EXPERIMENTAL STUDY OF INTERNAL GRAVITY WAVES OVER A SLOPE by David A. Cacchione A.B., Princeton University (1962) SUBMITTED IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY at the MASSACHUSETTS INSTITUTE OF TECHNOLOGY and the WOODS HOLE OCEANOGRAPHIC INSTITUTION September, 1970 Signature of Author Joint Program in Oceanography, Woods Hole Oceanographic Institution and Massachusetts Institute of Technology, Department of Earth and Planetary Sciences Certified by Thesis Supervisor Thesis Supervisor Thesis Accepted by C15 Chairman, Joint Oceanography Committee in the Earth Sciences, Massachusetts Institute of Technology - Woods Hole Oceanographic Institution EXPERIMENTAL STUDY OF INTERNAL GRAVITY WAVES OVER A SLOPE by David A. Cacchione Ph. D. Thesis MASSACHUSETTS INSTITUTE OF TECHNOLOGY and WOODS HOLE OCEANOGRAPHIC INSTITUTION Experimental Sedimentology Laboratory- Department of Earth and Planetary Sciences Massachusetts Institute of Technology Cambridge, Massachusetts 02139 USA Sponsored by the U. S. Office of Naval Research Nonr 1841 (74) Contract Authority NR 083-157 Reproduction in whole or in part is permitted for any purpose of the United States Government. Distribution of this document is unlimited. Report 70-6 November 1970 ft*? b 2-1 . -CIS" ABSTRACT ., CALIF, 93940 Experimental Study of Internal Gravity Waves Over a Slope by David A. Cacchione Submitted to the Department of Earth and Planetary Sciences in partial fulfillment of the requirement for the degree of Doctor of Philosophy. A series of laboratory experiments were conducted in a glass wave tank to investigate the propagation of internal gravity waves up a sloping bottom in a fluid with constant Brunt- Vaisala frequency. Measurements of the wave motion in the fluid interior were primarily taken with electrical conductivity probes; measurements in the boundary layer were made with dye streaks and neutrally buoyant particles. The results indicate that, outside of the breaking zone, the amplitude and horizontal wave number of the high-frequency waves increase linearly with decreasing depth; this is shown to agree with existing linear, inviscid solutions. A zone of breaking or runup is induced by these high-frequency waves well upslope. Shadow- graph observations show that, if the wave characteristics are coincident, or nearly so, with the bottom slope, the upslope propagation of the low- frequency waves causes a line of regularly spaced vortices to form along the slope. Subsequent mixing in the vortex cells creates thin horizontal laminae that are more homogeneous than the adjacent layers. These lam- inae slowly penetrate the fluid interior, creating a step-like vertical density structure. Available linear theoretical solutions for the velocity in the viscous boundary layer, determined to be valid for certain experimental conditions, are used to develop a criterion for incipient motion of bottom sediment in- duced by shoaling internal waves. The maximum sediment sizes that can be placed into motion, according to this criterion, are larger than certain mean sediment sizes on the continental margin off New England. This sug- gests that internal waves might induce initial sediment movement. Specula- tion about the geological effects of breaking and vortex instabilities is also given. These processes, not definitely measured in the field as yet, might also be conducive to sediment movement. Thesis Supervisors: John Southard Assistant Professor of Geology Carl Wunsch Associate Professor of Oceanography 111 TABLE OF CONTENTS ABSTRACT iii LIST OF SYMBOLS vii LIST OF FIGURES ACKNOWLEDGEMENTS xv 1. INTRODUCTION ± GENERAL x BACKGROUND Internal Wave Theory and Experiment 3 Field Observations 7 Wave-Induced Bottom Sediment Motion ,Q 2. EXPERIMENTAL SYSTEM 13 APPARATUS 13 Wave Tank and Mechanical Components .... 13 Wave Generating Apparatus 15 Power unit 15 Wave maker 17 DATA ACQUISITION 17 Detection of Wave Motion 17 Temperature 19 Specific Gravity 21 Wave- Maker Motion 2i Photographic Measurement 23 Data Recording and Display 26 EXPERIMENTAL PROCEDURE 2 8 3. DATA ANALYSIS 31 TEMPERATURE 31 MEAN DENSITY GRADIENT 31 DENSITY PERTURBATIONS 34 VELOCITIES AND BOTTOM SHEAR STRESS. ... 3? MISCELLANEOUS 41 TABLE OF CONTENTS (Cont) 4. EXPERIMENTAL RESULTS 45 GENERAL DISCUSSION 45 THEORETICAL DISCUSSION 49 Interior Wave Field 50 Boundary Layer 53 INTERIOR WAVE FIELD - RESULTS 55 Subcritical Case 55 Wave number 56 Wave amplitude 64 Critical and Supercritical Cases 77 Wave number 80 Wave amplitude 88 BOUNDARY ACTIVITY 9 5 Case A: y < c 95 Qualitative results 95 Quantitative results 107 Cases B and C: y = c and y > c 128 Qualitative 128 Quantitative 134 DISCUSSION 150 5. SEDIMENT MOVEMENT BY INTERNAL WAVES ... 151 GENERAL DISCUSSION 151 Oceanic Conditions 152 INCIPIENT MOTION CRITERIA 165 MODEL ANALYSIS AND SPECULATION 183 REFERENCES 197 APPENDIX A 208 APPENDIX B 213 BIOGRAPHY 226 vi LIST OF SYMBOLS A amplitude of the stream function a wave amplitude a input wave amplitude a measured wave amplitude m a fluid acceleration af effective fluid acceleration at level of effective fluid velocity c slope of internal wave characteristic D sediment diameter in Chapter 5 D mean sediment diameter m F_, fluid resistance force F F^ hydrodynamic drag F„ gravity force FT hydrodynamic lift Fp pressure force F M virtual or apparent mass force g gravitational acceleration h local fluid depth over slope h fluid depth over horizontal bottom k horizontal wave number k average local horizontal wave number, theoretical k measured horizontal wave number m N Brunt-Vaisala frequency n mode number in text; index of refraction in Appendix A p pressure R Reynolds number Rn particle Reynolds number Rw wave Reynolds number Rc boundary-layer Reynolds number s specific gravity when used with a subscript; otherwise, specific gravity ratio Vll T internal wave period t time C^ drag coefficient C lift coefficient C„M coefficient of virtual mass U average phase function, theoretical u magnitude of the instantaneous horizontal fluid velocity, Chapter 4; magnitude of the instantaneous velocity in boundary layer, parallel to the bottom, Chapter 5 v magnitude of the velocity of sediment particle w magnitude of the instantaneous vertical fluid velocity, Chapter 4; magnitude of the instanteous fluid velocity in the boundary layer, normal to the bottom x,z horizontal and vertical coordinates, respectively x , z absolute values of x,z, respectively x total horizontal projection of the slope a slope angle, degrees |3 angle between fluid resistance forces F„ and a line parallel to the slope bottom slope in radians y 6 boundary layer thickness 6 measured estimate of boundary layer thickness e ak; Stokes parameter e a k ; measured estimate of Stokes parameter m m m ^ 6 phase angle k bed roughness expressed as a particle diameter ix molecular viscosity v viscosity i/ kinematic viscosity 7, "rj coordinates along and normal to the slope p density Vlll p mean density; also amplitude of density over the flat bottom, theoretical p' perturbation density, theoretical p perturbation density computed from linear, inviscid solutions for progressive internal waves on a slope a electrical conductivity t shear stress t shear stress at the bed o t critical shear stress; i.e., that shear stress necessary to initiate sediment movement angle of repose of sediment particle ip stream function ip total stream function in boundary layer i/jT interior stream function ip. boundary -layer stream function co internal wave frequency to) critical internal wave frequency co-, semi-diurnal frequency b. d . subscript referring to boundary layer f subscript "f,T referring to effective kinematic fluid property or to physical property of fluid m quantity determined by experimental measurements max maximum value of quantity o subscript usually denotes those quantities pertinent to the input region; i. e., the portion of the wave tank with horizontal bottom s subscript Ms" referring to physical property of sediment T subscript referring to total value; i.e., interior value plus boundary layer value complex conjugate absolute value of quantity IX Fig. 1 Fig. 2 Fig. 3 Fig. 4 Fig. 5 Fig. 6 Fig. 7 Fig. 8 Fig. 9 Fig. 10 Fig. 11 Fig. 12 Fig. 13a Fig. 13b Fig. 14 Fig. 15 Fig. 16 Fig. 17 Fig. 18 Fig. 19 Fig. 20 Fig. 21 Fig. 22 Fig. 23 Fig. 24 LIST OF FIGURES Page Wave tank and filling system 14 Wave generating apparatus 16 Wave maker in actual location 18 Conductivity probe and probe assembly . . . . 2 0 Refractometer and mini-siphon 2 2 Shadowgraph system 2 5 Data acquisition system 27 Typical temperature - depth profile 32 Typical specific gravity - depth profile .... 33 Conductivity and wave-maker data 3 6 Sample per iodogram 38 Sample work sheet of dye-streak displacements . . 4 0 Motion of neutrally buoyant particles over the flat bottom 42 Motion of neutrally buoyant particles over the slope . 44 Classification of internal waves on a slope . . 4 6 Geometry and coordinate system 48 Upslope variation of wave number, y < c . . . . 57 Upslope variation of wave number, y < c . . . . 58 Upslope variation of wave number, y < c . . . . 6 0 Measured versus computed values of wave number, y < C 61 Measured versus computed values of wave number, y < C 62 Measured versus computed values of wave number, y < C 63 Measured versus computed values of wave number, y < C 65 Measured versus computed values of wave amplitude, y < C 67 Measured versus computed values of wave amplitude, y < C 68 XI LIST OF FIGURES (Cont) Page Fig. 25 Fig. 26 Fig. 27 Fig. 28 Fig. 29 Fig. 30 Fig. 31 Fig. 32 Fig. 33 Fig. 34 Fig. 35 Fig. 36 Fig. 37 Fig. 38 Fig. 39 Fig. 40 Fig. 41 Fig. 42 through Fig. 49 Fig. 50 Fig. 51 Fig. 52 Fig. 53 Fig. 54 Fig. 55 Fig. 56 > r < c > r < c > r < c > r < c > r < c > r < c Upslope variation of wave amplitude Upslope variation of wave amplitude Upslope variation of wave amplitude Upslope variation of wave amplitude Upslope variation of wave amplitude Upslope variation of wave amplitude Reflection of characteristics from a slope, y > c Upslope variation of wave number, y > c . Upslope variation of wave number, y > c . Upslope variation of wave number on log-log plot, y > c Upslope variation of wave number on log-log plot, y > c Wave number versus magnitude of zc/cxc, y > c Upslope variations of wave number, y = c Wave amplitude versus magnitude of zc/cxc, y> c . Wave amplitude versus magnitude of zc/cxc, y > c Upslope variation of wave amplitude, y > c . Upslope variation of wave amplitude, y = c Photographs of shadowgraph images of internal waves over a slope, y < c Diagram of initial vortex advancing upslope Position of breaking Theoretical increase in velocity components along the slope Estimates of boundary layer thickness and Reynolds numbers, y < c Vertical distribution of velocity near the bottom at station l,y c 129 Fig. 65 Onset of vortex motions as a function of wave frequency 13 5 Fig. 66 Maximum size of vortex diameters as a function of frequency 136 Fig. 67 Density-depth profile through streamers and layers . 138 Fig. 68 Average net velocities in streamers and intervening layers 139 Fig. 69 Estimates of boundary layer thickness and Reynolds numbers, y = c 141 Fig. 70 Estimates of boundary layer thickness and Reynolds numbers, y>c 142 Fig. 71 Velocity-time plots at selected level near the bottom, y = C 144 Fig. 72 Velocity and bed shear stress at five stations, y - c . 14 5 Fig. 73a Vertical distribution of velocity near the bottom at station 1, y > c 147 Fig. 73b Vertical distribution of velocity near the bottom at station 4, y>c 148 xm LIST OF FIGURES (Cont) Page Fig. 74 Velocity and bed shear stress at selected levels near the bottom, y > c 149 Fig. 75 Frequency spectra of horizontal kinetic energy density at site "D" 154 Fig. 76 Bathymetry off southern New England . . . . 155 Fig. 77 Internal wave measurements on continental margins . 156 Fig. 78 Idealized cross-section of continental margin . . . 158 Fig. 79 Characteristic slope and average bottom slope compared for various a> and N 16 0 Fig. 80 Definition sketch of sediment particles . . . . 166 Fig. 81 Resistance coefficients Cn for spheres .... 175 Fig. 82 Roughness effects on the resistance coefficients for spheres 176 Fig. 83 D./6 versus h/h 178 Fig. 84 Shields diagram 181 Fig. 85 Hypothetical size- frequency distribution of sediments on an oceanic slope 185 Fig. 86 Continental shelf model; high frequency waves . . 18 8 Fig. 87 Continental shelf model; tidal frequency waves . . 19 0 Fig. 88 Continental slope model; high frequency waves . . 192 Fig. 89 Continental rise model; tidal frequency waves . . . 193 Fig. Al Circuit schematic for conductivity sensor. . . . 210 XIV ACKNOWLEDGEMENTS I am deeply indebted to Professor John Southard and to Professor Carl Wunsch for their expert guidance throughout this study. Without their continuing technical and moral support this work would not have been pos- sible. Their patience to listen and their willingness to comment will not be easily forgotten. I would also like to thank Dr. Nicholas Fofonoff and Dr. Charles Hollister for their critical reviews of portions of this work; their comments aided the preparation of the final product. I gratefully acknowledge the assistance of Dr. Seelye Martin who provided the original circuitry for the conductivity probes and gave several helpful suggestions concerning the experimental techniques. Several persons provided invaluable assistance during the several phases of this study. I greatly appreciate the skilled technical assistance given by Mr. John Annese of the departmental machine shop during the con- struction of the experimental apparatus. The variety of computations and other tasks that were accomplished expeditiously by Mr. David Drummond during the analysis of data, and the fine illustrations which were provided by Mr. Valentin Livada were indispensable to the preparation of this work. I thank Mrs. Caroline Peterson for her careful typing of the final draft. Of course, without my wife, Virginia, there would not have been a final product. I am grateful to the Education Development Center in Newton, Massachusetts, for donating the glass tank used during the experiments. This work was supported by the Office of Naval Research under Contract No. Nonr 1841 (74). xv 1. INTRODUCTION GENERAL The purpose of this research is threefold: (1) to provide experimental results that describe the changing character of single-frequency, small -amplitude internal gravity waves as these waves propagate over a sloping bottom; (2) to compare these results with recent theoretical solutions for those conditions which are theoretically tractable, and to examine the phy- sical characteristics of those conditions which are difficult to treat theo- retically; (3) to apply the theoretical and experimental results toward the development of a simple analytical model that prescribes the conditions for the initiation of bottom sediment motion induced by shoaling internal gravity waves. The motivation for this work was rooted in speculation - a starting point not uncommon in experimental research. The speculation was stimu- lated by the results of previous work that had indicated the possible signifi- cance of internal waves as geological agents and by recent theoretical solutions that predict the intensification of internal wave motion along sloping bottoms. It was decided early in this research that a clearer physical under- standing of the shoaling process for internal waves was needed - particuarly knowledge of the nature of the bottom boundary layer - before the two-phase problem could be considered. The effects of internal waves on sediment particles of various shapes, sizes, and densities were investigated during a preliminary set of experiments. It was concluded from these kinds of experiments that the reliability of the results and the range of test condi- tions were limited by the experimental difficulties, mainly the difficulty of attaining large velocities in the small experimental system. The approach to this problem was changed to investigate the movement of sediment by in- ternal waves analytically, based on the results of a detailed laboratory study of shoaling internal waves over a smooth sloping bottom. Recent theoretical analyses had cast a framework for the experimental design (Wunsch, 1969; Keller and Mow, 1969); preliminary laboratory work had established the experimental techniques. Earlier experiments by other researchers had demonstrated the feasibility of some of the techniques (Mowbray and Rarity, 1967; Martin, Simmons, and Wunsch, 1969; Gibson and Schwarz, 1963), and had shown several stimulating results (Thorpe, 1966). The subsequent work is divided into three sections: (1) the experimental design, instruments, and techniques are dis- cussed in Chapters 2 and 3; (2) the results of the experiments and their comparison with theo- retical solutions of Wunsch (1969) are presented in Chapter 4. The results are separated into two basic types: experiments with high-frequency waves and experiments with low-frequency waves. The interior wave field and the boundary layer are discussed separately for each; (3) in Chapter 5 the balance of forces acting on a bed particle is used to derive a simple analytical model for the stability of the particle beneath a train of high-frequency, shoaling internal waves. The approach closely follows that used by Ippen and Eagleson (1955) and Eagleson and Dean (1959) in the development of an analytical model for sediment motion * The distinction between high and low frequency waves is discussed in more detail in Chapter 4. Basically the incident energy of the high fre- quency waves propagates upslope to the corner; i.e., toward the intersec- tion of the slope and the free surface, with no reflection. By contrast, the incident energy of the low frequency waves theoretically is back reflected from the slope. induced by surface waves. The applicability and limitations of the analysis of internal waves and sediment movement are discussed in Chapter 5, and some speculations are offered concerning the effects of breaking of internal waves and the action of boundary layer instabilities on bottom sediment in the ocean. BACKGROUND Internal Wave Theory and Experiment In two recent papers, Wunsch (1968, 1969) has developed normal- mode solutions for both standing and progressive internal waves in a wedge geometry. The solutions of interest here are for progressive waves pro- pagating into the wedge, i.e., over a linearly shoaling bottom. This theory assumes two-dimensional motion in a stably stratified, inviscid, Boussinesq ,2 e dp0 fluid with constant Brunt- Vaisala frequency N (N = — — - — ). The geometry and coordinate system that Wunsch considered is given in Figure 1. Wunsch's solutions are discussed in more detail in Chapter 4; these solu- tions are presented for comparison with the experimental results. One of the principal features in the solutions is the suggestion of a strong intensi- fication of velocity components along the sloping boundary. Wunsch (1969) conducted a simple experiment in which internal waves were propagated over a linearly sloping, rigid bottom. From photographic determinations of the changes in wave length up the slope, he showed that the results are in agreement with his solutions. The wave length was approximately pro- portional to the depth. Magaard (1962) presented solutions for a similar problem of standing internal waves except for the density distribution and shape of the bottom profile. He considered a density distribution that can be expressed analytically as: p(z) = c2 (2z + Cl)1/2 (1-1) where c-, c„ are constants. The bottom profile that he chose for computational facility and some realism was: h(x) = i (2bx)' (1-2) where c1 and b are constants. He developed a linearized characteristic equation for the velocity field and found that the equation was separable in the spatial coordinates only for the case of constant depth. He also showed that for a bottom slope y less than the slope of the characteristics c, a solution is possible which gives waves that have increasing amplitudes and decreasing wave lengths with depth. Magaard (1962) also found an appre- ciable intensification of the velocity components along the sloping bottom. Robinson (1970) has recently considered the effects of a corner (such as the point of transition between slope and flat bottom) upon an in- ternal wave train propagating toward the slope in an infinite medium (i.e., with free surface infinitely far from the horizontal bottom). Robinson provided solutions that consist of incident and reflected waves such that the reflected waves do not violate the radiation condition; i.e., they do not introduce energy from the far field (at x = +*>). Keller and Mow (1969) obtained an asymptotic solution to the problem of internal wave propagation in a horizontally stratified fluid of nonuniform depth by application of the principles of geometrical optics. Their general solutions apply to the linearized equations of motion, and the wave amplitude and fluid depth are assumed to vary very little over horizontal distances small compared to a wave length. They gave a partic- ular example in which N was assumed constant and the bottom was a small, linear slope. The solutions in this case show that the wave length is proportional to the depth, and the wave amplitude increases linearly with decreasing depth. Keller and Mow noted that their solutions also are in agreement with the experimental results given by Wunsch (1969). Hogg and Wunsch (1970; also personal communication) derived similar asymptotic solutions using a WKB approach with a two parameter expansion scheme. Their first order, linear solutions for progressive internal waves over a small, linear slope are equivalent to those of Keller and Mow (1969); both solutions are compatible with the normal mode solutions of Wunsch (1969) for small slopes (y«c). Hogg and Wunsch ex- tended their analysis to higher-order terms in the stream function and developed expressions for the wave-induced radiation stress. They showed that a significant set-up of the isopycnals induced by internal waves is pos- sible and demonstrated theoretically the plausibility of "longslope" currents for a line of breaking internal waves oblique to the bottom contours. From the higher order terms Hogg and Wunsch (1970) were also able to demon- strate that in the absence of mixing the net Lagrangian motion of water particles in a two-dimensional stably stratified fluid is necessarily zero for internal wave motion in the proximity of a sloping, rigid boundary. Fofonoff (1969) also examined analytically the unsteady motion in a horizontally stratified, uniformly rotating ocean. He considered the seaward propagation of an internal baroclinic wave that was generated at the continental shelf edge by the diurnal tide. By exam- ining the characteristics of the solution, he presented a numerical example in which the values of density, latitude, and depth were those measured at site "D" (39° 20'N, 70° W), located about 37 miles south of the 200 meter contour. Fofonoff (1966) showed that north of this location the bottom profile of the continental slope and the path of the characteris- tic were approximately coincident. Again there is implication of increased bottom shear along the slope. Sandstrom (1966) derived the reflection properties of internal waves from a sloping, rigid boundary. Both he and Robinson (1970) have noted that the normal- mode solutions derived by Sandstrom from his ray theory solutions for progressive internal waves propagating over a slope violate the radiation condition (i.e., the complete solution requires an anomalous set of reflected waves that introduce energy from the far field). Sandstrom predicted that upon reflection the waves change their wave number and amplitude; he found that the magnitude of the changes is dependent on a ratio equivalent to y /c. Sandstrom (1966) also conducted a set of experiments in which internal waves were generated by a flap-type wave maker and allowed to propagate over a planar, inclined bottom. His experimental results for a gently sloping bottom (y «c) showed qualitatively two features that were also found in the present study, and are discussed quantitively in Chapter 4: (1) amplification of internal waves traveling from "deep" into "shallow" water, and (2) intensification of motion near the slope. Sandstrom (1966) noted, however, that unfortunately the partic- ular experiment for which these features were observed was complicated by the presence of a surface mode that was also generated by the wave maker. Although his apparatus was apparently capable of generating the first internal mode, there are no experimental results in his work which indicate that he did this. Sandstrom discussed solutions for the case y >c (steep bottom slope), but did not present experimental results for this case. He concluded from the linear theory that in the case y ^ c, the motion close to the bottom becomes very large. Longuet-Higgins (1968) extended the discussion of reflection prop- erties of internal waves from rigid surfaces to account for various types of bottom roughnesses. He has shown that the important parameter is not the ratio of the scale of bottom roughness to wavelength but instead the ratio of the scale of roughness to the thickness of the oscillatory boundary layer. He pointed out that this suggests that small-scale irregularities can severly affect the transmittance and reflection conditions of the bottom. In a very thorough theoretical and experimental treatment of finite- amplitude effects on interfacial and internal waves in a horizontal channel, Thorpe (1968) also reported some qualitative observations of progressive internal waves on a slope. In his experimental arrangement, a linearly sloping, smooth bottom was mainly employed as a wave absorber. How- ever, during the course of the experimental runs for internal waves in a linearly stratified salt solution he noted several interesting features over the slope. These observations are summarized here as a preface to the results shown later in this study: (1) the wave length decreased upslope; (2) a large runup was apparent along the upper slope region, and the larger amplitude waves produced a larger runup; (3) the free surface remained undisturbed; (4) a "rotor" was observed near the free surface on one occasion; (5) breaking ("overturning") was observed well upslope on one occasion; (6) no reflections could be detected. Thorpe (1968) showed three photographs of the waves over the slope. His method of flow visualization, involving alternating dyed and undyed layers, did not permit detailed observations of the breaking or mixing, but his qualitative observations are quite instructive and essentially agree with the results shown later. It should also be mentioned here that excellent experimental studies of instabilities and turbulence in a stably stratified fluid, including breaking of interfacial waves, have been carried out by Pao (1968). Field Observations Two kinds of field observations are pertinent to this study: (1) measurements that suggest shoreward propagation of internal waves over the continental margin, and (2) observations that indicate, directly or in- directly, the motion of bottom sediment due to passage of internal waves. Measurements of the first kind have been presented by Lee (1961), Gaul (1961) and Ufford (1947), among others. Their investigations presented time-series measurements of temperature fluctuations taken simultaneously at three stations oriented in various triangular arrays in shallow water (depth <200 feet). The wide geographical distribution of these studies illus- trates the ubiquity of internal motions on the continental shelves. Common internal-wave periods were on the order of 5 to 20 minutes; wave propaga- tion was shoreward with speeds of 0.3 to 1.2 knots. Lafond (1962) has provided similar values. Longer period internal oscillations are described by Lee (1961), Boston (1964), and Summers and Emery (1963). The last study illustrated the refraction of internal waves of semidiurnal period as they propagated shoreward over the continental slope and shelf off Southern California. The estimated wave speed was 7 knots in deep water and slightly less than 1 knot over the shelf. Isolated asymmetrical internal temperature disturbances resembling the idealized solitary wave profile of surface waves (Ippen, 1966) have also been detected by Lee (1961), Gaul (1961), and Cairns (1967). The common period of this sort of disturbance seems to be approximately 10 minutes, and the propagation direction is again shoreward. Curves of spectral kinetic energy density versus frequency are available for velocity measurements of internal motion taken with vertical arrays of current meters at site "D" (Fofonoff, 1968). An example of these spectra is shown in Chapter 5. Direct observations which reliably link the two motions of internal waves and sediment movement are limited to a few isolated instances. This lack of evidence might suggest both the difficulty and the insufficiency of this kind of field measurement. Ideally, one would like to measure simul- taneously the internal wave velocity structure and the movement of bottom sediment beneath the wave motion. Either measurement alone is a challenge in the field. Lafond (1965), using underwater television to observe the sedi- ment motion and a vertical array of temperature sensors to measure the internal wave motion, found that the movement of sediment ripples in shal- low water (depth about 60 m) was correlated with large fluctuations in the temperature structure. He interpreted this as evidence for sediment trans- port by internal waves. Indirect field data that connect the two motions are similarly limited. Revelle (1939) inferred the connection in an attempt to explain an