  \p]  =  0(<p,  y)  -f-  ir((p,  \p) 
when  F(q>)  in  (12.2.14)  is  regarded  as  a  given  function.  That  is,  u> 
should  satisfy  all  of  the  conditions  formulated  above,  except  for  the 
free  surface  condition.  Later  on  we  shall  begin  our  iteration  process 
with  a  function  o^  which  has  the  period  2n/A  in  <p;  F(q>),  all  sub- 
sequent iterates,  and  the  solution  itself,  will  have  the  same  period. 
Since  we  expect  the  waves  to  be  symmetrical  about  a  crest  or  trough 


530  WATER   WAVES 

(indeed,  our  existence  proof  will  yield  only  waves  with  this  property), 
we  suppose  that  the  origin  is  taken  at  such  a  point,  and  hence  that  at 
any  stage  of  the  iteration  process  6(<p,  0)  =  6(<p)  would  be  an  odd  func- 
tion of  9?,  while  r(<p,  0)  —  r((p)  would  be  an  even  function  of  9?,  and 
both  would  have  the  period  2n/L  Thus  F(q>)  in  (12.2.14)  as  defined 
by  (12.2.12)  should  be  taken  for  our  purposes  as  an  odd  function  of  y 
with  the  real  period  2n/L 

If  we  were  to  work  at  the  outset  with  Fourier  series,  it  follows  that 
6  would  be  represented  as  a  sine  series,  and  F(<p)  also.  However,  we 
wish  later  on  to  carry  out  an  iteration  process  in  which  only  contin- 
uous, and  not  necessarily  differentiable,  functions  of  q>  are  employed, 
and  in  which  the  existence  of  a  certain  continuous  periodic  function 
ft)  (9?)  of  period  2;rc/A  is  first  proved;  this  function  will  furnish  the 
boundary  values  of  the  solution  a)(q>9  ip).  (Afterwards,  the  question  of 
the  existence  of  the  normal  derivative  6^(99,  0)  in  (12.2.14),  and  of 
other  derivatives,  will  be  dealt  with  separately.  )  In  doing  so,  we  shall 
have  occasion  to  approximate  such  continuous  periodic  functions  by 
finite  Fourier  series,  or  Fourier  polynomials,  a  process  justified  by  the 
Weierstrass  approximation  theorem  which  states  that  such  a  poly- 
nomial can  always  be  constructed  to  yield  a  uniform  approximation 
for  all  values  of  <p  and  any  arbitrary  degree  of  approximation. 

Suppose,  therefore,  that  F(<p)  in  (12.2.14)  had  been  approximated 
at  some  stage  in  the  iteration  procedure  by  a  function  g((p]  in  the 
form  of  the  following  finite  sine  series: 


(12.2.15)  g(<p)  =  J  bv  sin 

v=l 

and  we  seek  the  bounded  harmonic  function  6(99,  y)  which  satisfies 
the  boundary  condition  (12.2.14)  with  F  =  g.  For  this  purpose  we 
write  the  solution  6(<p,  y>)  =  Ste  00(99,  \p)  also  as  a  finite  Fourier  sum: 

(12.2.16)  0(q>9  \p)  =  J  avev**  sin  vhp. 

v=l 

Insertion  of  this  sum  in  (12.2.14),  with  F  =  g,  leads  to  the  following 
equations  for  the  determination  of  the  coefficients  av: 


v        v,  ..... 

7>-l)a,  =  6, 

and  thus  to  the  conditions 


LEVI-CIVITA'S  THEORY  531 


[  at  arbitrary,  fcj  —  0 


(12.2.17)  a    =  fc 


It  is  very  important  for  the  following  to  observe  that  the  Fourier  sine 
series  for  F(<p)  must  lack  the  first  order  term:  otherwise,  as  we  have 
remarked  above,  our  problem  would  have  no  solution:  a  term  of  the 
form  &!  sin  T^p  in  F(q>)  is  a  "resonance"  term,  the  presence  of  which 
would  preclude  the  existence  of  the  solution  of  the  nonhomogeneous 
problem.  The  unique  solution  for  0(<p,  y)  is 

n  ft 

(12.2.18)  0(<p.  \p)  =  ^ -—  e^  sin  vty  +  «i^sin  A<p, 

v=2^(v  ~   1) 

once  ax  is  prescribed  (cf.  Chapter  3)  and  0  is  assumed  to  be  an  odd 
function  of  (p.  The  harmonic  conjugate  r(cp9  y>)  of  0(<p,  \p)  is  obtained  by 
integrating  the  Cauchy-Riemann  equation  6V  —  —  r^,  with  the  result 

n  b 

(12.2.19)  T(Q?,  w)  =  V  -  -  ~ ev^  cos  vhy  +  a^  cos  Ay. 

vf2A(v  —  1) 

(A  possible  additive  integration  constant  is  set  equal  to  /ero  since 
r  ->  0  when  ip  ->  —  oo.)  Thus  we  would  have  for  co(y,  \p)  under  the 
assumed  circumstances  the  expression 

(12.2.20)  co(o>,  V)  =  *  2     ~~V e^-^ 

£2X(v  -  1) 

In  other  words,  if  F(<p)  is  given  as  in  (12.2.15)  by  a  finite  Fourier 
series  of  sines  which  lacks  its  lowest  order  term,  i.e.  is  such  that 

(12.2.21 )  I* 2n/*F((p )  sin  Xydy  =  0, 

j  o 

then,  as  we  see  from  (12.2.20)  and  the  discussion  preceding  it,  the 
function  a)(<p)  —  co((p,  0)  —  6((p)  +  ii((p)  given  by 

(12.2.22)  a>(<p)  -  i  2  -A  --  e-<"**  +  ia^"" 

yields  the  boundary  values  of  an  analytic  function  co(<p,  ip)  which 
would  satisfy  the  boundary  condition  (12.2.14).  Evidently,  o)  would 
also  satisfy  all  of  the  conditions  a)  to  e)  formulated  above,  if  the 
amplitude  ax  of  the  first  order  term  of  the  Fourier  series  is  chosen 


582  WATER   WAVES 

small  enough,  except  that  the  boundary  condition  b)  is  replaced  by 
a  linear  condition. 

It  is  clear  that  the  insertion  of  a  function  a)(<p)  as  given  by  (12.2.22) 
in  (12.2.12)  to  determine  a  new  function  F(cp)  in  order  to  continue  the 
iteration  process  would  not  yield  in  general  a  function  representable 
as  a  finite  Fourier  sum,  but  rather  to  one  representable  only  by  a 
Fourier  series.  However,  we  have  already  stated  that  we  wish  to 
carry  out  our  iteration  scheme  in  the  nonlinear  problem  within  the 
class  of  continuous  functions,  which  need  not  possess  convergent 
Fourier  series.  Nevertheless,  the  general  scheme  outlined  above  for 
determining  the  successive  iterations  can  still  be  used  once  it  has  been 
extended  in  an  appropriate  way  to  the  wider  class  of  functions.  For 
this  purpose,  and  later  purposes  as  well,  it  is  convenient  to  introduce 
the  terminology  of  functional  analysis.  Thus  we  speak  of  the  linear 
vector  space  of  elements  which  are  complex-valued  functions 
g(q>)  —  <x.((p)  +  ifi(<p)9  continuous  for  all  <p  and  of  period  2rc/A,  such  that 
a  is  an  odd  function  and  ft  an  even  function  of  <p.  The  scalars  are  the 
real  numbers.  This  space  is  made  into  a  normed  linear  space  by  intro- 
ducing as  the  distance  from  the  origin  to  the  "point"  g  the  following 
norm,  written  ||  g  ||: 

||  g  ||  =  ||  a  +  ift  ||  -  max  A/a2  +  ]82  -  max  |  g  |, 

9 

and  as  the  distance  between  two  elements  or  points  gl5  g2,  the  norm  of 
their  difference,  i.e.  ||  gl  —  g2  ||.  This  space,  which  we  shall  call  the 
space  JB,  is  complete,  i.e.  it  has  the  property  that  every  Cauchy  se- 
quence in  the  space  converges  to  an  element  in  the  space.*  By  a  Cau- 
chy sequence  gn  we  mean  a  sequence  such  that  ||  gm  —  gn  II  -*  0 
when  m,  n  ->  oo.  Since  the  norm  is  the  maximum  of  the  absolute 
value  of  g,  it  follows  that  a  sequence  gn  which  is  a  Cauchy  sequence  is 
uniformly  convergent  and  hence  has  a  continuous  function  as  a  limit. 
We  remark  also  that  the  notion  of  distance  thus  introduced  in  our 
space  has  the  usual  properties  required  for  the  distance  function  in  a 
metric  space,  i.e.,  the  distance  is  positive  definite: 

||  g  ||  ^  0,  and  ||  g  ||  =  0     implies  g  =  0, 
and  the  triangle  inequality 

ll*i+ foil  ^11  gill  +11  foil 
holds. 

*  We  remark  that  a  complete  linear  normed  space  is  called  a  Banach  space ; 
however,  such  properties  of  these  spaces  as  are  needed  will  be  developed  here. 


LEVI-CIVITA'S  THEORY  533 

We  introduce  next  the  subspaee  Bl  of  our  Banach  space  B  which 
consists  of  all  real  functions  g(<p)  given  by  finite  Fourier  sums  of  sines 
lacking  the  term  of  first  order: 

(12.2.23)  g(<p)  =  J  6vsinvA<p,         bv  real. 

v=2 

With  respect  to  this  set  Bl  of  functions  we  define  a  transformation 
S  as  follows: 

(12.2.24)  Sg(<p)  =  i  J  -  -  -  #**  e-i9**,      -  oo  <  \p  ^  0. 
Since  ip  ^  0  we  have  for  the  norm*  of  Sg  the  bound 


From  Cauchy's  inequality  the  following  inequality  for  ||  Sg  \\  is  then 
obtained  : 


the  last  step  resulting  from  (12.2.23)  because  of  the  orthogonality 
of  the  functions  sin  vA,q>.  Since 


it  follows  that  a  constant  K  exists  which  is  independent  of  g  and  n 
(though  not  of  A),  such  that 

(12.2.25)  \\Sg\\  <K  \\&\\,         torgCB,. 

Thus  S  is  what  is  called  a  bounded  transformation  in  Bl  since  it 
transforms  each  element  of  Bl  into  an  element  of  B  with  a  norm  bound- 
ed by  a  constant  times  the  norm  of  the  original  element.  Clearly,  S 
transforms  a  certain  class  of  boundary  data  given  in  terms  of  the  real 
function  g(<p)  into  an  analytic  function  defined  in  the  lower  half  plane. 
We  proceed  next  to  extend  the  domain  of  definition  of  the  trans- 
formation S  in  such  a  way  that  it  applies  to  a  certain  set  of  real 

*  By  the  norm  of  a  function  of  two  variables  we  mean  the  least  upper  bound  of 
its  absolute  value. 


584  WATER   WAVES 

functions  in  B  which  contains  the  set  B19  i.e.,  to  the  set  JB2  of  continu- 
ous real  functions  g  in  B  with  vanishing  first  Fourier  coefficients,  that 

is,  to  functions  such  that   j  *   g((p)  sin  h<pdy  =  0;  this  subspace  B2  is 

j  o 

also  complete,  with  the  same  norm.  The  extension  of  the  definition  of 
S  is  made  in  the  following  rather  natural  way:  Take  any  function  g 
in  B2  and  let  gn  be  a  sequence  of  functions  in  B^  (i.e.,  a  set  of  real  tri- 
gonometric polynomials  lacking  first  order  terms)  which  approximate 
g  uniformly.  That  such  a  sequence  exists  is  known  from  the  Weier- 
strass  approximation  theorem.  We  then  form  the  sequence  Sgn—  which 
is  possible  since  S  is  applicable  to  these  functions—  and  observe  that 


because  S  is  obviously  a  linear  transformation,  and  hence 

\\Sgm-Sgn\\  ^K\\gm~  gj| 

from  (12.2.25)  since  gm  —  gn  is  an  clement  of  J5X.  Thus  the  sequence 
Sgn  is  a  Cauchy  sequence,  for  1  1  gm  —  gn  \  \  ->  0  because  the  functions 
gn  are  assumed  to  furnish  uniform  approximations  to  g;  hence  the  se- 
quence Sgn  has  a  unique  continuous  limit  function  which  we  define 
as  Sg.  The  transformation  S  thus  extended  will  be  referred  to  by  the 
symbol  S.  S  is  easily  seen  to  be  a  linear  transformation  and  the 
inequality  (12.2.25)  holds  for  it  with  the  same  value  of  K  since  it 
holds  for  all  the  functions  gn,  independent  of  n. 

Once  the  definition  of  the  transformation  S  is  extended  so  that  it 
applies  to  functions  in  B2,  it  becomes  possible  to  widen  the  class  of 
functions  within  which  a  (generalized  )  solution  of  our  linear  problem 
can  be  sought,  and  at  the  same  time  to  reformulate  the  boundary 
problem  in  terms  of  the  the  following  functional  equation: 

(12.2.26)  o>(0>)  =  Sg(<p)  +  ia^r***, 

in  which  Sg(<p)  refers  to  the  above  extension  of  Sg((p),  with  g(q>)  C  1?2, 
evaluated  on  the  boundary  \p  —  0.  By  virtue  of  (12.2.20)  and  the 
definition  of  Sg(y>)  for  g(q>)  in  J52,  one  might  expect  that  <*>(<p)  would 
yield  correct  boundary  values  for  the  solution  a>(q>9  y).  We  proceed 
to  show  that  this  is  indeed  the  case,  i.e.  that  any  continuous  function 
<*>($)  which  is  given  by  (12.2.26)  for  gCB2  furnishes  the  boundary 
values  of  a  function  c*)((p,y))  defined  and  continuous  for  ip  ^  0  which  is 
analytic  in  the  lower  half  plane,  has  a  real  part  Q(<p,y)  with  a  continuous 
normal  derivative  Qv  in  the  closed  half  plane,  and  such  that  the  boundary 
condition  (12.2.14)  with  F((p)  ==  g(<p)  is  satisfied. 


LEVI-CIVITA'S  THEORY  535 

The  regularity  properties  of  the  function  co((p9  y)  on  the  boundary 
y  —  0  come  about  because  of  certain  smoothing  properties  of  the 
transformation  S,  which  we  proceed  to  discuss.  Consider  first  the 
special  case  in  which  g(<p)  is  given,  as  in  (12.2.23),  by  a  finite  Fourier 
sum.  The  function  a(<p)  +  $(<?)  =  Sg: 


Z  _L.  .-*       ?„        ^  b*  sin  yfo    .   v  v  6*  cos 

a  +  ?p  =  o£  =    7   --  ~r  ^  7 

^  p       8 


has  in  this  case  the  following  property: 

-  &  -  g  +  A5. 

