FINITE AMPLITUDE HARBOR OSCILLATIONS: THEORY AND EXPERIMENT by STEVEN R. ROGERS S.B., Massachusetts Institute of Technology ( 1972) Submitted in partial fulfillment of the requirements for the degree of Doctor of Science at the Massachusetts Institute of Technology (January, 1977) Signature redacted Signature of Author ...... - ...•...... ,\~r~ ...... . I'"\,., Departme~t of Ph}(S)i cs, January, 1977 Signature redacted Certified by . /' • , ~,,.., •. {J . . . . . .. -.-.- A 'fh~sis. S~p~r~i ;o; Signature redacted Accepted by • • • • • • • • ...-r -·/ • J ., -.-- ' ... ' .... - • • .- • . • . . . . • Chairman, Department of Physics FINITE AMPLITUDE HARBOR OSCILLATIONS: THEORY AND EXPERIMENT by STEVEN R. ROGERS Submitted to the Department of Physics on January 13, 1977 in partial fulfillment of the requirements for the Degree of Doctor of Science ABSTRACT A nonlinear theory of finite amplitude, time-periodic oscillations in a narrow harbor is developed, starting from the Boussinesq equations for shallow water waves. The narrowness of the harbor permits a linear treatment of radiation damping and a one-dimensional treatment of the nonlinear harbor response. The wave field in the harbor is expanded in harmonics, whose spatial dependence is governed by a set of coupled, nonlinear ordinary differential equations, subject to two-point boundary conditions. Solutions are obtained by analytical perturbation tech- niques and by an iterative numerical procedure. Results indicate sig- nificant nonlinear effects at large amplitudes. Harbor resonance experiments were carried out using a model harbor placed in a large, shallow wave basin. The effects of boundary layer dissipation, flow separation, and spurious reflections in the wave basin are analyzed. The experimental observations are found to agree reasonably well with the proposed nonlinear theory. -2- Finally, the nonlinear theory is applied to large-scale harbors and some theoretical predictions are presented. Thesis Supervisor: Chiang C. Mei, Professor of Civil Engineering -3- ACKNOWLEDGEMENTS The author wishes to express his appreciation to the Fannie and John Hertz Foundation for a generous fellowship which enabled him to devote full time to his studies, and to the Fluid Dynamics Branch of the Office of Naval Research (Contract N062-228) for pro- viding funds for laboratory supplies and computer time. Many members of the Ralph M. Parsons Laboratory for Water Resources and Hydrodynamics gave of their time and knowledge to assist in this endeavor. Among them are Mr. Edward F. McCaffrey (Research Engineer) and Mr. Stanley M. White, who designed the data acquisition system used in the experiments, and Mr. Roy G. Milley, who helped with the construction of the experimental apparatus. Mrs. Elizabeth A. Quivey did a superb job typing the manuscript. I would especially like to thank Mr. Dick K.-P. Yue and Mr. Demosthenes C. Angelides for many valuable discussions, and for their enthusiasm and fine sense of humor. Professor Chiang C. Mei, who supervised this work, has been a constant source of inspiration and guidance. I am deeply grateful for his time, his encouragement, and, most of all, his friendship. To my parents, who have taught me the importance of education, I am eternally grateful. Finally, I should like to dedicate this work to my wife, Rahel, whose patience, understanding, and devotion have really made it all possible. -4- TABLE OF CONTENTS Page TITLE PAGE 1 ABSTRACT 2 ACKNOWLEDGEMENTS 4 TABLE OF CONTENTS 5 LIST OF FIGURES 7 LIST OF TABLES 12 CAST OF CHARACTERS 13 1. INTRODUCTION 15 1.1 Description of the Problem and Review of Literature 15 1.2 Elements of Water Wave Theory 17 1.3 Linearized Theory of the Long, Narrow Bay 26 1.4 Scope of this Investigation 30 2. WAVE SYSTEM OF THE OCEAN 32 2.1 Motivation for a Linear Treatment 32 2.2 The Radiated Wave 33 2.3 The Impedance Boundary Condition 36 2.4 On the Validity of the Linear Approximation 40 3. RESONANT OSCILLATIONS OF A LONG, NARROW BAY 42 3.1 Reduction to One Space Dimension 42 3.2 Equations Governing Time-Periodic Solutions 44 3.3 Regular Perturbation Analysis for the Lowest Resonant Modes or for Very Small Amplitude Waves 48 3.4 An Energy Theorem and Relation to the Korteweg-de Vries Equation 61 -5- Page 3.5 Numerical Solution by Iteration 68 4. EXPERIMENTAL STUDY OF THE LONG, NARROW BAY 75 4.1 Equipment and Procedures 75 4.1.1 Wave Basin 75 4.1.2 Wave Gauges 77 4.1.3 On-line Digital Computer 77 4.1.4 Model Harbor 79 4.2 Multiple Reflections from Wavemaker 80 4.3 Real Fluid Effects 89 4.3.1 Viscous Dissipation in Wall Boundary Layers 89 4.3.2 Separation Loss at the Entrance 93 5. EXPERIMENTAL AND THEORETICAL RESULTS 101 5.1 Comparison of Theory with Experiment 101 5.2 Some Theoretical Predictions for a Large Scale Harbor 131 6. CONCLUDING REMARKS 162 6.1 Validity of the Various Approximations 162 6.2 On the Importance of Nonlinear and Frictional Effects 163 6.3 Suggestions for Future Research 164 LIST OF REFERENCES 167 APPENDIX A: Numerical Iteration Procedure for Solving a Nonlinear, Two-Point Boundary Value Problem 171 APPENDIX B: Fourier Analysis of Experimental Data 182 BIOGRAPHY 184 -6- LIST OF FIGURES Figure Page 1.1 Geometry of long, narrow bay. 27 3.1 Regular perturbation theory: Frequency response of large-scale harbor (h = 20 m., 2a = 100 m., L = 1000 m., A = .03). (a) First harmonic amplitude at the back wall. 52 (b) Second harmonic amplitude at the back wall. 53 (c) Zeroth harmonic amplitude at the back wall. 54 3.2 Regular perturbation theory: First resonant mode of large scale harbor (h = 20 m., 2a = 100 m., L = 1000 m., A1 = .03, = 1.41). (a) Spatial variation of the harmonics. 55 (b) Surface profiles at times t = mn/4, m = 0,l,2...7. 56 3.3 Regular perturbation theory: Second resonant mode of large-scale harbor (h = 20 m., 2a = 100 m., L = 1000 m., A1 = .03, z = 4.36). (a) Spatial variation of the harmonics. 57 (b) Surface profiles at times t = mn/4, M = 0,l,2...7. 58 3.4 Regular perturbation theory: Third resonant mode of large-scale harbor (h = 20 m., 2a = 100 m., L = 1000 m., A1 = .03, z = 7.36). (a) Spatial variation of the harmonics. 59 (b) Surface profiles at times t = mr/4, m = 0,1,2...7. 60 4.1 Experimental set-up. 76 4.2 Typical calibration curves of a wave gauge, at two attenuator settings. 78 -7- FigurePage 4.3 Theoretical study of multiple reflections from wavemaker: Placement of images. 