observed pattern of sediment distribution. Munk (1941) calculated the theoretical standing-wave characteristics in the Gulf of California and found that the nodal-antinodal separation coincided favorably with Revelle's (1939) ob- served distribution of sediment sizes. That is, the smaller sizes were noticeably lacking beneath the nodes, where larger instantaneous velocities are expected. However, in a subsequent detailed study of the sediment dis- tribution in the Gulf of California Van Andel showed that the sediment pat- tern described earlier by Revelle was not apparent in the later data. Emery (1956; 1960) suggested that deep, standing internal waves in certain California basins stir the bottom sediments and might produce the observed local concentrations of coarse sediments at the basin sills. More recent investigations of the large submarine sand waves (av- erage height =rf 8 meters) near the edge of the continental shelf southwest of Great Britain (water depths 180 meters) by Stride and Cartwright (1958), Cartwright (1959), Stride and Tucker (1960), and Carruthers (1963) indicate the possibility of interaction between internal waves and bottom sediment. Cartwright (1959) formulated a theory which suggests that these sand waves are maintained by a stationary internal- wave system of tidal period that exists during the presence of the local seasonal thermocline. Measure- ments of oscillations of the scattering layer (Stride and Tucker, 1960) and towed thermistor records (Carruthers, 1963) support this theory. Car- ruthers (1963) also measured the currents close to the crest of the sand waves at one-hour intervals over a seventeen- hour period. A plot of cur- rent vectors indicated that the predominant flow direction was normal to the crests, with a maximum current of 0.5 knot in this direction. Inter- estingly, the crests of the sand waves are parallel and adjacent to the shelf edge. Many investigators have studied the observed sediment distribution on the continental shelf and slope off the East Coast of the United States. Uchupi (1963), Emery (1966), and Schlee (unpublished manuscript) have summarized much of this work. Sand waves and local topographic highs 10 are abundant on the continental shelf. The type and size distribution of sediment varies both across and along the bottom contours. The seaward variations are generally sharper. For instance, Uchupi (1963) delineated four broad zones parallel to the shelf edge from Cape Cod to Hudson Can- yon. He noted that the most seaward zone (about 70 kilometers wide) con- sists of silty sand, sandy silt, and silt (water depth 60 to 135 meters). Uchupi (1963) claimed that from Hudson Canyon to Cape Hatteras the outer- most zone is composed of fine sand of approximately uniform grain size. The surface of the continental slope in these regions consists mainly of silt and clay; however, sand, silty sand, and gravel line the floors of the submarine canyons that cut the slope. Wave-Induced Bottom Sediment Motion The writer could find no published results on wave-induced bottom sediment motion in a stratified fluid. However, there are many published studies of sediment motion beneath surface waves in a homogeneous fluid. Most of these studies are empirical or semi-empirical. The obvious lack of theoretical treatment probably stems from the complex form of the gov- erning equations of motion, which are not well established. One appealing approach to a laboratory and analytical study of the problem was undertaken by Ippen and Eagleson (1955) and later extended by Eagleson, et al (1958), and Eagleson and Dean (1959). A statistical treat- ment of laboratory data that involved the motion of discrete spherical sedi- ment particles due to shoaling surface waves supplemented their theoretical presentation. Many of the complicating secondary factors, such as (1) nonuniform bed roughness, (2) shape variations of the particles, (3) mu- tual particle interactions, (4) nonuniform incident wave trains, (5) local channelization and fluidization of the bed, and (6) bed permeability were not included in their work. They were able to derive a force-balance equation for the individual particles. Ippen and Eagleson (1955) described two possible types of particle motion: incipient and equilibrium. Incipient sediment motion was defined by Eagleson et al (1958) as "an instantaneous 11 condition reached when the resultant of all active forces on the particle intersects the line connecting the bed particle contact points. " They also defined established sediment motion as "an oscillatory or quasi-oscillatory condition of motion reached when for some portion of each wave cycle the sum of the instantaneous active forces is greater than that value necessary to initiate motion. " The major difficulties in working with these definitions are: (1) the point of application of some of the individual forces is usually not known; (2) the various coefficients, such as those for drag, lift, and virtual mass, are not well defined for unsteady motion; (3) usually the effective fluid velocities acting on the bed particle must be estimated from an approximate theoretical equation, since the exact velocity profile near the rough bed in unsteady motion is not known. Despite these difficulties, Eagleson and Dean (1959) were able to arrive at a semi-empirical relation- ship for particle size, beach slope, and local wave characteristics. Eagle- son et al (1958) compared the analytical results with field data from eight stations on the Atlantic coast of the United States. As these authors put it, the agreement was "fortuitously" remarkable. The concept of an equilibrium diameter for given wave conditions was first proposed by Cornaglia (Miller and Zeigler, 1964) and has since been labeled the null-point theory. His arguments, as well as those of Ippen and Eagleson (1955) and Miller and Zeigler (1964), are basically simple. In an area of shoaling waves, every bottom sediment particle is affected by the existing sea state. For any given sea state and bottom slope, there will be one size class of particles in a state of equilibrium motion. That is, this size class will move back and forth along the bottom as successive troughs and crests pass, but will oscillate about some mean position. In general, particles larger than the null-point size will move offshore and those smaller will move onshore. Miller and Zeigler (1964) noted that in the field the null point should be regarded statistically, where- as in the laboratory it is convenient to regard it as single-valued. Their field study also indicated the limitations and the modified applicability of this theory in nature. However, at the same time their work showed that 12 the theory gives a satisfactory prediction of the gross sediment distribution of a nearshore sand bottom. Other laboratory investigations have produced relationships between the initiation of motion of a given particle size and the associated wave con- ditions (Manohar, 1956; Goddet, 1960; Vincent, 1958). These investigations were all conducted in a channel with a level bottom. Vincent (1958) observed that the fluid velocity just outside the boundary layer was almost constant at the onset of movement of equal- diameter grains. Abou-Seida (1965) noted that field measurements of radioactive sand tracers conducted by Sato et al (1963) compared favorably with the laboratory relationships de- duced by Manohar (1956) and Goddet (1960). A second approach to this problem is relatively new. Arguing from basic physical principles and certain assumptions, Bagnold (1963, 1966) derived a relationship between bed load transport and the dissipation of energy in the fluid motion. His argument rests on the assumption that a distribution of dispersive and tangential shear stresses exists in a bed of granular material that is subjected to shear deformation. Several sum- maries of the topic of wave-induced sediment motion are available (Inman, 1963; Abou-Seida, 1965; Raudkivi, 1967). Such presentations of incipient sediment motion usually include a discussion of the various criteria developed for open channel flow. In particular, the work of Shields (1936), White (1940), and Kalinske (1947) is often cited. 13 2. EXPERIMENTAL SYSTEM APPARATUS Wave Tank and Mechanical Components The rectangular wave tank illustrated in Figure 1 has an open top and 3/8-inch glass side walls and bottom. It is 4. 9 m long, 20. 5 cm wide, and 38. 4 cm deep and consists of two sections of similar dimensions joined end to end by an O-ring seal contained between plastic inserts attached to the sections with epoxy cement. A thin layer of clear silicone sealant (Dow Corning 701 Building Sealant) was applied along the inside seam to form a smooth transition between the two sections. All other joints were also formed with silicone sealant. The ends of the channel are acrylic plastic plates sealed by Neoprene gaskets. A 14-foot length of aluminum angle attached to five vertical aluminum angle supports behind the tank and 3/4 of an inch above the side wall served as an instrument mount. A filling technique similar to that described by others (Fortuin, 1960; Oster, 1965; Jaffee, 1968) was designed to establish the desired sal- inity gradients in the wave channel. Figure 1 shows a diagram of the filling system. Owing to the corrosive properties of salt water, the insides of the two 55-gallon containers were coated with epoxy paint, and all seams were caulked with silicone sealant. Wherever the liquid came into contact with bare metal, only stainless steel or brass was used. Flexible Tygon tubing is used throughout the filling system wherever tubing was required. The filling rate was controlled by a brass needle valve. Each drum was equipped with a Tygon sight tube for visual determination of the level in each barrel. A motor driven stainless stirring rod with two brass stir- ring vanes was mounted above the fresh water barrel. To prevent solid- body rotation of the water, the stirring rod lay off the axis of the barrel. The bottom slopes, one of which is shown in position in Figure 1, were cut from Plexiglas sheets. Each slope is beveled at one end to form 14 -Q -i g rH -p CD •H (d -P Cn MH £ (d C £ •H C C £ P 0 0 en -P p .c tn Q) rH ■H a >i M «J -P -H Eh <*h W en W vo fd ,£} CO 0> r- r- ^ Ti iH en W C n rd •H 0 P -H o) x & a x > o) a, o a) n)H3 (D rH £ ft to S ft h n nt in rd O (0 O -P -P • 0 e c CD — -P V) >1 e cn u Cn 00 C ro •H ■H X •H e HH u T3 o c CN rd X ,* c e (0 -p CT> • rd •^ w C • 0 rH •H cn • C tn CD •H e Pm •H T) .* C rd 4-> Q) > rd £ 15 a smooth junction with the bottom of the tank. Four slope angles were used, 7, 15, 30, and 45 degrees. During the experiments the slopes were supported by two parallel Plexiglas struts. These struts were positioned at the proper angle along the inner walls of the tank. Four Lexan plastic rods with threaded ends and plastic nuts were used to separate the struts and to maintain their position by forcing them against the channel walls. Neoprene strips were cemented to the outer surfaces of each strut to form a continuous and flexible seal with the side walls of the channel. All materials forming the beach slopes and supports were noncorrodible. Wave Generating Apparatus The wave generator had to meet five primary requirements: (1) smooth and sufficiently powerful motion; (2) desired frequency range and control; (3) variable amplitude; (4) compact size with components stably mounted on a plate above one end of the channel; (5) long running with low vibration level. In an effort to meet these requirements the apparatus was modified several times. Availability of various components such as the variable speed drive often aided the choice of parts. The end product was an outgrowth of the usual experimenter's dilemna: getting optimum performance at reasonable cost. Power unit. Figure 2 is a diagram of the power drive for the wave maker. The entire drive mechanism is rigidly supported above one end of the chan- nel by four 1-1/2-inch pipe legs with foot flanges. The base is a 1/4-inch steel plate. The assembly consists of a Graham variable-speed drive coupled to a 1/4-hp motor. The output shaft is joined to a fixed- ratio gear reducer (9.5:1) via a flexible coupling. This second speed reducer is con- nected to a Scotch yoke device that converts the rotary shaft movement into rectilinear sinusoidal motion of a vertical shaft. 16 U ^ CD u CH TS CD 0 •• 3 U -P •dm ,G -P rd CD w • -P -h Q) g T5 CO cr> -H rH t7 0 C -rH — ' 2: CL •H -l -P 10 U g rd O M i tP,Q TS CVH M T3 <1) C f0 -H C £ rd g m x: -h -p O rd >i CO g S-l c U -P u > 0 U CD U O TS 0 rd o -h rd g O t^tf) M f0 o M u B S • • • • • • • • in v>d r-> oo CO -P rd u (d &. a c •H -P rd u c CD tr dJ s 5 (N tn •H fa 17 Wave maker. Figure 3 shows the wave maker in actual location. Several additional considerations were involved in the design of this device. First, a unit was needed that could be inserted into position easily without appre- ciably disturbing the stratification after the tank had been filled. Secondly, the driving motion had to generate, to a good approximation, first- mode internal waves while minimizing mixing at the wave maker. In addition, the water motion in the section behind the wave maker had to be prevented from interacting with the fluid in the working section of the channel. Basically, the device consists of a plastic flapper plate free to pivot about its midpoint. Hinges anchor this plate to a supporting frame. A thin flexible rubber gasket covers the entire face of the plate and frame, thus preventing liquid from flowing between the two sections. This gasket is held in position by thin plastic covers cut to the same shape and dimemensions as the frame and flapper plate. The flapper plate is driven by a stainless - steel push rod connected to the vertical shaft of the Scotch yoke by a universal coupling that allows for minor misalignments between shaft and flapper plate. A closed- cell Neoprene strip was cemented to the sides and bottom of the frame. This afforded an adequate seal between the wave flapper assembly and the glass tank. A test showed that the wave maker could suc- cessfully maintain a head difference of at least 6 cm of water between the working section of the channel and the unused portion. DATA ACQUISITION This section describes the instruments used to record the experi- mental data. A more detailed discussion of the quality of the measurements, including calibration curves, error, and reliability estimates, is presented in the next chapter and in Appendix A. Detection of Wave Motion The changing properties of the internal waves as they propagated upslope were measured in the Eulerian sense with conductivity probes. 18 Fig. 3. Wave maker in actual location. Tank depth is 3 8.2 cm. 19 The basic feature of the water medium was its strong salinity stratification. Also, over the range of salinity values used during this study, salinity is approximately linearly proportional to conductivity. The temperature vari- ations in the water were found to be small (usually less than 0. 5°C over the total depth) except in the upper two centimeters, so that below this layer the total change of conductivity at a point is here a valid estimate of the total change of salinity: S = SQ + £ (a - aQ) (2-1) Da 1 DS Dt " v Dt E where Yj in %o/(ohm -cm)" S in%0 a in ohm cm Because the wave-induced fluctuations of conductivity are measured relative to some basic state of rest, only relative changes in conductivity at a point are needed to decode the waves. The conductivity sensor is a spherical platinum electrode that is the active arm of an A. C. Wheatstone bridge. Design and construction of the probe was inspired by Gibson and Schwartz (1963). Figure 4 shows a probe assembly, including electrode, glass housing, and connecting cable. A de- tailed description of its construction and an analysis of its performance are presented in Appendix A. Temperature Temperature of the room air and the water in the filling barrels was measured with a mercury-in-glass thermometer which can be read to 0.05°C. Vertical and horizontal temperature profiles of the stratified 20 5mm inner wire of coax spot weld 3 mm ////////////////r 5.3 cm glass annealed to Pt wire 0.3mm Fig. 4. Upper: Photograph of conductivity probe assembly and circuit box. Lower: Construction of conductivity probe tip. 21 water column were taken with a glass -coated thermistor bead that is the active arm in a D.C. Wheatstone bridge. Specific Gravity Specific gravity measurements of the water in the filling barrels were made with a lead- weighted glass bulb hydrometer. The hydrometer can be read to 0.0005 for any specific gravity in the range from 1. 000 to 1. 070. Its accuracy is about 0. 2 percent of the maximum value in the range. The basic specific gravity gradient was determined from measure- ments of the vertical distribution of index of refraction. Figure 5 shows the refractometer and mini- siphon that were used to obtain index of re- fraction at various depths in the stratified water. These measurements were converted to specific gravity values from calibration curves. The siphon consists of a 0.5 mm diameter stainless steel tube supported verti- cally by a brass bushing and two locking screws. The bushing is press - fitted into an aluminum bar. Centering pins allow repeatable positioning of the aluminum bar athwart the channel. Markings that are evenly spaced one centimeter apart are inscribed on the steel siphon for vertical positioning. Small-diameter Tygon tubing fits over one end of the siphon tube. A brass pin that can be inserted into the open end of the Tygon tubing is used to stop the flow. Flow rate through the siphon at mid- depth (about 16 cm of water) is about 0.2m? /sec. Wave -Maker Motion A rotary potentiometer in a D.C. voltage -divider circuit transduces the oscillating, rectilinear motion of the Scotch yoke into a fluctuating D.C. voltage signal. A precision gear meshes with a linear rack that is cut into the forward part of the upper vertical shaft of the Scotch yoke (Figure 2). The spur gear is fastened to the shaft of the rotary potentiometer. Vertical movement of the Scotch yoke causes a change in the voltage output of the sensing circuit. The sensitivity is approximately 0. 1 volt /cm. This tech- nique establishes a convenient and accurate frequency reference and time 22 Fig. 5. Refractometer and mini-siphon 23 base for the conductivity probe measurements. The accuracy of the circuit as a frequency reference is primarily dependent upon the linearity of the potentiometer. In this case, the linearity of the rotary pot is 0. 1 percent (Giannini Control Corporation). Amplitude calibration of the signal also provides a measure of the forcing amplitude of the wave flapper. The period of the motion is sensed by a microswitch mounted on the lower part of the vertical support plate for the Scotch yoke (Figure 3). This switch is closed once per cycle by a protrusion on the rotating arm, and provides a convenient trigger for a period counter. Photographic Measurement Photographic measurements of the wave motion were obtained by recording on film the behavior of neutrally buoyant particles and various dyes. Expandable polystyrene particles (Sinclair -Koppers Company) were chosen because their density was adjustable to the proper range by a tech- nique described by Bohlen (1969). In addition, their milky white color pro- vides an excellent contrast with the surrounding water when used with a black background and overhead lighting. After processing to achieve the desired density range, Bohlen found that their sizes varied over a consider- able range. Sieving produced a more uniform size distribution, with median size about 0.2 mm. The range of specific gravity for the lot was 0. 990 to 1. 040, with an obvious skew toward the heavier limit. The mean specific gravity was estimated at 1.020. These values were adequate for neutral buoyancy in the range of fluid densities used during this study. The particles were used to study orbital velocities and modal structure. Potassium permanganate (KMnO.) was used to make the vertical dye streaks to determine near-bottom velocities and bottom shear stress. When dropped into a water column, crystals of this substance leave a dis- tinct dark- red dye trace from surface to bottom. The vertical trace was then tracked photographically as it was deformed by wave motion. 24 Small quantities (one to two cc) of common liquid food coloring or a highly nondiffusive blue dye (Blue Dextran 2000) were injected into the fill- ing tube by means of a 16 -gauge hypodermic needle and syringe through a self- sealing rubber membrane at infrequent intervals. The membrane was fitted over one end of a glass T-joint in the Tygon filling line (Figure 1). These injections resulted in thin horizontal dye layers at various levels in the stratified water column. This enabled qualitative measurements of the internal waves by observation of the deformation of the dye layers by the propagating waves. The layers were typically 3 to 5 mm thick for the food dyes and only about 1 to 2 mm thick for the Blue Dextran 2000. The tech- nique also aided in observing flow conditions at the entrance to the main channel during filling. Excessive mixing at this point was indicated by rapid diffusion of the dye, as might be expected in an area of strong turbulence. In such cases, filling rate was decreased. Finally, the dye layers added some aesthetic value to the final stratification. Shadowgraph images were used to display instabilities in the boundary layer and the violent breaking of waves in the upslope turbulent zone. The shadowgraph system is shown diagrammatically in Figure 6. The two-way diffusive screen is sold by Edmund Scientific Company under the trade name Lenscreen. Its excellent dispersive properties and good transmittance pro- duced clear and evenly lighted images. The merits and techniques of shadow- graph flow visualization are discussed by several authors (Barnes, 1954; Goldstein, 1965; Edgerton, 1958). A Bolex 16 mm camera was used to take motion pictures of the shadowgraph images . The shadowgraph lighting was provided by a tungsten filament lamp (Beck Company) that contains a dial rheostat for light-intensity adjustments. It also has a circular portal lens that can be moved toward or away from the light source. This lens and an adjustable diaphragm are located in a collimating tube through which the light rays pass. A 500-watt combination flood- spot lamp provided illumination for the dye streak photographs. 25 4> ■o c C o Q> 1 to Xi a o x; 10 O U •H -P £ cu xi o CO IT (A O 26 Data Recording and Display Figure 7 shows some of the instruments used in this study. The oscilloscope at the left of the picture has provisions for dual-channel opera- tion (Tektronix Type 564 Storage Oscilloscope), and was used mainly to display the output of each conductivity bridge when null was set (Appendix A). The instrument at the right is a Brush dual-pen strip-chart recorder. The two charts operate on the same time base but with independent voltage scales. This recorder provided a real-time display of the transduced conductivity fluctuations during the internal-wave runs. Thermistor temperature mea- surements were also recorded in this way. The heart of the data recording system, a Wang model 2300 digital readout and recording system (Wang Laboratories), is at the center of the photograph in Figure 7. The fastest available sampling rate (1 cycle per second) was chosen for all phases of this investigation. The system was programmed such that each tape record was in card-image format (< 80 characters per record) and represented one second of real time. Included in each record were digital clock time to the nearest second and six sequen- tial channels of data. The data characters contained five measurement digits, decimal point location, and sign; each data point was recorded to the nearest millivolt. The recording format simplified use of the data in the digital computation routines on the IBM 360/67 system and in listing the raw data in decimal format on the IBM 1401. Several cameras were used to photograph the neutrally buoyant par- ticles and dye lines. General photographic work was done with a Nikon F single-lens 35 mm reflex camera. Another Nikon F camera with an auto- matic film advance attachment was used to photograph deformations of the dye streaks, and, in some cases, the motions of neutrally buoyant particles. The film advance was controlled by a synchronous timing device that provides continuous one-second trigger pulses to the camera at an accuracy of ±0. 1 percent of the timing interval. A Calumet view camera with fast Polaroid film was normally used to photograph the neutrally buoyant particles, which 27 Fig. 7. Data acquisition system Left: Tektronix type 564 stroage oscilloscope Center: Wang model 2300 digital data system Right: Brush dual-pen strip chart recorder 28 were illuminated by a flashing stroboscope. The strobe flashes were triggered externally by a signal generator (Wave-Tek). EXPERIMENTAL PROCEDURE Although the procedure for each experiment varied depending upon the purpose, the following basic procedure was employed throughout the prepara- tion and execution of all experiments. Several days (usually three) prior to the experiment, the two 55 -gallon barrels were filled to a predetermined level with hot tap water to which 7 to 10 teaspoons of Alconox laboratory detergent were added. Liberation of dissolved air accompanied the subsequent cooling to room temperature. The detergent inhibited formation of air bubbles along the walls of the drums. The water was also stirred manually as often as pos- sible. On the next day, a predetermined amount of salt was added to one drum. A submersible pump was used to mix the salt into solution and to minimize entrainment of air at the free surface. After the wave tank had been cleaned, the wave flapper assembled and tested, and the slope supports positioned at the desired angle, the siphon connecting the two barrels (Figure 2) was primed, and the stirrer in the fresh water barrel was started. The water temperature in each barrel and the room temperature were then recorded. Readings of specific gravity and index of refraction were taken in both barrels. During filling, the water level was recorded as a function of time, and dye was in- jected into the filling tube at selected intervals. Room and water temperatures were monitored at one -hour intervals. After filling was completed, about 10 hours later, the final temperature and specific gravity of the water in each barrel were measured. The slope was then inserted very slowly and smoothly by sliding it down the supporting struts until its beveled edge lay flush with the tank bottom. Very little mixing was observed, except at the edge, as the slope moved through the stratified fluid. The coherence of the dyed layers near the slope attested to the small degree of disturbance. The wave flapper, whose bottom had been positioned at mid- depth during the filling, was then inserted through the remainder of the fluid. Its driving rod had been fastened to the vertical shaft of the Scotch yoke prior to filling. This procedure also 29 caused very little noticeable mixing. The water depth was recorded, and the top of the tank was covered with a thin plastic sheet (Saran Wrap) to minimize evaporation losses at the free surface. (Prior experience showed that these evaporative losses could be quite substantial, 1-2 mm of water per day.) In addition, the conductivity probes were cleaned and platinized as necessary (Appendix A). Depth markings were checked and refurbished on all probes. On the next day, vertical profiles of temperature and index of re- fraction were taken over the flat bottom and at other stations along the slope. The depth was recorded, and the presence of an interface near the surface was investigated with the shadowgraph apparatus . The plastic covering was removed from the slope section of the tank, and the pertinent photographic apparatus was made ready. Prior to wave generation, the conductivity probes were given static calibration by raising and lowering them vertically by known increments while recording their output at each level; this allowed a comparison with previous calibration values and provided an updated sensitivity coefficient for each probe. Normally, the next group of measurements was taken with the probes all at the same depth over the flat-bottom region. Adjacent probes were separated by equal distances usually 5 to 10 cm, with the probe farthest from the slope at a distance of at least 200 cm from the wave maker. Runs were then conducted in which the forcing amplitude and frequency of the wave maker were varied to include most of the values to be investigated. These data provided a set of reference values with which the later measure- ments over the slope were compared. The procedure used to take the dye- streak measurements is summarized below. (1) Prior to the initiation of wave motion, a grid marked in milli- meters was gently lowered to the slope in the middle and parallel to the side walls of the tank, photographed in place, and gently removed. The position of the camera relative to the front wall of the tank was carefully measured; this position could be repeated to within 1 percent. 30 (2) After the disturbances introduced by the grid removal had sub- sided (usually 10 - 15 minutes), and several minutes after the wave-maker had been turned on, several dye pellets were dropped at various locations over the slope (normally four) and at one position over the flat bottom. The pellets were dropped through a small glass tube at the previous position of the grid. By choosing dye pellets that were small (diameter <0.3 mm) and nearly spherical, irregular motion during descent was minimized. During the rest of the experiment, photographic and conductivity measurements were taken at various locations and times. Each experiment usually lasted from two to four days, depending on the nature of the particular experiment and the physical endurance of the experimenter. 