The  validity  of  this  formula  for  *S  follows  from  the  fact  that  Sg(<p)  — 
a(9?>  V)  +  ^(9?,  y;)  as  given  by  (12.2.24)  is  an  analytic  function  in  the 
closed  half  plane  y  ^  0,  and  that  Sg  satisfies  the  nonhomogeneous 
boundary  condition  (12.2.14)  with  F  =  g,  when  6(<p,  \p)  is  identified 
with  a(<p,  \p).  This  implies,  because  of  (12.2.26)  and  the  triangle  in- 
equality, the  inequality 

||  £,11  ^AMIgH,     #!  =  constant, 

as  one  easily  sees.  If  g  is  any  function  in  J32,  it  now  can  be  proved  that 
Sg—  defined  for  functions  g  in  B2  in  the  manner  described  above—  is 
such  that  its  imaginary  part  /9  has  a  continuous  derivative  with  respect 
to  (p.  This  is  done  by  approximating  g  uniformly  by  a  sequence  gn  of 
finite  Fourier  sums  in  B^  The  corresponding  derivatives  f$ntp  form  a 
Cauchy  sequence  because  of  the  above  inequality  and  hence  would 
converge  to  a  continuous  function.  The  relation  —  ^  —  g+Aoc  also 
would  hold  in  the  limit  for  the  derivative  /^;  thus  ^  is  again  seen  to  be 
continuous.  It  follows,  therefore,  that  a  continuous  function  co((p)  = 
0(<p)+&(v)  giyen  by  (12.2.26)  has  the  property  that  r((p)  has  a  conti- 
nuous derivative,  and  in  addition  —^(9?)  =  g(q>)+AO(q>)-  We  observe 
next  that  6+ir  furnishes  the  boundary  values  of  an  analytic  function 
00(9?,  y>)  defined  for  —  oo  <  \p  <  0  and  continuous  in  the  closed  half 
plane:  this  follows  again  by  approximating  g(<p)  by  functions  gn(<p)  in 
J?!,  as  in  (12.2.23),  defining  the  corresponding  o>n(<p,  y)  by(12.2.20), 
and  making  the  passage  to  the  limit  to  obtain  a>(<p,\p  )=0(<p,y)  )  +ir((p,ip  ); 
that  the  functions  con((p,  ip)  converge  to  a  continuous  function  for 
y>  ^  0  we  know,  and  that  the  limit  is  analytic  at  interior  points  follows 
since  it  is  the  uniform  limit  of  analytic  functions.  Since  —  r^((p,  ip)  = 
Ov(<p,  y))fory)  <  0,  and  since  r^  is  itself  a  harmonic  function  with  con- 


586  WATER  WAVES 

tinuous  boundary  values  r^  it  follows  that  rv(<p9  y>)  ->  r^  as  y  ->  0, 
and  hence  that  0V  is  also  continuous  for  \p  =  0,  i.e.   6v((jp9  0)  =  0V; 
hence  the  condition  —  rv  =  g(<p)  +  A0,  which  we  have  proved  above 
to  hold  becomes  0V  —  A0  =  g(g?),  and  this  is  our  boundary  condition. 
We  have  therefore  shown  that  a  continuous  function  a)(<f>)  which  is 
given  by  (12.2.26),  with  g(^)  any  function  in  B2,  furnishes  the  boun- 
dary values  of  an  analytic  function  co(^,  y)  =  6  +  it  in  \p  <  0  which 
is  continuous  for  y  ^  0,  whose  real  part  0(^,  ip)  has  a  continuous  deri- 
vative 0V  in  the  closed  lower  half  plane  with  0V(^,  0)  —  A0(<£,  0)  =  g(^) 
or,  as  we  also  write  it:  0V  —  A0  =  g.  In  the  subsection  immediately 
following  we  shall  establish  the  existence  of  a  continuous  solution 
o}(<f>)  of  (12.2.26)  when  g(^)  is  not  given  a  priori,  but  depends  in  a  non- 
linear   way    on    c5(<£)  =  0(g)  +  ir(<f>),    i.e.    when    g(<£)  =  F(<f>)  = 
A(tf-3*sin0  -  0)  +  «?-3*sin  0  (cf.  (12.2.12)).  Assuming  this  to  have 
been  proved,  we  proceed  to  draw  at  once  further  conclusions  regarding 
the  properties  of  oj(<f>)  and  its  continuation  a)(<f>9  y>)  as  an  analytic 
function  in  the  lower  half  plane  ip  ^  0.  We  show,  in  fact,  that  the 
solution  a>(<f>)  in  B  of  our  nonlinear  functional  equation  will  not  only 
furnish  the  boundary  values  of  an  analytic  function  a)(<f>9  ip  )  in  \p  <  0, 
with  co  as  boundary  values,  but  that  a)  has  continuous  first  derivatives 
in  the  closed  half  plane  ip  ^  0,  and  is  as  a  consequence  then  seen 
actually  to  be  analytic  for  \p  —  0.  Thus,  in  particular,  cfl(^)  would 
possess  a  convergent  Fourier  series.  Consider  the  analytic  function 
F(x)  defined  in  the  lower  half  plane  y  <  0  with  boundary  values 
<%e  fi(%)  =  g(y)  for  \p  =  0  and  with  &(%)  bounded  at  oo.  The  boun- 
dary condition  0V  —  W  —  —  r^  —  A0  =  g(^)  satisfied  by  our  solu- 
tion a)((p<y>)  is  also  extended  analytically  into  the  lower  half  plane 
if  <  0  by  means  of  the   relation   3le(a)^  —  Aco)  =  3te  &(%)   in  the 
manner  used  frequently  in  Chapters  3  and  4;  hence  we  have 

ia>   —  Aco  = 


since  the  imaginary  parts  of  cox  and  co  both  vanish  for  y  =  —  oo.  We 
have  just  seen  above  that  co  has  an  imaginary  part  r  with  a  continuous 
derivative  TV  on  the  boundary  y  =  0.  The  fact  that  r^  is  continuous 
then  makes  it  possible  to  show  that  co((p,  y)  is  Holder  continuous  for 
y>  =  0.  This  follows,  in  fact,  from  a  classical  theorem  of  Privaloff  (Bull. 
Soc.  Math.  France,  Vol.  44,  1916)  which  states  that  a  function  which  is 
defined  and  continuous  in  the  unit  circle,  analytic  in  the  interior  of 
the  circle,  and  has  an  imaginary  part  which  is  Holder  continuous  on 


LEVI-CIVITA'S  THEORY  537 

the  boundary  of  the  circle,  is  itself  Holder  continuous  in  the  closed 
unit  circle:  in  other  words,  Holder  continuity  of  the  imaginary  part 
brings  with  it  the  Holder  continuity  of  the  real  part  of  the  function. 
This  theorem  is  made  applicable  in  the  present  case  by  mapping  one 
of  the  period  strips  of  the  solution  a)(%)  in  the  #-plane  conformally  on 
the  unit  circle  of  a  £-plane,  say:  we  know,  in  fact,  that  CD  has  the  real 
period  2n/L  The  part  of  the  boundary  of  the  strip  given  by  y)  =  0 
(i.e.  a  full  period  interval  on  the  boundary)  is  mapped  on  the  boundary 
of  the  unit  circle.  (Since  |  co  |  ->  0  as  \p  -+  —  oo,  the  infinity  of  the 
strip  is  mapped  on  the  center  of  the  circle.)  Thus  co  —  0  +  iv  has  an 
imaginary  part  with  a  continuous  derivative  rv  on  y>  =  0,  and  it 
follows  that  r  is  certainly  Holder  continuous  for  \p  =  0.  Consequently 
the  real  part  of  co(£),  hence  co(£)  itself,  is  Holder  continuous  in  the 
closed  unit  circle,  since  this  property  is  not  destroyed  by  the  conformal 
mapping.  The  real  part  of  F(%)  on  the  boundary  \p  =  0,  which  is  given 
by  g(<p),  is  now  seen  to  be  Holder  continuous,  simply  because  of  the 
way  g(<f>)  is  given  in  terms  of  0  and  r,  and  the  Holder  continuity  of 
the  latter  functions.  A  second  application  of  Privaloff 's  theorem,  this 
time  to  F(%),  then  leads  to  the  Holder  continuity  of  P(%)  for  \p  ^  0. 
The  relation  icox— Aco  =  %*(%)  thus  holds  for  y>  =  0  and  it  shows  that 
cox  is  continuous  for  \p  =  0,  since  both  co  and  F  have  this  property.  In 
other  words  co^  and  cov  are  both  continuous  for  \p  =  0.  Finally,  once 
co(%)is  shown  to  have  a  continuous  derivative  with  respect  to  %  on  the 
boundary,  we  could  make  use  of  a  theorem  of  H.  Lewy  [L.9]  to  show 
that  co(%)  is  actually  analytic  on  the  boundary. 

12.2d.  The  solution  of  the  nonlinear  boundary  value  problem 

The  nonlinear  problem  to  be  solved  here  differs  from  the  linear 
problems  discussed  above  because  of  the  fact  that  the  function  g(<p)9 
the  nonhomogeneous  term  in  the  free  surface  condition,  is  not  given 
a  priori,  but  rather  becomes  known  only  when  the  solution  co(q>,  \p) 
itself  is  determined.  On  the  other  hand,  we  have  seen  that  the  equa- 
tion (12.2.26)  furnishes  the  boundary  values  <o(<p)  for  co(<p,  y),  in  case 
g((jp)  is  a  known  function  in  the  space  B2.  To  solve  the  nonlinear  pro- 
blem we  now  reverse  this  process:  we  regard  the  equation  (12.2.26)  as 
a  functional  equation  for  the  determination  of  the  function  6>(<p)  when 
g(y>)  is  identified  with  the  function  F(<p)  in  equation  (12.2.12),  i.e. 
when  g(<p)  itself  depends  in  an  explicitly  given  way  on  co(<p).  The  dis- 
cussion of  the  preceding  subsection  shows  that  we  have  to  prove 


538  WATER  WAVES 

only  that  the  functional  equation  has  a  solution  o>  (y)  in  the  Banach 
space  B. 

The  existence  of  the  solution  6>((p)  of  the  functional  equation  will 
be  carried  out,  as  we  have  stated  earlier,  by  an  iteration  process 
applied  to  the  functional  equation.  To  this  end  it  is  convenient  to 
introduce  a  nonlinear  transformation  R  defined  for  any  function 
g  =  a  +  ifi  in  the  whole  space  B  by  means  of  the  relation 

(12.2.27)  Rg(<p)  =  X(e-M  sin  a  -  a)  +  ee~^  sin  a. 

In  order  that  the  transformation  S  defined  in  the  preceding  subsection 
should  be  applicable  to  Rg(<p)  we  require,  as  part  of  the  definition  of 
R,  that  e  should  be  so  determined  that  Rg(<p)  lies  in  B2  and  thus  lacks 

its  first  Fourier  coefficient,  i.e.  such  that  J  "  Rg((p)  sin  h<pd(p  =  0; 
this  leads  to  the  following  condition  on  e: 

I  "    A(£~3^  sin  a  —  a)  sin  hpdcp 

(12.2.28)  *  =  - 


e-w  sin  a  sin  /.<pd<p 

This  implies  that  R  is  defined  only  if  the  denominator  of  (12.2.28) 
does  not  vanish.  Clearly,  this  nonlinear  transformation  yields  always 
real  odd  functions. 

Consider  now  the  functional  equation 

(12.2.29)  <b(<p)  =  SRco((p)  +  iaf**> 

in  which  Sg((p)  refers  to  the  extension  of  Sg(y>)9  as  defined  in  the  pre- 
ceding subsection  for  functions  g(<p)  in  the  space  I?2,  on  the  boundary 
y  =  0  and  ax  is  a  given  real  constant.  Because  of  (12.2.12),  (12.2.14), 
and  the  discussion  of  the  preceding  paragraph,  it  follows  that  a  solu- 
tion co(%)  of  the  nonlinear  boundary  value  problem  will  be  established 
once  a  function  c5(<p)  in  B  is  found  which  satisfies  the  functional  equa- 
tion (12.2.29). 

In  carrying  out  the  existence  proof  it  is  convenient  to  introduce  a 
few  new  notations.  The  function  r(<p)  is  introduced: 

(12.2.30)  r(<p)  =  °^    -  &-<*, 


and  a  new  transformation  T  on  r  is  defined  by 
(12.2.31)  Tr  =  1  SRfafr  +  t> 


LEVI-CIVITA'S  THEORY  539 

In  other  words,  T  is  applied  only  to  those  functions  r  such  that 
ai(r  +  ie****)  is  in  the  domain  of  definition  of  R.  The  functional 
equation  (12.2.29)  is  now  seen  to  be  equivalent  to  the  equation 

(12.2.32)  r  =  Tr, 

and  we  seek  a  solution  of  it  in  the  space  B. 

We  shall  solve  (12.2.32)  by  an  iteration  process  which  starts  with 
a  function  i\  in  B  such  that  the  corresponding  function  /?ft>1  is  in  Bz 
(i.e.  such  that  R  is  applicable  to  o^),  inserts  it  in  r2  =  Tr^  etc.,  thus 
obtaining  a  sequence  rk  with  rk  =  Trk_v  In  order  to  make  sure  that 
(12.2.28)  holds  for  the  solution  we  stipulate  that  the  parameter  e  in 
(12.2.12)  be  fixed  at  each  stage  of  the  iteration  process  so  that  (12.2.28) 
is  satisfied;  this  is  done  by  setting 

I  n   X( 


(12.2.33)  sk  = 


I  e~~k  sn  a    —  a     sn 


f2*M  Q*  .  -  , 

e~*p*  sin  <x.k  sin  Acpckp 


At  the  same  time,  this  ensures  that  the  transformation  T  is  really 
applicable  to  the  members  of  the  sequence  rk.  Of  course,  it  will  be 
necessary  to  show  that  the  denominators  in  the  equations  (12.2.33) 
are  bounded  away  from  zero  and  that  the  sequence  ek  converges.  The 
existence  of  a  sequence  rk  converging  to  a  solution  r  of  (12.2.32)  will 
be  shown  by  disposing  properly  of  the  arbitrary  constant  %  (the 
amplitude  of  the  wave  in  the  linearized  solution  of  the  problem), 
i.e.  by.  showing  that  «t  7^  0  can  be  chosen  small  enough  so  that  the 
sequences  rk  and  ek,  each  of  which  is  a  function  of  ar  converges.  We 
note  in  passing  that  if  d)k  =  <xfc  -f-  ijjk  is  small  of  order  %,  then  ek  as 
given  by  (12.2.33)  is  also  of  order  al— later  on,  we  give  an  explicit 
estimate  for  it— so  that  the  quantities  ek  should  not  turn  out  to  be  of 
the  wrong  order. 

The  convergence  of  the  sequence  of  iterates  rk  to  a  solution  r  of 
r  =  Tr  will  be  shown  by  proving  that  all  of  the  functions  rk,  for 
values  of  ax  less  than  a  certain  fixed  constant,  satisfy  the  following 
conditions:  for  some  real  positive  constant  r]  and  real  positive  «  <  1 

I)  \\r\\^r,  implies  \\Tr\\  ^r], 

II)  llrJMIr.H^       implies  ||  JVj -2V.H  £x  ||  1^-r,  ||, 

for  any  pair  of  functions  rv  r2.  Condition  I)  says  that  the  transforma- 
tion T  carries  any  function  in  the  closed  "sphere"  of  radius  77  into 


540  WATER   WAVES 

another  function  in  the  same  "sphere",  and  Condition  II)  is  a  Lip- 
schitz  condition. 