83 4.4 Experiment (-) vs. theory (-) for the wave ampli- tude in the ocean (h = .5 ft., 2a = .33 ft., L = 1.211 ft., w = 4.067 sec-1, Lm = 31.17 ft.). Top: harbor closed. Middle: harbor open. Bottom: the radiated wave. (a) A1 = .015 86 (b) A1 = .027 87 (c) A1 = .040 88 4.5 Entrance loss: Normalizd first harmonic ampli- tude, |T/T0 1, vs. normalized friction factor, a. 100 5.1 Experiment vs. inviscid nonlinear theory (? = 0). (A,+,x): measured, first, second, and third har- monic amplitudes. (-): nonlinear theory. C----): linear theory for first harmonic amplitude. (a) L = 1.211 ft., A = .015 107 (b) L = 1.211 ft., A = .027 108 (c) L = 1.211 ft., A = .040 109 (d) L = 4.173 ft., A = .015 110 (e) L = 4.173 ft., A = .027 111 (f) L = 4.173 ft., A = .040 112 (g) L = 7.136 ft., A1 = .015 113 (h) L = 7.136 ft., A = .027 114 (i) L = 7.136 ft., A = .040 115 5.2 Experiment vs. inviscid nonlinear theory Cf = 0) with exact dispersion relation. (A,+,x): mea- sured first, second, and third harmonic amplitudes. (-): nonlinear theory. (----): linear theory for first harmonic amplitude. (a) L = 4.173 ft., A = .015 116 -8- Figure Page 5.2 (b) L = 4.173 ft., A1 = .027 117 Cc) L = 4.173 ft., A1 = .040 118 5.3 Experiment vs. nonlinear theory with separation loss (f = .35). (A,+,x): measured first, second, and third harmonic amplitudes. (-): nonlinear theory. (----): linear theory for first harmonic amplitude. (a) L = 1.211 ft., A1 = .015 119 (b) L = 1.211 ft., A1 = .027 120 Cc) L = 1.211 ft., A1 = .040 121 (d) L = 4.173 ft., A1 = .015 122 (e) L = 4.173 ft., A = .027 123 (f) L = 4.173 ft., A = .040 124 (g) L = 7.136 ft., A = .015 125 (h) L = 7.136 ft., A = .027 126 (i) L = 7.136 ft., A = .040 127 5.4 Experiment vs. nonlinear theory with separation loss (f = .35) and exact dispersion relation. (A,+,x): measured first, second and third har- monic amplitudes. (-): nonlinear theory. (----): linear theory for first harmonic ampli- tude. (a) L = 4.173 ft., A = .015 128 (b) L = 4.173 ft., A = .027 129 Cc) L = 4.173 ft., A = .040 130 5.5 Inviscid nonlinear theory (f = 0): Frequency response of large-scale harbor (h = 20 m., 2a = 100 m., L = 1000 m., A1 = .03). (a) Number of harmonics, N*, required in the numerical solutions. 137 -9- Figure Page 5.5 (b) First harmonic amplitude at the back wall according to nonlinear theory (A) and regular perturbation theory (-). 138 (c) Second harmonic amplitude at the back wall according to nonlinear thoery (+) and regular perturbation theory (-). 139 (d) Third harmonic amplitude at the back wall according to nonlinear theory (x). 140 (e) Zeroth harmonic at the back wall according to nonlinear theory (A) and regular pertur- bation theory (-). 141 5.6 Inviscid nonlinear theory C? = 0): First reso- nant mode of large scale harbor (h = 20 m., 2a = 100 m., L = 1000 m., A = .03, k = 1.41). (a) Spatial variation of the lower harmonics. 142 (b) Spatial variation of the higher harmonics. 143 (c) Surface profiles at times t = m7/4, m = 0,1,2...7. 144 5.7 Inviscid nonlinear theory (f = 0): Second resonant mode of large-scale harbor (h 20 m., 2a = 100 m., L = 1000 m., A1 = .03, z = 4.36). (a) Spatial variation of the lower harmonics. 145 (b) Spatial variation of the higher harmonics. 146 (c) Surface profiles at times t = m/4, m = 0,1,2...7. 147 5.8 Inviscid nonlinear theory C? = 0): Third reso- nant mode of large-scale harbor (h = 20 m., 2a = 100 m., L = 1000 m., A1 = .03, z =7.36). (a) Spatial variation of the lower harmonics. 148 (b) Spatial variation of the higher harmonics. 149 (c) Surface profiles at times t = mn/4, m = 0,1,2...7. 150 - 10 - Figure Page 5.9 First harmonic amplitude at the back wall of the large-scale harbor vs. incident wave amplitude. (----): linear theory. (A): inviscid nonlinear theory (? = 0). (x): nonlinear theory with separa- tion loss (f = .35). 151 5.10 Nonlinear theory with separation loss (? = .35): First resonant mode of large-scale harbor (h = 20 m., 2a = 100 m., L = 1000 m., A1 = .03, z = 1.41). (a) Spatial variation of the lower harmonics. 152 (b) Spatial variation of the higher harmonics. 153 (c) Surface profiles at times t = mnr/4, m = 0,l,2...7. 154 5.11 Nonlinear theory with separation loss C? = .35): Second resonant mode of large-scale harbor (h = 20 m., 2a = 100 m., L = 1000 m., A1 = .03, = 4.36). (a) Spatial variation of the lower harmonics. 155 (b) Spatial variation of the higher harmonics. 156 (c) Surface profiles at times t = mnTr/4, m = 0,1,2...7. 157 5.12 Nonlinear theory with separation loss (f?= .35): Third resonant mode of large-scale harbor (h = 20 m., 2a = 100 m., L = 1000 m., A = .03, = 7.36). (a) Spatial variation of the lower harmonics. 158 (b) Spatial variation of the higher harmonics. 159 (c) Surface profiles at times t = mir/4, m = 0,,2...7. 160 5.13 Response of large-scale harbor to nonmonochromatic incident wave (h = 20 m., 2a = 100 m., L = 1000 m., A2 = .050, A3 = .015, = 1.41). 161 - 11 - LIST OF TABLES Table Page 2.1 Orders of Magnitude of Incident, Reflected, and Radiated Waves in the Ocean 40 5.1 Composition of the Incident Waves Used in the Experiments 102 5.2 Exact and Approximate (Boussinesq) Wavenumbers in the Experiments 103 5.3 Resonant Modes of the Large Scale Harbor According to Linearized Theory 131 - 12 - CAST OF CHARACTERS Symbol a Half-width of narrow bay A Twice the incident wave amplitude en(x) Basis function in the numerical scheme E Energy density averaged over time fJ'f ,2 Friction parameters g Acceleration of gravity G Green's function h Depth of water H Interval size in finite difference mesh HO Zeroth order Hankel function of first kind0 k Wavenumber kc Minimum wavenumber for cross modes 9,L Dimensionless and dimensional harbor length MLM Dimensionless and dimensional length of model ocean ( )n the n-th harmonic of ( ) n S A vector normal to surface S N* Number of harmonics present in a numerical solution NPT Number of points in the finite difference mesh p Pressure P R Radiated power Pv Power dissipated is viscous boundary layer Q Source strength for radiated wave - 13 - Symbol r Radial coordinate Rc Critical Reynolds number Rns' 9ns Coupling constants in the nonlinear equations t Time coordinate T,T0 Coefficient of first harmonic, in linearized theory u Velocity vector U Horizontal velocity in harbor entrance x Horizontal coordinate x* Near field x-coordinate x A slow scale in x, e.g. x = EX y Horizontal coordinate y* Near field y-coordinate z Vertical coordinate Z Radiation impedance cx Normalized separation loss parameter 6 Dimensionless half-width of narrow bay 6 v Boundary layer thickness V Two-dimensional gradient operator = s Nonlinearity parameter TI Surface elevation A A relaxation parameter; also wavelength P2 Dispersion parameter v Viscosity p Fluid density Velocity potential; also a phase angle W Circular frequency -14 - CHAPTER 1 INTRODUCTION 1.1 Description of the Problem and Review of Literature A harbor is a partially enclosed body of water which is joined