31 3. DATA ANALYSIS Several kinds of experimental data were obtained; these included (1) temperature, (2) mean density gradient, (3) density perturbations, (4) boundary layer velocities and bed shear stress, and (5) miscellaneous data including mode number, net particle drift, and measurements of boun- dary layer instabilities. The instruments and some of the procedures that were used to obtain the various kinds of data have already been described; this section presents a summary of the data reduction techniques and the methods of analysis . TEMPERATURE Vertical temperature profiles at several horizontal locations were made at least once before and after each experiment, and usually several times in between. Typically the maximum temperature variations were relatively small (<0.5°C) over the fluid interior; more irregular changes often occurred in the near-surface layer. This layer will be discussed in more detail in the next section. Figure 8 is a representative temperature- depth profile. MEAN DENSITY GRADIENT A typical density gradient is shown in Figure 9. Measurements of index of refraction were made at about 2 -cm intervals in the vertical and at closer spacing near the free surface. Index of refraction was converted directly to specific gravity by applying a calibration constant that included a correction for the local temperature. Several times during the course of the experiments the specific gravity values obtained in this way were checked by carefully weighing a known volume and determining density in g/cc di- rectly. The maximum difference in the two methods was always found to be 3 less than 1 part in 10 (or about 0. 1 percent). 32 23.5 Temperature , Fig. 8. Typical temperature variation with depth Data from exp. group 4. 33 Z , cm 10 15 20 (Specific gravity — l.o) 25 I0~3 Fig. 9. Typical specific gravity variation with depth Data from exp. group 2; computed vertical distribution of Brunt Vaisala frequency is also shown. 34 The linearity of the curve in Figure 9 is quite good over the fluid interior but nonlinear tails are evident near the bottom and the free surface. Evaporative cooling at the free surface, when the plastic wrap was removed from the slope section of the tank during the conductivity runs, tended to increase the density of the water there, countering the effect of lower den- sity due to low salt concentrations. The overall result was a more homo- geneous density layer near the free surface. Weak convective motions were observed in this layer in the absence of wave motion by dropping potassium permanganate crystals through the water and tracking the motions of the dye traces in this layer. The amount of evaporation was reduced significantly (to < 1 mm per day) when the plastic cover was used. The variation of salt concentration with depth as a function of time for a similar experimental arrangement has been derived theoretically from the one- dimensional diffusion equation for a constant coefficient of diffusion and conditions of no flux of salt across the boundaries (Mowbray and Rarity, 1967). The theoretical concentration-depth curves were shown to have a linear interior and nonlinear tails that thickened in time. The point of maximum curvature located between the nonlinear and linear por- tions of the curves might explain the sharp shadowgraph image that was usually observed a small distance below the free surface and was inter- preted as a change in the density gradient (i.e., ap /az #0 at this point). An example of the vertical distribution of Brunt-Vaisala frequency N is shown in Figure 9. The effects of the nonlinear tails cause N to decrease sharply in the small layers. Photographic measurements of the modal structure and of the particle displacements indicate that the basic wave motion was not measureably affected by the small, nonlinear portions of the density distribution. DENSITY PERTURBATIONS Perturbations in the density field induced by internal wave motion were sensed with conductivity probes. Voltage variations relative to some 35 mean or null value and proportional to the conductivity changes were recorded digitally on magnetic tape. The quality of the data was checked by using a real-time display on a strip-chart recorder. Normally, measurements from all four probes were taken at least one minute before and after the wave motion to establish an instrument drift correction. Each of six data channels was sampled sequentially once per second (At = 1 sec); the channel allocations were four channels for probe data, one for the wave -maker signal, and one for a reference voltage. Samples of the raw data signals from one probe and wave- maker are shown in Figure 10. The length of each data record depended on the particular ex- perimental run. Record durations of least 50 wave periods, often up to 100, were obtained for quasi- steady conditions as determined from the strip- chart output. The data records on the magnetic tape were then processed on the IBM 360/65 as follows: (1) average and trend were computed, listed, and removed; (2) mean square of the data values was computed and listed; (3) the periodogram was computed for 2N data points using a Fast Fourier Transform algorithm (IBM Manual H20-0205-3) to obtain squares of the real Fourier coefficients a, , b. such that: k' k N-l v = l/2a0 + I {akcos(^) + bksin(f )} + £ a^-1)* J k=l 3 =0, 1, , 2N-1 (3'1} Normally, the starting point for each computation of the periodogram was selected when it was determined that the data were quasi- steady (transients and other irregular motions were no longer evident on the strip chart). Output listings and selected plots of the squares of the periodogram esti- mates a (in m v were obtained. 2 2 mates a (in cm ) and of the phase $ at each frequency f (in cps) 36 en en -en ^ i£j cn L±J -cn fT1 -CD > •H 4J O d co O 4J o co o d CO jj *d* ^ > ctj 6 0) J-i o d co i fd 0 co CD M > id (0 CD s • M rH P fd CO Cn II CD C P >i-H H3 M 0 3 P 14-1 T3 0 CD n3 > W G •H d 0 -P 0 CJ fd •H CD iH +J W CD t0 5-1 O jG O O CD H ItJ U CD rd rC P -P ^-s •H (3 CO 5 u •fl CD -P CD .Q (D -P e 0) -P 3 A O c w iH — - M D^ u CO 0 03 A > CD 5-1 rH fd d) u e rH •H Qj-P CD S U e r3 fd ■H CO D^fH CM •H &4 41 „ 3>t ],t-l j,t " At t = 1, 2, , T J = 1, 2, , L T is the number of streaks (usually equal to the nearest integer greater than the wave period). L is the number of vertical levels. j = 1 represents the lowest level for which the data of any particular set are considered reliable. As stated above, the spacing between the various vertical levels from j = 1 was constant in the boundary layer. (4) Estimates of bottom shear stress (one-second averages) were computed from the velocity values at j = 1: ul t T = » M Aif t = 1, 2, , T as before. A 77 is the distance between the bed and the level j = 1 measured normal to the bottom. MISCELLANEOUS To ascertain that the wave maker was in fact generating primarily the first mode, several photographs of the orbital motion of neutrally buoy- ant particles over an entire vertical section were taken at various positions along the tank. Figure 13a is a typical section over the flat bottom. The predominantly vertical motion near the center of the water column and the horizontal pattern near the free surface and at the bottom are characteris- tic of internal wave motion of the first mode. The photograph was taken with Polaroid film that was exposed by flashing a strobe through a thin slit 42 Fig. 13 a. Time sequence of neutrally buoyant particle positions over horizontal bottom. Modal structure approximately equal to the first internal mode (n = 1) . Pointer depth: 16.0 cm. Bright horizontal section on pointer: 1.0 cm, Brunt-Vaisala period: 6.5 sec. Wave period: 12.0 sec. (Strobe: 1 flash/sec; Polaroid 4x5 film, ASA 3000) 43 from above the tank at 2 cps. The illuminating flash was a rectangular beam about 1. 5 cm wide and approximately centered between the side walls of the tank. Similar photographs were taken at various positions over the slope. Figure 13b shows a close-up view of successive displacements of neutrally buoyant beads near the bottom of a relatively low slope (a^ 15 degrees). Note the amplification in the particle displacements nearer the bottom. Shadowgraph images were used to visualize the boundary layer in- stabilities and breaking of the high frequency waves. Both of these phenomena are discussed in Chapter 4; several photographs of the shadowgraph images are shown in that chapter. The net drift patterns of neutrally buoyant particles were determined by two methods: (1) The positions of selected particles were marked on the glass side walls of the tank over long durations (up to 30 minutes). (2) The positions of the particles were traced from 16 mm movies of the shadowgraph images that were projected onto a screen of large grid paper. By counting the frames between position marks, the approximate time that had elapsed was estimated and a net velocity computed. 44 Fig. 13 b. Time sequence of neutrally buoyant particle positions over slope (a=30 deg) . Note intensification of motion near bottom Brunt-Vaisala period: 6.0 sec. Wave period: 9.6 sec. Input wave amplitude: 0.2 cm. (Strobe: 1 flash/sec; Tri-x 35mm film.) 45 4. EXPERIMENTAL RESULTS GENERAL DISCUSSION This section describes the organization of the results and presents an overview of the hydrodynamics. Subsequent sections discuss the results in more detail and compare them with the predictions of recent theories. The results are separated into two basic categories: (1) one that describes the interior wave field (wave number and wave amplitude); (2) another that describes the activity at the boundaries (wave breaking and boundary layers). It is useful to define two distinct hydrodynamic regimes on the basis of the reflection conditions for internal gravity waves from a smooth, rigid slope (Phillips, 1966; Sandstrom, 1966). This subdivision is based on the ratio of bottom slope y to the slope of the input wave characteristics c. Figure 14 illustrates the classification and shows that regime n can be fur- ther subdivided into two cases. Phillips (1966), among others, has shown the slope of the characteristics to be a function only of the stability fre- quency N and the frequency of the input waves u>, provided that N is con- stant with depth: c = l/(N2/w2-l)1/2 (4-1) Since N and y were essentially constant during each experiment, the ratio y/c could be altered by changing co. In this way, both hydrodynamic regimes y/cl were investigated during several experiments. Tables Bl and B2 in Appendix B summarize the experimental param- eters for the various measurements that were made with the conductivity probes. Table Bl is divided into experimental groups that each have a simi- lar Brunt-Vaisala frequency and bottom slope. This table is a reference for Table B2, which lists the various input wave frequencies that were generated during the experiments. The experimental conditions for each frequency can be found by matching the second number in the first column of Table B2 with 46 Reflection Regime Case Designation r/c Properties I A subcritical < l transmissive; all energy propagates into corner n B critical = 1 marginally trans- missive; energy propagates parallel to bottom slope C supercritical > 1 nontransmissive; energy is back- reflected Fig. 14. Classification of internal waves based on reflection from a smooth, rigid slope. Slope of wave charac- teristics = c; bottom slope = y. 47 the number of the experimental group in Table Bl. Those frequencies which belong to hydrodynamic regime II are identified by a dash in column 7. An asterisk is located above those frequencies for which more than one ex- perimental run were made. The emphasis on two hydrodynamic regimes is supported by the re- sults. The subcritical case was examined extensively, since the results for this case are pertinent to the later analysis of sediment motion and since this case is important for certain oceanic conditions (Chapter 5). The data for this case suggest a further classification into two wave types based on the frequency of the input waves. This distinction is more justified for the measurements that were taken over the low slope angle (15 degrees) than those measurements taken over steeper slopes. The experimental re- sults indicate that the high-frequency input waves decrease in wave length and increase in wave amplitude as they propagate upslope until at some lo- cation near the corner and below the free surface they dissipate by breaking. The low-frequency waves in this case (yl) indicate that the interior wave motion is very complicated. The conductivity-probe measurements show that during most experimental runs the wave amplitude over the slope was a maximum at or near a depth z = -ex. (The geometry is defined in Figure 15.) In the critical case, a line of small, regularly spaced vortices forms along the slope. The vortices have diameters on the order of 1 cm and axes that are normal to the side walls of the tank. They oscillate along the slope with the wave motion, collapsing and reforming over a half-cycle. The distance from the vortex core to the slope is about 48 g W >1 in 0> -p id ■H T3 H O O O >i M -P (D g O traveling toward the corner is composed of two progressive waves: ip = A -iq£n(cx - z) -iq£n(cx + z) e"lcot (4-6) 2n7T C + y /. „\ where n is the mode number. This solution is singular at x = z = 0 for case A (y apQ/az w (4-10) (the prime on p has been dropped for convenience). 51 The vertical velocity w is obtained from the stream function in Equation (4 -6), w = a i///ax: -icjt w Aiqc 1 -iqCnM 1 -iq£nN "Me + N6 (4-11) M = ex - z N ex + z Applying Equation (4-11) into Equation (4-10) and taking the time average of the product of p and its complex conjugate p*, we obtain an expression for the square of the amplitude variation with coordinate position: Pf, = \pp* = iA2(S£)2(8p,yaz>2 W U) (4-12) X + ivr N 2 r , /% cos |qCn(^)} NM The comparative experimental value is the square of the spectral amplitude, 2 a , as determined from the periodogram of the conductivity-probe measure- m 2 2 ments. The method of comparison between a and p was discussed in m rw Chapter 3. It is useful to point out here that for y<c, Wunsch (1969) showed that the apparent singu- larity at z = -ex in Equation (4-6) could be removed. Wunsch rewrote q in the form q = r + is (4-15) and showed that the expression for the stream function for waves propagat- ing upslope could be rewritten -icot if) = A where (s - ir)£nM (s - ir)#nN e - e (4-16) r = 2n7r£n|A' [(£n|A'|)2 +tt2] -2m s = [Un|A|)2 +7T2] c - A' - ^-^ = 1/A C + y ' and n is a negative integer. (Note that £nN is complex for ex + z < 0.) The notation here does not conform to that of Wunsch (1969) in order to avoid later confusion. He wrote the complex form of q as q = r\ + is, and gave an analogous expression for Equation (4-16) in which the real part of the exponential group was written in a power of s(i. e., 5 in his notation). 53 In order to have the velocities and their first derivatives continuous, Wunsch derived the requirement that s > 2 or .„ > fialfJli + ! (4.17) 7T -n represents the lowest, nonsingular inviscid mode permissible for these solutions. The vertical velocity is obtained from Equation (4-16): w = Ac(s - ir) J_ (s - ir)0nM _1 (s - ir)£nN M e N e e"ia)t (4-18) 2 An expression for p analogous with that in Equation (4-12) can be obtained w in principle from Equation (4-10) and Equation (4-18) and compared to the 2 measured amplitude a for specific coordinate positions by the method described in Chapter 3. However, it is not simple in this case to obtain a physically meaningful phase function that is analogous to Equation (4-8) owing to the complicated wave forms in Equation (4-18). The amplitude dependence on the spatial coordinates indicates that these are not simple plane waves. By writing w = f(x, z)eig(x' Z' t} (4-19) where f and g are complex, a phase function U can be defined as U = tan"1 gg (4-20) where I(w) and R(w) are the imaginary and real parts of w, respectively. This form of U will be used in the later discussion of theoretical and ex- perimental results for cases B and C. Boundary Layer The full set of perturbation Equations (4-2) through (4-5) were used by Wunsch (1969) to obtain linear boundary-layer solutions for case A 54 (y < c). The solutions match those for the inviscid interior just outside the boundary layer. The stream function can be written: ^ = h+ *b.i. (4'21) where ^T and $ are the interior and boundary layer contributions, re- spectively. In terms of the coordinates J, rf shown in Figure 15, the expression for ^ 55 -icot e iAq V(?) (4-24) Fe-iqCn(BT) _ Q e-iqfin(D|) The behavior of the velocity component u parallel to the bottom with variations in J and rj" is obvious: (1) u grows to infinitely large values as { — 0; in fact, at rf = j = 0, the solution is singular, as might be ex- pected since the interior solution is singular at the origin for v < c; (2) for if = 0, J * 0, the solution reduces to that of the interior field; and (3) as "rj -* + °°, the boundary-layer part (u, ) becomes infinitesimally small and u -* u.. The solutions for w can be derived in principle from wT = a ip/dT, but are not presented here since the measured values of wT were very small in the boundary layer (this is discussed later in this chapter) . For cases B and C (y > c), Wunsch (1969) found that the boundary - layer equation for \\j, was nonseparable in the spatial coordinates. In fact, Wunsch showed that when the slope is near critical the scale of the -1/3 -1/2 boundary-layer thickness is proportional to R ' , in contrast to R J for > z/c.) 57 High Frequency Waves EXP group ao = 0 2 cm 1 0 2 □ 3 A 5 O X x2- 10' Fig. 16. Upslope variation of wave number, y < c 58 A© 0 Low Frequency Waves EXP group aQ= 0.2 cm 1 O 2 □ 5 0 6 0 3 4 X 1 — i — * — r 6 7 8 X2- z c 10' Fig. 17. Upslope variation of wave number, y < c 59 the data points in Figure 16 depart significantly from the theoretical curve. It is also evident that the value farthest to the right for each experimental group in Figure 16 is displaced toward higher wave number than that pre- dicted by theory. This trend is illustrated in Figure 18, which includes data for the high-frequency waves (u) = 0.094 cps) in Group 1 (Table B3). The upslope increase in wave number is similar for both input waves (a = 0. 2 cm and a = 0. 1 cm). However, the greater increase in wave number for the larger waves (a =0.2 cm) at the most cornerward position (i. e., closer to x = 0) suggests that the upslope variation in wave number might depend on amplitude. The experimentally determined value of e at this most cornerward position is approximately 0. 15; this might indicate that nonlinear effects are beginning to have a significant effect on the motion. In Figure 17 there is good agreement between theory and measure- ment for the low -frequency waves, although most of the data values tend to be lower than the theoretical curve in the right-hand side of the diagram (near the corner). This suggests that the wave length of these lower fre- quency waves decreases more slowly with decreasing distance to the corner than theory predicts. Another type of comparison between theoretical and experimental values of wave number is shown in Figures 19 through 21. The data are taken from columns 12 through 14 in Table B3 (Appendix B). Experimental wave numbers k and k were determined for each input wave from probe mo ^ measurements taken over the slope and over the flat bottom, respectively. The ratio k /k is the ordinate in Figures 19 through 21 (k = A0 /Ax). m/ o & & m m/ A comparative theoretical ratio is used as the abscissa in these figures; i. e., kw/kw where k is the theoretical wave number computed from Equation (4-9) for coordinates established by the probe locations over the slope, and k is similarly computed for coordinates at the entrance to wo the slope region (k = A0 /Ax). (The depth z used in the computation of k is equal to the probe depth during the determination of k and k .) o 60 1 A© q ax •I02 3 - m 2 - I - EXP group 1 oj - 0 094 cps do , cm 0 0.2 0 0. 1 *w 10' Fig. 18. Upslope variation of wave number, y < c 61 _L A®m Ko ' AX 6 - 5- 2 - High Frequency Waves EXP group do = 0.2 cm 1 O 2 □ 3 A 4 ♦ 5 O 1 r 0 ~ r i 4 _L A®w Kw0' AX Fig. 19. Measured versus computed values of wave number, y < c 62 A©m Ko AX 3- 2 - I - Low Frequency Waves EXP group ao = 0.2 cm 1 O 2 □ 5 0 6 <3> Kw, A©w AX Fig. 20. Measured versus computed values of wave number, y < c 63 _L A®. Ko AX m 4-i 3- 2 - I - EXP g roup 1 (X) = 0. 094 cps Oo , cm O 0.4 G 0.-2 I A® w Kwo A X Fig. 21. Measured versus computed values of wave number, y < c 64 Figure 19 and 20 also show better agreement between experiment and theory for the smaller wave numbers over the lower slope sections. Farther upslope the data for the higher frequency waves in Figure 19 tend toward larger values of wave number relative to the theoretical values; by contrast, the lower frequency waves in Figure 20 generally tend toward smaller wave number. The effect of input wave amplitude on wave number is indicated in Figures 21 and 22, for high and low frequency waves, re- spectively. The values for the larger amplitude waves in Figure 21 appear to deviate from theory more significantly than the smaller waves. Signifi- cant departures from theoretical estimates are also shown for the larger input waves of low frequency (Figure 22). In this case, however, the measured wave numbers are smaller than the predicted values at locations farthest to the right in Figure 22. This behavior corresponds to the pre- vious observations (Figure 17) that the wave lengths of the lower frequency waves apparently decrease less than theory predicts at positions well upslope. Wave amplitude. The results of the conductivity-probe measurements for case A, concerning the effects of shoaling on wave amplitude, are sum- marized in Table B4 (Appendix B). The experimental group numbers cor- respond to those in Table Bl. Experimental determinations of the period- 2 2 ogram amplitudes a and a for each input wave were made from probe measurements over the slope and over the flat bottom, respectively (for 2 more detail see Chapter 3). The value a was determined for each of the 2 2 ° four probes so that each ratio a /a represented data taken with a par- ticular probe. This procedure was convenient since it removed the necessity of introducing calibration constants relating one probe to another. That is, the measured amplitude ratios could be compared to one another without applying relative gain factors. This method of normalization also simplified the detection of measured changes in wave amplitude relative to 2 the input wave amplitude a . The comparative theoretical values p and p were computed from Equation (4-12) for coordinates of the probe loca- tions over the slope and at the entrance to the slope region, respectively. 65 A© E O EXP group 2 oj = 0 .062 cps ao , cm o 0.2 □ 0. 1 — I 50 0 0 2.0 30 4.0 A© w K wc AX Fig. 22. Measured versus computed values of wave number, y < c. 66 2 2 The ratio p fp was equally convenient, since its use circumvented a conversion from density p to the units of a and permitted direct com- 9 9 ^ parison with the experimental ratio a /a . nr o The results shown in Figures 23 and 24, for the high-frequency and low-frequency waves, respectively, suggest that the agreement between 2 theory and experiment is best for the lower values of (a /a ) , i. e., for the wave motion on the lower and middle sections of the slope. This is in agreement with the results for wave number. The solid lines in these fig- ures were drawn for reference and represent points where measurement and theory coincide. The deviations of the points from the straight line in Figure 23 are significant for (p /p ) > 20, but no consistent pattern is apparent in the variations. The values for lower frequency waves show earlier departures from theoretical values than do the higher frequency waves. For example, the measured and theoretical values shown for low- 2 frequency waves in Figure 24 are significantly different for (p /p ) > 10. w o There is strong indication in the trend of the points that the actual ampli- tudes are much lower than predicted on the middle and upper sections of the slope. At these slope sections it was also found that the measured values of steepness a k (or e ) are considerably less than those esti- m m v nr mated from theory. Measured amplitude growth is further compared with theoretical estimates in Figures 25 through 30. In these figures the variation of wave amplitude is examined as a function of horizontal distance x to the corner at various fixed depths z . The data are again taken from Table B4 (Appen- _. * c dix B). In each of these figures the theoretical and experimental ratios of * The actual wave amplitude in centimeters at a point along the slope can easily be found from the graphs. If the correct wave amplitude (in centi- meters) for the flat bottom region is multiplied by the square root of the measured value ((a /a ) ), the measured estimate of the actual amplitude is obtained. 67 ^2 Qm ao 6-i 5- 4- 3 - 10 -I El a High Frequency Waves EXP group Oo = 0.2 cm 1 O 2 □ 3 A 5 © 2 rO-,2 P 1 w P. 10" L' 0 Fig. 23. Measured versus computed values of wave amplitude, y < c. 68 - I2 Qm ao 4 -l 3- io- Low Frequency Waves EXP group Qo = 0.2 cm 2 □ 5 O 6 0 l_ ' 0-J Fig. 24. Measured versus computed values of wave amplitude, y < c. 69 r ^-,2 or L'OJ 8- 7 - 6- 4 3 2- I - r -.2 am a~o" © EXP group 1 ai = 0.094 cps G 00=0.2 cm □ ^ 0 1 1 3 4 r 5 1 6 i 7 1 8 Xc Xo I01 Fig. 25. Upslope variation of wave amplitude, y < c 70 'R, -, 2 or o- Qm ao -.2 4- 3 - EXP gi oup 2 ai = 0.102 cps * ao = 0.2 cm A ao = 0.4 cm O ft, 3 4 Xc Xo I01 i 7 8 Fig. 26. Upslope variation of wave amplitude, y < c 71 P n2 ,- ^2 P, L'oJ or ao 10' 5-1 4 - 2 - EXP group 2 oj = 0.062 cps 0 ao = 02 cm □ Pu, AO 8 Fig. 27. Upslope variation of wave amplitude, y < c 72 2 r -i2 EXP group 3 CJ = 0.094 c ps D do =0.2 cm A 0 o = 0.4 cm O A, ** I01 Xo Fig. 28. Upslope variation of wave amplitude, y < c 73 La or -.2 ao 10" EXP group 5 a; = 0 098 cps a) =0 082 cps -O- ao = 02 cm O ao=0.2 cm *> Pu □ ft, 5- 4 - 2 - 0 &-'°' Fig. 29. Upslope variation of wave amplitude, y < c 74 r- ~-,2 Po or o J 20-, 18- 16- 14- 12- 10- 8 - 6 - 4 - 2 - 0 r- -.2 Om do EXP group 6 cu , cp s Oo P 1 U) 0.062 O i 0 047 Q o 0.043 A 0 3 4 Xc Xo i 9 0' Fig. 30. Upslope variation of wave amplitude, y < c 75 2 2 amplitude variations (p /p ) and (a /a ) , respectively, are plotted along the ordinate; the ratio of the horizontal distance between the probe and the corner x to the total horizontal projection of the slope x is plotted along the abscissa. The depth of the probe measurements is constant for any one of the figures, but as shown in Appendix B, Figure B4, it generally varies for each experimental group. In Figures 25, 26, 27 and 30 the re- sults pertain to the slope angle of 15 degrees, for which the distinction be- tween high and low frequency waves appears most applicable. Two effects of the method used to display the results must be mentioned here: (1) the deviations between the theoretical and measured estimates of amplitude are exaggerated in the graphs, since squares are plotted as the ordinate; (2) the peaking of the waves is a function of the vertical position z of the probes as well as distance from the corner x . This latter effect is subtle; since the internal waves are of the first mode, the amplitude should be largest near the center of the water column at any position over the slope if the modal structure remains intact. Consequently, the amplitude is expected to vary along a horizontal line (constant depth) due to the modal structure as well as to the amplification during shoaling. To the right of the peak values (i. e., away from the corner) in Figures 25 and 26 (high-frequency waves), the variations are essentially parallel, although the experimental values are consistently higher than the predicted. The maximum measured values are slightly higher than pre- dicted, with an indication that the larger input waves (aQ = 0.4 cm, Figure 26) peak sooner than the smaller waves. There is also evidence *For this case (y < c), photographic measurements indicate that the mode is preserved at least over the lower and middle slope areas. 76 that the measured values fall off to lower values than the theoretical values to the left of the zone of peaking (i. e., toward the corner). The high-frequency waves over the steeper slope (Figure 28) con- form closely to the predicted behavior for x /x > 0. 35, but nearer the corner the amplitudes of these waves grow less than predicted. This is probably a result of the anomalously large growth of the theoretical solu- tions near the corner (recall that for y < c Equation (4-6) is singular for x = z = 0). The larger input waves (a =0.4 cm, Figure 28) show a tendency to peak earlier than predicted. The measured value at x /x =0. 35 for a = 0. 2 cm is suspiciously low. Figures 27 and 30 illustrate the behavior of wave amplitude during shoaling of the lower frequency waves for a low slope angle (15 degrees). The measured values obviously do not achieve the high predicted values well upslope (x /x < 0.4 in Figure 27 and x /x < 0. 5 in Figure 30). The departure from theoretical estimates is most noticeable for the very low frequencies (o> = 0. 047 cps and w = 0. 