The  iteration  scheme  for  solving  the  functional  equation  Tr  =  r 
proceeds  in  the  following  standard  fashion.  Take  any  function  rQ(<p) 
in  B  with  ||  r0  ||  <  r\  to  which  T  is  applicable  and  consider  the  iterates 
rn  defined  by  rw  =  Trn_±.  From  I)  we  see  that  all  such  functions  r 
have  a  bounded  norm.  We  have,  evidently: 

rn+i  -rn  =  Trn  -  Trn_v 
Since  II)  holds  we  may  write 

II  fVu  -  rw  ||  =  ||  Trn  -  Trn^  ||  ^  *  \\  rn  -  rn_,  \\. 
and  hence 


We  consider  next  the  norm  of  rm  —  rn,     m  ^  n: 

II  rm  -  rn  ||  =  ||  (rm  -  r^)  +  (rm_^  -  rm_2)  +  .  .  .  +  (rn^  -  rn)  \\ 


-  r 


1  - 


(The  triangle  inequality  is  of  course  used  here.)  Since  x  <  1  it  is  clear 
that  the  sequence  rn  is  a  Cauchy  sequence  and  hence  it  converges  to 
a  unique  limit  function  r  in  B  with  norm  less  than  r).  (The  uniqueness 
statement  holds  of  course  only  for  functions  r  with  norm  less  than  rj. ) 
That  the  limit  function  r  satisfies  Tr  =  r  is  clear,  since  the  sequence 
rn  is  identical  with  the  sequence  Trn_±  and  hence  both  converge  to 
the  same  limit  r. 

In  order  to  establish  conditions  I)  and  II)  for  the  functions  in  the 
sequence  rk9  and  hence  to  complete  our  existence  proof,  it  is  conve- 
nient to  introduce  certain  continuous  functions  Fl(N)9  F2(N),  .  .  ., 
which  are  defined  for  real  N  2>  0,  bounded  near  N  =  0,  and  increasing 
with  N. 

Suppose  that  rl  C  B  is  such  that  ||  rl  \\  ^rj.  We  set  OJ1  —  Ot  +  irl 

and  (cf.  (12.2.30))  recall  that  rl  =    l  +  lTl  -  ie-*<p  so  that  ||  o>,  ||  ^ 

<*i 

I  ai  I  (1  +  1?)  =  N.  In  what  follows,  however,  we  omit  the  circumflex 
over  0  and  r,  and  we  shall  also  omit  the  subscript  on  a. 

The  following  inequalities  hold  when  |  a  \  is  sufficiently  small: 


LEVI-CIVITA'S  THEORY  541 


1.     |  |<r**i  sin   ,|  | 

(12284)      2'     ll*-3T'sin0i- 

8.     \\e-^smei-e-^sinem\\^\a\\\ 

4.     1  1  e~3T«  sin  0Z  -  e-3r»»  sin  dm  -  (Bl  -  6m)\  \ 

^\a\\\ri-rm\\NFt(N). 

00 

These  inequalities  are  all  based  on  the  fact  that  if  A(f  )  =  J  &nfn  is 

o 
an  absolutely  convergent  power  series  for  all  real  f  ,  then  1  f  |  ^  N 

00 

implies  |  A(|)  \  ^  ^\  hn\  Nn.  We  derive  the  second  and  third  of  the 

o 
above  inequalities  as  typical  cases—  the  others  are  derived  in  a  similar 

way.  Consider  the  second  inequality;  we  write 

||  e-*  sin  0  -  0  ||  =  ||  0(*-3T  -  1)  +  e-3T  (sin  0  -  0)  || 


.. 


with  F2(Ar)  =  3^3N  +  Ate4Ar.  Consider  next  the  third  inequality.  From 
the  mean  value  theorem  we  have 


--3e-T  snT^rm+^r  cos         ,-m. 

in  which  0*,  T*  are  some  values  on  the  segment  joining  (0j,  rt)  and 
(0m,  rm).  From  this  we  have 
||  *-*.  sin  0,  -e-**m  sin  0m  ||  ^  ||  0t-0m 


with  F3(A^)  =  4e3JV,  in  view  of  the  definition  of  rl9  rm  given  in  (12.2. 
30). 

It  is  also  essential  to  give  an  estimate  for  ek  in  (12.2.33).  First  we 
obtain  a  lower  bound  for  the  denominator.  We  have,  from  the  defini- 
tion of  r  and  the  second  inequality  above: 


r 


sin  0  sin 

/•2^1/A  /»2 

—  (e~^r  sin  0  —  0)  sin  Ag?  dgp  +         0  sin 

Jo  Jo 

-f          |  a  |  (sin  Arc  +  ^?*  ^)  sin 
Jo 


~  [|  a  |  (1  -  27?)  - 


542  WATER   WAVES 

Subscripts  have  been  dropped  in  the  above.  Since  N  =  \  a  \(l  +  q) 
it  is  clear  that  for  r/  ^  J  and  |  a  \  sufficiently  small,  say  |  a  \  ^  a(1), 
the  resulting  expression  is  greater  than  k  \  a  |,  with  k  a  positive  con- 
stant depending  on  a(1).  Use  of  this  fact  together  with  the  second  in- 
equality above  in  the  definition  (12.2.33)  of  e  leads  at  once  to  the 
inequality: 

5.  |.| 


a 


Thus  e  is  of  order  |  a  \  if  TJ  ^  J,  since  N  is  of  order  |  a  \.  Thus  the 
quantities  ek,  as  defined  by  (12.2.33),  are  in  fact  of  the  correct  order. 
In  the  same  fashion,  by  using  all  four  of  the  above  inequalities,  one 
obtains 

(12.2.34)      6.        \sl-£m\^\\rl-  rm 

a 

We  are  now  in  a  position  to  show  that  the  conditions  I)  and  II) 
hold  once  proper  choices  of  17  and  a  have  been  made.  We  suppose  that 
||  r  ||  ^  r\  ^  J  and  choose  a  such  that  0  <  |  a  \  ^  a(1);  any  value  x 
in  the  range  0  <  K  <  1  is  taken.  As  before,  the  norm  of  the  function 
a)  defined  by  r  satisfies  ||  o>  ||  ^  |  a  |(1  +  77)  =  N  ^  ||  a  \.  Our  next 
objective  is  to  give  an  estimate  for  Tr  as  defined  by  (12.2.31).  We 
have,  in  view  of  (12.2.31),  and  (12.2.27)  and  (12.2.25): 


|  a  I  I  a  I  |  a 

and  this  in  turn  yields: 

|  Tr     <  A 


\a\  \a\ 

with  K  a  fixed  constant,  upon  using  the  first,  second,  and  fifth  of  our 
inequalities,  together  with  the  fact  that  N  is  of  order  a.  Thus  if 
a(2)  ^  a(1)  is  a  positive  constant  such  that  Ka(2)F7(^a(2})  ^  r\  it 
follows  that 

||  Tr  ||  ^  r\  if  ||  r  \\  <^  r\  ^  J  and  |  a  \  ^  a(2). 

This  establishes  the  condition  I).  The  proof  that  II)  holds  is  carried 
out  in  much  the  same  way.  Suppose  that  rl9  r2  are  such  that  ||  rl  ||, 
II  ^2  II  ^il-  We  have,  upon  using  the  inequalities  1.  to  6.: 


LEVI-CIVITA'S  THEORY  543 


— 
a 


sin  ei-f9r,  sin  02-0! +02) 


\a\ 

+e1(tf~3Ti  sin  dl— <r3T2  sin  02)-f  e~*T2  sin  6^—e^) 

K  JV3 

I  al       X      2  4  3    5      Tal      l    8 


with  ^  a  fixed  positive  constant.  If  «(3)  ^  <z(2)  is  a  fixed  positive  con- 
stant such  that  3?a(a>Fe(£a<8))  ^  *,  then 


and  the  condition  II)  is  verified. 

It  follows  that  an  iteration  process  starting  with  an  arbitrary 
function  r0  in  B,  such  that  Ra)0  lies  in  B2,  with  \\rQ\\  ^rj  ^  ±  will 
converge  to  a  solution  r  of  Tr  =  r  if  0  <  |  a  \  ^  a(3).  The  function 
a>  —  a(r  +  te""*^)  is  then  a  solution  of  the  functional  equation 
(12.2.29)  which  lies  in  /?,  and  which  is  furthermore  not  the  "trivial" 
solution  a)  =  0  (which  always  exists),  since  ||  cb  \\  ^  a(l  —  ||  r  \\)  ^ 
fa  since  \\  r  \\  fg  J.  This  concludes  the  proof  for  the  existence  of  a 
continuous  solution  co(^)  of  the  nonlinear  functional  equation.  Once 
this  has  been  done  we  have  seen  at  the  end  of  the  preceding  sub- 
section that  d>(<f>)  is  actually  analytic  in  <£. 

It  is  also  clear  that  the  quantities  ek  assigned  to  each  cbk  and  rk 
exist,  and  that  they  converge  since  the  ek  form  a  Cauchy  sequence  in 
view  of  the  sixth  inequality  above  and  the  fact  that  1  1  rm  —  rn  \  \  ->  0. 
If  we  set  e  —  lim  ek,  it  is  clear  that  the  resulting  value  of  A'  obtained 
from  (12.2.13),  in  conjunction  with  the  arbitrarily  prescribed  value  of 
A,  yields  the  propagation  speed  U  through  (12.2.10)  as  a  function  of 
the  amplitude  parameter  a.  Since  co(#)  has  the  period  2yr/A,  it  follows 
from  the  discussion  at  the  beginning  of  this  section  that  the  motion 
in  the  physical  plane  has  the  period,  or  wave  length,  2jr/AC7. 


Bibliography 


Arthur,  R.  S. 

[A.I]      Revised  wave  forecasting  graphs  and  procedure.  Scripps  Institution  of 

Oceanography  of  the  University  of  California.  Wave  Report  78,  1947. 

i.A.2]      Variability  in  direction  of  wave  travel.  Annals  of  the  New  York  Academy 

of  Sciences,  Vol.  51,  Art.  3,  1949,  pp.  511-522. 

[A. 3]      Refraction  of  shallow  water  ....  Transactions  of  the  American  Geophysi- 
cal Union,  Vol.  31,  1950,  pp.  549-552. 
A.4]      See   [M.14]. 
A.5]      See   [P.18]. 
Atkinson,  F.  V. 

A. 6]      On  SommerfeWs  "radiation  condition."  London,  Dublin  and  Edinburgh 

Philosophical  Magazine,  Ser.  7,  Vol.  40,  1949,  pp.  641-651. 
Abdullah,  A.  J. 

rA.7]      Cyclogenesis  by  a  purely  mechanical  process.  Journal  of  Meteorology, 
Vol.  6,   1949,  pp.  86-97. 

Baird,  E.  G. 

[B.I]      See   [E.5]. 
Baker,  B.  B.,  and  E.  T.Copson 

[B.2]      The  Mathematical  Theory  of  Huygens'  Principle.  The  Clarendon  Press, 

Oxford,  1939. 
Bakhmeteff,  B.  A. 

[B.3]      Hydraulics  of  Open  Channels.  McGraw-Hill,  New  York,  1932. 
Barber,  N.  F.,  and  F.  Ursell 

[B.4]      The  generation  and  propagation  of  ocean  waves  and  swell.  1.   Wave 
periods  and  velocity.  Philosophical  Transactions  of  the  Royal  Society 
.    of  London,  240,  1948,  pp.  527-560. 
Batcman,  H. 

[B.5]      Partial  Differential  Equations  of  Mathematical  Physics.  The  University 

Press,  Cambridge  (Eng.),  1932.  Chapter  11,  Diffraction  problems. 
Bates,  C.  C. 

;  B.6]      Utilization  of  wave  forecasting  in  the  invasions  of  Normandy,  Burma,  and 
Japan.  Annals  of  the  New  York  Academy  of  Sciences,  Vol.  51,  Art.  3, 
1949. 
Beach  Erosion  Board 

[B.7]      Technical  Report  No.  2.  A  summary  of  the  theory  of  oscillatory  waves. 

U.  S.  Government  Printing  Office,  1942. 
Bernard,  P. 

[B.8]      Sur  certaines  propriet^s  de  la  houle  6tudi6es  a  Vaide  des  enregistrements 
stismographes.   Bulletin  Institut  Oc£anographique,   Monaco,  Vol.  38, 
1941. 
Biesel,  F. 

[B.9]      Le  filtre  a  houle.  La  Houille  Blanche,  Vol.  4,  1949,  pp.  373-375. 
[B.10]    Study  of  wave  propagation  in  water  of  gradually  varying  depth.  U.S. 
National  Bureau  of  Standards.  Gravity  Waves.  NBS  Circular  521, 1952. 
Birkhoff,  G. 

[B.I I]    Hydrodynamics.  Princeton  University  Press,  1950. 

545 


546  WATER   WAVES 

Birkhoff,  Garrett,  and  Jack  Kotik 

[B.12]    Fourier  analysis  of  wave  trains.  Proc.  of  the  NBS  Semicentennial 

Symposium  on  Gravity  Waves,  1951. 
Blue,  F.  L.f  and  J.  W.  Johnson 

[B.I 3]    Diffraction  of  waves  through  a  breakwater  gap.  Transactions  of  the 

American  Geophysical  Union,  Vol.  30,  1949,  pp.  705-718. 
Bondi,  H. 

[B.14]    On  the  problem  of  breakers.  Great  Britain,  Admiralty  Computing  Service, 

WA-2304-13,  1943. 
Bouasse,  H. 

[B.I 5]    Houle,  rides,  seiches  et  mare'es.  Librairie  Delagrave,  Paris,  1924,  pp. 

92-145. 
Boussinesq,  J. 

[B.I 6]  Thtorie  de  V intumescence  liquide  appeUe  onde  solitaire  ou  de  translation 
se  propageant  dans  un  canal  rectangulaire.  Institut  de  France,  Academic 
des  Sciences,  Comptes  Rendus,  .June  19,  1871,  p.  755. 

[B.17]    Essai  sur  la  the*orie  des  eaux  courantes.  Institut  de  France,  Academic 
des  Sciences,  Memoires  present6s  par  divers  savants,  Vol.  23,  1877. 
Broer,  L.  J.  F. 

[B.I 8]    On  the  propagation  of  energy  in  linear  conservative  waves.   Applied 

Scientific  Research,  Vol.  A  2,  1951,  pp.  447-468. 
Bruman,  J.  R. 

[B.19]    Application  of  the  water  channel — compressible  gas  analogy.  North  Ameri- 
can Aviation,  Inc.,  Report  NA-47-87,  1947. 
Bjerknes,  J.,  and  H.  Solberg 

[B.20]  Life  cycle  of  cyclones  and  the  polar  front  theory  of  atmospheric  circulation. 
Geofysiske  Publikationer  (Kristiania),  Vol.  3,  1922,  pp.  3-18. 

Carr,  J.  H.,  and  M.  E.  Stelzriede 

[C.I]      Diffraction  of  water  waves  by  breakwaters.   U.S.  National   Bureau  of 

Standards,  Gravity  Waves,  NBS  Circular  521,  1952. 
Chrystal,  G. 

[C.2]      An  investigation  of  the  seiches  of  the  Loch  Earn  by  the  Scottish  lake 
survey.  Transactions  of  the  Royal  Society  of  Edinburgh,  Vol.  45,  1906, 
pp.  361-396. 
Cooper,  R.  I.  B.,  and  M.  S.  Longuet-Higgins 

[C.3]      An  experimental  study  of  the  pressure  variations  in  standing  water  waves. 
Proceedings  of  the  Royal  Society  of  London,  Ser.  A,  Vol.  206,  1951, 
pp.  424-435. 
Copson,  E.  T. 