to the sea by a relatively narrow entrance. The entrance, which is quite often a man-made system of breakwaters, serves to shield the harbor from the most common disturbances, namely wind-generated waves of the open sea. Because it limits the amount of energy exchange between the harbor and the sea, a narrow entrance occasionally has the damaging effect of trapping long waves, which may be originated by tsunamis, surf beats or other sources, and of creating within the harbor large oscillations which may persist for days. This is the phenomenon of harbor resonance or "seiching." The deleterious effects of these oscillations are well- known--ships are torn loose from their moorings causing damage to dock facilities and neighboring vessels, and strong currents in the harbor entrance prevent the safe passage of incoming and outgoing ships, cf. Wilson (1957). From a theoretical standpoint, the study of harbor oscillations begins with the determination of the frequencies at which resonance occurs and the wave profiles of the resonant modes. Much of the work to date uses linearized, inviscid wave theory, in which the wave amplitude is presumed so small that nonlinear effects are negligible. Rectangular harbors were studied by Miles and Munk (1961), Ippen and Raichlen (1962), Ippen and Goda (1963), Raichlen and Ippen (1965), and Mei and Unluata - 15 - (1976). Miles and Munk discovered the peculiar phenomenon, referred to as the "harbor paradox," that the amplitude of monochromatic resonant oscillations actually increases with decreasing width of the harbor entrance. For harbors of arbitrary shape, numerical solution procedures have been developed by Hwang and Tuck (1970), Lee (1971), and Chen and Mei (1974). The validity of all treatments based upon linearized wave theory is necessarily limited by the assumption of small wve amplitudes. Fre- quently, the resonated wave heights predicted by these theories are so large that it is no longer justifiable to neglect nonlinear effects. Some attempt to include nonlinear effects in the analysis of harbor resonance was made by Gaillard (1960) and Biesel (1963), but complete solutions were not obtained. The general properties of nonlinear waves, and of water waves in particular, have been studied extensively by Phillips (1960), Longuet-Higgins (1962), Benney and Luke (1964), Whitham (1965, 1967 a,b), Benjamin and Feir (1967), Chu and Mei (1970), Kim and Hanratty (1971), Mei and Unluata (1972), Bryant (1972), etc. A review article by Phillips (1974) is especially useful. For the most part, these authors restrict their attention to progressive waves travel- ling in an infinite medium having no obstacles in the path of propaga- tion. Finite amplitude standing waves in a basin have been treated by Tadjbakhsh and Keller (1960), Benney and Niell (1962), and by Verhagen and van Wijngaarden (1965), who also performed several experiments. Laboratory investigations of harbor resonance have been hampered by several difficulties, among them inadequate modelling of the ocean - 16 - domain and Reynolds number dissimilarity. Biesel (1954) discusses the similitude of scale models for the study of harbor oscillations. Ippen and Raichlen (1962) and Raichlen and Ippen (1965) investigate the prob- lem of coupling between a large but finite ocean basin and a smaller harbor basin. Ippen and Goda (1963) and Lee (1971) performed experi- ments to corroborate their theoretical work on harbor resonance. Much of this experimental work is indirectly predicated on the assumptions of linearized theory. For example, it is assumed that the response of a harbor to monochromatic incident waves is itself monochromatic so that wave records need not be Fourier analyzed. Further, the experiments are often conveniently performed in deep water, because the extrapolation to shallow water is trivial in the context of linearized theory. For nonlinear theory, this extrapolation cannot be made. 1.2 Elements of Water Wave Theory In a wide variety of water wave problems, the fluid motion is essentially incompressible, inviscid, and irrotational. For constant depth, the governing equations are V2 +ezz = 0 in fluid -h < z <_ n(x,y,t) z=0 on z=-h t + 4'xx + yfny = z on z = n(x,y,t) + 12 + (vfl2) + gn = 0 on z = n(x,y,t) (1.2.1) where - 17 - n(x,y,t) = displacement of the free surface from its undisturbed position U(x,y,z,t) = (y ,z) = fluid velocity (x,y,z,t) = velocity potential g = acceleration of gravity h = water depth Subscripts are used to denote differentiation. (1.2.la) is the conse- quence of continuity for an incompressible fluid and irrotationality; (1.2.lb) states that the vertical component of velocity must vanish at the bottom; (1.2.lc) is the kinematic free surface boundary condition which ensures that particles on the free surface move tangentially to it; and (1.2.ld) is the dynamic boundary condition stating that atmos- pheric pressure is constant (e.g., zero) on the free surface and that surface tension is negligible. Note that the two boundary conditions on the free surface are nonlinear because both p and n are unknown. In linearized wave theory, the wave amplitude is assumed to be infinitesimal and (1.2.lc,d) are replaced by 1t ~1 z =0 on z = 0 t + gn = 0 on z = 0 (1.2.2) In an infinite ocean of constant depth h, one solution to (1.2.la,b) and (1.2.2a,b) is the monochromatic progressive wave n(x,y,t) = Re {Aei(kx - wt)I p(x,y,z,t) = Re {-igA cosh k(z+h) ei(kx - wt)} (1.2.3)W cosh kh - 18 - where W2h/g = kh tanh kh =..1p2 (1.2.4) is the dispersion relation. In deep water, p 2 >> 1, the dispersion relation is approximately w2h/g = kh, and the phase velocity is twice the group velocity. In shallow water, p 2 «<1, 2h/g = (kh) 2[1 - ]-(kh) 2 + (kh) 4 (kh) 2 (1.2.5)3] the dispersion relation is approximately linear; the phase and group velocities are both equal to gi (1 + O(kh)2); and the waves are said to be non-dispersive. For intermediate values of p, w is a transcen- dental function of kh, and the waves are highly dispersive. When the assumption of infinitesimal wave amplitudes is removed, the nonlinearity parameter E_= A/h = wave height/water depth is also needed to characterize the wave. Different wave motions are 2possible depending on the relative magnitudes of E and p2. In deep water, the combination sp 2, which is a measure of the slope of the free surface, is found to be of great importance in determining when a wave will break. In shallow water, the combination E/ 2, known as Stokes' number, is used to identify three qualitatively different kinds of waves: /P2 << 1 : Linear, dispersive waves (Linearized theory) 6/12 n 0(1) : Nonlinear dispersive waves, including cnoidal and