043 cps) in Figure 27. This sug- gests that the waves increase in amplitude as they propagate upslope until at some point near the half-distance of the horizontal projection of the slope the rate of growth decreases. It was indicated earlier that these waves tend to surge as they approach the corner. Possibly a transformation from the sinusoidal wave form to a different wave structure (such as a solitary wave) that is continually losing energy to viscous dissipation along the slope accounts for the nature of the experimental data for these kinds of waves. In any event, the amplitude of these lower frequency waves is not as large as predicted, and there seems to be evidence of a more gradual dissipa- tion of these low frequency waves with the formation of runup. 77 Critical and Supercritical Cases The experimental runs for the critical and supercritical cases are indicated by a dash in column 6, Table B2 (Appendix B). A comparison of the input wave frequency with the critical frequency for each of these runs shows that o)< ct> , a necessary condition for this regime. A brief description of the reflection properties that define this regime was presented in the first section of this chapter. In particular, for case C (supercritical) two fea- tures are suggested by the reflection properties: (1) the presence of back reflected wave energy; (2) concentration of this back-reflected energy in a zone bounded by the bottom slope and the characteristic that intersects the corner. Figure 31 further illustrates these features. This figure shows that wave characteristics which are incident to the slope from above, i.e., those with a downward-pointing vertical component, upon reflection from the slope will remain below the characteristic (defined by z = -ex) that intersects the corner. The energy flux associated with these charac- teristics enters the slope region in a relatively broad band and exits in a confined, triangular zone (for example, the triangular area AOB in Figure 17). As Figure 31 indicates, there is a smaller influx of energy which, upon reflection from the flat bottom, propagates upslope within the narrow bottom zone (for example, the path of the characteristic "b" in triangle AOB). The characteristics associated with this latter energy flux have upward-pointing vertical components as they enter their slope regions; upon reflection they preserve the sense of the upward-directed vertical components. In both cases (characteristics "a" and "b" in Figure 31) once the characteristics reflect from the slope the horizontal direction of energy is away from the corner. Critical conditions (case B) were difficult to achieve experimentally for two reasons: (1) y - c required that the wave-maker frequency be tuned precisely; (2) it was difficult to determine the Brunt-Vaisala fre- quency accurately during an experiment. In addition, the experimental groups which involved the 15 degree slope angle had critical frequencies 78 0 Zc > CXc ( I) Zc >CXc (2) FIG. 31. Reflection of characteristics for case C, y > c where y = tan a, z = IzL x = Ixl and characteristics are labeled a, b, c. Char- acteristic c is "critical" i. e., passes directly to the corner at O in (1) and at F in (2). The increased concentration of charac- teristics in the triangular zones (hatched areas) is illustrated by the path of characteristic a. 79 that corresponded to long horizontal wave lengths over the flat bottom (on the order of the tank length). Data for only six runs at the critical fre- quency are considered reliable. During each of these runs the input wave frequency was held to within two percent of the calculated critical frequency. Table B5 (Appendix B) summarizes the experimental results from the conductivity-probe measurements for this regime (y > c). Like the organization in Table B3, the frequency, period, and probe coordinates (x , z )* are given for each experimental run. The experimental runs having the same N and y are again collected into groups; as before, the experimental parameters for each experimental group are listed in Table Bl. An asterisk in column 2 of Table B5 again indicates that the experi- mental run was repeated at least once. The local depth h is shown in column 6 of Table B5, and the parameter — ~ ~ in column 10 is x - (z cr analogous to the parameter in column 7 of Table B3. The experimental method of determining wave number and wave am- plitude was discussed in Chapter 3. These quantities are plotted as functions of x /x and z /ex in this section to examine their dependence on (1) hori- zontal distance from the slope at a fixed depth, and (2) nearness to z = ex (i. e., distance to the theoretical position of the critical characteristic). The pertinent theoretical curves were computed from Equations (4-18) and (4-20) and are presented in several of the figures for comparison with the data. It must be pointed out here that qualitative considerations of the linear, inviscid solution for y >c, given by Equation (4-16) suggest a priori that these solutions might fail to fit the experimental data. It can be shown that the velocity components derived from Equation (4-16) become anoma- lously large for large coordinate values (x, z). This behavior was not * In this discussion and in Table B5 x = |x| and z = |z| for x,z as shown in Figure 15. c c 80 observed in the experiments. In addition it is shown later that the theory (in particular, Equation (4-17)) selects |n| = 2 as the lowest nonsingular inviscid mode permissible over the slope region for each of the experimental conditions tested. In the particular experimental geometry, it was not ap- parent that the second mode was present over the slope; it was found most often that the modal structure was approximately that of the input mode (|n| = 1). In spite of these qualitative objections, the following discussion of the experimental results is done quantitatively by comparison with the theoretical predictions and by consideration of the reflection properties (Figure 31). It is felt that this method of presentation aids the interpreta- tion of the results. Wave number. Figures 32 and 33 illustrate the complicated behavior of the experimentally determined wave numbers for this case. Data from three experimental groups and for frequencies co= 0.051 cps and co= 0.066 cps are represented in these figures. The calculated positions of the critical characteristic (determined by z = ex ) for two of the experimental groups are shown along the horizontal axes. The two theoretical curves shown in each figure were computed for two different values of the mode number (see qualitative discussion at end of previous section). Equation (4-20) was used to determine the minimum permissible nonsingular mode number for the curves labeled n = 2. The mode number n = 1 was preselected for the theo- retical curves labeled n = 1. Although this latter choice violated the con- dition necessary for finite values of the velocity components and their first derivatives at z = -ex (Equation (4-20)), it nevertheless forces the theoretical mode number over the slope to be equal to the input mode of the experiments. It is apparent in the figures that there is no straightforward relation- ship between linear theory (n = 2) and the data. There is only qualitative indication that the curves for n = 1 have the same trend as the experimental results for x /x < 0.035 in Figure 32 and for x /x < 0.45 in Figure 33. c/ o to c' o It is obvious that the theoretical estimates for n = 1 grow toward very large values at the critical points (for example, x /x = 0.26, Figure 33). The 81 Km Ko 22i 20- 18- 16- 14- 12- 10- 8- 6- 4 2- 0 n = O n = 2 EXP group 3 OJ = 0.051 0 Qo = 0.2 cm — I T 1 1 1 1 1 1 1 0.2 0.3 0 4 0.5 0.6 0.7 0 8 0.9 10 Xo Fig. 32. Upslope variation of wave number, y > c. 82 Km Ko a; = 0. 066 ao = 0. 2 cm -O EXP group 4 -O- EXP group 5 0.1 0.2 0.3 0.4 05 0.6 07 0.8 09 1.0 Xo Fig. 33. Upslope variation of wave number, y > c 83 experimental values near the calculated critical points are apparently bounded. The dip in the data in Figure 32 for x /x ^ 0.30 is not apparent in Figure 33. Possibly the absence of the dip in the data from experimental group 4 (Figure 33) can be explained by considering nonlinear effects. The measured value of e (Table B5) for this particular data set is approximately 0.5 at x /x = 0.36. This suggests that nonlinear amplitude growth might be significant near these positions. By comparison, the measured value of e for the data in Figure 32 is only 0. 13 for x /x ^0. 2 (Table B5). It should also be noted that there is a significant shift in the peak value to the right of the calculated critical point in Figure 32. Possibly this shift indi- cates that the singularity is removed by nonlinear or viscous processes and that the linear inviscid solutions are not valid in this particular physical situation. Data from experimental groups 3 and 4 are plotted on log- log scales in Figures 34 and 35, respectively. Theoretical curves for n = 1 are again shown in these figures. The qualitatively similar trend between the theo- retical curve and the data is apparent, particularly in Figure 35. Values of wave number for the larger amplitude input waves (a =0.4 cm) tend to be higher than those for the smaller waves (a =0.2 cm) in experimental group 3 (Figure 34). Figure 36 shows the results of experimental group 3 plotted as a function of z /ex . It is obvious that the theoretical curve for n = 1 grows & c toward infinite values of k /k at z = ex ' the theoretical curve for n = 2 nv o c c7 (obtained from Equation (4-17)) is finite at this point. The trend in the data values reverses at about z /cx„ * 0.80, and again at z /ex =, 1. 0. c c ^ ^ No simple relationship between the data and the theoretical curves is apparent from this figure. Figure 37 summarizes the experimental results of wave number for case B (r = c). The log-log plot emphasizes the steep rise toward larger values of wave number for decreasing distance to the slope. It is interesting that the magnitudes of wave number for the larger input waves (a =0.4 cm) 84 0 9" 8- 2 Ko 6- 5- 4- 3- □ O ~ET O O B CD O □ o n = EXP g roup 3 CJ = 0. 051 cps do i cm 0 0.2 □ 0.4 0.2 0.3 0.4 0.5 Xc Xo Fig. 34. Upslope variation of wave number on log-log plot, y > c 85 io- 9 8 7- 6- 1 Km 2 Ko 4- 0 n = 3- 2- EXP g roup 4 U) - 0. 066 cps ao • cm 0 0.2 o 0.15 0.2 03 0.4 05 06 07 Xc Xo Fig. 35. Upslope variation of wave number on log-log plot, y > c 86 2.0 Fig. 36. Wave number versus magnitude of zc/cxc, y > c. 87 9- 1 Km 2 Ko 8- 7- Q 6- 0 4- 3- 2- □ O 0 □ □ A O □ EXP group 3 EXP group 4 oi - 0. 07 3 c p s w = 0 .079 c ps Zc = 5.5 cm Zc = II. 1 cm a o , cm do , cm 0 0.2 A 0 2 B 0.4 A □ 0.2 0.3 0.4 0.5 0.6 0.7 Xc Xo Fig. 37. Upslope variations of wave number, y = c. 88 in experimental group 3 are larger than the corresponding values of the smaller input waves for x /x < 0. 3, but for x /x > 0. 3 the wave numbers CO CO of the larger waves are consistently less than those of the smaller waves. In general, the wave number variations with distance to the slope for this case are similar to the variations shown in Figure 34 for the supercritical case. No theoretical comparison is shown, since for y - c the minimum permissible nonsingular inviscid mode over the slope, given by Equation (4 -17), is infinite. In summary, the results show a complicated relationship between wave number and position over the slope. There is evidence in Figures 34 and 35 that the wave number increases rapidly as the waves pass over the lower and middle sections of the slope for the supercritical case, but there does not appear to be a systematic pattern to the data near z = ex . The variations in wave number for the critical case are qualitatively simi- lar to those for the supercritical case; a large increase in the wave number for smaller x /x is apparent. The trends of the theoretical values of wave & o number computed for n = 1 are similar to the measured values for z /ex < 1. It is obvious that the unbounded theoretical values at z = ex for n = 1 are c c removed when the condition given by Equation (4-17) is satisfied. For both cases, the measured values of e (Table B5) are greater than 0. 1 for most of the measurements taken closest to the corner. This suggests that nonlinear effects might be important near these positions. Wave amplitude. The experimental estimates of wave amplitude were ob- tained from the periodograms of the conductivity-probe measurements (Chapter 3). The results are presented as ratios of the square of the wave amplitude over the slope to the square of the wave amplitude over the flat bottom for various frequencies of the input waves. Like the earlier presen- tation for wave number, these results are illustrated graphically with z /ex and x /x used as the horizontal axes, c c & o 89 The results for wave amplitude are easier to interpret in terms of the reflection properties of the slope than were the results for wave number. Earlier theoretical discussion indicated that near the characteristic defined by z = ex and in a zone between this characteristic and the bottom slope (Figure 31) there is a possible region of concentrated energy flux. Figure 38 shows that the measured values of wave amplitude are highest near z /ex =1.0 for each frequency tested in this group except u)= 0.039 cps. However, it is not conclusively shown in this figure that the amplitudes peak exactly at z = ex . This figure also shows that the ordinate values for the various frequencies at equal values of z /ex generally increase with de- creasing frequency. In the light of Figure 31, back-reflected wave energy m might contribute smaller amounts to the total energy at locations above the critical characteristic (i. e., at positions given by z < ex ) for frequencies nearer to critical. This in turn might account for the lower measured 2 levels of (a /a ) for the frequencies nearer to critical in Figure 38. Figure 39 is a similar plot for data from experimental groups 3 and 4; two input wave amplitudes are represented in the values from experimental 2 group 3. Also shown in Figure 39 are theoretical curves for (p /p ) de- rived from Equations (4-10) and (4-18) by the methods described earlier. The mode number |n| was determined from the condition in Equation (4-17) 2 for curve n = 2 in this figure. The theoretical values (a /a ) decrease b m/ o toward the slope and undergo a smooth transition through the critical point z /ex = 1.0, where they then begin to increase toward the slope. The theoretical curve with n = 1 has increasing values toward the slope until the critical point (z = ex ) is approached; a sharp dip in the theoretical values at this point followed by a rapid rise suggests that the solutions are not well behaved at this critical point. The ordinate values for the two input ampli- tudes of group 3 in Figure 39 agree within 10 percent of one another for values of z /ex < 1.0; in general, the larger input waves have slightly higher measured amplitudes in this range. It is also apparent that the larger input waves grow more rapidly as the value z /ex = 1.0 is approached. Owing to the larger values of e (Table B5) for the waves of larger input 90 a m 2 © 24- a 0 22- 20- o 18- -0- 16- 14- 12- Ar io- A <> 8" ta 0 6- A El 4- A A- A O 0 A 2- i s — 9— -El- El 0 I — P" 0 i 0 1 1 E i 1 Zc cXc 1 02 0.4 0.6 08 .0 1.8 20 EXP group 4 EXP group 5 Oo = 0.2 cm Oo = 0.4 cm cj , cps ai , cps O 0.073 O 0.066 A 0055 □ 0.051 0 0.039 -Er- 0.0 6 6 -A- 005 5 Fig. 38. Wave amplitude versus magnitude of zc/cxc, y > c 91 Cl» = 0 051 cps EXP group do , cm 3 O A 0.2 0.4 4 □ 0.2 Fig. 39. Wave amplitude versus magnitude of zc/cxc, y > c 92 amplitude (a =0.4 cm), the increased disparity in the measured amplitude values for input waves of 0.4 and 0. 2 cm, respectively, at values of z /ex > 1.0 in Figure 39 might be the result of nonlinear effects. Figure 40 shows the experimental results and theoretical curves for a) = 0.051 cps (group 3) on a log-log scale; the horizontal axis is distance x to the corner normalized by the total horizontal projection of the slope x . The rapid increase in amplitude upslope for 0. 28 < x /x < 0. 60 is readily apparent in the data; however, the theoretical values for n = 2 show an op- posite trend, toward lower values for decreasing distance to the slope. The measured amplitudes below the indicated position of the critical point remain relatively high; this suggests amplification of the motion near the bottom. Nonlinear effects also might be important near the corner (x /x = 0. 19) The results for the probe measurements of wave amplitude at critical frequencies are shown in Figure 41. The increase in amplitude over the range 0. 28 < x /x < 0. 60 is similar to the increase shown in Figure 38 for the same range. However, the maximum amplitude measured for these waves is considerably less than the corresponding value for the supercritical case (Figure 40). Larger viscous dissipation along the slope in the critical case might explain the smaller peak amplitude levels. In summary, the data indicate that the wave amplitudes generally increase toward the slope; the higher frequency waves (i. e., those closer to a? ) for these cases show rapid growth during shoaling over the middle sections of the slope. Peak levels of amplitude were generally found near z /ex m 1.0, with lower measured peak values for waves of critical fre- c' c ' * quency. In both cases the motion is relatively large near the slope. In general, the conductivity-probe measurements of amplitude agree qualita- tively with the reflection properties of internal waves from a rigid, sloping bottom. There is some indication from the measured values of e that m nonlinear amplification of the amplitude occurs for small values of x /x . 93 10.0 0m ao 12 A A O 5.0 A 0 n = i — - A O EXP group 3 u) = 0.051 cps ao , cm o 0.2 A 0.4 A O 20- .0 0.5- 0.2 n = 2 i — i — r 0.1 02 0.5 Xc Xo Fig. 40. Upslope variation of wave amplitude, y > c 94 -|2 5,0— 0m ao 0.5- A 2.0- 1.0" A O A A O 0 % I A 0 $ 0.2- EXP group 3 a» = 0.073 cps ao , cm O 0.2 A 0.4 T i — r "TT 0.2 0.5 Xo 1.0 Fig. 41. Upslope variation of wave amplitude, y > c 95 It might be added here that the bounded experimental values at z = ex suggest that viscous dissipation and nonlinear processes might act to remove the singularity at the critical characteristic. For example, the shift in the measured peak wave number in Figure 32, mentioned earlier in this section, and the formation of vortices near the bottom for these cases, discussed in the next section, suggest that the linear inviscid model is not properly tested by these experimental results. Nonlinear and viscous effects in the interior, particularly in the vicinity of the critical characteristic, should be examined in future analytical treatments of this problem. BOUNDARY ACTIVITY The discussion of the bottom boundary layer and the zone of breaking follows the procedure of separating the experimental results into two hydro- dynamic regimes (Iy < c and II y > c). The distinction between these two regimes was defined in Figure 14. Qualitative features observed during the experiments and illustrated by several photographs are presented at the outset as an introduction to the later discussion of quantitative results. In the discussion of quantitative results, experimentally determined velocities in the bottom boundary layer at several positions along the slope are shown to agree with theoretical values computed from the linear solutions in Equation (4-24) for case A. On the basis of the good agreement between experiment and theory for the conditions tested in case A, Equation (4-24) is used in Chapter 5 to develop a model for sediment movement induced by shoaling internal waves. Case A: y CD £ rd £ 0 £ •H M Eh fa OOOOOOOOLOOOOOOOLOO oooo(N(NHnH(NnmMm id 0 CD 10 iDLnminmLnooooooiOLnuoLOLO oocooooocccoyjvo^yj^voooooc rHrHiHrHrHrHiHiHr-lrHiH P, (U £ 3 53 ■H Ix, cu ^ T3 -P ■H CUrH ooooooooooooooooo fdXffd-Qfd^Ufd^Ofd^fdXJU ^3* ^3* ^J* ^3* ^J1 ^J* ^^* ^J* ^J1 Kj* ^* ^^ ^^ ^* ^3* ^3* ^J* 99 Fig. 42. Fig. 43. 100 Fig. 44 a Fig. 44 b 101 Fig. 45 a Fig. 45 b 102 Fig. 46 a Fig. 46 b Fig. 46 c 103 Fig. 47 a. Fig. 47 b Fig. 47 c 104 Fig. 48 a. Fig. 48 b. 105 Fig. 49 a Fig. 49 b, Fig. 49 c. 106 FIG. 50. Advancing initial vortex at successive locations A, B, and C Note observed reversal in circulation at position C. The outflow above the vortex balances the flow of water that is advected toward the corner along the bottom by the wave motion. As indicated by the well developed lamellar structures in Figures 49b and 49c, the steeper slopes appear to produce a high intensity of breaking. The formation of a surge with associated runup was not observed for the steeper slopes (30 degrees and 45 degrees) for any wave frequency tested. Figure 51 presents a summary of the observed positions of breaking on slopes of 30 degrees. The accuracy of the distance L, is low, owing to the difficulty in defining the point of breaking; however, the relative order of the values of L, for the waves of different frequencies shown in the figure is reliable. The general trend indicates that for a given N, the higher fre- quency waves break first, i. e., in deeper water. 107 FIG. 51. Position of breaking a = 30°, N = 0.97 -1 rr. Li, cm u), sec T, sec b? a = 0. 2 cm a =0.4 cm o o 0.722 8.7 13.4 14.0 0.668 9.4 12.5 12.9 0.593 10.6 10.6 11.1 0.561 11.2 9.9 10.5 0.537 11.7 9.0 9.5 0.515 12.2 8.2 8.6 0.495 12.7 7.7 8.2 Quantitative results. The experimental techniques that were used to mea- sure the motion in the boundary layer are discussed in Chapter 2. The results of these measurements are compared with the predictions of the linear, viscous boundary layer solutions in order to check the validity of these solutions for certain conditions and provide some empirical estimates of the motion where the theory does not apply. Since successive displacements of the dye streaks and neutrally buoy- ant particles were measured at one-second time intervals, the velocity values computed from these measurements are essentially one- second time averages in the Lagrangian sense. Equation (4-24) provides a linear, Eulerian solution for the velocity field parallel to the bottom in the viscous boundary layer; this solution is identical with the Lagrangian one to first order in e. The Lagrangian velocity uT can be determined in principle from the Eulerian velocity u„: uL (a, t) = uE (a,t) + Jto u(a, t') dt' ■ V^ u£ (4-27) where a is a position of the dye streak, say, at time to; t - t is small com- pared to the wave period T, and V is the gradient operator (Phillips, 1966). 108 — 2 — This expression for ul is correct to 0(e ) if u„ is correct to this order. Li Hi Let (D ■ A? (2) (4_28) UE = eUE + e UE where u is the linear solution for the velocity field. It is easy to show that there are two second order contributions to uT in Equation (4-27): (1) u ^2' and (2) (f* u dt') • Vu£. The first contribution (1) is due to the o nonlinear terms in the equations of motion; the second (2) is commonly re- ferred to as the Stokes velocity and can be computed in principal from Equa- tion (4-21). Hogg and Wunsch (1970) have shown that to order e the time average of u for this problem is zero. This implies that there is no net transport of water particles induced by internal waves near a sloping boun- dary. It also implies that since u„(l) = 0, then uL = -(£ uE(1)dt') • VuE(1) (4-29) o where the bar denotes a time average over one wave period. A complete analysis of the data was carried out for three wave fre- quencies in regime I (y c) are discussed in a later section. 109 A time-sequenced series of streak and particle positions from which velocity information was obtained is shown in Figure 12. The streaks are numbered according to consecutive, one-second time marks relative to the initial one. The numbers on the particles correspond to those for the streaks. The increase in displacement of the streaks nearer to the bottom agrees qualitatively with the amplification predicted by Wunsch (1969) for y 6 fv = local boundary layer Reynolds number 6+ = (2v/co) 1/2 = (2y/N)1/2 (o-/(o-2-sin2a)1/2 , a = u>/N = «iA a k sin a m m FIG. 53. Estimated of boundary layer thickness and Reynolds numbers, y < c. 112 the experiments in which oj = 0. 105 cps, a = 15 degrees. It appears that the boundary layer thickness given by linear theory gives the correct order, mm rather than cm, but the measured values tend to be significantly higher. However, it must be recalled that the definition of boundary layer thickness for the purposes of measurement, i. e., the distance between the bottom and the point of inflection in the velocity profiles, is somewhat arbitrary, and might account for the discrepancies with the theoretical values. It should also be mentioned here that at station 1, the computed value of e is relatively high (e a 0. 5); nonlinear effects are probably im- portant there. It was mentioned earlier that Wunsch (1970) recently showed theoretically that there can be a mean upwelling velocity independent of wave motion in a layer whose thickness is 6 w aw~(,*)1/4/N1/2 where v and k are the vertical diffusivities of momentum and heat, re- spectively. In terms of molecular processes, i/~102cm ' and 2 / -1 k ~ 10-3 cm /sec# Using the value of N given in Figure 53, N ~ 1 sec , we can estimate 8 ~ 0o 06 cm for this particular experiment. The curves in Figures 54 through 58 show velocity versus depth for all five stations. Each curve represents a one-second time average of the Lagrangian velocity parallel to the bottom., The consecutive numbers on the curves indicate the time history of this velocity over one wave cycle. The curves were computed only to the vertical level nearest the bottom for which the streak data was considered reliable. The horizontal velocity scales are uniform in each of the figures except for the expanded scale of the flat bottom results (Figure 58) and the condensed scale at station 1; the vertical scales are the same for all figures. It is important to note that at each of these stations the observed motion in and near the boundary layer was essentially parallel to the bottom. 113 O _ CO | ** _ CO < < * — O 0^ •H 0) + -P T3 10 -P o W co 00 4-> II cd o §tf ~"~ -p +J CO 0 - d £ en a a) u <0 -C d -P LT) IT) M iH *- (d • d II CvJ >1 -P 2 d •H U - 0 W D rH 0, Q) U > CO IT) M-l O d 0 -H d 1 C o 0 •H II -P to 3 ^ d i 53 M CO -P d U) * i •h g T! U o .H (N 7~ (13 • u o -H °o ■P II ~ U Q) o ^ > A3 i . to ■^ in 1 , CO tJi •H 1 Cm o <\i 1 CO 114 m 6 CN c - <*- O cr- d ■H CD -P 73 (C -P o to en fO -P II id 6 o ° -p -p 0 - C\J P. d CD U & -P IT) in U rH id . QJ C — c d II >1 -p a •H o •> 0 CO -H C ;> CO u > m m o 0 rH C o 0 — ■H || 6 i -p 1* >-i -P w - £ CM •h r o d TJ 0 j o iH (N • (0 • ro U O (N ■H ro ■P II II d CO o O i > tf X • in o c o •H -p u -p to •H o o CM (T5 U •H -P U 0) > m o U «0 X 116 rr - tr> C i2 -p •H ■> O W o a > .H U in o IP r-\ o • o C O II -H -P 53 •H 5-1 -P W - £ •H £ 0 m T3 O o b ■H cd x • r- m • tn -H P>4 o i 117 m iZ •H - U CO o a rH u m o 4-1 rH o • o o •H -p X! •H 5-1 -P W -H II U rH CN O rrj . o O O rH •H -P II II to > oo in o u fO X •H fa 118 Only very small velocities normal to the bottom were detected from the motion of the plastic beads, and this vertical motion was usually confined to regions adjacent to and in the zone of breaking. A reason for the anomalous behavior of the velocities at station 4 (Figure 57) was hypothesized earlier in this section. An alternative ex- planation is that the abrupt change in bottom geometry acts as a source of diffracted waves which locally affect the pattern of motion. However, since the density did not vary linearly with depth at the bottom (for example, Figure 9), but instead gradually became more homogeneous in a small layer near the bottom, the dynamics in this transition region are probably quite complicated. Visual observations of the motion of neutrally buoyant particles and dye patches that were placed at the transition in bottom geom- etry during several experiments (for 15 and 30 degree slope angles) did not disclose any significant patterns other than those already mentioned. Figure 59 shows a comparison between the theoretical and experi- mental RMS values of the velocities at selected vertical levels shown in Figures 54 through 58. The experimental RMS value at each level was computed as: RMS " lUMS; where \\ MS I (up/r , i = l, 2, and T was chosen as the smallest integer greater than the wave period. For this particular set of data, T = 10 (wave period a 9. 5 sec). For stations on the slope, theoretical RMS values were obtained from the Eulerian expression for the boundary layer velocity component parallel to the slope (Equation (4-24)): URMS(the0ry) = (1/2»E * ue)1/2' 119 These RMS values are listed in Figure 59 as first order theoretical estimates of the RMS Lagrangian velocities. Qualitatively, we might expect that the measured values are more accurately represented by the sum of the first order Eulerian values and the second order values consisting of the Stokes velocity component and the nonlinear component, particularly farther upslope where e become larger. When the various sources of uncertainty in the measurements of velocity and experimental parameters (AQ, N, w, h, etc.) are considered, a rather good comparison would require that the theoretical and experimental values fall within 15 percent of one another at each level tj • The results in Figure 59 show that excluding station 4 there is excellent agreement (with- in 10 percent) between the measured RMS values and the corresponding first order, theoretical values. The excellent comparison at station 1 is particularly remarkable, considering that the estimated value of e is approx- imately 0. 