[C.4]      On  an  integral  equation  arising  in  the  theory  of  diffraction.  Quarterly 

Journal  of  Mathematics,  Oxford  Scries,  Vol.  17,  1946,  pp.  19-34. 
[C.5]       The  asymptotic  expansion  of  a  function  defined  by  a  definite  integral  or 
contour  integral.  British  Admiralty  Computing  Service,  Report  No. 
S.R.E./ACS  106,  1946. 
Also  see  [B.2] 
Cornish,  Vaughan 

[C.6]      Waves  of  the  Sea,  and  Other  Water  Waves.  T.  Fisher  Unwin,  London,  1910. 
[C.7]      Ocean  Waves  and  Kindred  Geophysical  Phenomena.  Cambridge  University 

Press,  London,  1934. 
Coulson,  C.  A. 

[C.8]      Waves.  Fourth  Edition,  Oliver  and  Boyd,  Edinburgh,  1947. 
Courant,  R.,  and  K.  O.  Friedrichs 

[C.9]  Supersonic  Flow  and  Shock  Waves.  Interscience  Publishers,  Inc.,  New 
York,  1948. 


BIBLIOGRAPHY  547 

Courant,  R.,  and  D.  Hilbert 

[C.10]    Methoden  der  Mathematischen  Physik.  2  Volumes,  Interscience  Publish- 
ers, Inc.,  New  York,  1944. 
Courant,  R.,  E.  Isaacson,  and  M.  Rees 

[C.ll]    On  the  solution  of  nonlinear  hyperbolic  differential  equations  by  finite 
differences.  Communications  on  Pure  and  Applied  Mathematics,  Vol.  5, 
1952,  pp.  243-255. 
Crossley,  H.  E.,  Jr. 

[C.I 2]    The  analogy  between  surface  shock  waves  in  a  liquid  and  shocks  in  com- 
pressible gases.    Experimental   study   of  hydraulic  jump   interactions. 
California  Institute  of  Technology  Hydrodynamics  Laboratory  Report 
No.  N-54,  1,  1949. 
Cray  a,  A. 

C.13]    Calcul  graphiques  des  regimes  variables  dans  les  canaux.  La  Houille 

Blanche,  N.S.  Annee  1,  1946. 

C.14]    The  criterion  for  the  possibility  of  roll  wave  formation.  U.S.  National 
Bureau  of  Standards,  Gravity  Waves,  NBS  Circular  521,  1952,  pp. 
141-151. 
Charney,  J.  G. 

I  C.I  5]    Dynamic  forecasting  by  numerical  process.  Compendium  of  Meteorolo- 
gy, American  Meteorological  Society  Boston,  1951,  pp.  470-482. 
Carson,  Rachel  L. 

[C.16]    The  Sea  Around  Us.  Oxford  University  Press,  New  York,  1951. 

Daily,  J.  W.,  and  S.  C.  Stephan  Jr. 

[D.I]      The   solitary   wave.    Massachusetts    Institute  of  Technology,  Hydro- 
dynamics Laboratory,  Technical  Report  No.  8,  1952. 
Dane!,  Pierre 

'D.2J      On  the  limiting  clapotis.  U.S.  National  Bureau  of  Standards,  Gravity 

Waves,  NBS  Circular  521,  1952,  pp.  35-38. 

[D.3]      Hydraulic  Research   Work.  Edited  by  La  Houille  Blanche. 
Darbyshire,  J. 

[D.4]     Identification  of  microseismic  activity  with  sea  waves.  Proceedings  of  the 

Royal  Society  of  London,  Vol.  202,  Ser.  A.,  1950,  pp.  439-448. 
Davies,  T.  V. 

[D.5]      Symmetrical,  finite  amplitude  gravity  waves.  U.S.  National  Bureau  of 

Standards,  Gravity  \Vaves,  NBS  Circular  521,  1952,  pp.  55-60. 
Deacon,  G.  E.  R. 

[D.6]      Recent  studies  of  waves  and  swell.  Annals  of  the  New  York  Academy  of 

Sciences,  Vol.  51,  Art.  3,  1949,  pp.  475-482. 
[D.7]      Analysis  of  sea  waves.  U.S.  National  Bureau   of  Standards,  Gravity 

Waves,  NBS  Circular  521,  1952,  pp.  209-214. 
Dean,  W.  R. 

[D.8]     Note  on  waves  on  the  surface  of  running  water.  Proceedings  of  the  Cam- 
bridge Philosophical  Society,  Vol.  43,  1947,  pp.  96-99. 
Deymie,  P. 

[D.9]      Propagation  d'une  intumesence  allonge'e  (probleme  aval).  International 
Congress  for  Applied  Mechanics,  5th  Proceedings,  September  1938, 
John  Wiley  and  Sons,  New  York,  1939. 
Dirichlet,  P.  L. 

[D.10]    Untersuchungen  uber  ein  Problem  der  Hydrodynamik.  Journal  fur  die 

reine  und  angewandte  Mathematik,  Vol.  58,  1861,  pp.  181-228. 
Dixon,  G.  F. 

[D.ll]    Theses  prteenttes  a  la  Faculty  des  Sciences  de  VUniversitt  de  Grenoble, 
1949. 


548  WATER    WAVES 

Dressier,  R.  F. 

[D.I 2]    Mathematical  solution  of  the  problem  of  roll-waves  in  inclined  open 

channels.  Communications  on  Pure  and  Applied  Mathematics,  Vol.  2, 

1949,  pp.  149-194. 
[D.I 3]   Stability  of  uniform  flow  and  roll-wave  formation.  U.S.  National  Bureau 

of  Standards,  Gravity  Waves,  NBS  Circular  521,  1952,  pp.  237-241. 
Dressier,  R.  F.,  and  F.  V.  Pohle 

[D.14]    Resistance  effects  on  hydraulic  instability.  Communications  on  Pure  and 

Applied  Mathematics,  Vol.  6,  1953,  pp.  93-96. 
Dubreuil-Jacotin,  M.  L. 

[D.15]    Sur  les  ondes  de  type  permanent  dans  les  liquides  he'te'rogenes.  Rendiconti 

della  Accadcmia  Nazionale  dei  Lincei,  Ser.  6,  Vol.  15, 1932,  pp.  814-819. 
|p.l5aj£wr   la   determination   rigoureuse   des   ondes  permanentes  pdriodiques 

d'ampleur  finie.  Journal  de  Math£matiques  Pures  et  Appliqu^es,  Ser.  9, 

Vol.  13,  1934,  pp.  217-291. 
Dean,  \V.  R. 

[D.I 6]    On  some  cases  of  the  reflection  of  surface  ivaves  by  an  infinite  plane 

barrier.  Proceedings  of  the  Cambridge  Philosophical  Society,  Vol.  42, 

1945,  pp.  24-28. 

[D.17]  On  the  reflection  of  surface  waves  by  a  submerged  plane  barrier.  Proceed- 
ings of  the  Cambridge  Philosophical  Society,  Vol.  41,  1945,  pp.  231-238. 

Eckart,  C. 

[E.I]      The  ray-particle  analogy.  Journal  of  Marine  Research,  Vol.  9,  1950, 

pp.  139-144. 

[E.2]  Surface  waves  on  water  of  variable  depth.  Scripps  Institution  of  Oceano- 
graphy, Lecture  Notes,  Fall  Semester  1950-51.  Ref.  51-12.  Wave 
Report  No.  100,  1951. 

[E.3]      The  propagation  of  gravity  waves  from  deep  to  shallow  water.  U.S. 
National  Bureau  of  Standards,  Gravity  Waves,  NBS  Circular  521, 1952, 
pp.  165-173. 
Egiazarov,  I. 

[E.4]      On  the  Daily  Regulation  of  Hdryoelectric  Stations  (Experimental  In- 
vestigations of  the  Negative  Wave).  Leningrad,  1931.  (In  Russian,  with 
an  English  summary.  Cited  in  Favre   (F.2J.) 
Einstein,  H.  A.,  and  E.  G.  Baird 

[E.5]      Progress  report  of  the  analogy  between  surface  shock  waves  on  liquids 
and  shocks  in  compressible  gases.  California  Institute  of  Technology, 
Hydrodynamics  Laboratory,  Report  No.  N-54,   1946. 
Emde,  F. 

[E.6J      See  [J.I]. 

Favre,  H. 

[F.I]      Contribution  a  Vttude  des  courants  liquides.  Publication  du  Laboratoire 
de  Recherches  hydrauliques  annex£  a  1'Ecole  Polytechnique  F6d6rale 
de  Zurich,  1933. 
[F.2]      Etude  thtorique  et  exptrimentale  des  ondes  de  translation  dans  les  canaux 

dtcouverts.  Dunod,  Paris,  1935. 
Finkelstein,  A. 

[F.3]      The  initial  value  problem  for  transient  water  waves.  Dissertation,  New 

York  University,  1953. 
Fjeldstad,  J.  E. 

[F.4]  Observations  of  the  internal  tidal  waves.  U.S.  National  Bureau  of  Stand- 
ards, Gravity  Waves,  NBS  Circular  521,  1952. 


BIBLIOGRAPHY  549 

Fleishman,  B.,  J.  J.  Stoker,  and  L.  Weliczker, 

[F.5]      Floating  breakwaters  in  shallow  water.  New  York  University,  Institute 

of  Mathematical  Sciences,  Report  No.  IMM-192,  1953. 
Forchheimer,  P. 

[F.6]      Hydraulik.  Encyklopadie  der  mathematischen  Wissenschaften,  Vol.  IV, 

Teil  8,  pp.  324-472,  B.  G.  Teubner,  Leipzig,  1901-1908. 
Forel,  F.  A. 

[F.7]      Le  Leman,  monographic  limnologique.  Bibliotheque  Nationale,  Vol.  2, 

F.  Rouge,  Lausanne,  1895. 
Frank,  P.,  and  R.  v.  Mises, 

[F.8]      Die  Differential-  und  Integralgleichungen  der  Mechanik  und  Physik. 

F.  Vieweg  und  Sohn,  Braunschweig,  1925-27. 
Freeman,  J.  C.,  Jr. 

[F.9]      An  analogy  between  the  equatorial  easterlies  and  supersonic  gas  flows. 

Journal  of  Meteorology,  Vol.  5,  1948,  pp.  138  —  146. 

[F.10]    The  solution  of  nonlinear  meteorological  problems  by  the  method  of  charac- 
teristics. Compendium  of  Meteorology,  American  Meteorological  Society, 
1951,  pp.  421-4.33. 
Friedrichs,  K.  O. 

[F.ll]    On  the  derivation  of  the  shallow  water  theory.  Appendix  to  The  formation 
of  breakers  and  bores  by  J.  J.  Stoker,  Communications  on  Pure  and 
Applied  Mathematics,  Vol.  1,  1948,  pp.  81-85. 
Friedrichs,  K.  O.,  and  H.  Lewy 

[F.I 2]    The  dock  problem.  Communications  on  Pure  and  Applied  Mathematics, 

Vol.  1,  1948,  pp.  135-148. 
Friedrichs,  K.  O.,  and  1).  II.  Ilycrs 

[F.I 3]    The  existence  of  solitary  waves.  Communications  on  Pure  and  Applied 

Mathematics,  Vol.  7,  1954,  pp.  517-550. 
Friedrichs,  K.  O. 

[F.14]    Water  waves  on  a  shallow  sloping  beach.  Communications  on  Pure  and 

Applied  Mathematics,  Vol.  1,  1948,  pp.  109-134. 
[F.15]    See  [C.9]. 
Fleming,  R.  H. 

[F.16]    Sec   [S.32a]. 

Goldstein,  E.,  and  J.  B.  Keller 

[G.I]      Reflection  of  waves  in  shall  wv  water.  (Unpublished.)  Sec  E.  Goldstein, 

Master's  Thesis,  New  York  University,   1950. 
Guthrie,  F. 

[G.2]      On  stationary  liquid  waves.  London,  Edinburgh  and  Dublin,  Philosophi- 
cal Magazine,  Ser.  4,  Vol.  50,  1875. 
Gwyther,  R.  F. 

[G.8J      The  classes  of  progressive  long  waves.  London,  Edinburgh  and  Dublin, 

Philosophical  Magazine,  Ser."  5,  Vol.  1,  1900.  pp.  213,  308,  349. 
Goldstein,  E. 

[G.4]      See  [K.8]. 
Gerber,  R. 

[G.5]      Sur  les  solutions  exactes  des  Equations  du  mouvement  avec  surface  libre 
d'un  liquide  pesant.  (Thesis,  Univ.  of  Grenoble,  1955.)  To  appear  in 
Journal  de  Mathematique. 
Goldner,  S.  R. 

[G.6]  Existence  and  uniqueness  theorems  for  two  systems  of  nonlinear  hyperbolic 
differential  equations  for  functions  of  two  independent  variables.  Doctor's 
thesis,  New  York  University,  1949. 


550  WATER   WAVES 

Hadamard,  J. 

[H.I]     Sur  les  ondefi  liquides.  Institut  de  France.  Academic  des  Sciences, 

Comptes  Rendus,  Vol.  150,  1910,  pp.  609-611,  772-774. 
Hamada,  T. 

[H.2]     Breakers   and   beach   erosion.    Report    of    Transportation   Technical 

Research  Institute  (Report  No.  1)  Tokyo,  1951. 
Harleman,  D.  R.  F. 

[H.8]     Studies  on  the  validity  of  the  hydraulic  analogy  to  supersonic  flow.  Parts  I 
and  II.  Massachusetts  Institute  of  Technology,  1950.  (Technical  report 
to  U.S.  Air  Force). 
Haskind,  M.  D. 

[H.4]      Waves  arising  from  oscillation  of  bodies  in  shallow  water.  Prikladnaya 

Matematika  i  Mechanika,  Vol.  10,  1940,  pp.  475-480. 

[H.5]      The  hydrodynamic  theory  of  the  oscillation  of  a  ship  in  waves  (in  Russian). 
Prikladnaya  Matematika  i  Mechanika,  Vol.  10,  1946,  pp.  39-66. 
Oscillation  of  a  ship  on  a  calm  sea  (in  Russian),  Izv.  Ak.  Nauk,  Ot.  Tex. 
Nauk,  No.  1,  1946,  pp.  23-34.  Translated  by  V.  Wehausen  as  "Two 
papers  on  the  hydrodynamic  theory  of  heaving  and  pitching  of  u  ship", 
Technical  Research  Bulletin  1-12,  Society  of  Naval  Architects  and 
Marine  Engineers,  New  York,  1953. 
Haurwitz,  B. 

[H.6]     Dynamic  meteorology.  McGraw-Hill,  New  York,  1941. 
Havelock,  T.  H. 

[H.7]      The  wave-making  resistance  of  ships.  Proceedings  of  the  Royal  Society 

of  London,  Ser.  A,  Vol.  82,  1909,  pp.  276-300. 
[H.8]      Wave  resistance  theory  and  its  application  to  ship  problems.  Society  of 

Naval  Architects  and  Marine  Engineers,  New  York,  1950. 
[H.9]     The  propagation  of  groups  of  waves  in  dispersive  media.  The  University 

Press,  Cambridge  (Eng.),  (Cambridge  Tracts,  No.  17),  1914. 
Hayashi,  T. 

[H.10]   Mathematical  theory  of  flood  waves.  Japan  National  Congress  for  Applied 

Mechanics,  First.  Proceedings.  1951. 
Heins,  A.  E. 

[H.ll]   Water  waves  over  a  channel  of  finite  depth  with  a  submerged  plane  barrier. 