solitary waves (Boussinesq and Korteweg-de Vries equations) - 19 - E/y2 >> : Nonlinear, nondispersive waves (Airy's equations) In treating harbor resonance, we deal almost exclusively with shallow water, or "long" waves. This is because the wavelengths of resonant harbor oscillations are on the scale of the horizontal dimen- sions of the harbor, which are normally much greater than the water depth. For example, a harbor might be two kilometers long and only twenty meters deep, so that p 2 10~4 for the lower resonant modes. It is useful then to have a theory in which the assumption of shallow water is built-in, but the assumption of infinitesimal wave amplitudes is not. To derive such a theory, we introduce scaled variables so that the relative magnitudes of terms in the equations of motion appear ex- plicitly. For a shallow water wave of frequency w, an appropriate choice of variables is t' = ot (x',y' = tO (xy) Z' = z/h (1.2.6) -n =/h U' = u//gh $' = o4/gh where primes denote dimensionless variables. Henceforth, all variables are dimensionless, and for convenience, the primes will be omitted. Substituting (1.2.6) into (1.2.1), the equations of motion take the dimensionless form: 2 1v P (- 0) (1.2.20) n2 = Re {EA2 ei(k2x - n2t)} (n2, k2 > 0) where, from (1.2.17), k1 = n1 + 0( 2) (1.2.21) k2 = n2 + o(V2 Through nonlinear interaction, a third wave is formed T3 = Re {EA3 ei[(k1+k2)x - (n+n2)t] + ei[(k1-k2)x- (n1-n2)t (1.2.22) Since k1 +k2 = n1 -1+ n2 + 0(p2 ) n3 represents a propagating wave which travels with approximately the same phase velocity as n and n2; thus, the three waves maintain the same relative phases as they propagate. The magnitude of n3 is found to increase linearly with distance so long as the three waves are in phase and jn3j « 1'n111n21. This is known as a resonant three-wave interaction. Eventually, the influence of dispersion limits the distance of coherent propagation to 0(1/u2). The net result is that nonlinear interactions of 0(s2) add constructively -over a distance of 0(1/p2), and the magnitudes of sA3 and EB3 are - 25 - found to be sO(s/p2). The Stokes parameter, E/p2, therefore governs the extent of harmonic generation. The subject of resonant nonlinear interactions is discussed at length in Chapter 3. When e/p2 << 1, higher harmonics are generated in profusion, and the horizontal length scales used in (1.2.6) are no longer appropriate. This is the situation just before breaking. The Boussinesq equations cannot be expected to apply to such cases because the scales used in their derivation are no longer the correct physical scales. 1.3 Linearized Theory of the Long Narrow Bay The harbor we shall consider is a long, narrow bay of width 2a, length L, and constant depth h. The geometry of the ocean-harbor system is shown in Fig. 1.1. The coastline and harbor walls are taken to be rigid, impermeable, vertical surfaces along which the normal component of velocity must vanish, viz. u=xgySt)-n= 0 (1.3.1) where (x5,yS) is a point on the surface, and 'As is a vector normal to the surface. The presence of finite boundaries in the fluid introduces new length scales, in addition to those which characterize the wave length and amplitude. We shall see that the dimensionless parameters 2 L(w/gii) (1.3.2) 6 2 vr-h - 26 - L' a f & & & 1 .0 A 4 - .-e--- -- 4 Ft' F cr N ~ (7 FFUFF grr ,j UFV' 4 HARBOR INC-tDENT WAVE OCSAN Fig. 1.1. Geometry of long, narrow bay 07c. I determine the linear resonance properties of the harbor basin. For a narrow bay (a/L << 1), resonant oscillations typically occur when z = 0(1) and 6 = t-(a/L) << 1. Because 6 << 1, the ocean-harbor system can be split into three distinct regions: the far field of the ocean, r =/x2 + Y2 >> 6; the near field of the entrance, r r 0(6); and the far field of the harbor, -z < x << -6. In the far field of the ocean, the fluid motion varies on the scale of the wavelength and (x,y) are suitable horizontal coord- inates. In the near field of the junction, the scale of motion is 6 and the coordinates (x*,y*) = (x,y) / 6 (1.3.3) are most appropriate. Finally, in the far field of the harbor, the natural choice for coordinates is (x, y*). Before embarking on a nonlinear treatment of the resonance proper- ties of a long, narrow bay, we summarize known results from linearized theory. The geometry of Fig. 1.1 has been studied by Miles and Munk (1960) and Unluata and Mei (1973). Starting from linearized water wave theory, Unluata and Mei obtain asymptotic expansions for the wave field in each of the three regions. By matching the asymptotic solutions in adjacent regions, they determine the response of the harbor in terms of the incident wave amplitude and frequency. For an incident wave of the form n (x,y,t) = Re(4A e i(k1x-t) (1.3.4) where k1 is the dimensionless wave number satisfying the dispersion relation (1.2.19), the response in the far field of the harbor is found - 28 - to be H (xt) = Re (T cos k1 (x + z)e-it uH(x,t) = Re ( T sin k 1(x + z)e-it k11 where T E A[cos k I- sin k k]~(1.3.6)A[o k1 k1 and (k1 6)[l + -in k16 + zn )] (1.3.7) zny E .577216 = Euler's constant Z is the radiation impedance of the harbor entrance; its real and imaginary parts are the radiation damping and mass reactance, respective- ly. Letting Z1 = Z R + iZ 1P, we have from (1.3.6), Max nH(x,t) = jT1 I x,t - A1 I[(cos kz + Zjj sin k1k)2 + (ZiR sin k 1)21-1/2 (1.3.8) When kI satisfies the "resonance condition" cos k1zk + Z11 sin k1 z = 0 (1.3.9) k (1.3.8) yields - 29 - Max n(xt) = ZR IAI kT (1.3.10) x~t I1Rsn 1; k, so that the incident wave amplitude is amplified by a factor proportional to 1/ZlR, which is in turn proportional to 1/6. As the width of the harbor entrance is made smaller, 6 decreases and the resonant amplifica- tion increases; this is the so-called "harbor paradox.' Since Z 1 ' u 0(6 zn) << 1, the resonance condition (1.3.9) can be written as cos (k1 Z - Z 1) 0 (1.3.11) F-k1 which yields the resonance criterion k1 = (m + 1)r +-Z m = 0,1,2... (1.3.12) 1 The lowest mode, m = 0, is often called the quarter-wave mode because 1]_ 2w7T 1kz i/2 or z~ (A)(T) = 4-, where xl is the dimensionless wave- length. Notice that, in (1.3.5)., nH and uH are independent of y; that is, the far field of the long, narrow bay is essentially one-dimensional. This allows a great simplification in the nonlinear treatment of the harbor, as will be seen in Chapter 3. 