5 there. The high measured values at station 4 are a consequence of the anomalous motion that has already been discussed. The results for the maximum velocity |u| max parallel to the bottom at the selected levels (columns 6, 7, 8, Figure 59) also show good agreement between experiment and theory. Theoretical estimates of the maximum values of this velocity component were again obtained from Equation (4-24). The experimental values are generally higher than the theoretical first order values. The large deviations between experiment and theory at station 4 are again obvious. The results for the other analyses of this kind show similar good 120 O) LO o CM s O CO 0) >> T5 X a >> 5h O 0> C- E- t- tr- CO CO CD CO LO IO LO LO l-H l-H T-l T-t CD CD CD CD CO CD CO CO rtHHH CO CO CO 00 o o o o ee CO CO CD CM O i-f CM CM CM CM o o o o X c- c- c- c- CO CD CD CD C- t- t- t- 00 00 CO CO O O O O ^ * H rf CM CO rf ^ CM t- i-i O) CM CM CO CO rj< CO 00 LO CO LO t- t- CM CO 00 O) l-t i— 1 i-l 1-1 O O O O O O O O o o o o o o o o o CO If o CO K 3~ o 0) H u d G •ft TjH O) O) i-H LO CO t- O) O CO O tJh CM CM CO CO CO CD O) CM i— 1 H i-H CM O) CM Tf LO O tH i-H i-H O i-H i-i O o o o o O O O O O O O O O O O O o o o o o o o o CO O) CO i-< LO LO D- CO 00 O CD O) CM CM CM CM CO LO O) O) LO i-l i-l i-l CM H O) CD CO CM CM CO CO O) i-t CM CM O i-t i-H iH O O O O O O O O o o o o O O O O o o o o s IP ifi NO) ^ H CM CM Tf IO CM O) Tf i-< CM CM rj< LO CM O) rj< i-l CM CM Tf LO CM O) rj< i— I CM CM rj< LO CM OS ^ r-t CM CM Tf o o o o O O O O o o o o O O O O o o o o Station No. i-H CM CO Tt« LO 121 agreement between theory and experiment. The velocity values for station 4 with the 15 degree slope are more in agreement with the theoretical values (within 25 percent) than are the results in Figure 59 and did not show anomalously large velocity values for this station. The asymmetry in the velocity profiles at station 1 (Figure 54) contrasts with the more symmetrical curves for the other stations. Station 1 is located well upslope and only a few centimeters from the zone of breaking. The skewness toward positive velocities suggests that there might be a measurable net upslope motion of the water particles over one wave cycle. Although the maximum velocity is directed downslope (curve 1, Figure 54), the magnitudes of the upslope component are quite high (approximately 1 cm/ sec) for a considerable fraction of the wave cycle. The local measured value of e is about 0. 5 at this station, which suggests that nonlinear effects might be important. It is also interesting to note that <5 at this station shows a decrease from that at station 2 (Figure 53). Experimental estimates of net drift at the outer edge of the boundary layer were computed at each station by algebraically averaging the mea- sured velocity values at the estimated level of 6 in Figures 54 through 58, respectively. The average is for one wave cycle (sampling interval of one second) and represents a single estimate of mean streaming at each station. No statistical significance can be inferred from this method, but the mea- sured results in Figure 60 show that the streaming is insignificant at all slope stations except station 1. Here the net drift is approximately 1.3 mm/ sec. This drift is probably induced by mixing in the nearby (3-5 cm) zone of breaking. As discussed earlier, protrusions of thin fluid laminae slowly advect into the interior over the slope from the zone of breaking. The direc- tion of the drift measured at station 1 is toward the breaker, possibly indi- cating that this flow compensates in part for the mass efflux in the protru- sions. The observed direction net drift near the free surface above this station is also toward the zone of breaking. 122 FIG. 60. Measured values of mean streaming at ij = 6 stations 1 through 4 on slope. Station No. x , cm |", cm u , cm/sec 1 8,7 10.05 +0.13 2 23.0 26.6 -0.07 3 36.2 41.8 -0.003 4 51.0 58.9 -0.023 123 Owing to the poor reliability of the raw streak data very near the wall (<0.07 cm), the best experimental estimate of bed shear stress, t , ' o' was obtained by extending the reliable portion of the velocity-depth curves to the zero point at the bottom. The shear stress was then computed from a simple straight-line fit between the zero point and the lowest point on the velocity curve. This gives a low estimate of the bed shear stress, since the actual shape of the velocity curves in the boundary layer is more exponential. The maximum bed shear stress A- /max measured at each location is listed in column 7 of Figure 59; the time variations in r over o one wave cycle corresponding to the time variations in the velocities in Figures 54 through 58 are shown in Figure 61(a - d). The theoretical values of bed shear stress were computed from Equation (4-24): tq = m au/a^ , rf = 0 where " = 0.01 —62^. cm-sec The maximum theoretical value is shown in column 8 , Figure 59, for comparison with the corresponding measured value. Since the theo- retical estimates were made from the linear Eulerian expression of velocity at each station, they correspond to the experimental Lagrangian estimates only to first order in e. It can be shown that \i varies by only 4 percent over the density range for the particular experimental conditions shown in Fig- ure 53. In spite of the poor comparison of the estimates of boundary layer thickness given by linear theory and experiment, the theoretical estimates of maximum bed shear stress at stations 1-3 compare reasonably well with the measured values. This is apparently the consequence of the higher ex- perimental values of maximum velocity as compared to the first-order theoretical values for levels near the boundary. Figure 61 also shows the excellent fit of the theoretical curve computed from first order solutions to the experimental values at stations 1, 2, and 3. 124 Shear stress, dynes/cm' 7.0- 6.0- 5.0- 4.0- 3.0 2.0 1.0- -1.0 -2X) -3i0 -4.0 -5.0 -6.0 -7.0 O measured — theoretical curve Fig. 61 a. Experimental and theoretical values of bed shear stress t0 at station 1. The experi- mental values were obtained for each second of time t during one wave cycle. 125 Shear stress, dynes/ cm 3.0-1 © me a sured — theoretical curve Fig. 61 b. Experimental and theoretical values of bed shear stress tq at station 2. The experi- mental values were obtained for each second of time t during one wave cycle. 126 Shear stress, dynes /cm 2.5-1 © measured — theoretical curve Fig. 61 c. Experimental and theoretical values of bed shear stress x0 at station 3. The experi- mental values were obtained for each second of time t during one wave cycle. 127 Shear stress , dynes /cm 2 5t time , sec © measured — theoretical curve Fig. 61 d. Experimental and theoretical values of bed shear stress tq at station 4. The experi- mental values were obtained for each second of time t during one wave cycle. 128 Cases B and C: y = c and y > c The description of the boundary -layer motion in this section follows a format similar to that presented in the discussion of boundary activity for case A (y < c). The qualitative remarks deal mainly with the observations and photographs of shadowgraph images, whereas the quantitative results were derived from the measurements of the motion of dye streaks and neutrally buoyant particles. Qualitative During those runs for the critical case (y = c) with moder- ately steep slopes (30 and 45 degrees), the zone of breaking near the cor- ner was characterized by intensive mixing as evident in the shadowgraph images, similar to that described for the high-frequency waves on a 30 degree slope (see earlier discussion for y c is decreased as compared with the experi- mental observations of these processes for the subcritical case. Figures 62 through 64 illustrate in a very striking way that the in- ternal wave motion for these cases (y > c) is associated with the growth of an instability along the slope. The shadowgraph observations of the insta- bility illustrated in these figures indicate that a line of regularly spaced vortices forms near the bottom. These vortices grow and decay over each half cycle as they oscillate along the slope at the frequency of the input waves. It appeared that on the downslope propagation during the wave cycle, the vortices would "flatten", indicating that they form only during the upslope movement of the water particles. This is indicated in Figures 63a and 63b. The circulation of adjacent vortices has the same sense; 129 CD ^~, & tP 0 0) rH T3 W ' 4-i P fi •H W 0) en rd B •H ,C T3 D. 0 rd •rH M U s-~. tn 0) U 5 PM (D 0 cn 13 0) v-* rd > £3 rd w !S 4-1 0 w X! Q) a • TS rd w 3 u a 0 -H 1 a. .— x Cn Q) e 0 cJ > o p ■H rd — & £ CD 0 T) -H -P •H rH d 3 0 a O M-l g H • *JD x: tr> u 3 0) 0 XI u B x: 3 p £ CN . V£> •H • p4 W tr> •H fc o o o o o o oo oo oo oo co oo r- r- o o rH m cn KO VD VD VD VD 130 Fig. 62 a Fig. 62 b 131 Fig. 63 a Fig. 63 b 132 Fig. 64 a Fig. 64 b. 133 this circulation is indicated diagrammatically below. The axis of each vortex is oriented approximately normal to the side walls of the tank, and the maximum distance observed between the core of adjacent vortices was was about 2 cm. Unlike the vortices which form beneath the shoaling, high-frequency waves well upslope and which precede the violent breaking (Figure 45), these vortices retain their cyclic growth and decay pattern for the duration of the input wave motion. Their regular spacing was much shorter than the wave lengths of the input waves, particularly at the middle and lower slope regions. Circulation in Vortices Shortly after the formation of the line of vortices, thin streamers or wisps appear near the cusp of each vortex and penetrate horizontally back into the interior of the fluid (note the thin lighter bands that alternate with the darker bands in Figures 62 through 64). Qualitatively, it appears that fluid of slightly different densities from vertical levels adjacent to each vortex cell is mixed in the cell and returns to the interior as a thin streamer at a new equilibrium level. The result of this process is alternate layers of thin, nearly homogenous streamers and thicker, linearly stratified laminae. 134 Quantitative (1.) Vortices and Layering. The onset of vortex activity for any particular run normally occurs within 10 to 20 wave periods following the arrival of the first wave motion at the corner region of the slope. The initial vortex generally appears somewhere along the middle section of the slope, and the instability spreads gradually toward the corners. The ini- tiation of vortex activity was observed as a function of input frequency for experimental groups 4 and 5; the results are shown in Figure 65. These results indicate that, for this set of experiments, (1) the vortices occur earliest (in time measured from the start of the wave maker) for a> =* u> even though the lower frequency waves have a higher group velocity; (2) there is some w < a; for which the vortices do not appear, and (3) there is a small range of co slightly greater than a> (i. e., waves in the sub- critical range) for which vortices form. The diameters of the vortices were difficult to measure accurately from the shadowgraph images because of the constant process of either growth or decay during each half cycle. From data taken from various photographs, the size of the vortex diameters is largest for co =* co and generally decreases for lower and higher frequencies. Figure 66 shows the size variation of maximum vortex diameters measured at midslope positions during experimental group 4. The corresponding input wave frequencies are listed in column 1. The position of the vortices relative to the bottom also varies over each half cycle of the wave motion. As a vortex grows in diameter it also appears to move away from the boundary; conversely, as its diameter diminishes its distance from the bottom decreases. The maximum distance measured between the center of a vortex and the bottom was approximately 1.2 cm. * It is worth noting here that the alternate motion toward and away from the boundary by a diminishing and enlarging vortex cell gives the impression of a periodic pumping in a direction normal to the bottom. 135 tc t 1.0 1 0.9- 0.8- 0.7- 0.6- 0.5- 0.4- 0.3- 0.2- 0.1 0.8 O o o o o o o o o o 09 1.0 I.I OJc c^- = O. ^s sec tc = 250 sec. delimit the range of observed vortex formation. FIG. 65. Onset of vortex motion as a function of input wave frequency. 136 -1 it), sec T, sec d, cm 0.546 11.5 0.32 0.547 11.7 0.39 0.519 12.1 0.43 0.495 12.7 0.52 0.487 12.9 0.51 0.480 13.1 0.52 0.468 13.4 0.47 0.458 13.7 0.47 0.430 14.6 0.43 FIG. 66. Maximum size of vortex diameters as a function of frequency a) = 0.495 sec c N =0.99 sec -1 -1 137 Shadowgraph images of the thin, horizontal streamers (for example, Figure 62b) usually intensify with time if the wave motion continues. Their initial appearance generally varies with frequency much in the same way as the onset of the vortices. The streamers are usually sharply outlined as shadowgraph images near the vortices, but within short horizontal distances from the slope (order of 2 -4 cm) their images weaken. Although they can be detected out to distances of 20 cm from the bottom from the shading variations on the shadowgraph, the contrast between the images of the streamers and adjacent layers becomes very indistinct at distances greater than 20 cm from the slope. Their deepest observed penetration of the interior was approximately 25 cm, measured horizontally away from the slope. This particular measurement was made after one hour of continu- ous wave motion at the critical frequency for a run during experimental group 4 {a = 30 degrees). The thickness of the streamers during this particular run was 2 mm near the boundary, and the vertical distance separating two streamers was typically 5 mm. Normally the vertical separation between streamers is approximately equal to the diameter of the nearest vortex. The shadowgraph images of the streamers are most distinct for wave frequencies near oj . c Several conductivity probe traverses through short vertical sections (3 and 4 cm) were made in order to detect changes in the linear density- depth profiles that might be associated with the streamers. Figure 67 shows the measured structure for a typical vertical traverse of a probe after the formation of the streamers. Small scale irregularities in the curve (after the wave motion) are typical of these traverses. The values along the horizontal axes represent changes in output voltage relative to the null po- sition that was chosen approximately at the center of the traverse; the vertical axis is actual tank depth in cm. Changes in equivalent density de- termined from the calibration curves are also shown along the horizontal axis. Although sharp changes in density at the edges of the streamer are apparent from the photographs, some conductivity probe and specific gravity 138 -1.0 "0.5 0.5 1.0 Ayo- I03, g/cc Z.Cm -0.5 -0-4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4 0.5 8- 9- 10- 12- I L J 1 I L J I I L AV, volts Fig. 67. Density-depth profile through a region of streamers (numbered) and intervening layers 139 measurements suggest that the density is not homogeneous within the streamer, but has a small, nonzero gradient there. This latter remark is based on only six measurements within two streamers (not shown in Figure 67), and these were taken near the associated vortex cell close to the slope, where the streamer could easily be identified from shadowgraph observa- tions. The apparent density inversion at layer "3" in Figure 67 is not significant. The net movements of neutrally buoyant particles in the streamers and in the invervening layers were observed for several runs. The typical net motion (away from the slope in the streamers and toward the slope in the intervening layers) is indicated diagrammatically in Figure 68. Aver- ages of the net velocity measurements in the streamers and layers are listed in the table in Figure 68. Each individual determination of net mo- tion that went into the computation of this average involved tracking particles for a minimum of 10 wave periods. The differences among the various num- bers are not significant; however, a magnitude of about 0. 1 mm/sec in each zone is reliable. Streamer o .^Intervening Layer Case Average net velocities in mm/sec Streamer Intervening Layer Critical (y = c) 0.14 0.21 Super Critical y > C 0.17 0.13 Fig. 68. Average net velocities in streamers and intervening layers. 140 (2) Boundary- layer description. Owing to the complicated near- bottom motion caused by waves of frequencies at or near critical, it was difficult to choose a level that accurately described the physical extent of the boundary layer. The height above the bottom of the inflection point in the streak curves was selected as an approximate measure of the boundary- layer thickness. Admittedly, the choice is somewhat arbitrary; however, the change in curvature in the streaks was generally well defined at this point during at least the total time of sampling. In any event, this estimate of 5 should be regarded as only approximate. In addition to the values of 6 , other estimates of the boundary -layer thickness are presented in Figure 69 for comparison. The values of s„ and 6„ represent two estimates of the boundary- layer thickness that result from scaling considerations. Wunsch (1969) found that for w^w there is a balance in the governing boundary- layer C -1/3 equation over a distance proportional to R ' . He also showed that the equation is nonseparable in the spatial coordinates. This type of balance -1/3 (R ' ) is represented by <5 0 in column 8. The value of s„ represents a -1/2 boundary-layer thickness of Stokes type (<5 aR ' ) and is presented for comparison in column 9. Firm conclusions based on quantitative com- parisons between the scaling estimates and s are not justified here since 6 is only approximate for each slope station. Similar results for a representative run of the supercritical cases are shown in Figure 70. For these measurements, the boundary- layer motion at each station was very regular, and the dye- streak profiles gen- erally resembled those described for the subcritical case. There was no evidence of vortex activity with the associated generation of thin streamers The estimates of 6j and <5o differ approximately by a factor of 2; this is obvious from their definitions. 141 /y = local wave Reynolds number m R B a u) 6 Iv = local boundary layer Reynolds number hi ¥ (2y/u>)1/2 -1/3 = h R = h R, -1/2 _ (v/N) 1/2 = a k sin a m m FIG. 70. Estimates of boundary layer thickness and Reynolds numbers, y > c. 143 or breaking along the entire slope. Unlike the subcritical case, with slopes of 30 degrees, the displacements of dye streaks and neutrally buoyant par- ticles at station 4 did not show any irregularities or anomalous behavior. The measurements of the boudary-layer thicknesses at each station are considered to be as reliable as those given for the subcritical case (Figure 53). (3) Velocity and Shear Stress. One-second velocity averages were again computed from the streak displacements; for the critical case these values are shown only for one vertical level: that described above for the measurements of s . This level was chosen since it normally was the m J vertical position closest to the slope for which the streak curves remained coherent over the entire wave cycle during sampling. The dye streaks usually became quite distorted above this level (probably due to the motion in the wisps and streamers), and generally the streaks were very faint adjacent to the slope (probably due to mixing). These velocity measure- ments for the various stations on the slope are presented in Figure 71; the experimental conditions at the time these values were determined are listed in the upper right-hand section of Figure 69. The values of velocity in Figure 71 are plotted along the vertical axis in cm/sec, and the successive seconds of time during the wave cycle are shown along the horizontal. The phase has been adjusted artificially so at time t = 1 the velocity is nearly zero. The average distance above the bottom for which these measurements were taken is indicated for each station. This distance is approximately equal to <5 (Figure 69). A sum- mary of the maximum and RMS values of velocity is given in Figure 72 for each station (including the velocity variations on the flat bottom not shown in Figure 71). 144 velocity , |£c - 1.0 -0.5 o-^ --0.5 --I.0 1.0 05 o_ ^-0.5 -- 1.0 1.0 0.5 Station I O G O O -©- o ~0~ G G G o G Station 2 G G G G G -e- G G G G O G © G Station 3 G G G G G --05 -■10 - 1.0 - 0 5 G G G G Station 4 G 0 G G G ^o — e- --0.5 -©- G G G G G G ~i 1 1 1 1- 3 4 5 6 7 time , sec FIG. 71 1 1 1 1 1 1 T" 8 9 10 II 12 13 14 145 The last column of Figure 72 gives rough estimates of the maximum bed shear stress, , I _ u max t max =* a J—L- 1 o1 r 6 where ~ju = 0. 01 gm /cm-sec, that were computed at the level of the inflec- tion point (rf = 6 ) described above. This kind of estimate assumes a laminar velocity distribution that remains essentially linear with distance above the bottom over the wave cycle. This assumption is not accurate for two reasons: (1) the actual profiles of velocity are probably not linear, and (2) the observed dissipation of the dye streaks below 6 indicates pos- sible turbulence in this region. The estimates of It are intended & ' o ' max only as approximate lower bounds on the local maximum bed shear stress; the actual values are probably larger due to a nonlinear velocity distribu- tion with superimposed turbulent fluctuations near the bed. Fig. 72. Summary of velocity and bed shear stress values for critical case (y^ c) measurements. Station (see Fig. 11) 1 u 1 max cm/sec URMS cm/sec rolmax dynes /cm 2 (XlO+2) cm <5m 1 0.65 0.38 1.80 0.36 2 0.68 0.44 1.74 0.39 3 0.72 0.45 1.76 0.41 4 0.35 0.25 0.85 0.41 5 0.18 0.13 0.64 0.28 The velocity measurements show that the magnitude of the maximum oscillatory motion at the selected level of measurement is nearly uniform at the middle and upper slope stations. The value of lul . on the lower slope rc r ■ ' max (station 4) is smaller than that farther upslope (stations 2, 3, and 4) by about 146 a factor of 2. The increase in maximum velocity between the flat bottom and station 1 is approximately a factor of 4. This is considerably lower than that found between similar locations for the subcritical-case measure- ments shown in Figure 59. The measured bed shear stress (as defined above) at station 1 is greater by only a factor of 3 than that for the flat bottom. A striking feature in the results is the smooth sinusoidal appear- ance of the velocity- time curves in Figure 71. The apparent lack of irreg- ularities in these curves suggests that in spite of the vortex- streamer production, the basic harmonic motion of the input waves is still quite recognizable. The Lagrangian velocity measurements for the supercritical case are represented by the results in Figures 73a and 73b. These velocity- depth profiles are analogous to those shown in Figures 54 through 58 for y a x 4-J CL a) X) 6 i i 5>i u •H O O CM i— 1 O ctf « ro t- O i O M fa 148 b ro d CVJ d o i CO d i ro d d I o A c o 4-i td 4-i 0) 4-i Cfl CO 4-1 O S-t a 4-> a a> i 4J •H O O r-H at > ,0 CO O 149 FIG. 74. Velocity and bed shear stress at selected levels t max 1 o1 Station No. V cm 1 'max cm/sec Iu'rms cm/sec (xlO2) dynes/cm 1 0.11 0.18 0.08 1.64 0.26 0.26 0.14 0.38 0.29 0.17 0.45 0.28 0.17 2 0.16 0.32 0.22 2.00 0.30 0.44 0.30 0.45 0.48 0.33 0.66 0.50 0.33 3 0.13 0.20 0.13 1.54 0.24 0.30 0.20 0.38 0.35 0.24 0.53 0.36 0.24 4 0.13 0.25 0.17 1.92 0.24 0.35 0.23 0.37 0.38 0.25 0.51 0.35 0.23 5 0.12 0.05 0.03 0.42 0.21 0.07 0.04 0.32 0.11 0.06 0.46 0.11 0.07 150 DISCUSSION The experimental results can be summarized briefly. The measure- ments of the wave motion in the subcritical range y c, no breaking was seen near the corner. The vortices and streamers persisted until the wave frequency reached low values, below which the vortex production subsided. The motion in the boundary layer was very regular in the absence of the insta- bilities. In the fluid interior the measurements show that the wave motion intensifies near the critical characteristics, whose position is given by z = -ex. 151 5. SEDIMENT MOVEMENT BY INTERNAL WAVES GENERAL DISCUSSION This chapter attempts to assess the role of shoaling internal gravity waves in the initiation of bottom -sediment movement in the ocean, in the light of the theoretical and experimental results discussed in the previous chapters. Although the concept of internal waves as an effective geological agent in the ocean is not new, very few published studies have dealt speci- fically with the erosion, transportation, or deposition of sediments on sub- marine slopes. Chapter 1 has provided some historical background on this subject. Several explanations for the apparent lack of attention are possible: (1) our knowledge of the mechanics of sediment movement beneath the more familiar oceanic surface waves is still incomplete; (2) the wide spatial separation of available data on bottom sediment types on the continental margins hampers understanding of local sedimentary processes; (3) reli- able measurements of oceanic internal waves, particularly those taken in conjunction with a sediment- sampling program, are few; (4) until very recently, theoretical and experimental studies of internal waves passing over a bottom with variable depth were virtually nonexistent. In spite of these problems, there have been inferences to the potential importance of this process (Lafond, 1962; Hulsemann, 1968). There exists a body of internal-wave measurements (see Chapter 1) which indicate that internal waves propagate shoreward normal to the bottom contours. (Wunsch, 1969, has also shown theoretically that these waves can refract, in the sense that their crests tend to align parallel to the bottom contours.) These measure- ments, together with the results of the previous chapters, promote the fol- lowing discussion about the interaction of these waves with bottom sediment. gfc Defant (1961) gives two illustrations of the experimental work of Zeilon (done in 1934) in which interfacial waves incident onto a model shelf- slope configuration were studied. 152 The first section in the following discussion presents various examples of internal-wave measurements in the ocean and discusses quali- tatively some features of sediment distribution on the continental margins, particularly off the northeastern United States. The purpose of this section is to provide some background material on ocean conditions relevant to the later discussion and to demonstrate the wide range of values of the various hydrodynamic and sediment properties. The second section develops a criterion that describes the initiation of bottom -sediment motion. This criterion is basically an extension of the analysis developed by Eagleson, et al (1958) for incipient motion of discrete bottom- sediment particles in- duced by surface waves. The analysis for shoaling internal waves over a constant bottom slope is then specialized for the case of small slopes (y < which is shown to be particularly appropriate for high-frequency internal waves in the ocean. Another criterion (the Shields curve, with a later extension by Vanoni, 1964), which represents the more traditional approach to the problem of incipient motion of bottom sediment and is based primarily on dimensional analysis and empirical results, is also discussed. This latter criterion is shown to be approximately equivalent to the first. The analysis is then applied to specific oceanic measurements, and the results are discussed in a final section, in light of the various limita- tions. Qualitative remarks concerning the effect on sediment movement by breaking internal waves and critical slope conditions (y^c) are also given in the last section. Throughout the subsequent discussion, consideration of the interaction between internal waves and sediment particles is only an initial attempt to evaluate the possibilities of incipient motion induced by this interaction; the actual importance, or even existence, of this process under real conditions must await future field studies. Oceanic Conditions Chapter 1 provides several references to actual measurements of internal wave motion in the ocean, including studies of the directional 153 properties of these waves (for example, Summers and Emery, 1963). In most cases the measurements were made either by recording fluctuations of temperature or horizontal velocity at fixed positions from a moored buoy or anchored ship or by sampling the vertical distribution of temperature with devices like the bathythermograph. Since the frequency content of the internal wave field is not known a priori, a common means of distinguishing the contributions to the measured changes by the various frequency compo- * nents is the energy spectrum. Figure 75 illustrates kinetic energy spectra at several levels computed from horizontal velocity data taken at site TT> (location shown in Figure 76) and described by Fofonoff (1969). The large values of spectral energy density at the inertial and tidal frequencies are typical of this location. Frequencies and amplitudes of vertical tempera- ture fluctuations have been measured by others (Chapter 1). Figure 77 demonstrates the wide range of amplitudes and frequencies represented in previous measurements of internal waves in the ocean. The measurements by Gaul (1961) and Ufford (1947) taken over the outer continental shelf and continental slope suggest that wave amplitudes of 1 meter to 5 meters are not unreasonable for the higher frequency internal waves. * Time series of measurements of vertical velocity or temperature fluctua- tions can be used to compute frequency spectra of potential energy. Mea- surements of horizontal velocities like those taken at side MD" (location given in the following text) are the basis for kinetic energy spectra. Fofonoff (1969) discussed the spectral ratio of potential energy to kinetic energy in terms of a linear internal wave model and found that this ratio is approximately unity for low frequencies (near inertial frequency), but ap- proaches infinity for co -* N. He noted that this latter behavior is probably caused by the exclusion of the advective terms in the linear model. Voor- his (1968) measured vertical internal motions near site "D" with neutrally buoyant floats. After comparing the frequency spectra of potential energy and vertical kinetic energy (obtained from the measurements) with the spec- tra of horizontal kinetic energy computed from the site "D" records, Voorhis found that there is an approximate equipartition of potential and kinetic energies for frequencies between inertial and Brunt- Vaisala. He interpreted this equipartitioning to indicate the predominance of internal wave motion in the measurements. 