Canadian  Journal  of  Mathematics,  Vol.  2,  1950,  pp.  210-222. 
[H.12]   Water  waves  over  a  channel  of  finite  depth  with  a  dock.  American  Journal 

of  Mathematics,  Vol.  70,  1948,  pp.  730-748. 
Henry,  Marc 

[H.13]  Note  sur  la  propagation  des  intumescences  dans  un  canal  rectangulaire. 
Communique  a  la  seance  du  5  nov.  1937  du  comit£  technique  de  la 
soci^te"  hydrotechnique  de  France. 
Hinze,  J.  O. 

[H.14]  Die  Erzeugung  von  Ringwellen  auf  einer  FlUssigkeitsoberfldche  durch 
periodisch  wirkende  Druckkrdfte.  Zeitschrift  fur  angewandte  Mathematik 
und  Mechanik,  Bd.  16,  1986. 
Hogner,  E. 

[H.15]  A  contribution  to  the  theory  of  ship  waves.  Arkiv  for  Matematik,  Astro- 

nomi,  och  Fysik,  Vol.  17,  1922. 
Hopf,  E. 

[H.16]  On  S.  Bernstein's  theorem  on  surfaces  z(x,  y)  of  nonpositive  curvature. 
Proceedings  of  the  American  Mathematical  Society,  Vol.   1,   1950, 
pp.  80-85. 
Hyers,  D.  H. 
See  [F.18] 


BIBLIOGRAPHY 


551 


Hanson,  E.  T. 

[H.17]  The  theory  of  ship  waves.  Proceedings  of  the  Royal  Society  of  London, 
Ser.  A,  Vol.  Ill,  1926,  pp.  491-529. 

Isaacson,  E. 

[I.I  ]  Water  waves  over  a  sloping  bottom.  Communications  on  Pure  and  Applied 
Mathematics,  Vol.  3,  1950,  pp.  1-32. 

[1.2]  Waves  against  an  overhanging  cliff.  Communications  on  Pure  and 
Applied  Mathematics,  Vol.  1,  1948,  pp.  201-209. 

[I.3J       See  [C.ll]. 
Isaacson,  E.,  J.  J.  Stoker,  and  B.  A.  Troesch 

[1.4]  Numerical  solution  of  flood  prediction  and  river  regulation  problems. 
Report  2.  Numerical  solution  of  flood  problems  in  simplified  models  of  the 
Ohio  River  and  the  junction  of  the  Ohio  and  Mississippi  Rivers.  New 
York  University,  Institute  of  Mathematical  Sciences,  Report  No. 
IMM-205,  1954. 

[I.4al  Numerical  solution  of  flood  prediction  and  river  regulation  problems. 
Report  3.  Results  of  the  numerical  prediction  of  the  1945  and  1948  floods 
in  the  Ohio  River,  of  the  1947  flood  through  the  junction  of  the  Ohio  and 
Mississippi  Rivers,  and  of  the  Jtoods  of  1950  and  1948  through  Kentucky 
Reservoir.  New  York  University,  Institute  of  Mathematical  Sciences, 
Report  No.  IMM-NYU-235,  1956. 
Isaacs,  J.  D. 

Ll.31       See   M.7J. 
Iverson,  H.  W. 

ij.6]       Laboratory  study  of  breakers.  U.S.  National  Bureau  of  Standards, 
Gravity  Waves',  NBS  Circular  521,  1952. 


Jahnkc,  K.,  and  F.  Kmde 

[J.1J       Tables  of  Functions.  Dover,  New  York,  1945,  p.  138. 
.Jeffreys,  Harold 

[J.2]       The  flow  of  water  in  an  inclined  channel  of  rectangular  section.  London, 
Edinburgh  and  Dublin,  Philosophical  Magazine,  Ser.  6,  Vol.  49,  1925, 
pp.  793-807. 
Jeffreys,  H.  and  Bertha  S.  Jeffreys 

!J.3j       Methods  of  Mathematical  Physics.   Cambridge  (Eng.),  The  University 

Press,  1948. 
John,  F. 

["J.4J       \Vavett  in  the  presence  of  an  inclined  barrier.  Communications  on  Pure 

and  Applied  Mathematics,  Vol.  1,  1948,  pp.  149-200. 

[J.5]      On  the  motion  of  floating  bodies  I,  II.  Communications  on  Pure  and 
Applied  Mathematics,  Vol.  2,  1949,  pp.  13-57;  Vol.  3,  1950,  pp.  45-100. 
Johnson,  J.  VV.  and  M.  P.  O'Brien 

[J.6]       The  engineering  aspects  of  waves  in  harbor  design.  International  Asso- 
ciation of  Hyd.  Str.  Research,  Grenoble,  1949. 
Johnson,  J.  VV.,  M.  P.  O'Brien,  and  J.  D.  Isaacs 

[J.7]       Graphical  construction  of  wave  refraction  diagrams.  Hydrographic  Office, 

Navy  Department,  Publication  no.  605,  1948. 
Johnson,  J.  W. 

[J.8]       See  (B.I 3]. 
Johnson.  M.  YV. 

[J.9]       See   [S.32a]. 


552  WATER   WAVES 

Kampe  de  Feriet,  J.,  and  J.  Kotik 

[K.I]     Surface  waves  of  finite  energy.  Journal  of  Rational   Mechanics  and 

Analysis,  Vol.  2,  1953,  pp.  577-685. 
von  Karman,  T. 

[K.2]     Eine  praktische  Anwendung  der  Analogic  zwischen  Ubcrschallstromung 
in  Gasen  und  uberkritischer  Stromung  in  offenen  Gerinnen.  Zentralblatt 
fiir  angewandte  Mathematik  und  Mechanik,  Vol.  18,  1939. 
Karp,  S.  N. 

[K.3]      Wiener-Hopf  techniques  and  mixed  boundary  value  problems.   Com- 
munications on  Pure  and  Applied  Mathematics,  Vol.  3,   1950,   pp. 
411-426. 
Keller,  J.  B.,  and  P.  Lax 

[K.4]     Finite  difference  schemes  for  hyperbolic  systems.  LAMS  1201,   1950. 
Keller,  J.  B. 

[K.5]     The  solitary  wave  and  periodic  waves  in  shallow  water.  Annals  of  the  New 

York  Academy  of  Sciences,  Vol.  51,  Art.  3.  1949,  pp.  345-350. 
[K.6]      The  solitary  wave  and  periodic  waves  in  shallow  water.  Communications 

on  Pure  and  Applied  Mathematics,  Vol.  1,  1948,  pp.  323-339. 
[K.7]     Scattering  of  water  waves  treated  by  the  variational  vncthod.  U.S.  National 

Bureau  of  Standards,  Gravity  Waves,  NBS  Circular  521,  1952. 
Also  see  [W.8] 
Keller,  J.  B.,  and  E.  Goldstein 

[K.8]      Water  wave  reflection  due  to  surface  tension  and  floating  ice.  Transact ioiks 

of  the  American  Geophysical  Union,  Vol.  34,  1953,  pp.  43-48. 
Keller,  J.  B.  and  M.  Weitz 

[K.9]      Reflection  and  transmission  coefficients  for  waves  entering  or  leaving  an 
icefield.  Communications  on  Pure  and  Applied  Mathematics,  Vol.  6, 
1953,  pp.  415-417. 
Kellog,  O.  D. 

[K.10]    Foundations  of  potential  theory.  J.  Springer,  Berlin,  1929. 
Kelvin,  Lord  (Sir  W.  Thomson) 

[K.ll]   On  the  waves  produced  by  a  single  impulse  in  water  of  any  depth,  or  in  a 
dispersive  medium.  Proceedings  of  the  Royal  Society  of  London,  Ser.  A. 
Vol.  42,  1887,  pp.  80-85. 
Keulegan,  G.  H. 

[K.12]    The   characteristics   of  internal  solitary  ivaves.  (Abstract  only.)  U.S. 
National  Bureau  of  Standards,  Gravity  Waves,  NBS  Circular  521 ,  1952. 
Keulegan,  G.  H.,  and  G.  W.  Patterson 

[K.13]   Mathematical  theory  of  irrotational  translation  waves.   U.S.   National 

Bureau  of  Standards  Journal  of  Research,  Vol.  24,  1940. 
Kotschin,  N.  J.,  I.  A.  Kibel,  and  N.  W.  Rose 

[K.14]   Theoretische  Hydrodynamik.  Akad.  Verlag,  Berlin,  1954. 
Korteweg,  D.  J.,  and  G.  de  Vries 

[K.I  5]   On  the  change  of  form  of  long  waves  advancing  in  a  rectangular  canal  and 
on  a  new  type  of  long  stationary  waves.  London,  Dublin  and  Edinburgh, 
Philosophical  Magazine,  Ser.  5,  Vol.  39,  1895,  p.  422. 
Korvin-Krukovsky,  B.  V.,  and  E.  V.  Lewis 

[K.I 6]   Ship  motions  in  regular  and  irregular  seas.  Stevens  Institute  of  Technol- 
ogy, Experimental  Towing  Tank,  Technical  Memorandum  no.  106, 1954. 
Kotik,  J. 

[K.I 7]   Existence  and  uniqueness  theorems  for  linear  water  waves.  Massachusetts 

Institute  of  Technology,  1952. 
[K.18]   See  [K.I],  [B.12]. 
Kreisel,  G. 

[K.19]   Surface  waves.  Quarterly  of  Applied  Mathematics,  Vol.  7, 1 949,  pp.  21-44. 


BIBLIOGRAPHY  553 

Krylov,  A.  N. 

[K.20]   A  general  theory  of  the  oscillations  of  a  ship  on  waves.  Transactions  of  the 

Institute  of  Naval  Architects  and  Engineers,  Vol.  37,  1896. 
Keldysh,  M.  V. 

[K.21]  Remarks  on  certain  motions  of  a  heavy  fluid.  (Russian)  Tckhnich. 
Zametki  Tsagi,  no.  52,  1935. 

Laitone,  E.  V. 

[L.I]      A  study  of  transonic  gas  dynamics  by  the  hydraulic  analogy.  University  of 
California,  Berkeley,  California,  Institute  of  Engineering  Research,  1951. 
Lamb,  H. 

[L.2]      On  deep  water  waves.  Proceedings  of  the  London  Mathematical  Society, 

Ser.  2,  Vol.  2,  1904,  p.  388. 
[L.3]      Hydrodynamics.    Dover   Publications,    New   York,    1945;    Cambridge 

University  Press,  1932. 
Lavrcnticff,  M. 

[L.4]      On  the  theory  of  long  waves.  Akad.  Nauk  Ukrain.  R.S.R.,  Zb.  Proc.  Just. 

Math.,  Vol.   1,  1948.  In   Ukrainian. 
Lax,  Anneli 

[L.5]      Decaying  shocks.  Communications  on  Pure  and  Applied  Mathematics, 

Vol.   1,   1948,  pp.  247-257. 
Lax,  Peter 

[L.6]      See   [K.4J. 
Levi-Civita,  T. 

[L.7j      Determination  rigoureuse  des  ondes  permanentes  d'ampleur  finie.  Mathe- 

matische  Annalen,  Vol.  93,  1925,  pp.  264-314. 
Lcwy,  H. 

^L.8]      Water  waves  on  sloping  beaches.  Bulletin  of  the  American  Mathematical 

Society,  Vol.  52,  1946. 

[L.9J      A  note  on  harmonic  functions  and  a  hydrodynamical  application.  Proceed- 
ings of  the  American  Mathematical  Society,  Vol.  3,  No.  1,  1952. 
[L.9al    On  steady  free  surface  flwv  in  a  gravity  field.  Communications  on  Pure 

and  Applied  Mathematics,  Vol.  5,  1952. 
[L.10]    See   [F.12], 
Lichteustein,  L. 

[L.ll]  Vorlesungen  uber  einige  Klassen  nichtlinearer  lutegralgleichungen  und 
Integro-Differentialgleichungen  nebst  Anwendungen.  J.  Springer,  Berlin, 
1931. 

[L.12]    Grundlagen  der  Hydromechanik.  J.  Springer,  Berlin,  1929. 
Longuet-Higgins,  M.  S. 

[L.I 8]    A  theory  of  the  generation  of  microseisms.  Philosophical  Transactions  of 

the  Royal  Society  of  London,  Ser.  A,  Vol.  243,  1950,  pp.  1-35. 
Longuet-Higgins,  M.  S.,  and  F.  Ursell 

[L.14J    Seawaves  and  microseisms.  Nature,  Vol.  162,  1948. 
Longuet-Higgins,  M.  S. 

[L.15]    See  [C.3]. 
Lowell,  S.  C. 

[L.16]    The  propagation  of  waves  in  shallow  water.  Communications  on  Pure  and 

Applied  Mathematics,  Vol.  2,  1949,  pp.  275-291. 
[L.17J    The  Propagation  of  Waves  in  Shallow  Water.  Ph.  D.  Thesis,  New  York 

University,  June  1949. 
Lunde,  J.  K.,  and  W.  C.  S.  Wigley 

[L.18]  Calculated  and  observed  wave  resistances  for  a  series  of  forms  of  fuller 
midsection.  Quarterly  Transactions  of  the  Institute  of  Naval  Architects, 
London,  Vol.  90,  1948. 


554  WATER   WAVES 

Lunde,  J.  K. 

[L.19]    On  the  linearized:  theory  of  wave  resistance  for  displacement  ships  in 
steady  and  accelerated  motion.  Society  of  Naval  Architects  and  Marine 
Engineers,  1951. 
Lewis,  E.  V. 

[L.20]    See  [K.16]. 

MacDonald,  H.  M. 

[M.I]     A  class  of  diffraction  problems.  Proceeding  of  the  London  Mathematical 

Society,  Ser.  2,  Vol.  14,  1915,  pp.  410-427. 
MacRobert,  T.  M. 

[M.2]     Functions  of  a  Complex  Variable.  MacMillan,  London,  1947. 
Martin,  J.  C.,  W.  J.  Moyce,  W.  G.  Penney,  A.  T.  Price,  and  C.  K.  Thornhill 
[M.3]     Some  gravity  wave  problems  in  the  motion  of  perfect  liquids.  Philosophical 
Transactions  of  the  Royal  Society  of  London,  Ser.  A,  Vol.  244,  1952. 
Mason,  Martin  A. 

[M.4]     Some  observations  of  breaking  waves.  U.S.  National  Bureau  of  Standards, 

Gravity  Waves,  N.  B.  S.  Circular  521,  1952. 
Massau,  J. 

[M.5]     Me'moire  sur  V integration  graphique  des  equations  aux  dcrive'es  partielles. 

Annales  des  ing6nieurs  sortis  de  Gand,  Tome  XII,  1889. 
McGowan,  J. 

[M.6]     On  the  solitary  wave.  London,  Dublin  and  Edinburgh,  Philosophical 
Magazine,  Ser.  5,  Vol.  32,  p.  45,  1891.  On  the  highest  wave  of  permanent 
type.  Ser.  5,  Vol.  38,  p.  351,  1894. 
McNown,  J.  S. 

[M.7]      Waves  and  seiches  in  idealized  ports.  U.S.  National  Bureau  of  Standards, 

Gravity  Waves,  N.  B.  S.  Circular  521,  1952. 
Miche,  A. 

[M.8]     Mouvements  ondulatoires  de  la  mer  en profondeur  constante  ou  dfaroissante. 
Annales  des  ponts  et  chaussees,  1944,  pp.  25-78,  131-164,  270-292, 
369-406. 
Michell,  J.  H. 

[M.9]      The  wave  resistance  of  a  ship.  London,  Dublin  and  Edinburgh,  Philoso- 
phical Magazine,  Ser.  5,  Vol.  45,  1898. 
Milne-Thompson,  L.  M. 