1.4 Scope of this Investigation ' The theoretical part of this study analyzes the response of a long - 30 - narrow harbor to time-periodic incident waves. In Chapter 2, the wave field of the ocean is shown to be of sufficiently small amplitude that it may be treated as linear. Assuming the entrance width to be small, e.g. << 1, the interaction of the ocean with the harbor is then approximated by an impedance boundary condition applied at the harbor entrance. The advantage of this approach lies in its mathematical simplicity; its disadvantage is that higher harmonics have larger values of knS and are therefore less accurately represented. Where greater accuracy is required, a semi-numerical method, such as that of Lee (1971) or Chen and Mei (1974), should be used. In Chapter 3, the nonlinear response of a long, narrow harbor is examined. With the assumption 6 << 1, the wave field far from the harbor entrance is shown to be essentially one-dimensional, allowing a simplification of the nonlinear equations. The field variables are Fourier analyzed into time harmonics, and an infinite set of coupled, second-order, ordinary differential equations is found to govern the spatial variation of the complex harmonic amplitudes. Perturbation techniques and direct numeri- cal methods are used to solve these equations. Laboratory experiments were carried out using a model harbor placed in a large shallow wave basin. Chapter 4 deals with the experimental apparatus and procedures, and with the effects of friction and multiple reflections from the walls of the wave basin. Results of the theoretical and experimental work are compared in Chapter 5. In addition, some theoretical predictions for large scale harbors are presented. A summary of the main conclusions and suggestions for future research are found in Chapter 6. - 31 - CHAPTER II WAVE SYSTEM OF THE OCEAN 2.1 Motivation for a Linear Treatment The wave field of the ocean consists of three parts; the incident wave, n(x,y,t); the wave reflected from the coastline, nr(x,y,t); and the wave radiated away from the harbor entrance, nR(x,y,t). To a first approximation, the total wave field, nT(x,y,t), is simply a linear superposition of these three, viz. nT i + nr + nR (2.1.1) +T The total velocity field, u (x,y,t), satisfies a similar relation. In analyzing the ocean, we begin with the linear approach of (2.1.1). After the linear solution is obtained, the magnitude of the nonlinear terms will be estimated, and the assumption of linearity will be reexam- ined. We shall find that, when the harbor entrance is narrow (6 << 1), the radiation damping is 0(6). It is then possible for an incident wave of O(sa) to excite a resonant harbor response of 0(E). To determine a solution that is accurate to 0(62), one need only account for nonlinear effects in the harbor, not in the ocean. This is the motivation for a linear treatment of the ocean. Deleting nonlinearterms in (1.2.16) and (1.2.17), we have for the n-th harmonic -inpn + v-u = 0 n n -in n + nn = 0 (2.1.2) n - 32 - and (72 + k2))n = 0 (2.1.3)nfn 2 2 1 2 2 where k2 = n (1 + n P ). Note that n (x,y) is proportional to then 3n velocity potential for t (x,y). From (2.1.2b), $n = 0 so that n (x,y)veoiyptnilfrun 0 0 is a constant everywhere in the ocean. By properly defining the water depth, the constant can be set to zero, to the leading order, without loss of generality. The term t0(x,y) represents a steady current and will be presumed to be zero, to the leading order. Thus, the zeroth harmonic may be deleted. 2.2 The Radiated Wave If the harbor entrance were sealed, the boundary condition at the coastline would be T u= .(o,y) = < y<) (2.2.1) k n and the total wave field would be given by Tii n (xy) = n (xy) + ny(-x,y) (2.2.2) where n1(-x,y) is simply the reflection of the incident wave about then y-axis. The function n4(x,y) can be any solution of the Helmholtz equation in the domain x > 0, - co < y < o. For example, n(x,y) might be an obliquely incident wave -iaX' iyn ( 2 n 2) r$(x,y) = cne nXe (n'n > 0, an +$g 2 = k2 (2.2.3) - 33 - in which case T nisTn(x,y) = 2cn cos ax e (2.2.4) represents a standing wave whose amplitude varies sinusoidally in the y direction. With the harbor entrance open, the boundary condition on the y-axis is more complicated: TT= ian T(o,y) 0 jyj > 6 ux(oy)x 1 (2.2.5) kn Un(y) ly < 6 where Un(y) is the unknown velocity distribution in the harbor entrance. The total wave field nT(x,y) = n (xy) + n (-xy) + n (x,y) (2.2.6) contains a radiated wave component which is the solution to the following boundary value problem: (V2 + k2)n = 0n n an (o,y) 0 |yl > 6 n x f .A2 (2.2.7)9X ik2 nnU(Y) lyI < 6 RR lim 1,r ('-n ik ) = 0ar n n where r x2 + y Equation (2.2.7c), known as the radiation condi- tion, ensures that the radiated wave will propagate away from the har- bor entrance, as required by causality. In essence, the radiation - 34 - condition replaces the initial conditions that would be present in a fully time-dependent treatment of wave scattering. For a detailed derivation of the radiation condition, see Stoker (1957), pp. 174-181. A formal solution for the radiated wave can be constructed using Green's functions. The appropriate Green's function, which satisfies (2.2.7) with Un(y) replaced by the unit singularity 6(y-y'), is found to be Gn(xyy') = ( n2)( H x2 + (-y'))) (2.2.8) (1)where H is the zeroth order Hankel function of the first kind. Since 6 Un(y) =f Un(y) 6(y-y') dy' (2.2.9) -6 Rthe solution for qn(x,y) is given by k2 6 (xyH(k + (-y) 2)U (y)dy' (2.2.10)n2n 0on / + __)ny -6 Let us examine (2.2.10) in the "far field" of the ocean, that is, 2 22 where r = x + y2 >> 6. Substituting the Taylor series (1) (1) H)(kn x2 + (-y1 2) = H 1 (knr) + y'[Hl + 2.2.11) (2.2.11) into (2.2.10), we have R k2 6 aH(1) 6 n (xy) = --- {H)(knr)f U(y')dy' + [0 ] { y'U=(y'dy'+... -6 -6 (2.2.12) - 35 - The first term on the right hand side represents the field of a mono- 6 pole whose strength is proportional to the total flux, {Un(y)dy. The second term is that of a dipole whose strength is proportional to the 6 dipole moment of the velocity distribution, f YUn(y)dy. When 6 << 1, -6 the dipole term is smaller than the monopole term by a factor of 6, and higher terms in the multipole expansion are still smaller. Defining tte average velocity by U n Un(y)dy' (2.2.13) -6 we have (-~k( ) kn6U )Hl) (knr)-[l + 0(6)] (2.2.14) Note that, if U is 0(s), the radiated wave is O(skn6) in the farn n field. We turn our attention now to the near field of the harbor entrance. 