154 10 SITE "D" (39'20'N, 70* W) 2143 1004m 0CC.1966 (Temp.) 2204 2203 51 1 m 101 Jrn #*•. Feb, i96r reb.i9S7 4500 (velocity) (velocity) cm /sec CO2 2204 I 1 !:| jfs 2 20 3 2143 10 -2 10 -4 10*2 10( FREQUENCY ( CyC^ S /hour ) Fig. 75. Frequency spectra of horizontal kinetic energy density at site "D". 155 T3 C (0 -H tr> C w 0) c i ■C -P id CQ V£> 0> •H Cm 156 FIGURE 77 INTERNAL WAVE MEASUREMENTS ON CONTINENTAL MARGINS Source Water Depth (meters) Wave Height (meters) Wave Period Celerity* (knots) 1. Lafond (1963) 20 7 (max.) 7 min. to 12 hr. 0.11-0.6 2. Boston (1964) 20, 33 8 12 hr. 3. Gaul (1961) 62 1 - 3 10 min. 0.5 - 1.5 4. Ufford (1947) 40, 100 1 - 30 9 min. - 2hr. 0.08-0.68 5. Emery (1956) 1350 130 - 200 12 hr. (standing) 6. Lee (1961) 20 3 - 7 5-15 min. 7. Summers and Emery (1963) 10 - 1500 30 12 hr. 1 (shallow) 7 (deep) *In all cases, the dominant direction of propagation was shoreward. 157 In order to describe internal wave motions in the ocean, an estimate of the vertical distribution of Brunt-Vaisala frequency is needed. The aver- age vertical distribution of Brunt-Vaisala frequency for locations seaward of the continental slope (near site "D", for example) is typically character- ized by relatively large values of N in the seasonal thermocline, if present, and in the deeper main thermocline. Figure 78 shows measured values of Brunt-Vaisala frequency N versus depth for stations near site "D"; the hydro- graphic data were provided by Volkmann (personal communication). An idealized distribution of Brunt-Vaisala frequency with depth is also shown in this figure. The idealized case consists of three distinct vertical layers; in each layer the density gradient is assumed to be constant (i. e., N ^ constant). This hypothetical vertical distribution of N is later used with the idealized topography also shown in Figure 78 as a basis for various models of sedi- ment movement induced by internal waves. The generalized topographic cross-section in Figure 78 is not con- strued to be representative of any particular area, but is taken as an idealized case in which the slopes are linear and the boundaries between the three provinces (shelf, slope and rise) are well defined. Each idealized spatial province in this figure is similar to the experimental conditions described in previous chapters in the sense that N and y are constant in each. The actual bottom topography of the various provinces comprising the continental margins of the oceans not only varies markedly with geo- graphic position but also can be quite complicated locally. Featureless and even bottom slopes are probably rare. In his summary of the continental shelves and slopes of the world, Shepard(1963) remarks that it is exceptional to find a shelf that has a constant slope out to the shelf edge; constancy of slope usually implies widely spaced sample points. The values for bottom slopes shown in Figure 78 are recommended averages taken from Shepard (1963) and Heezen, et al (1959). More detailed descriptions of the physiog- raphy, structure, and sedimentology of the continental margin off the east 158 i 1 o rtf ^ o O 00 O "-H r- C C C P +J • «J g (fl &ig E i-3 C ^ 0 53 0 iH U g J 0 fQ 0 > -H (D ^ «■— ' fd N 4-J S C ■H 5 0 ^— » H id - - 03 c fd -P c r- u 0 0) KJ . . Q) •H T3 T3 C •H "C C fO H nJ 2 •p o < • C ■H tr >-i • fd C r » o H £ fd w -p c c i 0 o C c T5 i 0) +J N c •H 3 H u fd m 0) t) m H 0 • CO r*- 6 M L^ 159 coast of the United States is presented by Uchupi (1963, 1968), Emery (1966) Pratt (1968), Schlee (unpublished manuscript), and Emery, e_t aj. (1969). At this point it is instructive to compare average values of the oceanic slopes y with hypothetical estimates of the slopes of the internal wave characteristics c in order to determine which case (y c) in Figure 1, if any, is predominant. Figure 79 summarizes the relative values of c and y for typical values of Brunt- Vaisala frequency N, wave frequency a>, and a constant value of Coriolis frequency (f = 0.042 cph). The results show that the continental shelves are approximately critical (i. e., y a c) for internal wave frequencies near inertial for each of the values of N shown; for frequencies greater than the frequency of the semi-diurnal tide u) — , the shelf is subcritical. The assumption that the average gradient of the continental shelf is small compared to the slope of characteristics c for normally incident internal waves (i. e., y <0. 1 cph and N > 1. 0 cph, orifa)>0.5 cph and N = 6. 0 cph, then y < c for oceanic conditions; this is done later and only touched upon here. The ex- perimental results show that for internal wave frequencies equal to or slightly less than critical, vortex instabilities form along the slope. If this occurs along submarine slopes, the mixing in these vortices, with consequent formation of thin streamers that might advect fluid horizontally 160 .001 Characteristic slope c (ordinate) computed for various frequencies w and selected Brunt Vaisala frequencies /V Average oceanic slopes Y are also shown. 161 back into the interior, might provide a means of stirring the fluid near the bottom and moving the very fine suspended matter seaward. This agitation might inhibit or prevent deposition of fine suspended matter, and thus main- tain a relatively turbid and well-mixed fluid layer near the bottom. How- ever, it is dangerous to extrapolate the laboratory observations to unobserved natural phenomena. The intention of the above discussion of average gradients of the oceanic slopes was to present the gross features of these slopes in relation to an idealized hydrodynamic structure (Figure 79); local bottom features such as animal markings, sediment ripples, depressions, and rock outcrops were intentionally left out of consideration. It is realized that these additional features might locally affect the internal wave motion by causing complicated reflection patterns of the internal wave characteristics (Longuet-Higgins, 1969) and by influencing the flow in the boundary layer. The actual scales of bottom roughness on oceanic slopes are quite variable. Spatial variations of these scales have been illustrated by several excellent photographic and echo-sounding studies of submarine topography. A series of camera stations spaced two miles apart on a line extending south of Martha's Vineyard and lying approximately normal to the bottom contours was described by Northrup (1951). Sediment samples were taken at most of the camera stations. Northrup noted that there was good evidence for the action of bottom currents on the outer shelf. Others (for example, Uchupi, 1963) have described topographic features such as sand waves with heights of 5 to 25 meters on the continental shelf. Owen and Emery (1967) cited photographic evidence for intensive erosion of the continental slope south of of Martha's Vineyard by strong bottom currents. They found several un- usual features in their photographs including small masses of sediment that protruded a centimeter or so above the general level of the bottom; in this instance they described the overall appearance as "sand-blasted". Bowin, et al (1967) made an extensive photographic and sediment survey of the western and southern slopes of Plantagenet Bank (near Bermuda) in water 162 depths of 60 meters to 2000 meters. The height scales of the various sand ripples that are clearly shown in some of their stereo-photographs appear to be on the order of 10 cm. (The ripples were observed at various depths, but especially between 950 and 1200 meters.) The effects these local variations in bottom roughness actually have on the internal wave motion have not been determined. In fact, what role internal waves might have in forming or maintaining features such as ripples or sand waves is uncertain, and must await field studies of this problem. The analysis and discussion in the subsequent sections deals only with the simpler problem of a linearly sloping bottom, mainly for reasons of analy- tical tractability. It is probably true that the wide variations in the hydro- dynamic and topographic features of the ocean include this case; its consider- ation might provide some insight into more complex situations. A general discussion of the distribution of sediments on the various continental shelves and slopes is given by Shepard (1963). For the purposes of this discussion, only a brief summary of the sediment types in the area shown in Figure 76 will be given. Uchupi (1963) and Schlee (unpublished manuscript) presented detailed descriptions of the bottom sediments in this region. In his report Schlee noted that although the sediments on the shelf are dominantly quartzose sands that are moderately sorted, many of the sediment samples from the continental shelf (about 20 percent) showed sizable silt-clay fractions. Schlee also noted that there is some evidence for a seaward reversal in the median sand sizes (i. e., fine to coarse in a seaward direction) at limited sections of the shelf break southeast of Cape Cod. This seaward reversal has been found by others. Several studies have indicated that the sediments on the outer shelf shown in Figure 76 are predominantly Tertiary, and that modern sediments are either lacking or the result of reworking of the older material (Donahue et al 1966). The dynamics of the by-passing of the shelf by much of the silt and clay of modern sediments, particularly on that part of the shelf below surface wave base, are not well known. Emery (1966) states that waves 163 and currents on the continental shelf produce enough motion to cause much of the silt and clay of modern sediments to travel across the outer shelf and come to rest in deeper waters. It is suggested later in this chapter that near-bottom velocities induced by internal waves might be partly re- sponsible for the sediment transport in this region. Pratt (1968) described the continental slope off the eastern coast of the United States as a complex feature whose surfac e is generally more irregular on the upper parts . In the area under consideration the m edian diameter of the sur - face sediments is in the silt range; quartzose sands are locally predominant on the upper slope (Schlee, unpublished manuscript). The continental rise has been de- scribed as a depositional apron sloping seaward from depths of about 2000 m to about 5000m at a gradient of between 1:100 and 1:700 (Heezen, etal 1959). The surface is covered with fine well-sorted sands and coarse silts interbeddedwith clays. Emery (1966) suggests that these sediments were deposited by turbidity currents; however, Heezen, etal (1966) maintain that ocean bottom currents flowing parallel to the contours are responsible for most, if not all, continental rise sedimentation. Actual sediment data that are later compared with the analytical results were obtained from the work of Hollister (personal communication), Stetson (1939), Pratt (1968), and Emery and Ross (1968). The geographic locations of the sampling sites are shown in Figure 76. These results are particularly useful, since the sample locations are on lines extending off- shore near site "D". The hydrodynamic data at and near site "D" provide convenient inputs to the models considered in a later section in which incipient- motion criteria are discussed for the various sediment data. Before proceeding to the analysis of sediment movement by internal waves, some consideration of the boundary layer along oceanic slopes would be worthwhile. Studies of sediment movement in rivers and beneath surface waves have indicated the importance of the flow characteristics (laminar or turbulent) in this layer in determining the motion of bottom sediment. Linear solutions due to Wunsch (1969) for the boundary layer beneath internal waves on a sloping bottom were discussed in Chapter 4. Wunsch (1970) also 164 considered theoretically oceanic boundary mixing for quasi- steady conditions in a stratified fluid adjacent to a sloping boundary, and showed that a mean vertical velocity is induced at the sloping boundary to satisfy the no-flux condition. He found that the effect is confined to a boundary layer of thick- ness , vl/4ml/2 8 = {v K) ' /N ' where v is the eddy momentum coefficient k is the eddy heat coefficient N is the Brunt-Vaisala frequency, as before. He also pointed out that the various estimates of the mixing coefficients 2 v , k in the ocean are unreliable, and showed that for v , k ~ 1 cm /sec, -3 -1 and N = 2 X 10 sec , then 6 ~ 20 cm. Attempts to define v for various oceanic processes are abundant (Defant, 1961); however, near a sediment- strewn bottom that is disturbed by periodic and quasi-steady currents, such a definition is elusive. Based on measurements of suspended matter near the bottom of the continental slope east of Chesapeake Bay (Ewing and Thorndike, 1965), Ichiye (1966) computed an eddy diffusivity from a linearized version of the Fickian dif- 2 fusion equation. He arrived at a figure of 0. 12 cm /sec for the eddy viscosity. Others have shown photographically that the bottom along the submarine slopes can be quite turbid; high concentrations of suspended mat- ter in this near-bottom zone might increase both the viscosity and the den- sity of the fluid. By contrast, many photographs show remarkable clarity and suggest low concentrations of suspended material near the bottom. It therefore appears that the local conditions along oceanic boundaries are sufficiently variable to preclude a single, all-inclusive value of v • The above discussion suggests that many processes are operative along the oceanic slopes (currents, animal stirrings, etc.); internal waves are only 165 one possible kind. The actual viscosity near the bottom must be the result of the composite effects of the processes. INCIPIENT MOTION CRITERIA The various approaches to the formulation of analytical criteria that adequately describe the initiation of motion of bottom sediment due to water waves are limited owing to the complexity of the process. Several authors (Inman, 1963; Raudkivi, 1967; Abou-Seida, 1965) have summarized these approaches and discussed some of the limitations of each method. Prior to the development of an analytical model, it might be helpful to present a dimensional analysis of the problem of initiation of sediment motion induced by internal waves, in order to show the large number of independent variables and dimensionless groupings of these variables that are possible. The specific problem is to relate functionally the incipient diameter of the movable sediment particles to the pertinent hydrodynamic, geometric, and sediment variables. It is assumed that internal gravity waves of a given mode number in a continuously stratified fluid propagate over a bottom slope that has a sedimentary bed. The density stratification is approximately linear and stable, and the bottom slope is constant. The bed is assumed to consist of noncohesive particles with mean diameter D ; the roughness is arbitrarily described by a length scale k . Figure 80 shows the idealized geometry and particle relationships; this figure is similar to that given by Eagleson and Dean (1959). Incipient diameter is the largest size within a local size-frequency distri- bution of particles that can be placed into motion by the action of the hydro - dynamic and gravitational forces. The condition of incipient motion induced by surface waves on beaches has been described as follows. At some point on the offshore slope, the instantaneous hydrodynamic forces may become sufficiently large to cause static instability of bed- sediment particles of a given size. The diameter of this size is referred to as the incipient diam- eter (Johnson and Eagleson, 1966). 166 V °r FIG. 80. Definition sketch of sediment particles; forces and geometry. symbol name definition point of application F rG submerged 3 ttD /6 p g (s -sf) center of weight gravity - buoyancy mass 2,2 F F fluid CF pf ttD /8 uf J. resistance dynamic pressure and top VM pressure force virtual mass force viscous shear acting over surface of particle ttD /6 p. a r — o instantaneous pressure center gradient force present in mass fluid whether or not particle is present. Equal to the inertial force of the displaced fluid. CM ttD /6 pf af force required to accel- top erate fluid due to presence of particle of 167 Eleven independent variables are selected a priori as being important in determining the incipient diameter D. for this problem. The functional re- lationship can then be expressed as D. = f(a, u), N, n, Pf, v, h, r, Dm, Pg, K) (5-1) (The symbols are defined in the front section of this paper.) By dimensional analysis the functional relationship can be expressed equivalently in terms of eight independent dimensionless groups: VDm= f(a/h' W/N' n' ^c' awDm/y ' aw/[(s-DgDm]1/2, V) (2) (3) (4) (5) (6) (5-2) ^[(s-Dg/Dj1/2, DmA) (7) (8) ' where s = s /s„. s/ f The first four expressions on the right in Equation (5-2) relate to the * local wave field ; the remaining groups relate either to the wave-sediment interaction or to the sediment properties alone. Dimensionless groups (5) through (8) resemble those developed by Carstens, et al (1969) in his analy- sis of incipient motion due to surface waves. The large number of dimension- less groups indicates the complexity of the process and the potential difficulties in developing accurate dynamic models of this process. Several other vari- ables have purposefully been omitted; these include measures of sorting and packing. These factors would increase the complexity of the process. If the problem is simplified to consideration of the balance of forces acting on a single, protruding bed particle such as shown in Figure 80 , an * a/h is proportional to the local Stokes parameter e , since k =* cmr /k and e = ak (Chapter 4). 168 expression can be derived for the incipient diameter in terms of the local, wave- induced velocity field in the viscous boundary layer, certain resistance coefficients, bottom slope, and sediment properties. Furthermore it is postulated that incipient conditions hold when the sum of all hydrodynamic and gravitational moments acting on the particle at any instant is equal to zero. Such a condition can be represented (similar to Eagleson and Dean, 1959) by Y ! M = 0 = (F sin ^ sin 0 ± FF cos |3 (1 + cos ) /5„v ± FyM (1 + cos (f>) ± Fp cos - FQ sin ( +a)) ^ Brief definitions of each force and its assumed point of application for this model are given in Figure 80. The points of application follow from the discussion in Ippen and Eagleson (1.955) and Eagleson and Dean 0959). Ippen and Eagleson (1 955) give a more detailed discussion of the various forces and their representation. The fluid resistance can be rewritten in terms of components normal to and parallel to the bed; these components are commonly referred to as lift and drag, respectively (Schlichting, 1955): FL = Fp sin 0 = CL Pf ttD2/8 uf2 FD = FF cos 0 = CD Pf ttD2/8 uf2 To obtain an expression for the incipient diameter from Equation (5-3), the following assumptions are made. (1) The magnitude of the acceleration is small compared to the magnitude of the effective velocity (|a |, |af| << |uf |). If the advective accelerations are neglected (this is a linearized model), then the instanta- neous accelerations are approximately equal to u> times the instantaneous velocity. Their neglect is reasonable for internal waves in the ocean (say o)^ 10 sec ). 169 (2) The lift coefficient is approximately equal to the drag coefficient. The actual relationship used in this model (CT « 0. 85 C„) was obtained from Li U Chepil (1958). It is recognized that this relationship was found for particles in a turbulent boundary layer; its applicability to laminar conditions has not been verified. Applying assumptions (1) and (2) above and the lift and drag expres- sions for the fluid resistance in Equation (5-3) and solving for D, the incipi- ent diameter is expressible as u2C 3 f n Di-f-V5,!1-! (5"4) where r = ± (1 + cos ) + 0.85 sin (sg/sf - 1) sin (0 Tff) The sign convention necessitates the use of absolute value for r; if u. is directed upslope the choice of the bottom signs would give a negative value of r. The value of D. represents the largest diameter that can be moved for particular sediment and internal-wave conditions. The quantity r is dependent on<£, s /sf, and a- Since emphasis is S J placed here on noncohesive particles in sea water, the choice of s -2. 65 (quartz) and sf = 1.027 appears reasonable. The density variation in the water column over the continental margins off the northeastern United States is < 1 percent. The angle of repose <£ for both well-rounded and very angular grains of sizes > 0. 3 mm was shown graphically by Albertson, et al (1963). There is strong evidence from the nature of the curves for rounded and angular grains of small diameters (< 0.3 mm) that <£ approaches 30 degrees (± 3 degrees). Eagleson, et al (1958) showed that for the idealized situation similar to Figure 80 this angle can be derived geometrically: tan ^ = s : t-/9 (5-5) [(D/Kr + DA- 1/3] l/l 170 If the diameters of the bed grains are approximately equal to the equivalent diameters of the roughness (D/k m 1), tan § = 0. 53 or <* 28 degrees. If these values of s , sr, and 6 are used and a is assumed to be small, such s f that a << ), Uf CD D. = 2.31 (downslope incipient motion) (5-6a) Uf CD D. = 1.51 (upslope incipient motion). (5-6b) The problem now remains to obtain a value uf in Equation (5-6a) or (5 -6b) that is the maximum velocity which effectively acts at the uppermost point of the grain surface (77" = D.). The boundary-layer velocity at this point can be expressed as the real part of Equation (4-24) and uf = 4^ [{F sine1 - G sine2} - e"V6 X {F sin^ + D./5) - G sin (0 + D^)}] / 2v» \1/2 2 M2 . 2 \a; -N sin of/ (5-7) (5-7a) where 4 cnA ' c - y 61 = qjfn(By) + wt , e2 = qen(Pf) + wt For the case of small slopes in the sense y << c, Wunsch (1969) has shown that q « — — . Furthermore, for distances well away from the corner (cx>> |z |) the coefficient A can be written in terms of the input wave motion 171 a co o 2k also k = CriTT (5-7b) o where a and k are the amplitude and wave number of the internal waves oo over the flat bottom, respectively. This expression is true for an input area whose top and bottom are horizontal and rigid. If the condition that the vertical velocity of the wave motion at the top and bottom of the input channel is approximately equal to zero, then Equation (5-7b) is approximately true for an input channel with nonrigid upper and lower boundaries. It is important to note that these approximations of q and A imply a smooth transition in flow conditions from the input region to that over the slope (i. e., in two dimensions this implies a smooth matching of stream lines at the transition between the two regions). For conditions of small slope (y << c) and a smooth bottom transi- tion, several simplifications are possible in Equation (5-7). For small slope, y « a « sina and for high enough wave frequencies, co/N>> sin ex. In addition, M a c: N =* c c' C so that u„ <* - a omTT -D./6 sin03 - e smie^ + D^/b) (5-8) where 0„ = q£n(c£) + cot ; 6 - (^) CO V2 Since k = -r — and h o h o — y X r £ for small y, then a co uf " cTh/h0) -D./5 sine,, - e sin (0„ + 0^/5) (5-9) with n = 1. 172 After a short amount of manipulation it can be shown that the maximum value of the terms inside the brackets on the right side of Equation (5-9) can be approximated as ^2 D./6 to order (D./6) . For waves of 1-hour -3 l l -2 2 period, co =* 1.7 x 10 rad/sec, and for y ^ 10 cm /sec, 6 ^ 10 cm. Typical sediment diameters on the outer shelf are on the order of 10 cm, indicating that the above approximation is reasonable to about one percent. Since for incipient motion conditions it is the maximum value of the velocity at "rj = D. that is the important quantity, then from Equation (5-9) a a) |u,| a ?u/u c \2 D./6 (5-10) 1 f'max c(h/h ) * v for the following assumptions: (1) small slopes such that y << c and a <<<£; (2) D.<<6. It is also worth recalling here that the results of the laboratory experiments described in the earlier chapters for the subcritical case (y < c) suggest that, at least for the conditions tested, the velocity field in the boundary layer was adequately described by the linear theoretical solutions along the slope to the point of breaking. For any particular case in which y <vk- The problem of selecting a repre- sentative value of y was discussed at the end of the previous section in this chapter. The basis for other choices of viscosity here lies in an "intuitive feeling" that natural oceanic boundary layers are turbulent. The actual viscosity is probably a result of several processes at work, including in- ternal waves. It is worth recalling, however, that the experimental runs with linear, smooth slopes showed no appreciable net transport of fluid in the boundary layer outside of the breaking zone for y c are two possible instances for which a choice of y >y, might be made, at least for the laboratory conditions. The other apparent difficulty in the direct application of Equation (5-6a) or Equation (5-6b) to oceanic conditions is the choice of the drag co- efficient C . Considerable experimental evidence shows that Cn is largely dependent on a suitable Reynolds number R-. for solid bodies immersed in a moving fluid (Prandtl, 1952; Batchelor, 1967). Usually the relationship be- tween Cn and Rn is given for a steady flow past a smooth and regular body such as a sphere in an unbounded fluid. In the case of an irregularly shaped particle resting on a rough bed in an oscillatory flow, the relationship be- tween C and Rn is not well known. Eagleson et al (1958) showed 174 measurements from which the drag coefficients for spheres rolling down a smooth, inclined boundary immersed in water or oil were evaluated from the relation CD pf— Vs = FG Sln a where v is the measured translational velocity of the sphere of diameter D. Their results are shown in Figure 81. Line A in this diagram is the con- ventional curve for C~ °c r for spherical particles moving in an un- bounded fluid (Batchelor, 1967). Eagleson et al (1958) later extended these results experimentally to account for the presence of a rough bottom. Figure 82 shows that for constant values of particle diameter D, decreasing roughness size k (i.e., increasing ratio D/k) causes a reduction in the drag coefficient determined for a smooth bottom from Figure 81. In their studies of incipient motion of bed particles due to surface waves, Eagleson and Dean (1959) chose CD = 19. 2/RD with D/k =* 1 and RD = uf D/y. On the basis of a large number of experiments in which C~ was computed from an equation similar to Equation (5-4) for various measured values of uf, D., <£, a, and s /sf, Eagleson and Dean (1959) concluded that there was no qualitative disagreement between the values of Cn obtained in this way with those values obtained from Figures 81 and 82. It is not obvious, however, that their results show conclusively that the two methods of computing C are equivalent for incipient motion of the particles. On the other hand, their results are quite convincing that the two methods are in agreement for discrete particles that are already in motion. In view of the apparent importance of the local particle Reynolds number Rn on the value of C , it is useful to establish some upper bound on R for the oceanic problem. Sediments on the outer shelf have median -2 diameters on the order of 10 cm. Lacking specific measurements of 175 io O m C^ O m cr. •-{ c fd CO 0 ii ^ cr c u en 4J C ~~~ Q) •H o -H 4-1 14-1 U) 00 •H W b *o ro CsJ O O C\Jw U_° > <|C\J Q^ ii a C J> 00 M Cm 176 1.0- 08- 0.6- smooth R = const 0.4- 0.2- o.i- .08" .06- incipient motion .04 T 2 4 6 8 10 D/k FIG. 82. Effect of bed-particle geometry on the resistance coefficients for spheres (after Eagleson and Dean, 1959) 177 internal-wave velocities in the bottom boundary layer, a value of 100 cm/sec * is selected as a crude upper estimate. -2 2 If v = vk * 10 cm /sec, then RD a 100. For larger values of y or smaller uf, RD becomes less than 100. This upper bound suggests that, according to Figure 81, CD » 192/RD for a hydraulically smooth bottom. If the additonal assumption that D/k ~ 1 is made, then C = 192/R (Figure 82). If this value of CD is applied to Equation (5-11) with D u D. Rn = and v ^ 8 co/2, then D/*=31-36iSTE7y (5"12) for incipient downslope motion on small slopes. It is obvious from Equation (5-12) - and not unexpected - that the local incipient diameter is a function of the boundary layer thickness (or alternatively, the viscosity v) as well as the local depth and input wave conditions. The value of D. increases linearly with decreasing depth, sug- gesting progressive sorting of sediment. Equation (5-12) predicts that sed- iment particles having diameters smaller than D. will initially move down- slope so that there should be a trend toward smaller mean diameters in the seaward direction. Since the model does not describe the subsequent history of bottom particles once they are initially moved from their roughness hol- lows, it is not possible from this analysis to trace their net motion after initial movement. However, one might speculate that once the particles are set into motion they might be advected to different locations as bed load *A few values of instantaneous near-bottom velocities on the continental margins are available. For example, Emery and Ross (1968) reported a maximum value of 70 cm/sec on the continental slope south of Martha's Vineyard; the flow direction was parallel to the bottom contours. 