[M.10]    Theoretical  Hydrodynamics.  MacMillan,  London,  1938. 
Moyce,  W.  J. 

[M.ll]   See  [M.3J. 
Munk,  W.  H. 

[M.12]    The  solitary  wave  theory  and  its  application  to  surf  problems.  Annals  of 

the  New  York  Academy  of  Sciences,  Vol.  51,  Art.  3,  1949,  pp.  376-424. 

[M.I 3]   Supplement  to  breakers  and  surf,  principles  in  forecasting.   Scripps 

Institution  Oceanography.  Wave  Report  No.  49. 
Munk,  W.  H.,  and  R.  S.  Arthur 

[M.14]    Wave  intensity  along  a  refracted  ray.  U.  S.  National  Bureau  of  Standards, 

Gravity  Waves,  N.  B.  S.  Circular  521,  1952. 
Munk,  W.  H.,  and  M.  A.  Traylor 

[M.15]  Height  of  breakers  and  depth  of  breaking.  Scripps  Institution  of  Oceano- 
graphy, Wave  Report  No.  47,  1945. 
[M.I 6]   Refraction  of  ocean  waves:  a  process  linking  underwater  topography  to 

beach  erosion.  Journal  of  Geology,  Vol.  15,  1947. 
Munk,  W.  H. 

[M.17]    See  [S.33],   [S.84],   [S.35],  [S.36],   [S.37]. 


BIBLIOGRAPHY  555 

Nekrassov,  A.  I. 

[N.I  ]     On  steady  waves.  Izv.  Ivanovo- Voznesensk.  Politekhn.  In-ta,  No.  3,  1 921 . 
[N.la]    The  exact  theory  of  steady  waves  on  the  surface  of  a  heavy  fluid.  Izdat. 

Akad.  Nauk,  SSSR,  Moscow,  1951. 
Nirenberg,  L. 

[N.2]      Waves  of  finite  amplitude  created  by  an  obstacle  in  a  stream.  Unpublished 

manuscript  (New  York  University). 
Nakano,  N.,  S.  Unoki,  and  Y.  Kuga 

[N.3]     On  the  result  of  wave  observations  at  Jogashima  Island  and  its  application 
to  wave  forecasting.  Meteorological  Research  Institute,  Tokyo,  1953. 

O'Brien,  M.  P. 
[O.I]      See  [J.Q]. 

[O.2]      See   [J.7]. 

Patterson,  G.  W. 

[P.I]      See  [K.131. 
Penney,  W.  G. 

[P.2]      See   (M.3] 
Penney,  W.  G.  and  A.  T.  Price 

[P.3]      Diffraction  of  sea  waves  by  breakwaters.  Directorate  of  Miscellaneous 
Weapons  Development  History  No.  26-Artif  icial  Harbors,  Sec.  3D,  1 944. 
Peters,  A.  S. 

fP.4]      A  new  treatment  of  the  ship  wave  problem.  Communications  on  Pure  and 

Applied  Mathematics,  Vol.  2,  1949,  pp.  123-148. 
[P. 5]      The  effect  of  a  floating  mat  on  water  waves.  Communications  on  Pure 

and  Applied  Mathematics,  Vol.  3,  1950,  pp.  319-354. 

[P.6]      Water  waves  over  sloping  beaches  and  the  solution  of  a  mixed  boundary 
value  problem  for  A(p  —  k*<p  —  0  in  a  sector.  Communications  on  Pure 
and  Applied  Mathematics,  Vol.  5,  1952,  pp.  87-108. 
Peters,  A.  S.,  and  J.  J.  Stoker 

[P.7]      The  motion  of  a  ship,  as  a  floating  rigid  body,  in  a  seaway.  Institute  of 
Mathematical  Sciences,  Report  No.  IMM-203,  New  York  University, 
1954. 
Pierson*  W.  J.,  Jr. 

[P.8]      The  interpretation  of  crossed  orthogonals  in  wave  refraction  phenomena. 
Technical    Memorandum    No.    21,    Beach    Erosion    Board,   Corps  of 
Engineers. 
Picrson,  W.  J.,  Jr.,  and  M.  St.  Denis 

[P.9]      On  the  motion  of  ships  in  confused  seas.  Society  of  Naval  Architects  and 

Marine  Engineers,  1953. 
Pierson,  W.  J.,  Jr. 

[P.  1 0  ]    A  unified  mathematical  theory  for  the  analysis,  propagation ,  and  refraction 
of  storm  generated  ocean  surface  waves.   Part  I,   Part  II,  Reports  of 
College  of  Engineering,  New  York  University,  1952. 
Pohle,  F.  V. 

[P.ll]    The  Lagrangian  equations  of  hydrodynamics:  solutions  which  are  analytic 

functions  of  the  time.  Thesis,  New  York  University,  1950. 
[P.12J    Motions  of  water  due  to  breaking  of  a  dam,  and  related  problems.  U.  S. 
National  Bureau  of  Standards,  Gravity  Waves,  N.  B.  S.  Circular  521, 
1952. 

[P.13]    See  [D.14J. 
Price,  A.  T. 

[P.14]    See  [P.8]. 
[P.15]    See  [M.3]. 


556  WATER   WAVES 

Preiswerk,  E. 

[P.16]    Anwendung  gasdynamischer  Methoden  auf  Wasserstromungen  mil  freier 
Oberflache.  Mitteilungen  aus  dem  Institut  fur  Aerodynamik,  Eidg.  Tech- 
nische  Hochschule,  Ztirich,  1038. 
Putman,  H.  J. 

[P.17J    Unsteady  flow  in  open  channels.  Transactions  of  the  American  Geophy- 
sical Union,  Vol.  29,  1948. 
Putnam,  J.  A.,  and  R.  S.  Arthur 

[P.  18]    Diffraction  of  water  waves  by  breakwaters.  Transactions  of  the  American 

Geophysical  Union,  29  (4),  1948. 
Peters,  A.  S.,  and  J.  J.  Stoker 

[P.  19]  A  uniqueness  theorem  and  a  new  solution  for  Sommerf eld's  and  other 
diffraction  problems.  Communications  on  Pure  and  Applied  Mathema- 
tics. Vol.  7,  1954,  pp.  565-585. 

Rayleigh,  Lord 

[R.I]      On  waves.  London,  Dublin  and  Edinburgh,  Philosophical  Magazine, 

Ser.  5,  Vol.   1,  1876. 

[R.2 j      The  form  of  standing  waves  on  the  surface  of  running  water.  Proceedings 

of  the   London   Mathematical   Society,   Vol.    15,    1883-4,  pp.  69-78. 

[R.3]      On  the  theory  of  long  waves  and  bores.  Proceedings  of  the  Royal  Society 

of  London,  Ser.  A,  Vol.  90, 1914. 

IR.4]      The  Theory  Of  Sound.  Dover  Publications,  New  York,  1945. 
Re    R. 

[R.5]      titude  du  lacher  instantane  d'une  retenue  d'eau  dans  un  canal  par  la 

m&hode  graphique.  La  Houille  Blanche    N.  S.  Vol.  1,  1946. 
Rees,  M. 

[R.6]      See  [C.ll]. 
Rellich    F. 

[R.7]      Vber   das   asymptotische   Verhalten  der  Losungen  von  Au  -f-  u  =  0  in 
unendlichen  Gebieten.  Jahresbericht  der  deutschen  Mathematiker  Ver- 
einigung,  Vol.  53,  1943,  pp.  57-65. 
Riabouchinsky,  D. 

[R.8]     Sur  V analogic  hydraulique  des  mouvemente  d'un  fluide  compressible. 
Institut  de  France,  Academic  des  Sciences,  Comptes  Rendus,  Vol.  195, 
1932,  p.  998. 
Rose,  N.  W. 
See  [K.14] 
Roseau,  M. 

[R.9]      Contribution  a  la  the'orie  des  ondes  liquides  de  gramle  en  profondeur 
variable.    Publications    Scicntifiques  et  Techniques   du  Ministere  de 
1'Air,  No.  275,  Paris,  1952. 
Rouse,  H. 

[R.10]    Fluid  mechanics  for  hydraulic  engineers.  McGraw-Hill,  New  York,  1938. 
Rouse,  H.,  editor 

[R.  1 1  ]    Engineering  Hydraulics.  John  Wiley  and  Sons,  New  York,  and  Chapman 
and  Hall,  Ltd.,  London,  1950.  (Especially  Gilcrest,  B.  R.,  Flood  rout- 
ing, pp.  635-676.) 
Ruellan,  F.,  and  A.  Wallet 

[R.12J    Trajectoires  internes  dans  un  clapotis  partiel.  La  Houille  Blanche,  Vol.  5, 

1950. 
Rubin,  H. 

[R.13]    The  dock  of  finite  extent.  Thesis,  New  York   University,   1953. 
Russell,  S. 

[R.14]    Report  on  waves.  British  Association  Reports,  1844. 


BIBLIOGRAPHY  557 

St.  Denis,  M.,  and  G.  Weinblum 

[S.I]      On  the  motions  of  ships  at  sea.  Society  of  Naval  Architects  and  Marine 

Engineers,  1950. 
[S.2]       See   fP-91. 
Saint  Venant,  B.  de 

[S.3]      Thtorie  du  mouvement  non  permanent  des  eaux.  Institut  de  France, 
Academic  des  Sciences,  Comptes  Rendus,  Paris,  July  1871,  Vol.  73, 
pp.  147,  237. 
Schonfeld,  .J.  C. 

fS.4]       Propagation   of  Tides   and  Similar   Waves.   Doctor's   Thesis,   Delft, 

Holland,  1951. 
FS.4a]     Distortion  of  long  waves;  equilibrium  and  stability.  Union  Geod£sique  et 

Geophysique  Internationale,  Vol.  4,  1951. 
Sch winger,  J.  S. 

[S.5]       Fourier   transform   solution   of  integral   equations.    M.I.T.    Radiation 

Laboratory  Report. 
Scripps  Institution  of  Oceanography 

[S.6  j       Waves  in  shallow  water  (I).  Change  in  wave  height  along  Scripps  Institu- 
tion pier  with  special  emphasis  on  breaker  cliaracteristics.  Wave  Report 
No.  13. 
[S.7'       Effect  of  bottom  slope  on  breaker  characteristics  as  observed  along  the 

Scripps  Institution  pier.  Wave  Report  No.  24,  1944. 
Seiwell,  H.  R. 

[S.8J       Military  oceanography  in  World  War  II.  Mil.  Eng.  39  (259),  1947. 
I^S.9]       The  principles  of  time  series  analyses  applied  to  ocean  wave  data.  Proceed- 
ings of  the  National  Academy  of  Sciences,  Vol.  35,  1949,  pp.  518-528. 
S.10]    Sea  surface  roughness  measurements  in  theory  and  practise.  Annals  of  the 
New  York  Academy  of  Sciences,  Vol.  51,  Art.  3,  1949,  pp.  483-501. 
Sneddon,  I.  N. 

[S.11J     Fourier  Transforms.  McGraw-Hill,  New  York,  1951. 
Solberg,  II. 

See   [B.20] 
Sommerfeld,  A. 

[S.I 2]     Mathematische  Theorie  der  Diffraktion.  Mathematische  Annalen,  Ud. 

47,  p.  317,  1896. 
[S.13]    Vorlcsungen  iiber  theoretische  Physik,  II,  Mechanik  der  deformierbaren 

Medien.  Akademie  Verlag,  Leipzig,  1949. 
Southwell,  R. 

[S.14]     Relaxation  Methods.  Oxford  Press,  1940. 
Sretenski,  L.  N. 

[S.15J    Ship  waves  for  circular  courses  (in  Russian).  Bulletin  dc  TAcaddmie  des 

Sciences  de  TURSS,  1946. 
Stelzriede,  M.  E. 

[S.16]     See   [C.I]. 
Stephan,  S.  C.,  Jr. 

[S.17]     See   [D.I]. 
Stoker,  J.  J. 

[S.I 8]    Surface  waves  in  water  of  variable  depth.  Quarterly  of  Applied  Mathe- 
matics, Vol.  5,  1947,  pp.  1-54. 
,S.19]     The  formation  of  breakers  and  bores.  Communications  on  Pure  and 

Applied  Mathematics,  Vol.  1,  1948,  pp.  1-87. 
[S.20]     The  breaking  of  waves  in  shallow  water.  Annals  of  the  New  York  Academy 

of  Sciences,  Vol.  51,  Art.  3,  1949,  pp.  360-375. 

S.21]    On  radiation  conditions.  Sixth  Symposium  on  Applied  Mathematics. 
American  Mathematical  Society,  1953. 


558  WATER   WAVES 

[S.22]    Unsteady  waves  on  a  running  stream.  Communications  on  Pure  and 

Applied  Mathematics,  Vol.  6,  1953,  pp.  471-481. 
[S.23]    Numerical  solution  of  flood  prediction  and  river  regulation  problems  I : 

Derivation  of  basic  theory  and  formulation  of  numerical  methods  of  attack. 

New  York  University,  Institute  of  Mathematical  Sciences,  Report 

No.  IMM-200,  1953. 
[S.24]    Dynamical  theory  for  treating  the  motion  of  cold  and  warm  fronts  in  the 

atmosphere.  Institute  of  Mathematical  Sciences,  Report  No.  IMM-195, 

1953,  New  York  University. 
[S.25]    See  [F.5]. 
[S.26]    See  [1.4]. 
[S.27]     See  [P.7]. 
[S.27a]  See  [I.4a]. 
Stokes,  G.  G. 

[S.28]  On  the  theory  of  oscillatory  waves.  Transactions  of  the  Cambridge  Philo- 
sophical Society,  Vol.  8,  1847,  and  Supplement.  Scientific  Papers, 

Vol.  1. 
Struik,  D.  J. 

[S.29]    Determination  rigoureuse  des  ondes  irrotationnelles  periodiques  dam  un 

canal  a  profondeur  finie.  Mathematische  Annalcn,  Vol.  95,  1926,  pp. 

595-634. 
Suquet,  F. 

[S.30]    Remarks  on  graphical  computation  of  wave  refraction.   International 

Association  for  Hydraulic  Research,  Grenoble,  1949. 
[S.31]    Etude  experimental  du  deferlement  dc  la  houle.  La  Houille  Blanche, 

Vol.  5,  1950. 
Sverdrup,  H.  U. 

[S.32]    Oceanography  for  Meteorologists.  Prentice- Hall,  Inc.,  X.  Y.,  1948. 
Sverdrup,  H.  U.,  M.  W.  Johnson,  and  R.  H.  Fleming 
|S.82n]  The  Oceans.  Prentice-Hall,  New  York,  1946. 
Sverdrup,  H.  U.,  and  W.  H.  Munk 

[S.33]     Theoretical  and  empirical  relations  in  forecasting  breakers  and  surf. 

Transactions  of  the  American  Geophysical  Union,  Vol.  27,  1946. 
[S.34]    A  study  of  progressive  oscillatory  waves  in  water;  a  summary  of  the  theory 

of  water  waves.   Technical   reports  No.   1,  2,   Beach  Erosion  Board, 

Office  of  Chief  of  Engineering,  U.S.  War  Department. 
[S.35]     Breakers   and  surf.    U.S.    Navy   Department,    Hydrographic   Office, 

Publication  No.  234,  1944. 
[S.36]     Wind,  sea9  and  swell:  Theory  of  relations  for  forecasting.  U.S.  Xavy 

Department,  Hydrographic  Office,  1947. 