2.3 The Impedance Boundary Condition In the near field of the harbor entrance, r is 0(S) and, assuming k n6 << 1, H l)(knx2 + (y-)y) ) 1 + 2n( 2 + 7-y-)2) + O(kn6znk6)2 (2.3.1) where any 2 Euler's constant = .57722... . Substitution of (2.3.1) into - 36 - (2.2.10) yields k n yT, (x3,Y) ~- 1I+ in( /x2 + G-y_ )2)U (y')dy'* (2.3.2) -6 Defining the average wave amplitude in the entrance by 6 Tin {R(o,y)dy (2.3.3) -6 we have 2 = {I + n( f )(y)dy'y(2.3.4)n n n 1T 26jf j 2 ''n -6 -6 SZ n Un where k2C 6 yk U (y') Z n = ( 2 ){l J{n( 2ny-y') dy'dy} (2.3.5)n -2 6 -6n is called the "radiation impedance" of the ocean-harbor junction. The real and imaginary parts of Zn are the radiation resistance and mass reactance of the junction, respectively. Note that only the mass reactance depends upon the details of the velocity profile, Un(y). Introducing the variable v = y/6, and using (2.2.13), we arrive at the following expression for the radiation impedance: k . Zn ( f)(kn6) [1 + -v (Zn kn6 + cn)] (2.3.6) where 1 1 cn =zin + < nlv-v'(Un(V')/n)dv'dv (2.3.7) -37 - Observe that the radiation resistance and mass reactance are O(kn6) and O(kn6znk n6), respectively. In deriving (2.3.6), we assumed that k 6 << 1. This assumption is progressively worse for the higher har- monics, which have larger value of k n6. We accept this deterioration of accuracy in the hope that the magnitudes of the higher harmonics are progressively smaller. Where greater accuracy is required in treat- ing the high harmonics, the impedance approach can be abandoned in favor of a semi-numerical method such as that of Lee (1971) or Chen and Mei (1974). To predict the essential features, we choose to avoid these elaborate alternatives. The form of the velocity profile, Un(y), depends upon the geometry of the harbor entrance and upon the wave field in the harbor, which is, of course, still undetermined. In the linearized theory of harbor resonance, the wave field in the harbor is expressible as an integral over the Green's function for the harbor, in exact analogy with (2.2.10). The requirement that rn(x,y) be continuous at x=o leads to a linear integral equation for Un(y). The integral equation can then be solved by variational approximation or a numerical quadrature procedure. This approach is only feasible when the harbor is treated as linear, because the use of Green's functions relies upon the principle of superposition. To solve the nonlinear problem, a different approach must be found. When flow spearation at sharp corners is not significant, one may exploit the assumption that kn << I and determine an approximate form for the velocity profile, without prior knowledge of the wave field of the harbor. This has been recognized long ago by Rayleigh and is easily demonstrated as follows: In the near field of the junction, - 38 - the appropriate length scale is 6, and the terms in the Helmholtz equation are in the ratio, (k )n v2) ' 2 O(kn6)2 (2.3.8)'nan''n n It follows that V2 n = 0 + O(kn6)2 (2.3.9) so that Laplace's equation governs the motion in the near field with a relative error of O(kn&)2. This is called the "quasi-static" approximation. The name derives from the fact that an observer in the near field sees steady flow and is unaware that it is being driven by oscillatory wave motion. To solve Laplace's equation, it is only necessary to specify the geometry of the harbor entrance; the details of the adjacent wave fields are irrelevant. For the geometry of Fig. 1.1, the potential flow field can be found by applying a Schwartz-Christoffel transforma- tion. Details are given in Unluata and Mei (1973). The radiation impedance is then found to be Z2 . )f(kl6)[l +- nkn6 + zn ](2.3.10) where Zny = Euler's constant = .5772. From (2.2.6), the average total wave amplitude at x=o is given by T= 2H1(0) + = A + Zn U (2.3.11) n n n nn where - 39 - A =2n 1(O) 2n i r(o,y)dy (2.3.12) -6 Equation (2.3.11) will be referred to as the "impedance boundary condi- tion." Observe that it was derived without even specifying the geometry of the harbor basin. 2.4 On the Validity of the Linear Approximation Having obtained the linear solution for the wave field in the ocean, we now estimate the magnitude of the neglected nonlinear terms. The following table shows the orders of magnitude of the n-th harmonic incident, reflected, and radiated waves in both the near and far fields of the ocean. Table 2.1 i r R n n n Far Field A A k 6U n n n n Near Field A A (k 6znk 6)U n n n n n The magnitude of Un/An is thus far arbitrary. In section 1.3, Eq. (1.3.6), we have seen that, when the n-th harmonic is resonated, Un/An 10(g16) (2.4.1) n This result can be anticipated if one thinks of the harbor as a damped - 40 - harmonic oscillator, with damping coefficient k 6, being driven by n an externally applied force of magnitude An. The response of the nn oscillator U at its natural frequency, is then given by (2.4.1). Of course, a harbor is a continuous system and therefore has many resonant frequencies, rather than just one. In analyzing the harbor, we shall require that the maximum value of U n be 0(c). Equation (2.4.1) then constrains An to be O(cekn6). From Table A, the largest wave amplitude in the ocean occurs in the near field, whereTi is O(ekn6znkn6). Thus, the largest nonlinearn n n terms are O(ekn6znkn 6)2 in the ocean and 0(c2) in the harbor. It is consistent to treat the harbor as nonlinear and the ocean as linear only if (skn6znkn6)2 «E2 (2.24.2) which is the case whenever kn6 << 1. The conclusion, therefore, is that, so long as kn << 1, it is consistent to neglect nonlinear effects in the ocean, and to include leading order nonlinear effects in the harbor. - 41 - CHAPTER III RESONANT OSCILLATIONS OF A LONG, NARROW BAY 3.1 Reduction to One Space Dimension The harbor domain, -k < x < 0, consists of two regions: the near field, lx| O 0(6), and the far field, -z < x << -6. We shall show that the waves in the far field are essentially plane waves varying only in the x-direction. Intuitively, this is because oscillations in the y- direction have a minimum cutoff frequency given by a = 7r/2. Since the frequencies under investigation are far below this cutoff, the y- dependence of the near field disturbance cannot propagate outward into the far field of the harbor. A more rigorous demonstration of the fact that, for low frequency oscillations, the far field of the harbor is one-dimensional, proceeds as follows: Let us define a velocity potential q(x,y,t) for the depth-averaged horizontal velocity u(x,y,t) by the relation u E v. Integrating the momentum equation (1.2.12b) over space, the Boussinesq equations take the form t + VY.