178 if there is net motion of the surrounding fluid. The net motion of the fluid need not depend on mixing induced by internal waves, but can be caused by other currents. It is emphasized that this analysis pertains to initiation of motion of noncohesive bed particles for the conditions assumed; subsequent motion as established bed load or suspended load is only speculative. The inverse relationship between the ratios D./6 and h/h can be conveniently illustrated for different choices of a , N, and oj; Figure 8? is an example of this relationship. Each straight line in the diagram represents a particular input wave amplitude a for internal waves that o are assumed to be normally incident to the slope. The larger values of a are permitted only for h large in the sense a /h << 1. (The linear in- ternal wave theory used in this analysis is valid only for small -amplitude waves.) Finite-amplitude effects have not been considered explicitly. Con- sequently, the results are valid upslope to a point where the Stokes param- eter e becomes order one; the exact value of e for this limiting condition is not precisely defined. A discussion of e for the small-slope approxi- mation is given in Chapter 4. The experiments show that for y < c, the high-frequency waves induce a zone of breaking for values of e > 0. 5. A value of e > 0. 5 might be used to establish a limiting depth ratio h/h for this model; i. e., at depth ratios shallower than h/h the linear analysis is invalid. If a = mh where m << 1 and positive, then from Equation (4-14) h/hQ = (f ctt)1/2 (5-13) where mode n = 1. For example, for relatively high-frequency, first-mode internal waves, such that that that w* N/2 and a = 0.01 h (m = 0.01), it follows from Equation (5-13) / -2\1/2 179 o.ooi- 0.000 0.00001 QQ- 0.002 FIG. 83. Ratio of incipient diameter D. to boundary-layer thickness S for various positions on the slope, qiven by h/h , and for selected amplitudes a.f of :he o incident internal waves. 180 If e = 0. 1, h/hQ * 0.4; if e = 0.5, h/hQ a 0.2. Figure 83, shows the limiting depth ratios derived in these examples. Before discussing the results of this model in terms of oceanic conditions it is instructive to show that the form of Equation (5-12) is in some sense compatible with the familiar Shields parameter S, S " (*sVl>Vs (5"14) where D is equivalent to D. and r is the shear stress necessary to pro- S 1 c duce initial motion of noncohesive bed particles in a steady flow. Shields (1936) presented experimental results relating this parameter to a partic- ular Reynolds number. Others (for example, Vanoni, 1964) provided sub- sequent experimental results that established a single curve relating this Shields parameter S to the particle Reynolds number R*n, where R* = S, withu* =\Jr/pf D v The diagram in Figure 84 shows the classical Shields curve with some later experimental results plotted on it; the Shields parameter is the ordi- nate, and the particle Reynolds number is the abscissa. Two basic prob- lems arise in the application of this curve to the present problem: (1) the experimental measurements represented by the curve were made for boundary layers that were fully turbulent; and (2) in most cases the mean flow conditions were steady (channel flow). White (1940) derived a similar semi-empirical result for the critical stress at the bed for steady flow condit ions and a laminar boundary layer: S = 0.18 tan 0 (5-15) where S is given by Equation (5-14). According to White, tan is about unity which gives S « 0. 18, as compared to about 0.03 to 0.06 for S ob- tained by Shields (Figure 84). White found, however, that the value of S 181 i - 1 o o -fi ■ r o o o* m> «0 ■ * ■ ^» O ■ * o ■ c • • II i 1 1 1 1 •<* H 1 ' t {Tf 1 1 «. 1 ._+!: _:[.. ■H 1 1 p 1 J 1 — i .... r — t — T i rl- c . 0 c 1 1 i 1 | l ! '< i > i 1 ' i ; • ! ! ; CD ! i i i 1 i i M-l 1 ill _. __ 4-J-J • i u 1 1 r— T-i— • " — ts:~:l4Tixiii~': ~> &i 4- „ (T (0 1 ! i 1 J i _-»._JIUJ £ r'i1 tf- 3 •H T3 - UI CO Tin i i D I ! ! !<> • J 10 •0 i i M 1 |7T^ O . . 1 i m | ' : 1 i » H 0) •H n : i ! 1 1 I ! i i ; I _l 'l => ] — o '■ i "T *'V 7 CO 5. {. • i v. ; ■ 00 i I i ! ' j i __ ! ._. •". jjj_u j_l£> • 1 * ! - p lr ! 1 ' H i CO — o— | h;rrt / i ! l! J? f i , 1! 1 1 1 i 1 ! i E o a *0 h- tf"> tf"> «"> lO lO ic c io ;' o i / S\ — — c»T UJ — -~ — e. 2 .t<< 1 i".'.- ■ 0 •% • 5? • * o o ■o o o * o c c o FULLY DEVELOPED TURBULENT VELOCITY PROFILE cc UI >- o o 182 for turbulent flow in the boundary layer was about one-half that for laminar conditions. The remaining discrepancies in the results of White for turbu- lent flow with those of Shields were attributed to the more intense turbulence in the flows with fully developed velocity profiles which Shields used (Vanoni, 1964). The difference between the two results according to Vanoni was that for both turbulent and laminar flow conditions in the boundary layer, White found that t was apparently independent of R*n- By contrast, Figure 84 shows that t is dependent on R* to values of R* near 10 where the Shields parameter approaches a constant value of about 0.06. The mean conditions of steady flow that are implicit in Figure 84 are not unreasonable for low-frequency internal wave phenomena, since the advective accelerations are small, as was assumed in the earlier model. The compatibility of the earlier model with White's equation for laminar flow conditions (Equation (5-15)) can be demonstrated simply. If the bed shear stress is written as t = n au/a^T and if au/a rj" is approximated by a linear fit: au/aii" ~ Au/6 at au>/c6 1/2 where Au * aco/c, then for 6 = (2y/u)) ; (i. e., small slopes) and ju = v pf the bed shear stress can be rewritten approximately o t = ypf act)/c6)=(6/2)aco /c . (5-16) On small slopes it was shown earlier that a = a h /h (correct to first order in e ). (5-17) Combining Equations (5-16) and (5-17) it follows that 2 183 Applying this expression for bed shear stress into White's result for incipient motion in a laminar boundary layer (Equation (5-15)) an expres- sion is obtained for D./s: 2 aoa> V6 = K gc(hAo) (5-19) K = 2. 78 " (Pg - Pf) tan 0 The functional form of this result is equivalent to Equation (5-12). A closer comparison between this equation and Equation (5-12) is not justified because: (1) Equation (5-17) does not include a Reynolds number dependence; (2) Equation (5-19) assumes a linear velocity profile in the boundary layer; and (3) Equation (5-19) does not specify the point of application of the effective velocity acting on the bed grains. A brief summary of the preceding analysis might be helpful here. The foregoing model deals with conditions of incipient motion of noncohesive bed particles induced by internal gravity waves shoaling over a small linear slope. The static equilibrium of these particles on the bed was defined in terms of an incipient diameter D.. The analysis predicts that the net in- stantaneous hydrodynamic and gravitational moments acting on the particle about the contact points with other bed particles will cause those particles with diameters less than D. to move downslope initially. Equation (5-12) relates D. to the internal wave parameters (a , qj, c), a local depth ratio (h/h ), and the boundary layer thickness (s ). It is also shown that Equa- tion (5-12) is compatible with the results of Shields (1936) and White (1940). MODEL ANALYSIS AND SPECULATION As defined above, the criterion in Equation (5-12) essentially defines the condition of neutral equilibrium set up by instantaneous hydro- dynamic forces of internal waves and that of gravity acting on discrete bed particles that are resting on a rough bed. A different dynamic balance is 184 needed to define a condition of equilibrium for particles that have already been placed in motion. This balance must depend on the net transport cur- rents influencing the particle motion, including any net currents that might be induced by the internal waves, and the downslope pull of gravity. How- ever, as was indicated in Chapter 4, steady transport of water particles induced by internal waves, near a sloping boundary in a stably stratified fluid, other than that caused by diffusion, theoretically cannot occur in a plane normal to the depth contours unless other processes of mixing are active. Consequently, it is not possible to predict a type of oscillating equilibrium for onshore-offshore motion of sedimentary particles for which D< D.. In the case of surface waves, Ippen and Eagleson (1955) and Eagle- son and Dean (1959) have shown that a balance between the forces on a par- ticle due to the mass transport current in the bottom boundary layer, directed upslope, and the downslope component of gravity creates a situation where sediment particles of increasing size in the onshore direction are in oscil- lating equilibrium. The diameter of the local equilibrium size is often referred to as the "null" or "equilibrium" diameter D . The location of this type of equilibrium for a particular sediment size is called the "null" point. Miller and Zeigler (1964) and Eagleson and Johnson (1966) discuss sediment sorting by surface waves and give some comparisons with field data. The only available direct evidence for net fluid transport induced by internal waves on a slope are the laboratory results that were presented in Chapter 4. In view of the absence of other evidence for the existence of mass transport by shoaling internal waves, only the effects of the incipient condi- tion which was derived in the preceding section are considered here. Curves similar to that shown in Figure 83 can be used to determine the hypothetical distribution of D. along a linear bottom slope if 6 is known. Suppose that the initial size-frequency distribution of noncohesive bed particles is given by the solid curve in Figure 85 at each position along the slope. After the internal waves have acted to produce incipient motion of sediment having 185 Frequency nit ia I distribution sma Sediment size large FIG. 85. Hypothetical size-frequency distribution of sediments on an oceanic slope. 186 diameters less than D. at each depth according to Equation (5-12), the final shape of the size-frequency distribution at each location might eventually have a spike at D. and appear as shown by the broken line in Figure 85. The sediment size at which the spike occurs decreases in the offshore direction. It is unreasonable to require this simple analysis, with Equation (5-12) as its tidy result, to explain the complicated sedimentary pattern found on the continental margins. Instead, it is instructive to postulate various simple models to determine at what locations and for which hydro- dynamic conditions, if any, the predicted values of D. according to Equa- tion (5-12) might exceed the observed natural sediment diameters. Models are constructed for three provinces defined by the idealized bottom geometry and the simplified distribution of Brunt-Vaisala frequency shown in Figure 78. Each layer (i. e., the three fluid layers intersecting the shelf A, slope B, and rise C) with constant N is assumed to support small-amplitude, single- frequency internal waves of the first mode that propagate shoreward at approximately normal incidence to the slopes. Furthermore, only examples that satisfy the condition y << c were con- sidered, except for the hypothetical instance of waves with semi-diurnal frequency over the continental shelf and continental rise. In these cases, it is shown that the ratio y/c is on the order of one-tenth. Values of the pertinent hydrodynamic parameters are listed on each diagram. In addi- tion, since the depth range over which the analysis is valid is limited to small values of e (as discussed in the initial section of this chapter), the depth ratios h/h at which e = 0. 1 and e = 0.5 are indicated on each diagram. It should be recalled that the experiments for y < c showed breaking of in- ternal waves at e > 0. 5. On this basis, it was suggested that the linear solutions might be invalid if local values of e exceed 0.5. The sources for the sediment data are listed in each diagram; the approximate locations of the sampling areas for each source are shown in Figure 76. 187 Figure 86 treats a continental- shelf model with relatively high- frequency internal waves (co m N/2, N = 6 cph); the pertinent assumptions in the model are specified in the figure. The ordinate is mean diameter Dm (for actual sediment data) or incipient diameter D. (computed from Equation (5-12)); the horizontal axis is the depth ratio h/h . As mentioned above, values of h/hQ computed from Equation (5-13) for two selected values of e and for the particular wave conditions (a , w, c) of the model are marked long the horizontal axis. The variations of D. with h/h are shown i ' o as theoretical stright-line curves for selected values of H: H =-2- A >/2 o k where a is input wave amplitude in meters, v is viscosity in cm /sec, A = 1 m, o -2 2 v, = 10 cm /sec. (Ao„k1/2 = 10 cm2 • sec"1/2.) For example, the curve in Figure 86 labeled H = 5 represents 1/2 2 -1/2 a v =50 cm • sec ' , or if v - v^, then a = 5 meters. This kind of log-log plot simplifies presentation of the analytical results; the theoreti- cal curves are straight lines (Equation (5-12)), and D. can readily be obtained for various combinations of a and v (or 5). The field data of mean sedi- o ment diameters in Figure 86 show a general trend toward decreasing sizes in an offshore direction. This trend reverses for h/h > 0.4 in Stetson's data and for h/h > 0. 5 in Hollister's data. Interpretation of the relationship between the actual mean values D and the theoretical distributions of D. was described earlier and will be 1 summarized here. D. represents the maximum sediment diameter that is in 188 h/ho Model pqrometers N = 6 cph u) - 3 cph c = 0 577 f = 0.062 cph A0 = I m uK - I0~2 cm2 /sec H = a0 v 1/2 W2 u,2-f2 -i 1/2 LN2-a,2 Sediment data O Ho Mister A Stetson # 4 FIG. 86. Continental shelf model; relatively high frequency waves . 189 neutral equilibrium at slope positions given by h/h for hypothetical internal- wave types and idealized bottom conditions. The analytical model predicts initiation of motion of bed particles in a downslope direction if D < D . m ^ i In Figure 86 incipient motion of measured sizes D is predicted for H > 50 (for example, a = 5m,!/ = 1 cm /sec) and h/h > 0.48 (for the cut-off condition imposed by e = 0. 5). If H = 5 (a = 5m, v= v-J) and if O K. e = 0. 5 is the limiting value for the linear analysis, then all values of D m from Stetson's data for 0. 25 < h/h < 0. 45 could be placed into motion by the waves. It might be noted here that the nonlinear effects which are sig- nificant farther upslope (where e is order one, say) could make the case for possible sediment movement even stronger. For instance, the experiments for y < c showed that the measured velocities in the boundary layer near the zone of breaking were significantly larger than corresponding measure- ments taken farther downslope. Consequently, one might expect that the presence of nonlinearities in the velocity field might increase the values of D. for positions well upslope. The consideration here of a strictly linear analysis is in this sense a "worst" case: if initiation of motion of measured diameters is predictable from the linear model, then the inclusion of non- linear effects strengthens the argument. We might also recall that values of v >v* are included mainly on the supposition that the oceanic boundary -layer dynamics might involve an eddy viscosity. Figure 87 is another shelf model similar to Figure 86 except that the internal-wave frequency is assumed to be semi-diurnal. In this model, the requirement y/c <<1 is necessarily relaxed to y/c 1 cm /sec (see H = 50, for example). The values of D^ along theoreti- cal curves H = 1 and H = 5 are below the mean diameters; this suggests that waves of 5-meter amplitudes with v = v, will not move the measured mean sizes. 190 Model parameters N = 6 cph H = aQ v cu= 0. 083 cph c = 0.01 f = 0.062 cph c = ~u>2- f2~ 1/2 Ao= 1 m _N2-u,2 i/R = I0~2 cm2/sec h/ho Sediment data O Hollister A Stetson # 4 FIG. R7. Continental shelf model ; wave of tidal frequency (semi-diurnal) . 191 The continental slope model in Figure 88 assumes relatively high- frequency waves (a = N/2; N = 3 cph) because critical conditions (r * c) are approximately achieved along the slope for waves of lower frequencies, such as the frequency of the semi-diurnal tide. The data shown in this figure were obtained from various sources; the lines of sample locations are shown in Figure 76. Initial movement is predicted for sediment along the lower extremity of the slope (h/h > 0. 8) for H = 10. For H = 50 (a = 5m, 2 v = 1 cm /sec), initiation of sediment motion is possible for each of the measured values along the slope. A smaller number of data points are shown in the continental rise model in Figure 89. The available values of mean diameter tend to be in the range of fine to very fine silt; the ordinate (D. or D ) has been scaled accordingly. It is worth recalling the earlier discussion of sedimentary types on the continental rise. Several previous workers (Heezen, et al 1966, for example) noted that clean quartzose sands and coarse silts inter- bedded with the clays can be locally dominant. These larger sizes are not represented in the data plotted in Figure 89. It is obvious from Pratt's data, however, that the mean diameters found on the lower parts of the rise are generally larger than farther upslope. The computed incipient diameters exceed the mean diameters for H = 5 (a = 5m, y - yA on the upper rise. O K. Initial movement of the larger measured sizes on the lower rise requires larger wave amplitudes (or larger viscosity), such that H > 50. The effects of cohesive forces acting on the smaller sizes (very fine silts and clays) have not been considered; these forces might inhibit initial movement and thereby require larger values of H for incipient motion. The results of this section can be summarized briefly. It has been shown that, within the constraints of the models, the predicted values of incipient diameters D. can exceed the values of mean diameters D mea- r i m sured at various positions on the continental margin southeast of New England. The condition for incipient motion of bed particles induced by internal waves shoaling over a small slope whose average gradient is linear is given by 192 Model parometers N= 3 cph oj= 1.5 cph c = 0.5768 f = 0.062 cph Ao= I m v*- I0-2 cm2/sec H = QoV C 1/2 AoV* 1/2 C = a,2-f2 nl/2 LN2- W2J e h/ho Sed iment dc ita 0 Stetson # 5 □ S t etson # 6 A Pratt O Emery and Ross (averac, e of 3 samples ) FIG. 88. Continental slope model; relatively high-frequency waves . 193 \ \s \ \ ■ \ Di \ \H=50 ■ \ \ \ \ \ \ \H=I0 \ \ \ \ \ \ \ \ \ \ \ \ \ \ H=5 \ - 0.1 \ \ \ \ o \ - 0 \ ©\ 0 \ . \ H=l \ © \ - \ \ \ \ \ \ _ 0.05 © \ \ \ - \ \ \ - \ \ \ symbol h/ho € a0 , m \ \ \ \ a 0.0 1 1 0.5 I b 0.025 0. 1 I c 0.025 0.5 5 d 0.057 0- 1 5 \ \ 0 05 1 , ■ , 01 1 0.5 . 1 1 b,c Model parameters N = 1 cph a> = 0.083 cph c • = 0.056 f = 0.062 cph A0= Im H = c = a0z/ 1/2 A0i/, 1/2 h/ho Sediment data © Pratf i/ = 10 2cm2 /sec a;2-f2 -l 1/2 N2-oj2 FIG. 89. Continental rise model; waves of tidal frequency (semi-diurnal) . 194 Equation (5-12). It is obvious from Equation (5-12) that larger values of wave amplitude and viscosity produce larger incipient diameters at specified values of h/h . Incipient motion of the measured sediment is predicted for specific instances in the models considered above: (1) on the outer continental shelf (h > 50 m, say) with N = 6 cph, to = 3 cph, and H > 5; (2) on the outer continental shelf with N = 6 cph, semi-diurnal frequencies, and H > 50; (3) on the lower continental slope (h/h > 0. 8), for N = 3 cph, a; = 1. 5 cph and H = 10, or on the middle sections of the continental slope withH= 50; (4) on the upper and middle sections of the continental rise for N = 1 cph, oj = 0. 083 cph, and H > 5, or on the lower rise with H > 50. The internal wave amplitudes suggested in the discussion of the various models are representative in that they fall within the range measured in the ocean. The various proposed values of viscosity v (or boundary layer thicknesses 6) are speculative. The available data for specific sediment distributions on the conti- nental margins do not correspond in a simple way to the predicted distri- butions; the natural setting is too complex to fit these idealized models. However, what is not answerable at this time is whether or not the internal waves do in fact initiate motion of bottom sediment: for certain conditions, the models predict that they can. It must be noted that an additional attri- bute of the sediments on the slope and rise that is probably important in increasing the local forces necessary to dislodge the grains is the existence of cohesive forces acting between particles of clay and very fine silt sizes. The incipient-diameter criterion in Equation (5-12) assumed noncohesive bed particles. Other factors, such as bed roughness scales greater than granular, stirring by bottom organisms, bed fluidization, grain angularity, and packing of bed grains, are probably significant locally in the transport 195 of bed particles. Their effects are difficult to treat analytically; future field studies might establish their actual importance. In the light of the experimental results presented above, the effects of two other phenomena induced by internal waves must be considered. Breaking internal waves of high frequency were observed to produce turbu- lent conditions and increased instantaneous velocities near the bottom. The breaking often generated filimentous horizontal layers which penetrated the adjacent flu id interior. Secondly, experimental runs with y> c showed a vortex instability along the slope; the mixing in the vortices caused a ver- tical structure of alternating horizontal streamers and layers within which the net fluid transport was away from and toward the slope, respectively. If these processes occur in nature (there are no definitive field measure- ments of either process) speculation about their possible impact for sedi- ment processes on the continental margins raises some interesting questions. Wherever the breaking might occur on the slope, sediment that is well-sorted and coarse relative to nearby material probably will be associ- ated with this zone. The finer materials might be winnowed out and trans- ported from the area by currents. Hogg and Wunsch (1970) have shown that for sufficiently large values of eddy viscosity and slightly oblique incidence, in theory the breaking internal waves can generate "longslope" currents. Hence, a littoral drift of fine materials similar to that observed for surface waves and supplementing the effects of the contour currents (Heezen, et al 1966) might be possible. If the position of breaking along the slope is a function of time, i. e., if the stratified layer containing the waves shifts its vertical position over a time scale of days or longer, as seasonal or local if. The laboratory measurements suggest that the breaking occurs well up- slope (for r < c). In fact, when the stratified section was beneath a less strongly stratified or homogeneous layer, the breaking occurred near or at the base of this upper layer. 196 thermoclines might, then a time- sequenced winnowing out of fine materials might occur along the bottom over a wide offshore traverse. Deposition of recent sediments entering the area from the continental sources might be prevented, and local size-frequency distributions might be skewed toward the coarser sizes. A dynamic process like this might help to explain the observed scarcity of recent sediments on the outer shelf. Finally, speculation is ventured for the effects of breaking internal waves at the edge of the continental shelf. Suppose that relatively large- amplitude, high-frequency internal waves were generated at infrequent intervals (for example, by the passage of severe surface storms) in the seasonal thermocline when the upper extent of the thermocline was at the approximate depth of the shelf edge. Sudden substantial increases in bot- tom turbidity due to the suspension of bed materials by breaking of these large-amplitude waves might increase the density of the water along the shelf-slope transition to values high enough to generate density currents. This mechanism might provide a means of continuing erosion of the slope face at irregular times and places by relatively weak small-scale turbidity currents. The vortex instabilities and streamers that were created in the lab- oratory, generally for y > c, might also provide a means of increasing suspension of bottom materials, by increased turbulence within the eddies near the bed, and by transporting fine suspended materials away from the slope. The net transport in the streamers might move the finer sediment (clay and very fine silt) away from the continental slope to deeper sites of deposition on the rise or abyssal plain. None of these mechanisms has been observed in the ocean. The case for increased mixing along the oceanic slopes is not new; Munk (1966) and Wunsch (1970) have called attention to this process. It is hoped that these experiments have illuminated some of the hydrodynamic features of shoaling internal waves, and strengthened the case for internal waves as geological agents on the oceanic slopes. 197 REFERENCES Abou-Seida, M.M., 1964, Sediment transport by waves and currents: University of California at Berkley, Hydraulic Engineering Laboratory, Rept. HEL 2-7. Albertson, M.L., Barton, J.R. , and Simons, D. C, Fluid Mechanics for Engineers, Prentice-Hall Inc., Englewood Cliffs, New Jersey, 1960. Barnes, N. F. , 1954, Optical techniques for fluid flow: J. Soc. Motion Pictures and Television Engineering, 5, High Speed Photography. Batchelor, G. K. , 1967, An Introduction to Fluid Dynamics: Cambridge University Press, London. Bohlen, W. F. , 1969, Hotwire anemometer study of turbulence in open- channel flows transporting neutrally buoyant particles: Massachusetts Institute of Technology, Department of Earth and Planetary Sciences, Experimental Sedimentology Laboratory, Rept. 69-1. Bagnold, R.A. , 1963, Part I. Mechanics of marine sedimentation: The Sea, 3, (M.N. Hill, ed.), Interscience, New York, 507-528. Bagnold, R.A. , 1966, An approach to the sediment transport problem from general physics, U.S. Geol. Surv. Prof. Paper 422-1. Boston, N. E.J. , 1964, Observations of tidal periodic internal waves over a three-day period off Panama City, Florida: Texas A. and M. Uni- versity, Department of Oceanography and Meteorology, A & M Project 286-D, Ref. 64-20T. Bowin, CO., Chase, R. L. , and Hersey, J.B., 1967, Geological applica- tions of sea-floor photography: Deep Sea Photography, (J.B. Hersey, ed.), Johns Hopkins University Press, Baltimore. Cairns, J. L. , 1967, Asymmetry of internal tidal waves in shallow coastal waters: J. Geophys. Res., 72(14), 3563-3565. 198 Carruthers, J.N. , 1963, History, sand waves and near-bed currents of La Chappelle Bank: Nature, 197(4871), 942-946. Carstens, M. R. , Neilson, F. M. , and Altinbilek, H.D. , 1969, Bed forms generated in the laboratory under an oscillatory flow: analytical and experimental study: U.S. Army, Corps of Engineers, Coastal Engi- neering Research Center, Tech. Memo. 28. Cartwright, D. , 1959, On submarine sand waves and tidal lee-waves: Proc. Roy. Soc. London, A, 253, 218-241. Chepil, W.S. , 1959, Equilibrium of soil grains at the threshold of move- ment by wind: Proc. Soil Sci. Soc. Amer., 23, 422-428. Defant, A. , 1961, Physical Oceanography, 2, Pergamon Press, Oxford. Donahue, J.G. , Allen, R. C. , and Heezen, B.C., 1966, Sediment size distribution profile on the continental shelf off New Jersey: Sedimentology, 7, 155-159. Eagleson, P.S. , and Dean, R.G. , 1959, Wave- induced motion of bottom sediment particles: Proc. Amer. Soc. Civ. Engrs. , J. Hydraulics Div., 85, HY10, 53-79. Eagleson, P.S. , Dean, R. G. , and Peralta, L.A. , 1958, The mechanics of the motion of discrete spherical bottom sediment particles due to shoaling waves: U.S. Army, Corps of Engineers, Beach Erosion Board, Tech. Memo. 104. Edgerton, H.E., Handbook of High Speed Photography: General Radio Company, West Concord, Massachusetts. Emery, K.O. , 1966, Atlantic continental shelf and slope of the United States: geologic background: U.S. Geol. Surv. Prof. Paper 529-A. Emery, K.O. , 1956, Deep standing internal waves in California Basins: Limnol. Oceanogr. , 1, 35-41. 199 Emery, K.O. , Uchupi, E., Phillips, J.D. , Bowin, CO., Bunce, E.T., and Knott, S.T. , 1970, Continental rise off eastern North America: Bull. Amer. Ass. Petrol. Geol. , 54(1). Emery, K. O. , 1960, The Sea off Southern California: Wiley, New York. Emery, K.O. , 1965, Characteristics of continental shelves and slopes: Bull. Amer. Ass. Petrol. Geol., 49(9), 1379-1384. Emery, K.O. , and Ross, D.A. , 1968, Topography and sediments of a small area of the continental slope south of Martha's Vineyard: Deep-Sea Res. , l!5, 415-422. Ewing, M. , and Thorndike, E.M. , 1965, Suspended matter in deep ocean water: Science, 147(3663), 1291-1294. Fofonoff, N. P., 1966, Internal waves of tidal period: Woods Hole Oceano- graphic Institution, unpublished manuscript. Fofonoff, N. P., 1968, Current measurements from moored buoys, 1959- 1965: Woods Hole Oceanographic Institution, Ref. 68-30. Fofonoff, N. P„ , 1969, The role of the NDBS in future natural variability studies of the North Atlantic: Proc. of the First U.S. Coast Guard National Data Buoy System Scientific Advisory Meeting, U.S. Coast Guard Academy, New London, Connecticut, May 12-14, 50-61. Fofonoff, N. P. , 1969, Spectral characteristics of internal waves in the ocean: Deep-Sea Res. , 16 (Supplement), 58-71. Fortuin, J.M.H. , 1960, Theory and application of two supplementary methods of construction of density gradient columns: J. Polymer Sci. , 44(44), 505-515. Garrison, L.E., and McMaster, R.L., 1966, Sediments and geomorphology of the continental shelf off New England: Marine Geol. , 4(4), 273-289. Gaul, R.D. , 1961, Observations of internal waves near Hudson Canyon: J. Geophys. Res., 66(11), 3821-3830. 