[S.37]     Empirical  and  theoretical  relations  between  zvind,  sea  and  swell.  Trans- 
actions of  the  American  Geophysical  Union,  Vol.  27,  1946. 
Sretenski,  L.  N. 

[S.38]     Waves.  In:  Mekhanika  v  SSSR  za  tridtsat'let  1917-1947.  Moscow,  1950, 

pp.  279-299. 

Thomas,  H.  A. 

[T.I]  The  propagation  of  stable  wave  configurations  in  steep  channels.  Carnegie 
Institute  of  Technology,  Pittsburgh,  Pa. 

[T.2]  Hydraulics  of  flood  movements  in  rivers.  Carnegie  Institute  of  Tech- 
nology, Pittsburgh,  Pa.,  1937. 

[T.8]  Propagation  of  waves  in  steep  prismatic  conduits.  Proceedings  of  Hydrau- 
lics Conference,  University  of  Iowa,  Studies  in  engineering,  Bulletin  20, 
1940. 


BIBLIOGRAPHY  559 

Thorade,  H.  F. 

[T.4]      Probleme  der  Wasserwellen.  1931,  from  Probleme  der  kosmischen  Physik, 

Bd.  13-14. 
Thornhill,  C.  K. 

[T.5]      See  [M.3]. 
Titchmarsh,  E.  C. 

[T.6]      Introduction  to  the  Theory  of  Fourier  Integrals.  Second  Edition,  The 

Clarendon  Press,  Oxford,  1948. 
Tray  lor,  M.  A. 

[T.7]      See   [M.15]. 
[T.8]      See   [M.16]. 
Troesch,  B.  A. 
[T.9]      See  [1.4]. 
[T.9a]    See  [I.4a]. 
Thompson,  P.  D. 

[T.10]    Notes  on  the  theory  of  large-scale  disturbances  in  atmospheric  flow  with 
applications  to  numerical  weather  prediction.  Airforce  Cambridge  Re- 
search Center,  Geophysical  Research  Papers,  No.  16,  1952. 
Tepper,  M. 

[T.ll]  A  proposed  mechanism  of  squall  lines:  the  pressure  jump  line.  Journal  of 
Meteorology,  Vol.  7,  1950,  pp.  21-29. 

U.S.  Navy  Hydrographic  Office 

[U.1J      Wind,  ivaves  and  swell:  Principles  in  forecasting.  H.  O.  Misc.  11,  1944. 
[U.2]      Breakers  and  surf,  principles  in  forecasting.  H.  O.  234    1944. 
Ursell,  F. 

[U.3]      The  effect  of  a  vertical  barrier  on  surface  waves  in  deep  water.  Proceedings 

of  the  Cambridge  Philosophical  Society,  Vol.  43,  1947.  pp.  374-382. 
[U.4]      On  the  waves  due  to  the  rolling  of  a  ship.  Quarterly  Journal  of  Mechanics 

and  Applied  Mathematics,  Vol.  1,  1948,  pp.  246-252. 
[U.5J      Surface  waves  on  deep  water  in  the  presence  of  a  submerged  circular 

cylinder,  I,  II.  Cambridge  Philosophical  Society,  Vol.  46,  1950,  pp. 

141-152,  153-163. 
[1T.6]      On  the  application  of  harmonic  analysis  to  ocean  wave  research.  Science, 

Vol.  Ill,  1950. 
[LI. 7]      Discrete   and  continuous  spectra  in  the  theory  of  gravity  waves.  U.S. 

National  Bureau  of  Standards,  Gravity  waves,  N.B.S.  Circular  521, 

1952,  pp.  1-5. 

[U.8J      Trapping  modes  in  the  theory  of  surf  ace  leaves.  Proceedings  of  the  Cam- 
bridge Philosophical  Society,  Vol.  47,  1951. 
[U.9]      On  the  rolling  motion  of  cylinders  on  the  surface  of  a  fluid.  Quarterly 

Journal  of  Mechanics  and  Applied  Mathematics,  Vol.  2,  1949. 
[U.10J    On  the  heaving  motion  of  a  circular  cylinder  on  the  surface  of  a  fluid. 

Quarterly  Journal  of  Mechanics  and  Applied  Mathematics,  Vol.  2,  1949. 
[U.11J    The  long-wave  paradox  in  the  theory  of  gravity  leaves.  Proceedings  of  the 

Cambridge  Philosophical  Society,  Vol.  49,  1953. 
[U.12J    See  [B.4J. 
[U.13J    See  [L.14]. 
Unaki,  G.,  and  M.  Nakano 

[U.14]    On  the  Cauchy-Poisson  waves  caused  by  the  eruption  of  a  submarine 

volcano,  (I)  and  (II).  Oceanograph.  Mag.  4,5;  1953. 

Vedernikov,  V.  V. 

[V.  1  ]  Conditions  at  the  front  of  a  translation  wave  disturbing  a  steady  motion  of  a 
real  fluid.  Doklady  (Comptes  rendus)  Akademiya  Nauk  SSSR,  Vol.  48, 
1945. 


560  WATER   WAVES 

[V.2]     Characteristic  features  of  a  liquid  flow  in  an  open  channel.  Doklady 

(Comptes  rendus),  Vol.  52,  1946,  pp.  207-210. 
Vergne,  Henri 

[V.3]     Ondes  liquides  de  gravite*.    Actualites    Scientifiques  et  Industrielles, 

Hermann  et  Cie.,  Paris,  1928. 
de  Vries,  G. 

[V.4]      See  [K.15]. 

Wallet,  A. 

See  [R.12] 
Watson,  G.  N. 

[W.I]     A  Treatise  on  the  Theory  of  Bessel  Functions.  The  University  Press,  Cam- 
bridge, 1944. 
Wilkes,  M.  V. 

[W.2]    Oscillations  of  the  Earth's  Atmosphere.  The  University  Press,  Cambridge, 

1949. 
Weinblum,  G.  P. 

[W.3]     Analysis  of  wave  resistance.  David  W.  Taylor  Model  Basin  (Washington, 

D.  C.),  Report  710,  1950. 
[W.4]     See  [S.I]. 
Weinstein,  A. 

[W.5]     On  surface  waves.  Canadian  Journal  of  Mathematics,  Vol.   1,  1949, 

pp.  271-278. 

[W.6]  Sur  la  vitesse  de  propagation  de  Vonde  solitaire.  Reale  Acadcmia  dci 
Lincei,  Rendiconti,  Classe  di  scienze  fisiche,  matematiche,  naturali, 
Ser.  6,  Vol.  3,  1926,  pp.  463. 

[W.7J    Sur  un  probleme  aux  limites  dans  une  bande  inde'finie.  Institut  de  France. 
Comptes  Rendus  de  1' Academic  des  Sciences.  Vol.  184,  1927,  p.  497. 
Weitz,  M.,  and  J.  B.  Keller 

[W.8]     Reflection  of  water  waves  from  floating  ice  in  water  of  finite  depth.  Com- 
munications on  Pure  and  Applied  Mathematics.   Vol.   3,   1950,  pp. 
305-318. 
Also  see  [K.9] 
Weliczker,  L. 

[W.9]     See  [F.5]. 
Wells,  Leon  W.  (formerly  L.  Weliczker) 

[W.10]  Some  remarks  on  shallow  water  theory.  New  York  University,  Institute 

of  Mathematical  Sciences,  Report  No.  IMM-198,  1953. 
Westergaard,  H.  M. 

fW.ll]  Water  pressure  on  dams  during  earthquakes.  Transactions  of  American 

Society  of  Civil  Engineers,  Vol.  93,  1933. 
Whitham,  G.  B. 

[W.12]  Dynamics  of  meteorological  fronts.  New  York  University,  Institute  of 

Mathematical  Sciences,  Report  No.  IMM-195,  1953. 
Wigley,  W.  C.  S. 
[W.13]  See  [L.18]. 


Author  Index 


Abdullah,  A.  J.,  373,  388  Freeman,  J.  C.,  Jr.  375,  377,  384,  388, 

Arthur,  R.  S.,   132,   133  408 

Friedriclis,  K.  O.,  27,  32,  73,  74,  75,  76, 

Baird,  E.  G.,  412  108,  293,  300,  343,  344,  345,  849,  371, 

Bakhmeteff,  B.  A.,  451  407,  470,  529 
Bateman,  II.,  117 

Bates,  C.  C.,   138  Gerber,  R.,  522 

Biesel,  F.,  364,  305  Gilcrest,  B.  R.,  451,  462,  482 

Birkhoff,  G.,  342  Goldstein,  E.,  447 
Bjerknes,  J.,  376 

Boridi,  H.,  72  Hamada,  T.,  359,  372 

Bouasse,  H.,  423  Hanson,  E.  T.,  72,  89 

Boussinesq,  J.,  342,  351,  451  Harleman,  D.  R.  F.,  412 

Broer,  L.  J.  F.,  53  Haskind,  M.  D.,  248,  259 

Bruman,  J.  R.,  412  Haurwitz,  B.,  376 

Haveloek,   T.   II.,   219,   242,   246,   248, 

Carr,  J.  H.,   132,  445,  448  253,  256 

Carson,  R.  L.,  XIV  Heins,  A.  E.,  74,  108,  141 

Cauchy,  A.  L.,  35,   154  Hinze,  J.  O.,  160 

Charney,  J.  G.,  375  Hogner,  E.,  220,  242 

Chrystul,  G.,  424  Hopf,  E.,  44 

Cooper,  R.  I.  B.,  138  Hyers,  D.  H.,  32,  343,  345,  349,  371,  529 
Copson,  E.  T.,   117,  181 

Cornish,  V.,  466  Isaacs,  J.  D.,   133 
Courant,  R.,  293,  300,  407,  470,  476,  477    Isaaeson,  E.,   73,   74,   451,   476,   477 

Crossley,  H.  E.,  412  Iversen,  H.  W.,  359 

Daily,  J.  \V.,  351  Jeffreys,  II.,  352,  422 

Danel,  P.,   138  John,  F.,  76,  113,  146,  175,  206,  249, 

Darby  shire,  J.,  137  431 

Davies,  T.  V.,  374,  522  Johnson,  J.  W.,   133 

Deacon,  G.  E.  R.,  137,   171  Johnson,  M.  \V.,  133 

Deymie,  P.,  505 

Dressier,  R.  F.,  466,  467,  468  Kampd  de  Feriet,  J.,  187 

Dubreuil-Jaeotin,  M.  I,.,  522  von  Kdrman,  T.,  410,  482 

Karp,  S.  N.,  117,  141 

Eekart,  C.,  73,   136  Keldysh,  M.  V.,  203 

Einstein,  II.  A.,  412  Keller,  J.  B.,  32,  141,  146,  343,  351,  371, 

447,  476,  481 

Finkelstein,  A.,   153,   187  Kelvin,  W.  T.,  163,  219 

Fjeldstad,  J.  E.,  147  Keulegan,  G.  H.,  370 

Fleishman,  B.,  438  Korteweg,  D.  J.,  342,  343 

Fleming,  R.  H.,  133  Korvin-Krukovsky,  B.  V.,  260 

Forchheimer,  P.,  451  Kotik,  J.,   153,  187 

Forel,  F.  A.,  423  Kotsehin,  N.  J.,  61 

561 


562  WATEE  WAVES 

Kreisel,  G.,  146  Re,  R.,  818 

Krylov,  A.  N.,  248  Rees,  M.,  476,  477 

Rellich,  F.,  113,  175 

Laitone,  E.  V.,  412,  418  Riabouchinsky,  D.,  25 

Lamb,  H.,  28,  27,  57,  58,  180,  421  Roseau,  M.,  74,  75,  76,  78,  95,  146 

Lavrentieff,  M.,  344  Rouse,  H.,  451,  462,  468,  482 

Lax,  A.,  368  Rubin,  H.,  146 

Lax,  P.,  476,  481  Ruellan,  F..  41 

Leray,  J.,  522  Russell,  S.,  342 
Levi-Civita,  T.,  17, 21, 343, 345, 346, 347, 

374,  513,  522,  524  St.  Denis,  M.,  248,  261 

Lewis,  E.  V.,  260  Schwinger,  J.  S.,  117,  146 

Lewy,  H.,  61,  72,  73,  74,  76,  79,  108,  537  Seiwell,  H.  R.,  138 

Lichtenstein,  L.,  515,  522,  529  Solberg,  H.,  376 

Littman,  W.,  522  Sommerfeld,  A.,  52,  77,  116,  117 

Longuet-Higgins,  M.  S.,  137,  138  Southwell,  R.,  426 

Lowell,  S.  C.,  136  Sretenski,  L.  N.,  234 

Lunde,  J.  K.,  246,  248,  257  Stelzriede,  M.  E.,  132 

Stoker,  J.  J.,  Ill,  451,  482 

Macdonald,  H.  M.,  117  Stokes,  G.  G.,  96,  373 

McNown,  J.  S.,  426,  428  Struik,  D.  J.,  17,  21,  342,  343,  344,  347, 
Massau,  J.,  482  522 

Mason,  M.  A.,  372  Suquet,  F.,  133 

Miche,  A.,  72,  137  Sverdrup,    II.   U.,    133,   357,   369,   371 
Michell,  J.  H.,  248,  253,  256,  263,  285 

Munk,  W.  H.,  69,  133,  352,  357,  369,  Tepper,  M.,  375 

870,  371  Thomas,  II.  A.,  451,  462,  467,  468,  482 

Thompson,  P.  D.,  375 

Nekrassov,  A.  I.,  522  Thorade,  H.  F.,  368 

Nirenberg,  L.,  344,  522  Thornhill,  C.  K.,  513,  522 

Traylor,  M.  A.,  133 

O'Brien,  M.  P.,  133  Troesch,  B.  A.,  451,  487,  505 

Patterson,  G.  W.,  370  Ursell,  F.,  96,  137,  146,  342,  372 

Penney,  W.  G.,  513,  522 

Peters,  A.  S.,  74,  75,  76,  78,  95,  96,  98,  de  Vries,  G.,  342,  343 

102,   103,   111,   124,  224,  242,  245 

Pierson,  W.  J.,  Jr.,  133,  138,  248  Wallet,  A.,  41 

Pohle,  F.  V.,  513,  515  Weinblum,  G.  P.,  248,  261 

Poincare,  H.,  181  Weinstein,  A.,  40,  87,  208,  ,342 

Poisson,  S.,  35,  154  Weitz,  M.,  141 

Preiswerk,  E.,  407,  411,  413,  482  Wells,   L.   W.,   434,   444,   449 

Putnam,  J.  A.,  132  Whitham,  G.  B.,  368,  377,   378,   388, 

389,   395,   399,   404,   487,   505 

Rankine,  W.  J.  M.,  76  Wigley,  W.  C.  S.,  246,  257 

Rayleigh,  J.  W.,  49,  321,  342,  351  Wilkes,  M.  V.,  147,  375 


Subject  Index 


Aerodynamics,  411 
Angle  of  trim,  286 
Archimedes'  law,  254,  277 
Atmosphere,  gravity  waves,  374 

tidal  oscillations,  374 

waves  on  discontinuity  surfaces,  375 

Backwater  curves,  456 

Backwater  effects   in   long   rivers,  461 

Beaches.  (See  also  Sloping  beaches.) 