[(1 + T)0V] = 0 $t+ n +12tt +1V 2 =C(t) (3.1.1) The integration constant C(t) can be absorbed into the term it by re- defining c; hence we may set C(t) equal to zero, without loss of generality. The boundary conditions in the far field of the bay are - 42 - x (-z,y,t) = 0 y (x,6,t) = 0 (3.1.2.) y)(x,-6,t) = 0 (3.1.2b,c) suggest that we rescale y, viz. y* = y/6 From (3.1.1), with C(t) = 0, we then have ((1 +nf*)*2 = 62{t + [0 +n)( ] } (3.1 .3) 1 2 = 2 + 1 2 + 12 7 * = -62 ft + n + } V 24tt + x Integrating (3.1.3a) over y* and using the boundary conditions (3.1.2b,c) we find 4Y)(xy*,t) = 0 + 0(62) (3.1.4) so that, to 0(62), (x,y*,t) is independent of y*. From (3.1.1b), it follows that n(x,y*,t) is also independent of y*, to 0(62). We conclude, therefore, that to 0(62), n, 4, and u = +xare independent of y and that the transverse velocity component v = 4Y is zero. To be consistent with the neglect of 0(62) terms in the quasi- static approximation of the near field (section 2.3), we shall terminate the demonstration of the one-dimensionality of the far field at 0(62 - 43 - 3.2 Equations Governing Time-Periodic Solutions The one-dimensionality of the far field of the harbor greatly simplifies the equations governing harmonic interactions. In one dimension, equations (1.2.16) and (1.2.17) reduce to -inn + u' + (nsun-s (3.2.1) 2 1 -inun + k(n + (Uu0 an and 1" + k 2 -in I snn n n 2 s n-s sl sn-s)" (3.2.2) where primes denote differentiation with respect to x. From (3.2.1), u'= inn*(1 + 0(c)), and u - n (1+ 0(sp)), for n 0. Ton n n nsn within the accuracy of the Boussinesq approximation (nsuYn-s s su-s + n'un-s5 -(i(n - s)n sTin-s s 1 (usun-s s (3.2.3) i - . . ) s-s (n f s) =(usus + sn-s (3.2.4) ~ n-s s n-s - s(n - s)fsfn-s s (n /s) so that (3.2.2) can be written as - 44 - Til 2 =1 n -5 s2)nn I1Iz(n +s),_,_ (3.2.5) n"+k ln(2 2 sn-s ~2s n-s s n-s (n f s) (3.2.5) is a second-order ordinary differential equation governing the spatial variation of the n-th harmonic. To solve (3.2.5), we require two boundary conditions on fn. The no flux condition at x = -Z provides one boundary condition: n(-z) = 0 (3.2.6) The other boundary condition is obtained by requiring that, at x = 0, the wave field in the harbor match smoothly with that in the ocean. This matching will only be performed in an average way. Equation (2.3.11), which expresses the average wave amplitude at x = 0 in terms of the incident wave amplitude and the horizontal fluid velocity at x = 0, yields the "impedance boundary condition," in(0) = An + ZnUn(0) (3.2.7) Since Zn is O(knSznkn6) and un = in- + 0( 2), we have kn (0)=A nk +0(sk nk) (3.2.8) n (3.2.5), (3.2.6), and (3.2.8) comprise the boundary value problem for link(). The zeroth harmonic can be found directly by integrating (3.2.1), viz. - 45 - 1 u0(x) = - jns(x) u-s(x) +C0 s (3.2.9) n0(x) = -1 us 2 + D0 us (x) Since all velocities, us(x), vanish at x = -z, it follows that C = 0 and u (x) is at most of O(c2). In fact, by the following argument, u0(x) is seen to be at most of 0(s3): Multiplying (3.2.la) by n- and (3.2.1b) by u-n and summing n from -co to CO, we have: I -in (nnf-n + unu-n) + I (nu- =0(s 3,52C 2) (3.2.10) nn n where we have used the fact that k = n2 + 0( 2). Replacing the n summation index n by -p, we see that Y -in (fn n, + uu)-n =X pNGpn + u uP) = 0 (3.2.11) n pp- so that S ('nUn)' = 0(63) (3.2.12) n Integrating over x and using the boundary condition un(-k) = 0 for all n, it follows that I (nnu-n) = OCs3)(3.2.13) n and, from (3.2.9a), with CO = 0, u=(x) - 0(33) (3.2.14) The constant D in (3.2.9b) is determined by the impedance condi- tion at the entrance, (3.2.7). In section 2.1, we noted that, by - 46 - properly defining the mean depth in the ocean, A0 = 0. From (2.3.5), Z= 0, so that i1(0) = 0 and D0 = IY us(0)I2 (3.2.15) s From (3.2.9b), nO(x) is then ,n(x) = Jus(0)1 2 - I Ius(x) 2 (3.2.16) s s Thus n0 is Q(c2). At the back wall of the harbor, all velocities us(-) vanish and no takes on its maximum value, viz. IO(-k)= I us(0)j > 0 (3.2.17) which shows that there is a mean set-up at the back wall. Since n0 and u0 are strictly 0(c2) and O(c3) , they do not have any feedback to other harmonics up to 0( 2). Thus, they can be calculat- ed after all the other harmonics are obtained. For this reason, we shall discuss the zeroth harmonic separately. Equation (3.2.5) represents a system of coupled, nonlinear, second-order differential equations for the harmonic amplitudes fl 1(x), n2(x), '3(x),... . When -n (x) is 0(s), the coupling terms are 0(s2) and are thus relatively small. By earlier theories, i.e., Mei and Unluata (1972) and Bryant (1972), it is known that weak nonlinearity does not produce first order effects within a distance of 0(1). There- fore, for the first few resonant harbor modes, for which z = 0(1), the effect of weak nonlinearity may be treated as a perturbation, as in the next section. - 47 - 3.3 Regular Perturbation Analysis for the Lowest Resonant Modes or for Very Small Amplitude Waves Substituting the perturbation series nn(x) = En (x + s2rj2) (x) + .. (3.3.1) into (3.2.5), we obtain a sequence of linear problems. At 0(E), (1) " + -k (1 ) = 0 n n n (1) n '(-z) = 0 (1) (0) = An - Zn1l)'(0) (3.3.2) n For simplicity, we shall assume that the incident wave is monochromatic, so that only A1 is nonzero. Then, n1 (x) = T1 cos k1(x + z) 1 1 1 T = [cos k1zY - (i/k1 )ZI sin k k]~ A1 n () = 0 n=2,3,4... (3.3.3)n This is just the solution of linearized theory, which exhibits resonances whenever cos kIz + (IM(Z)/k 1 ) sin k2 = 0. At ( 2) "(2) + (2)= 0 T11 + k11 "(2) + k2 n(2) _ 3 (1) (1) 3 (1 ) ,(1) 2222 2 I l ~2 l (3.3.4) =3 2 2 + 211'+24 - T [cos2 k(x+ k) - k2 sin2 k(x +01 - 48 - l"(2) +k (2) =0 n n n = 3,4,5... with the homogeneous boundary conditions (2)( 0 n ri(2) (0) + Zn( 2)-(0) = 0 k n n The solution to (3.3.4) and (3.3.5) is (2)rj =0 n n = 1,3,4,5... (2) T2 {f(x) - ( f(0) + 2i . 2 2f -'O) ) cos k2x + cos k2Z - 2- Z2 sin k2z 2 (3.3.6) I+ k222 1-k2 f(x) 2 2 2) cos 2k (x + ) + 2 k2 .4k1 2k2 The solution for n2)(x) is comprised term, a forced oscillation having the form oscillation having the form cos k2(x + O)j. of three parts: a constant cos 2kI(x + z), and a free Since k - 4k2 = 4(l + Aw2) - 4(l+p 2) = 4p2 k2 -2k ~=2(1 + i2) - 2(1 + y2 =2 (3.3.7) it follows that the constant term is 0(p2) and the forced term is 0(1/12). Thus, the magnitude of e n2 )(x) is generally of order E a(E/2) , in which case the perturbation series of (3.3.1) is valid only for e/p 2 < - 65 - = c7()e in(x-t) + 0~(i)e-in(x+t) E E F +(,x-t) + e F~(x,x+t) (3.4.17) where F (x,) E 0 c (R)e ine (3.4.18) -00 Using the fact that (n + s)cscn-s = I (n - s)csc n-s + I 2scscn-s s S S = 3 1 scs c(3.4.19) s we find from (3.4.9) + 2 3F 3+ aF FF + +3F =0 (3.4.20)6s E: 3 2 as (3.4.20), which is known as the Korteweg-de Vries (or KdV) equation, describes the propagation of nonlinear, weakly dispersive, shallow water waves progressing in one direction. In the limit of no dispersion, p2 + 0, (3.4.20) becomes + 1 F a = 0 (3.4.21)2 as which is a quasi-linear, first order partial differential equation possessing the general solution F = g(e - F X) (3.4.22) - 66 - where g is an arbitrary