200 Gibson, C.H. , and Schwarz, W.H., 1963, Detection of conductivity fluctuations in a turbulent flow field: J. Fluid Mech. 16(3), 357-364. Goddet, J. , 1960, Etude du debut d'entrainment des materiaux mobiles sous Paction de la houle: La Houille Blanc e, 15(2), 122-135. Haurwitz, B. , Stommel, H. , and Munk, W.H., 1959, On the thermal un- rest in the ocean: The Rossby Memorial Volume, (Bolin, B. , ed.), Rockefeller Institute Press, 79-94. Heezen, B.C., Hollister, CD., and Ruddiman, W. F. , 1966, Shaping of the continental rise by deep geostrophic contour currents: Science, 152(3721), 502-508. Heezen, B.C., Thorp, M. , and Ewing, M. , 1959, The floors of the oceans; I. The North Atlantic: Geol. Soc. Amer. Spec. Paper 65, 122 p. Higgins, L.L., 1963, Measurement of temperature, salinity, and velocity of water through electrolytic conductivity measurements: Space Tech- nology Laboratory, Tech. Rept. 8609-01000-RU-000. Hogg, N. G. , and Wunsch, C.I. , 1970, Isopycnal displacement and Lagran- gian current generation by internal waves: Trans. Amer. Geophys. Union, 51(4) (abstract). Hulsemann, J. , 1968, Morphology and origins of sedimentary structures on submarine slopes* Science, 161. Ichiye, T. , 1966, Turbulent diffusion of suspended particles near the ocean bottom: Deep-Sea Res. , 13, 679-685. Inman, D. L. , and Bow en, A.J. , 1963, Flume experiments on sand trans- port by waves and currents: Proc. Eighth Conf . on Coastal Engng. , 137-149. Inman, D. L. , 1963, Part n . Mechanics of marine sedimentation: The Sea, 3, (M.N. Hill, ed.) Interscience, New York. 201 Ippen, A.T. , and Eagleson, P.S. , 1955, A study of sediment sorting by waves shoaling on a plane beach: U.S. Army, Corps of Engineers, Beach Erosion Board, 63. Johnson, J.W. , and Eagleson, P.S. , 1966, Coastal processes: Chapter 9 in Estuary and Coastline Hydrodynamics, (A.T. Ippen, ed.) McGraw- Hill, New York. Jost, W. , 1960, Diffusion, Academic Press, New York. Kalinske, A. A. , 1947, Movement of sediment as bed in reviers: Trans. Amer. Geophys. Union, 2$ 615-620. Keller, J.B. , and Mow, V. C. , 1969, Internal wave propagation in an in- homogeneous fluid of non-uniform depth: J. Fluid Mech. , 38(2), 365-374. Kelley, R.E. , 1970, Wave induced boundary layers in a stratified fluid: J. Fluid Mech., 42(1). Krauss, W. , 1966, Interne Wellen: Part II: Gebnider Borntraeger, Berlin. LaFond, E.C. , 1961a, Boundary effects on the slope of internal temperature waves: Indian J. Meteorol. Geophys., 12, 335-338. LaFond, E.C. , 1961b, Internal wave motion and its geological significance: Mahadevan Volume, Osmania University Press, 61-77. LaFond, E.C, 1962, Internal waves, Part 1°. The Sea, (M.N. Hill, ed.) Interscience Publishers, New York and London, 731-751. LaFond, E.C, 1965, The U.S. Navy Electronics Laboratory's oceanographic research tower: U.S. Navy Electronics Laboratory Rept. 1342. LaFond, E.C , and Lee, O.S. , Internal waves in the ocean: Navigation, 9, 231-236. Lee, O.S. , 1961a, Effect of an internal wave on sound in the ocean: J. Acoust. Soc. Amer., 33(5), 677-680. 202 Lee, O. S. , 1961b, Observations on internal waves in shallow water: Limnol. Oceanogr. , 6(3), 312-321. Longuet-Higgins, M.S. , 1969, On the reflexion of wave characteristics from rough surfaces: J. Fluid Mech. , 37(2), 231-250. Magaard, L. , 1962, Zur Berechnung interner Wellen in Meeresraumen mit nicht-ebenen Boden bei einer speziellen Dichteverteilung: Kiel. Meeresforsch. , 18, 161-183. Manohar, M. , 1956, Mechanics of bottom sediment movement due to wave action: U.S. Army, Corps of Engineers, Beach Erosion Board, Tech. Memo. 75. Martin, S. , Simmons, W. F. , and Wunsch, C.I. , 1969, Resonant internal wave interactions: Nature, 224(5223), 1014-1016. Miller, R. L. , and Zeigler, J. M. , 1958, A model relating dynamics and sediment pattern in equilibrium in the region of shoaling waves, breaker zone, and foreshore: J. Geol. , 66(4). Miller, R. L. , and Zeigler, J. M. , 1964, A study of sediment distribution in the zone of shoaling waves over a complicated bottom topography: Papers in Marine Geology: Shepard Commemorative Volume, (R. L. Miller, ed.), Macmillan, New York, 133-153. Mowbray, D. E. , 1967, The use of schlieren and shadowgraph techniques in the study of flow patterns in density stratified fluids: J. Fluid Mech., 27(3), 595-608. Mowbray, D. E. , and Rarity, B.S.H. , 1967, A theoretical and experimental investigation of the phase configuration of internal waves of small am- plitude in a density stratified liquid: J. Fluid Mech. , 28(1), 1-16. Munk, W. , 1941, Internal waves in the Gulf of California: J. Marine Res. , 4(1), 81-91. Munk, W.H., 1966, Abyssal recipes: Deep -Sea Res. , 13, 707-730. 203 Northrup, J. , 1951, Ocean-bottom photographs of the neritic and bathyal environment south of Cape Cod, Massachusetts: Bull. Geol. Soc. Amer., 62, 1381-1384. Oster, G. , 1965, Density gradients, Sci. Amer., 213(2), 70-76. Oster, G. , and Nishijima, Y. , 1964, Moire patterns: their application to refractive index and refractive index gradient measurements: J. Opt. Soc. Amer., 54(1), 1-15. Owen, D.M. , and Emery, K.O. , 1967, Current markings on the continental slope: Deep -Sea Photography, Johns Hopkins University Press, Baltimore. Pao, Y-H. , 1968, Undulance and turbulence in stably stratified media: Boeing Scientific Research Laboratories, Document DI-82-0742. Pao, Y-H. , 1968, Vortex streets in stably stratified fluids: Boeing Scientific Research Laboratories, Document DI-82-0736. Phillips, O. M. , 1966, The Dynamics of the Upper Ocean: Cambridge University Press, London. Prandtl, L. 1952, Essentials of Fluid Dynamics: Hafner Publishing Co. , New York. Pratt, R.M. , 1968, Atlantic continental shelf and slope of the United States ■ Physiography and sediments of the deep sea basin: U.S. Geol. Surv. Prof. Paper 529-B. Raudkivi, A.J. , 1967, Loose Boundary Hydraulics, Pergamon Press, New York. Revelle, R.R. , 1939, Sediments of the Gulf of California (abstract); Geol. Soc. Amer. Bull., 50(2), 1929. Robinson, R. M. , 1970, The effects of a corner on a propagating internal gravity wave: J. Fluid Mech. , 42(2), 257-267. 204 Sandstrom, H. , 1966, The importance of topography in generation and propagation of internal waves: Ph.D. dissertation, University of California, San Diego, Department of Oceanography. Sato, S. , Ijima, T. , Tanaka, 1963, A study of critical depth and mode of sand movement using radioactive glass sand: Proc. Eighth Conf. on Coastal Engng, 304-323. Schlichting, H. , 1955, Boundary Layer Theory, McGraw-Hill, New York. Schlee, J. , 1969, Atlantic continental shelf and slope of the United States: sediment texture of the northeastern part: unpublished manuscript. Shields, A. , 1936, Anwendung der Xhnlichkeits-Mechanik und der Turbulenzforschung auf die Geschiebebewegung: Preussische Versuchsanstalt fur Wasserbau und Schiffbau, Berlin, Mitteilungen, No. 26. Stetson, H.C. , 1938, The sediments of the continental shelf off the eastern coast of the United States: Papers in Phys. Oceanog. Meteorology, 5(4). Stetson, H.C. , 1949, The sediments and stratigraphy of the East coast continental margin; Georges Bank to Norfolk Canyon: Papers in Phys. Oceanog. Meterol. , 11(2). Stokes, R. H. , Mills, R. , 1965, Viscosity of Electrolytes and Related Properties: Pergamon Press, Oxford. Stride, A.H. , and Cartwright, D. , 1958, Large sand waves near the edge of the continental shelf: Nature, 81, 41. Stride, A.H. , and Tucker, M.J. , 1960, Internal waves and waves of sand: Nature, 188(4754), 933. Summers, H.J. , and Emery, K.O. , 1963, Internal waves of tidal period off Southern California: J. Geophys. Res., 68(3), 827-839. Tareyev, B.A. , 1965, Internal baroclinic waves in a flow around irregu- larities on the ocean floor and their effect on sedimentation processes: Oceanology, 5(1), 31-37. 205 Thorpe, S.A. , 1968, A method of producing a shear flow in a stratified fluid: J. Fluid Mech. , 32(4), 693-704. Thorpe, S.A. , 1969, Experiments on the instability of stratified shear flows: immiscible fluids: J. Fluid Mech. , 39(1), 25-48. Thorpe, S.A. , 1966, Internal Gravity Waves: Ph.D. Dissertation, Trinity College, Cambridge. Uchupi, E. , 1963, Sediments on the continental margin off eastern United States: Art. 94 in U.S. Geol. Surv. Prof. Paper 475-C, 132-137. Uchupi, E. , 1968, Atlantic continental shelf and slope of the United States - Physiography: U.S. Geol. Surv. Prof. Paper 529-C. Ufford, C.W. , 1947, Internal waves in the ocean: Trans. Amer. Geophys. Union, 28(1), 79-86. Ufford, C.W. , 1947, The theory of internal waves: Trans. Amer. Geophys. Union, 28(1), 96-101. Vanoni, V. A. , 1964, Measurements of critical shear stress for entraining fine sediments in a boundary layer: California Institute of Technology, W. M. Keck Laboratory of Hydraulics and Water Resources, Rept. KH-R-7. Vincent, G. E. , 1958, Contribution to the study of sediment transport on a horizontal bed due to wave action: Proc. Sixth Conf. on Coastal Engi- neering, 326-354. Voorhis, A.D. , 1968, Measurements of vertical motion and the partition of energy in the New England slope water: Deep Sea Res. , _15, 599-608. White, CM. , 1940, The equilibrium of grains on the bed of a stream: Proc. Roy. Soc. London, A, 174(958), 322-338. Wunsch, C. , 1968, On the propagation of internal waves up a slope: Deep Sea Res. , 25, 251-258. Wunsch, C. , 1969, Progressive internal waves on slopes: J. Fluid Mech. , 35(1), 131-144. 206 Wunsch, C, 1970, On oceanic boundary mixing: Deep Sea Res., 17, 293-301. Zeilon, N. , 1912, On tidal boundary waves and related hydrodynamical problems: Svenska Vetenskapsakademien, Handlingar, 47, 1-46. Zeilon, N. , 1934, Experiments on boundary tides: Goteborgs Hogskola, Oceanografiske Institutet, Middelanden, B, 3(10), 1-8. 207 REFERENCES (ERRATA) American Optical Company, informal table 53. Sodium Chloride, 1969. Ives, David J. G. , and Janz, George J. , 1961, Reference Electrodes, Academic Press, New York. Wolbarscht, M.L. , MacNichol, E. F. , and Wagner, H.G. , 1960, Glass insulated platinum microelectrodes, Science, 131, 1309-1310. 208 APPENDIX A The use of conductivity cells in A. C. circuits to detect fluctuations of conductivity in electrolyte solutions has been briefly summarized by Gibson and Schwarz (1963). In this section we describe only the particular sensors and circuitry that were used during the set of experiments discussed in the text. DESCRIPTION The sensor is a single- electrode, platinum conductivity probe whose construction is shown diagrammatically in Figure 4. The rounded platinum tip and the L shaped glass housing minimized flow disturbances induced by the probes. The probe assembly was oriented so that the horizontal portion of the glass housing containing the probe tip was approximately parallel to the crests of the internal waves (i.e., perpendicular to the side walls of the tank). This prevented the wake of the vertical shaft from washing by the electrode and introducing spurious values. The electrical path was between this electrode and the solution ground (a large copper strip placed along the tank bottom beneath the slope). The platinum tip was normally platinized by a method similar to that described by Wolbarscht et al 0 960): (1) the probe was immersed in a 1 percent solution of Chloroplatinic acid to which lead acetate (80 percent by weight) had been added; (2) current from a 10 v D. C. source in series with a 10, 000 ohm resistor was passed between the probe (cathode) and a platinum wire (anode) either for about 15 seconds or until a thin stream of bubbles was emitted from the probe tip. The result was a thin grayish-black coating on the platinum probe. The effects of this coating on probe performance are discussed in Ives and Janz (1961); essentially, the platinization minimizes drift in the output of the sensor. The probe is the active arm in an A.C. Wheatstone Bridge (Figure Al). Potential imbalances across the bridge caused by fluctuations of con- ductivity at the probe are amplified, detected (diode rectification), and 209 recorded by the A-D conversion described in Chapter 2. The high excitation frequency (60 kc) and low excitation voltage (0. 5 v rms) inhibited polariza- tion of the electrolyte at the probe tip. This was essential in order to mini- mize signal drift and capacitance problems. PERFORMANCE Four complete circuits like that shown in Figure AT were used. It was found that the levels of signal output changed by as much as 15 percent when probes were interchanged, owing to nonuniform it ies and irregular platinization of the probe tip; this usually necessitated a calibration check before each experiment. Normally this calibration consisted of recording output voltage as a function of probe depth before each run. This provided an updated sensitivity value that could be checked with previous values. If it was observed that sensitivity changed markedly (> 10 percent) between runs, the probe tip was checked for fouling and cleaned as necessary. The cleaning process is like that described by Ives and Johns (1961). Prior to initial platinization, or when the platinized coating was cleaned off with a soft tissue, the following procedure was used: (1) wash in warm concentrated HNO„ for 30 seconds, (2) wash in distilled water, (3) polish platinum tip with glass rod, (4) treat polished electrode with warm 50 percent solution of Aqua Regia (4 Volumes HO, 3 Volumes HCp , 1 Volume HNOJ for 15 seconds, (5) wash in distilled water, (6) wash in warm HNO„ for 30 seconds, (7) wash in distilled water, (8) dry in air, (9) place in distilled water when not in use. Normally, the output signal was acceptable up to three hours of continuous use; however, shortly thereafter degradation of the signal to noise ratio 210 - o> T3 *- O >» O LO Q> cr c o • f « ■ ^ 2 Q — w Q. 6 J5* o tn O c o o> o o Q c o o <» O O < £ 2 c ^ - o o. E < 0D O to c CD (/> CD -Q O © w l_ e o i_ o o JC w o O O o > IT) 6 0 W c Q) en & -H > •H -P U 'O G 0 u M 0 M-l O •H -P tO g o w 4J •H O n •H u •rH Cm 211 and erratic output values necessitated cleaning of the probe tip (steps (5) through (7) above). The following list summarizes the average performance of the probes determined from experimental data and calibration tests. The length units are presented for convenience and are based on: where P = P0 (l-0z) 3 = 0. 001 g/cc/cm (1) sensitivity: 0. 3 volt/0. 001 g/cc (0. 3 v/cm) (2) resolution: 0. 5 x 10~ g/cc (0. 5 mm) (3) accuracy: 5 percent of maximum vertical displacement of probe relative to stratified fluid at rest for displacements < 0. 5 cm. (4) linearity: 2 percent of maximum vertical displacement of probe relative to stratified fluid at rest for displacements < 0. 5 cm. (5) drift: variable, depending on quality of platiniza- tion. (Typically 20 millivolts/hour.) It might be useful to present functional relationships between some of the physical variables for sodium chloride solutions that were assembled during this study. Several sources were used to obtain the relationships (L.L. Higgins, 1962; International Critical Tables, 1929; American Opti- cal Company, 1969). (1) n = 1.3330 + 0.244 (p-1); T = 20°C (accurate to approximately 1 percent) (2) S = 1.39 (p-1) x 103; T = 20°C (accurate to approximately 1 percent) (3) a = 1.902 (p-1); T = 20°C (accurate to approximately 5 percent) 212 where n is index of refraction, 3 p is density on g/cm S is salinity in °/oo, cr is electrical conductivity in (ohm-cm) -1 213 APPENDIX B 214 TABLE Bl Group N N O r-t r-> r-l H NM in Ol in O) >-l CO 00 CM «*" *-< in tt to to oo in »-i in N CD CO CO CO o Ol *■* CM 00 ^f CO t oo C- CO CO NCOOl 00 O CO tOH ^ J'. CM CO Ifi 1 I- © CO i co n in o oi ^r 1 CO CO CO 1 t- COM C- 00 CO 1 00 CO t> 1 00 o> CO CO C- CO 1 H JJOl i r- o as < ii O »H I— 1 i-l CM CO O O — i »-< CM CM <-> o 00 O "-■ co o o t— co co rr T — oo to oi 00 tJ- co CM t^ CM £ E ° — ~ o- Ed 2 ^> O II * o CO Ol [>- in i-( o 00 O CM CO 01 CO co in oo E ° ~ C V ~* CM CM © •* ■ . o £, cr* £ o ~- ~> o it 2S o M oi t- in co >«< t- t- <5 CM ocor- * me- CO 00 O CM CM t»< I CO CO t- I 00 O CM O «-• »-< H H CO CM CM O co cm in CO CO o* co cm in CM CO C- eo CQ W J 03 5 a" "" ~* ^ CM E © ~ CJ ll 2S o as cm oo ^»< CO «-< C- © CO in t- ao co o oo •»»< © oo to co © t- © m i cm o m i oi co ft i oo oi co i in o c- O m oi c— ^< ai ■^ oi «— < «-«' i-< CM 00 00 CO CO O CM V Ol o CM CO CM CO E 2 o £ N 5 ,y E cm oi in co <-l oo o co in in oi c- 00 CM TT 00 •"-! CO CM CO CO 0 Ol TJ< 01 CO 00 IH .-H i-H i-l CM CM cm co in CO * H CO © Tf Ol CO © CO O Ol I 00 00 t- I CO © o i oi cm in i oi in cm in c- rr oo co © i o © »■* i m »— i co i oi r-t ■*? i c- co ao co oo © i* coco o co m in cm c- © t- ~ Ol CO « ,-1 <-< ,— .-h CM CM COHf © o o CO «-• CO o »-< «-' O CM CO 00 CO CM in oi co oo o oo in in oi co *— < © co r-t i— i Oi O in H(DO i-ioo©eo co >-< in cm »-< oo © co ao »-< m cm m o © ^< -> CM CO CM VO C- CM tj> O) <— < ^< CM CO CO t- CO co in .-c in "HO i in r- co i cm ■<»• o> ^t* oo oo in co cm co —i o E < ooio) I c- c- t- i t- win o *+ <-< co o) o) I •»»< O) O) i in ^< cm ■— l CM CO 00 CO CO l OlO) O) I H Nt- CO CO Tf a* I ^ CM So 2 — U II c- c- co I t-rtCO i co in cm £ F CJ o J* *-H OJ cm a* fi d 2 •^ CJ ii > o w ■^ c o E U E CJ ~' .* 1-1 *"■ cm"" CO CQ O) a* £ d 2 o II X Ul o —' J rt m o < rj> t— CO h in co I t- NO I f CO O I 00 «-H Tt< I tO H O) O r-l i-l H N CN C- 00 CM NHt- I NOO) I CO CD r- 1 I t— CM T}< I rr CO •— I t- ~* O i o in tj< I COO) N CM CM CO y* t»< O I NH p) i o *-h in t- ■* I O H I ^< CM 00 C- O I Tf CD ^T I CO H CM 00 C- «-" O C- CO ^ ^ o O 00 T»< 00 i-i CO I CO »-l Tf I O O) CO i o co i-t i o co in i-H i-l r-* CM CM CO co O O) co O) co rfOO c-r-ao i O) «-i ao ioooc- icmoco i co n v i oh t- i co cm c— iincMin i in tt t— H N N CMCOin CM CO ^T -* CM CO 1 r— 1 1 CM~" y. CJ O .— 1 X h as in Tt< i— i co i oo cm in i in o i-" I 00 I— 1 co I co i— ' in o »-< •— i H NN CO t> 00 00)t- T}- 00 CO Oi CM O TTC-^r I >-l tO tO IOC-CM I Tf CO 00 l © »-i i-i l CO O CO ITt*0)CD I V tOCO ICOCMCD I P) l/)0) IOC— CO CJ cm n cmcotj< ^ to O Tj-inc— co^rr- inocOTj" co Tf co h co m* m co *t- o © c— ©in .— i in © co O co to m m t- co co t- c- to co co in in CM CM CM CM CO O 00 C- CMC-00CM H cot- CM c- in Tf Tf in tj< co co v co m h cD'-itO'-' co •— i co •— ■ in CO CM CM CM CMH rt CM CM — -+ oo cm in o TJ- CO CM CM CJ f_H OJ o o c- CO CD C- 3 a CM CD O © o en CM oo © o a w a ^ o 218 g S w J* cm o • < ° tONt- ^H o t- H (O ffl O * 1-H ^ t-oo 00 »-« <35 a g o * CM go 2 § u co ffi W a CQ < fi o .* «-i cm e^ cr g d 2 ^m u ll X o *" g g u 1A »-i 1-4 fg^ .cr g d 2 <-< u ll X < O CM t- l t"MD I CO CO CM NNH h n* (ON* m oo t- t— oi* o) h to coh m cm m >-i o co o i co to in i noio i tcooo iococm CO t- CM I CO t- 00 It-COt- PTfOOO I in CM t- H NW H N Tf J-HCOOJ H N N iH P) CO ' CM s 2 CO Ol CO I CM CO 00 I CO t- CO in OS t- I O O co I Ol * Tf < c cohon t- in in co t- in in co t- in t- c- t- m t- t- irtt-iftto co * cm co cm ■* cm co cm co in o Ncomo co in *j< co co m * co co in * n * co cm cm »c co n n o o o o o o o d o o d d o o o o o o o o i-li-l.-l.-l CO CO CO CO CO CO CO CO OOOO OOOO »-< r- 1 .— I .-t OOOO OOOO .— I »-l .-< i— I H H H p> < co co in in co oo co o co co o o * win o ^ n in o CO Nino tOOTOl too^oi t-t-ooco t-t-ooco Tf CO CM CM *MNH *co NH * CO N CM Tf CO N N Tf CO o TT o H to i—i 00 CM i-i (A «-H T-H (M O CM H f-H i~ I .-1 i-H CO co O 00 00 CM 3 oo o> t- a> 00 O O o o o O o o o d a ^ * 3 c O O o * in 219 E < rt co in oo in co in CO CM CO in co o CO Tf CO CO O T>< CO CD CD Tf 000 t- O CO y "* oT^ so 2 y ll X Eh y cm e^ so' 2 O ll X o ^" at h o in rr i-i O t-H (O Cft ^ i— ' in t- cm co O m t- CO CD in in o co cn in o E U E o ^-^ -« 7 -. CN^ CO CQ ST ^ Ed 2 1 — * •H O || X o CQ < H o < •»»< h o CO 00 CO OS CO Oi N if CO H H O cn rr o h o CM •-I co co o> tp *-< N h (M S 2 ° x v no **< CO CM CO (MO) **■ o if 1 CD t- O 1 t- O) 0> c- h m 1 t- H H 1 C- H c- < X ° E t»H NO C— i-i OS O t-H Nf incDcooo uocdcooo in co co c- COC0CDtJ< coco CO if co co CO if H O o o H O o O i- o o o in in LO in in in in m in in in to o* a o O) o> o> a o> o o OS ,u E cn cn 0> OS Cft CO O) cn cn cn cn en r- CM H CM CM 00 CO f MM t-H CO CM CD t— CM CM CM CO CO o CO CO r ) o co CM H <1> I/) CO I-I CO H CM Ol CM t- CO [fl CO T Tf 3 a o O o o o o § a- Ed 6 Z CO 220 TABLE B4 Expt. Group (j) T xc zc xc/xo K/fh) knA?) *nfij fe/a( cps sec cm cm ao=0ilcmao=0.Zcmao=0, 1 0.094 10.67 130.6 9.5 1.101 0.716 _ _ 2.642 100.6 0.848 1.716 3.016 3.242 80.6 0.680 3.261 3.666 4.474 70.3 0.593 4.488 3.045 5.123 70.9 0.598 4.410 3.038 5.202 55.9 0.471 6.060 4.246 7.996 45.6 0.385 4.791 3.782 3.412 40.6 0.342 2.544 1.412 2.679 0.055 18.28 130.6 1.101 0.704 0.884 100.6 0.848 1.780 0.723 0.867 80.6 0.680 3.650 1.268 1.602 70.3 0.593 5.380 2.712 6.550 70.9 0.598 5.260 -- — 55.9 0.471 8.735 1.633 3.687 45.6 0.385 9.758 0.802 1.357 40.6 0.342 8.678 0.837 1.221 2 0.102 9.85 132.6 6.6 1.085 0.735 _ _ _ _ _ . 102.6 0.840 1.901 -- 1.614 1.723 82.6 0.676 4.081 -- 4.202 3.782 72.6 0.594 6.258 -- 8.062 11.820 72.6 0.594 6.258 4.410 8.108 11.238 57.6 0.471 12.428 6.083 15.490 14.774 47.6 0.389 19.295 9.378 21.441 25.845 41.6 0.340 23.432 -- 25.005 24.242 51.0 0.417 16.479 14.210 -- 24.370 43.3 0.354 22.419 10.942 14.205 23.332 37.1 0.304 24.437 10.138 25.256 19.539 32.3 0.264 19.940 6.214 14. 178 10.839 0.062 16.00 132.6 6.6 1.085 0.729 — __ 102.6 0.840 1.942 0.843 1.715 82.5 0.674 4.354 1.956 5.083 72.3 0.591 6.974 2.001 7.272 72.8 0.596 6.8071 -- -- 57.9 0.474 14.670 4.954 10.724 48.0 0.393 25.431 12.842 25.578 42.9 0.351 33.443 1.471 26.616 51.0 0.418 21.520 6.122 19.504 43.3 0.355 32.771 12.645 13.248 37.1 0.304 43.212 4.571 12.368 32.3 0.265 47.519 2.082 11.488 221 TABLE B4 (Cont) Group 00 T xc z c XCA0 2 2 (Pw/Po) (am/ao) (an/ao) (am/ao) cps sec cm cm ao=0.1cm aQ=0.2cm ao=0. 4cm 3 0.094 10.67 54.1 5. 5 0.905 1.427 2.728 3.034 33.8 0.565 8.702 11.192 6.322 25.5 0.426 24.331 23.630 22.024 21.0 0.351 46.975 33.112 71.648 26.3 0.440 21.816 21.608 25.424 21.7 0.363 42.216 42.892 44.450 16.7 0.279 91.863 52.984 60.728 11.6 0.194 175.545* 73.624* 86.540 0.129 7.76 26.3 21.7 16.7 11.6 0.440 0.363 0.279 0.194 17.754 30.243 48.736 26.475 16.136 42.114 56.270 20.924 4 0.098 10.24 48.3 32.3 25.5 20.5 11. 1 0.853 0.571 0.450 0.362 1.770 6.238 10.014 17.103 0.086 11.64 48.3 32.3 25.5 20.5 0.853 0.571 0.450 0.362 1.877 8.990 22.063 83.440 0.090 11.13 46.8 30.8 24.0 19.0 10, 3 0.827 0.545 0.425 0.336 2.071 9.424 20.394 57.874 0.078 12.80 46.8 30.8 24.0 0.827 0.545 0.425 2.279 16.582 74.749 19.0 0.336X 1946.208 5 0.098 10.24 47.4 37.2 28.5 23.0 11. 0 0.427 0.335 0.257 0.207 2.038 4.535 9.121 12.364 1.485 3.221 11.402 15.100 0.082 12.19 47.4 37.2 28.5 23.0 0.427 0.335 0.257 0.207 2.241 6.078 18.656 50.672 1.881 8.563 19.068 37.403 222 TABLE B4 (Cont) Expt. Group 00 cps T sec xc cm zc cm xc/xo (Pw/Po) (am/ao) aQ = 0. 1cm (a ao a ) m o = 0. 2cm 2 a =0. 4cn 6 0.062 16. 00 177.9 112.9 82. 9 62.9 9. 5 1. 351 0.861 0. 632 0.480 0.318 1.713 4.777 9.839 2.227 4. 292 7. 008 0.047 21. 33 177.9 112. 9 82. 9 62.9 1. 357 0.861 0.632 0.480 0.318 1.784 5.729 15.146 1.747 4.472 4.090 0.043 23.26 177.9 112. 9 82.9 62. 5 1. 357 0. 861 0. 632 0.477 0.295 1.820 6.260 18.788 1.364 7. 346 6.042 223 w ° co © o co © co *r co ■*»• co cm en co »-• cm Orti-iC- I t— CO NnO N o> ^r o> to > o co inm wis o o o -- co en ^* co en ^r C- O) cm nt-N CM co ■»r d E O O — 00 CO CD •-< V t O CO t— in co in cn oo h cn * co co co v en -I o cn t~ co co co in ONH en tt ~+ CDOI' CD <-h CD t— TT rf f- MincO ocoiOHCi CO CO TT CO CM en in ^^ Cn cn Cn m o>" CJ CO CD O CO CO C- D- OCOC-^ CN CD O) CO CO C- CJ ^- CO ^ ^* CM O © X o> co inn in no) ocm^tco co ^« ~* ~* oi co tt co c- co cnococo CJ N^incDMOO CDt-Cnco rfCDOOO "tt-COOlO" WO) •" Ol o <6 a <6 d d d -«' d d d —' d d d «-i d d d d — ' —' d d — ■ — N Eco cn co co cm cm co oincncM ao cn -^ ^ ooffir-n co^j- co cn ~* *r — o oot-mco'-'Ot— ^r m oo co cdc-tt— co r- in tt cm -h cot- v< w CM — i — i-H -H — i ,-1 .-I CM .-I .-I .-I CM — -* i— -H ~* CM — — .-i m in in in m in in in in in in .^ ^h »— • •-« ^h ^* ^h »-h ^* ^* co co co co o E o y in in in in in in tn in m in in ,— i^h^h^* .-h ~* ^h ^* ■-* ^h oooo O o to co co in in i1 ^ co co oi in t- in co in r- © in cn co mt- coco X en in tc f co co cN ^ co CN h co in ^ co co in in ^ co co ao in tt co <^ .... .... .... o ooooooo oooo oooo oooooo oooo X h cor- mo h fl< cot- c- co co co in in co co co in h in co co in in o E < n tp co co m i-i oi ir co ^- co ^* oocMino cocMooincMO oocmuoo lOCO NMCM-Hr- CMCM-i H TT CO CM CM TJ- CO CM CM CM CM V CO N CM S 2 30 CO CD CD o g- 5 wo *Eo 224 ^ <« <*> o m o — i HOIHOI o> ■-! o in O 00 t-oooc- o) men cm •^"cot-cja in t- oi cm O — i-i *- 1 OHH M OOOO 000"-i oocts^h^ co O) ^ fl- ^ in in co *t m in co OOOO OOOO in t- in co in t- in co ^ toon ^©oh X CO m •*J- CO 00 m ■* CO CO CO in ■* 00 CO m TT u X o o o o O O o o o o o O o o o O > E u CO CO in m CO CO in in ■<»> CM in o •» CM m o CO CM CO in CM O CM CO •O" CM CO m CM O CM CO CO CM CO CM c- co CO CM CO CM 3 a 5 m Wo 225 co cm co uo c-r-^r^ ^o-^Tf o eo^ o CJJ © O CO .-i oo o co O •"» i- i CM -<° —< co — ' co cm -h t— co to uo cm t- .-I —i 00 O CO-h N oo o cm co Ninon o" o ■-< --< — ' — ' cm i-' -H ^> ««< -H N(0 in « t- oo oo co o o o tt E 2 co co o o uot-mco CD CD CO en uo cr> -*r CM CO CM ^« C> CO I/O «-« co ^« co ^ inoi'H omt-o CO Ol 00 o *r t- o in cm v oi © O u Eo U II -h 00 — O) 00 CM CM —i •-! CO CM CM O V -" i/o OQ W J m < sr < i-l i-< CM (DOH I CO I/O CO 1 if) Ift H I O0 CO CO I ^» ^*» © CM CM 00 9-8 W o 226 BIOGRAPHICAL SKETCH The author was born, August 4, 1940, and raised in Erie, Pennsyl- vania. After receiving an A. B. degree in geology from Princeton University in June, 1962, he entered the Naval Officers Candidate School in Newport, Rhode Island. Following commissioning into the U.S. Navy in December, 1962, he spent three years aboard the U.S.S. Saint Paul where he served as an engineering and a weapons officer. During this time he married the former Virginia Louise Beardsley in San Diego, California. He did grad- uate study in oceanography at the U. S. Naval Postgraduate School in Monterey, California for one year before transferring to M.I.T. in Septem- ber 1966. The author is a Lieutenant on active duty in the U.S. Navy, a member of the American Geophysical Union, and a member of the Society of Sigma Xi. He has two children, David Scott (5) and John Daniel (3). Unclassified Security Classification DOCUMENT CONTROL DATA -R&D (Security classification of title, body of abstract and indexing annotation must be entered when the overall report is clattll 1. ORIGINATING activity (Corporate author) Massachusetts Institute of Technology 77 Massachusetts Avenue Cambridge/ Massachusetts 02139 2a. REPORT SECURITY CLASSIFICATION Unclassified 2b. GROUP 3. REPORT TITL EXPERIMENTAL STUDY OF INTERNAL GRAVITY WAVES OVER A SLOPE 4 DESCRIPTIVE NOTES (Type of report and inclusive dates) Doctoral thesis 5 AUTHORiSi (First name, middle initial, last name) David A. Cacchione 6. REPOR T D A TE November 197 0 la, TOTAL NO. OF PAGES 226 76. NO OF RE FS 111 Ba. CONTRACT OR GRANT NO. Nonr 1841 (74) b. PRO J EC T NO . Contract Authority NR 083-157 9a. ORIGINATOR'S REPORT NUMBER(S) 70-6 9fc. OTHER REPORT NOiS) (Any other numbers that may be assigned this report) DISTRIBUTION STATEMENT Distribution of this document is unlimited. 11. SUPPLEMENTARY NOTES 12. SPONSO RING Ml LI T AR Y ACTIVITY Office of Naval Research ABSTRACT Propagation of internal gravity waves over a slope in a fluid with constant Brunt-Vaisala frequency was studied in a wave tank, motion in the fluid interior was measured with conductivity probes, and in the boundary layer, with dye streaks and neutrally buoyant particles Outside the breaking zone, the amplitude and horizontal wave number of high-frequency waves increase linearly with decreasing depth; this is in agreement with existing linear, inviscid solutions. Well upslope, a zone of breaking or runup is produced by the high-frequency waves, the wave characteristics are closely parallel to the bottom slope, upslope propagation of low-frequency waves produces a line of regularly spaced vortices along the slope, which slowly penetrate into the fluid interior; mixing in these vortices sets up thin horizontal laminae that are more nearly homogeneous than adjacent layers. A criterion for incipient motion of bottom sediment under shoaling internal waves was developed using available theoretical solutions velocity in the viscous boundary layer, found to be valid for certain experimental conditions. The predicted maximum sediment sizes that can be moved by typical oceanic internal waves are larger than typical sediment sizes on the continental margin off New England, breaking and vortex formation, not yet observed in the ocean, might also be conducive to sediment movement. DD FORM I NO V 61 1473 Unclassified So i ' :- Unclassified Internal gravity waves incipient movement of sediment LINK A Unclassified Security Classification Thesis C15 131803 Cacchione Experimental study of internal gravity waves over a slope. Thesis CIS 131803 Cacchione Experimental study of internal gravity waves over a slope. thesC 15 E,^,mental study of/ '■ssssk- *i