waves  breaking  on  shallow,  352 
Bernoulli's  law,  9 
Bernstein,  S.,  theorem,  43 
Bifurcation  conditions,  529 
Bifurcation  phenomenon,  343 
board,  as  fixed  breakwater,  432 

as  floating  breakwater,  436 
Bore  307,  315,  326,  368 

development  351 

Tsien-Tang  River,  320,  368 
Boundary  conditions,  10,  19 

dynamical,  55 

fixed  boundary  surface,  11 

free  surface,  11 

kinematic,   16,  56 

small  amplitude  theory,  19 

tidal  theory,  422 
Breaking  of  a  dam,  313,  333,  513 

discharge  rate,  338 

resulting  bore,  334 
Breaking  of  waves,  69,  307,  315 

at  crests,  369 

in  shallow  water,  351 

induced  by  wind  action,  372 

on  shallow  beaches,  352 
Breakwaters,  429 

dispersion  induced,  443 

fixed  board,  432 

floating  board,  436 

floating  elastic  beam,  438 

reflection  of  energy,  446 


Cauchy-Riemann  equations,  345 
Cavitation,  310 


Characteristic^ ),    curves,    294 

differential  equations,  294 

envelope,  307,  314,  355 

intersection,  355 

method,  293,  469 

propagation  of  discontinuities  along, 

473 

Chezy  formula,  466 
Circulation,  7 

Cliff,  waves  against  a  vertical,  84 
Cnoidal  waves,  342 
Cold  front,  380 
Comparison  of  predicted  and  observed 

floods,  498 
Compressibility,  3 
Contact  discontinuity,  318 
Continuity  equation,  453 
Convolution  theorem,  143 
Coriolis  acceleration,  383 

force,  381 

Crests,  breaking  of  waves  at,  369 
Critical     speed,     inappropriateness     of 
linear  theory  at,  217,  344 

instability  of  steady  flow  with,    344 
Cyclone,  376,  399 

Dam,  breaking,  313,  333,  513 
discharge  rate  on  breaking,  338 
shock  resulting  from  breaking,  334 

Diffraction  around  a  vertical  wedge,  109 
problem  of  Sommerfeld,  109 
theory,  physical  verification  of,  132 

Dipoles,  13 

Discontinuities,  propagated  along  char- 
acteristics, 473 

Discontinuity  conditions,  314 
surfaces  in  the  atmosphere,  375 

Dispersion,  51 

Divergence  theorem,  6 

Diverging  system  of  waves,  237 

Dock  problem,  74 
two-dimensional,  108 

Domain  of  dependence,  298 
of  determinancy,  299 

Dynamic  boundary  condition,  16,  55 


563 


564 


WATER  WAVES 


Eigenvalue  problems,  424 
Elastic  beam,  used  as  floating  break- 
water, 438 
Energy,  13 

average,  50 

balance  across  a  shock,  318 

flux,  13 

rate  of  change,  13 

reflected  by  a  breakwater,  446 

transmission  by  progressing  waves,  15 

transmission    by    simple    harmonic 
waves,  47 

velocity  of  the  flow  of,  49 
Envelope  of  characteristics,  307,  314, 

355 
Engineering    methods    in    flood    wave 

problems,  485 

Equation  of  continuity,  7,  453 
Equations  of  flow  in  open  channels,  452 
Equations  of  motion,  4 

Eulerian  form,  6 

Lagrangian  form,  4 
Equations    of   shallow    water    theory, 
nonlinear,  24 

validity  beyond  the  breaking  point, 

362 

Euler  variables,  5 
Exact  free  surface  condition,  513 
Experimental,  wave  tanks,  71 

solitary  wave,  351 

waves  on  sloping  beaches,  71 

Finite  difference  methods,  296,  424,  474 

convergence  of,  477,  481 
Floating,  bodies  in  shallow  water,  431 

breakwaters,  414 

elastic  beam,  438 

rigid  body,  245 
Flood  prediction,  482 
Flood  routing,  461 

Flood   waves,   in   the   Mississippi   and 
Ohio  Rivers,  458,  483,  494 

monoclinal,  462 
Flow,  around  bends,  405 

between  two  walls,  410 

in  open  channels,  451 

of  energy,  13 

over  obstacles,  344 

through  a  sluice,  407 
Forced  oscillations,  55 
Fourier  integral  theorem,  153 
Fourier  transform,  35,  155 
Free  natural  vibrations  of  a  lake,  424 
Free  surface,  11.  (See  also  Surface.) 
Free  surface  condition,  20 


exact,  11,  513 

linearized,  11,  12,  35 
Free  surface  elevation,  16 
Friction,  451 
Friction  slope,  455 
Front,  378 

cold,  374,  380 

occluded,  381 

stationary,  378 

warm,  374,  380 
Front  of  shock,  320 

Gas  dynamics  analogy,  25 
Geometrical  optics,  133 
Gravity  waves  in  the  atmosphere,  374 
Green's  function,  280 
time-dependent,  187 
Group  of  waves,  51 
Group  velocity,  51,  170 

Harbors,  design  of,  420 

model  studies,  429 

oscillations,  414 
Heave,  250,  255 
Heaving,  278 
HelmhoUVs  theorem,  7 
Higher-order  approximations  in  shallow 

water  theory,  28,  32 
Hump,  352 

Hydraulic  analogy,  412 
Hydraulic  jumps,   307,  324,  407.    (See 
also  Bore.) 

interaction  of,  412 
Hydraulic  radius,  454 
Hydraulics,  mathematical,  451 
Hydrostatic  pressure  law,  24,  31 

in  meteorology,  374,  382 

Influence  point,  228 
Initial  characteristic,  302 
Initial  steepness  of  a  wave,  357 
Instability  of  steady  flow  with  critical 

speed,  344 

Interaction  of  two  hydraulic  jumps,  412 
Internal  waves,  147 
Intersection  of  characteristics,  855 
Irrotational  flow,  9 
Iteration  process,  539 

Jump,  hydraulic.  See  Hydraulic  jumps. 
Junction  of  the  Ohio  and  Mississippi 

Rivers,  457,  509 
flood  wave  through,  494 

Kelvin's  theory  of  ship  waves,  219 
Kinematic  boundary  condition,  16,  56 


SUBJECT  INDEX 


565 


Lagrangian  form  of  the  equations  of 

motion,  4 

Lake,  free  natural  vibrations  of,  424 
Levi-Civita's  theory,  513,  522 
Linear  theory,  35 

derivation  of,  19 

free  surface  condition,  21 
Local  speed  of  small  disturbance.  See 

Wave  speed. 
Long  waves,  theory  of,  23,  291 

Much  lines,  40U 

Manning's  formula,  454 

Manning's  roughness  coefficient,  457 

Margules'  law,  387 

Mass  flux,  0 

across  a  shock  front,  318 
Mathematical  hydraulics,  451 
Meteorology,  374 

hydrostatic  pressure  law  in,  374,  382 
Michell's  type  ship,  257 
Microscisms,  origin  of,   137 
Mississippi  Hivcr,  509 

flood  waves,  458,  484 

model,  482,  509 

junction  with  the  Ohio  River,  457,  509 
Models    of   the    Ohio    and    Mississippi 

Rivers,  482,  508 
Model    studies    of   harbors,    429 
Momentum,  conservation  of,  3 
Monoclinal  flood  wave,  4(52 
Motions.    (See   also   Flow;    Wares.) 

steady,  199,  201 

uniqueness  of  unsteady,  187 

unsteady,  149 
Moving  pressure  point,  217 

Non-existence  of  depression  shock,  323 
Nonlinear  free  surface  condition,  11,  513 
Non linearity  of  breaking  phenomena,  71 
Nonlinear  shallow  water  theory  equa- 
tions, 24 

Numerical  solutions  for  sloping  beaches, 
73,  75 

Obstacles,  flows  over,  344 

waves  due  to,  35 
Occluded  front,  381 
Oceanography,  133 
Ocean  tides,  421 
Ohio  River,  505 

flood  waves,  458,  484 

junction  with  the  Mississippi  River, 
457 

model  of,  482 


Open  channel  flows,  451 

unsteady,  409 
Optics,  geometrical,  133 
Oscillations,  forced,  55 

free,  55 

in  harbors,  414 

of  a  lake,  423 

of  the  atmosphere,  374 

pitching,  250 

rolling,  250 

simple  harmonic,  37 

small,  35 

yawing,  250 
Overhanging  cliff,  73 

Particle  derivative,  5 
Periodic   impulse,   waves   due   to,    17 1 
Periodic   surface  pressure,   57 
Periodic  waves.  (See  also  Waves.) 

existence  of,  522 
Perturbation  procedure,  19,  269 
Phase  speed,   170 
Pitching  oscillation,  250,  278 
Point  source  (or  sink),   12 
Potential,  singularities  of,   12 
Potential  flow,  9 
Pressure,  3.  (Sec  also  Surface  pressure.) 

periodic,  57 

waves  caused  by  moving,  219 
Privaloff's  theorem,  536 
Profile,  of  a  river,  458 
Progressing  waves,  57.  (See  also  Wares.) 

of  fixed  shape,  461 

simple  harmonic,  45 
Propagation    of    discontinuities    along 

characteristics,  474 
Propagation  speed.  See  Wave  speed. 

Radiation  condition,  174,  209.  (See  also 

Sommerfeld  entries. ) 
Range  of  influence,  299 
Rnyleigh-Uitz  method,  426 
Reflection  of  energy,  446 

of  shock,  330 

of  waves,  71,  95 
Refraction  along  a  coast,  133 
Resistance  force,  453 
Resonance,  58 
Rigid  body,  floating,  245 
River  profile,  458 
Rivers,   backwater  effects  in,  461 
Rolling  oscillations,  250 
Roll  waves,  466 
Roughness  coefficient,  454,  466 

Manning,  457 


566 


WATER  WAVES 


Running  stream,  waves  on  a,  198 

Schauder-Leray  theory,  522 

Schmidt,  E.,  bifurcation  theory  of,  529 

Seiche,  423 

Seismology,  137 

Shallow  water,  floating  bodies,  431 

long  waves,  291 
Shallow  water  theory,  22,  291 

accuracy,  27 

equations  of,  24 

for  sloping  beaches,  75 

higher-order  approximation  in,  28,  32 

linear,  25,  75,  414 

linear,     compared     with     numerical 
solution,  75 

linear  sound  speed,  419 

mathematical  justification,  31 

reformulation  of  equations,  292 

systematic  derivation  of,  27 

validity  beyond  the  breaking  point, 

362 

Ship,  as  a  floating  rigid  body,  245 
Ship  wave  problem,  219,  224 
Ship  waves,  diverging  system,  237 

in  water  of  finite  depth,  243 

method  of  stationary  phase,  219 

transverse  system,  237 
Shock,  317 

advancing  into  still  water,  323 

back  of,  321 

conditions,  314 

constant,  326 

energy  balance  across,  318 

front  of,  321 

mass  flux  across,  318 

non-existence  of  depression,  323 

reflected  from  a  rigid  wall,  330 

resulting  from  the  breaking  of  a  dam, 
333 

turbulence  at  front  of,  320 
Simple  harmonic  waves,  37 

energy  transmission,  47 

progressing,  45 
Simple  wave,  300,  469 

applications  to  problems  of  meteoro- 
logy, 391 

centered,  311 
Singularities  of  the  velocity  potential, 

12,  13 
Sink,  12 

Slenderness  parameter,  250 
Sloping    beaches,    69,    369.    (See    also 
Beaches.) 

experiments  on,  71,  73,  75,  373 

numerical  solutions,  73,  75 


Sluice   in   a   dam,    flow   through,    407 
Small  amplitude  theory,  19.  (See  also 

Linear  theory.) 

Small  oscillations.  See  Oscillations. 
Solitary  wave,  327,  342,  370 

approximation,  343 

experimental  work,  351 
Sommerfeld's  diffraction  problem,  109 
Sommerfeld's  radiation  condition,   59, 

65,  111,  113,  175 

Sound  speed,  26.  (See  also  Wave  speed.) 
Source,  12 

Standing    waves,     cylindrically     sym- 
metric, 41 

simple  harmonic,  37 

three-dimensional,  41 

two-dimensional,  38 
Stationary  front,  378 
Stationary  phase,  163,  219 

justification,  181 
Steady   flow,    supercritical,    405 

with  critical  speed,  344 
Steady  motions,  199 
Steady  state  problems,  unnaturalness, 

175 

Stoke's  phenomenon,  117 
Stoke's  theorem,  8 
Stream,  waves  on  a  running,  198 
Surge,  250 
Sway,  250 

Subcritical  flow,  305,  406 
Supercritical  flow,  304,  406 
supersonic  flow,  304 
Surface,  11.  (See  also  Free  surface.) 

condition,  exact,  11,  513 

disturbance,  motions  due  to,  156 

pressure,  confined  to  a  segment,  58 
periodic,  57 
simple  harmonic,  55 
Surfaces    of   discontinuity    in    the    at- 
mosphere, 375 
Surface  waves,  18 

typical  problem,  15 

Tidal  oscillations    of   the  atmosphere, 

375 
Tidal  theory,  22 

boundary  conditions,  422 
Tides  in  the  oceans,  421 
Transverse  system  of  ship  waves,  237 
Trim  of  a  ship,  251 
Turbulence,  453 

at  a  shock,  320 

Undertow,  71 
Uniqueness,  150,  187 


SUBJECT  INDEX 


567 


Velocity,  group,  170 

of  flow  of  energy,  49 
Velocity  potential,  9 

singularities  of,  12 
Vertical  cliff,  three-dimensional  waves 

against  a,  84 
Vertical  wedge.  See  Diffraction  around 

a  vertical  wedge. 
Viscosity,  3 
Vibrations,  of  a  lake,  424 

Warm  front,  380 
Water  table  experiments,  412 
Wave  motions.  (See  also  Flow;  Motions; 
and  Waves. ) 

in  open  channels,  451 

on    discontinuity    surfaces     in    the 

atmosphere,  375 

Wave  refraction  along  a  coast,  133 
Wave  resistance  integral,  284 
Waves.  (See  also  Flew;  Motions.) 

against  a  vertical  cliff,  67,  84 

breaking  of,  69,  307 

breaking  of  at  crests,  369 

breaking  of  on  shallow  beaches,  352 

breaking  point,  354 

centered  simple,  311 

cnoidal,  342 

depression,  306,  352 

diverging  system,  237 

due  to  a  moving  pressure  point,  219 

due  to  disturbances  from  rest,  35 

due  to  harmonic  surface  pressure,  49 

due  to  obstacles  in  a  running  stream, 
35 

due  to  periodic  impulse,  174 


energy  transmission,  47 

existence  of  periodic,  522 

experimental  work  on  solitary,  351 

experiments  on  sloping  beaches,  71, 
73,  75 

group,  51 

initial,  steepness,  357 

in  open  channels,  451 

internal,  147 

in  the  atmosphere,  374 

of  small  amplitude,  19 

on  sloping  beaches,  69,  369 

past  obstacles,  69 

progressing,  57,  67 

progressing,  of  fixed  shape,  461 

progressing,  simple  harmonic,  45 

reflection  from  shore,  71,  95 

roll,  466 

ship.  See  Ship  waves. 

simple,  300,  469 

simple  harmonic  standing,  37 

solitary,  327,  342,  870 

standing.  See  Standing  waves. 

steady,  207 

transverse  system,  237 

unsteady,  210 
Wave  speed,  26,  293,  299,  473 

in  linear  shallow  water  theory,  419 

in  meteorology,  404 
Wave   tanks,   experiments   in,   71,   73 
Wetted  perimeter,  454 
Wiener-Hopf  technique,  108,  141 
Wine  glass  effect,  74 

Yawing  oscillation,  250