differentiable function. (3.4.22) is an implicit relation for F. The characteristic curves, along which F(x,e) is constant, are given by 3 C = g(e - Z- ci) (3.4.23) where c is a constant. If the function g(e) possesses a maximum at a finite value of e, then the characteristics of (3.4.23) can be shown to intersect. At the point of intersection, say (x0,e0), F takes on two or more different values. In physical terms, a shock is formed at (X0,o0). Therefore, in the limit of no dispersion, the KdV equation leads to the formation of shocks. When 12 is not zero, (3.4.20) is a third-order partial differential equation, and its solutions are radically different from those of (3.4.21); in fact, no shocks have been found in all the known solutions, analytical or numerical. One special class of solutions, known as perm- anent waves, is of the form F(x,e) = G(e - yx) (3.4.24) where y is a constant, and, from (3.4.20), G(o) satisfies the ordinary differential equation dG 2 3 3 d -Y + =4p-4+ G d o = 0 (3.4.25) (3.4.25) can be solved exactly in terms of the Jacobi elliptic function "cn"; hence the name "cnoidal waves." These waves typically have sharp crests, shallow troughs, and phase velocities which depend on amplitude. - 67 - From (3.4.20), the equation for FC does not involve F_ and vice versa. This means that the oppositely directed waves F+ and F do not interact resonantly to 0(s) in our perturbation analysis. Benney and Luke (1964) have used this fact to suggest that nonlinear standing waves can be constructed by superposing two cnoidal waves of equal amplitude travelling in opposite directions. Such waves differ from linear stand- ing waves in that they do not have fixed nodal points and the free sur- face is at no time perfectly flat. The solution to (3.4.9), which satisfies the complicated impedance boundary conditions at x = 0, in general will bear no simple resemblance to cnoidal waves or nonlinear standing waves. In the next section, we present a numerical procedure for obtaining the solution for the wave field in the harbor. 3.5 Numerical Solution by Iteration The infinite set of differential equations (3,2.5) is truncated at the N-th harmonic, yielding the system n + k n = R ns'n- +)Y S '%s s s n'(-z) = 0 nnnilO) = A - =3 Z -'Z~(Q) n n k 2 n n n = 1,2,3...N (3.5.1) where - 68 - k2 En 2(l 1-P2 2kn ( 3~- S 1 2 2 R =ns= (n -s) Si (n + s) ns2 n - s Zn = k6 [ + 2i en k 6 - Zn ) ( ) (3.5.2) (3.5.1) is as a nonlinear, two-point, boundary value problem. Unlike an initial value problem, which can be solved by direct numerical inte- gration, the solution of (3.5.1) requires an iterative procedure. At the (p + l)-th iteration, we solve the linear boundary value problem (p+l)" + k2 (p+l) = R n(p)I(p+) V Sn'(P)n(P+l) n n n ns s n-s + s ns s n-s ryjp+-) (-s)4=O0 n (p+) (0) = A - (ln)Z p+i) n n = 1,2,3...N (3.5.3) using finite difference methods. The rate of convergence of successive solutions can be increased by using a relaxation technique, which consists of replacing n (Pl)(x) by 7-(p-I)(x) = Xn (x) + (1 - X)fn (x) (3.5.4) With X = 1, convergence of successive iterates was found to be slow and oscillatory; with X = .5, convergence was much faster and, in most cases, - 69 - monotonic. The iterative procedure is terminated when successive solu- tions for nn(x) differ by less than .001, that is, when Max ) (x) - 4P(x)I < .001 x n = 1,2,3...N (3.5.5) If the maximum value of InN(x)I exceeds .001, truncation at the N-th harmonic is deemed inadequate and the system of differential equations is expanded to include the (N + 1) harmonic. The expanded system is then solved by iteration. This procedure is continued until a "penultimate" solution is found, for which Max In,*(x)I < .001 (3.5.6) x where N* is the highest harmonic in the solution. In the finite difference solution of (3.5.3), the domain -z < x < 0 is represented by NPT equally spaced points located at positons x = -P + (J - l)-H, J = 1,2,3...NPT, where H is the interval size H = Z/(NPT - 1) (3.5.7) When the solution requires many harmonics, the choice for H must be made correspondingly small. For example, if we insist that one wave- length of the highest harmonic shall be represented by at least ten points, then the requirement on H is 2w kN*H> 1, so that, in the domain 0 < x < zm all Hankel functions may be represented by their asymptotic forms. Then CO 2inkp, E(x) -/fQeikx ei m (4.2.11) n=1 v'2nkkm + kx - 82 - yj qJ 0QS K C 4 JhA d 0 1. 0I d E 0 6 A K (*4M ) 94 Fig. 4.3. Theoretical study of multiple reflections from wavemaker: Placement of images. (0 Except in the case k'm = (integer)-w, for which the sum in (4.2.11) diverges, we may approximate E(x) by Co i(2kz )v E(x) ~ Qedikx d e r (4.2.12)Q7jdv 2kkmv + kx m The integral in (4.2.12) can be evaluated explicitly in terms of Fresnel integrals; but, for the purpose of estimation, we simply integrate once by parts to obtain 2ikt E(x) 7 Qeikx f(2ikz) e } + 0(k.m)-5/2 (4.2.13) m 2kkM + kx From (4.2.9), the net result of all multiple reflections is i(2kz + kx) i(2kz - kx) ii(x,o) ~ {L " m e m )+0k -/ /ZXo) k m /2kim + kx v2kz -kx m (4.2.14) The interesting feature of (4.2.14) is that, whereas the effect of a single reflected wave is O(kmF)- 1/2 , the combined effect of all reflec- tions is only of O(kkm)- Our conclusion therefore is that, so long as kk >> *I (4.2.15) kkm t (integer)-w successive reflections from the wavemaker add destructively to produce a net effect whose magnitude is only 0(kM)3/2 When (4.2.15) is satisfied, the total wave field in the laboratory "ocean" is - 84 - n(x,y) = (si k )cos kx + QH(l)(kr) + 0(kzm)-3/2 (4.2.16)n~~) sin ky 0m (..6 and the effect of multiple reflections is small. When either of the conditions in (4.2.15) is violated, the effect of multiple reflections can be quite significant. When this occurs, the harbor basin and ocean basin are said to be strongly coupled. Condition (4.2.15) states that strong coupling will occur unless (a) the ocean basin is very much larger than a wavelength and (b) the frequency of oscillation is not close to a resonant frequency of the ocean basin. Figs. 4.4 (a),(b),(c)show experimental measurements of the wave field in the model ocean, along the centerline x = 0. The experimental values Lm = 31'2" o = 2f/(1.545 sec) = 4.067 sec 1 h = .5' yield km = 32.96 = 10.49n so that (4.2.15) is satisfied. The graph at the top shows the standing wave amplitude ns(x,o) in the ocean, when the harbor entrance is closed; the solid curve is (4.2.5). The middle graph shows the total wave field when the harbor entrance is open, and the first resonant - 85 - .0 Co0 2f 'itrrf Sir /Orr Fig. 4.4a. Experiment (.) vs. theory (-) for the wave amplitude in the ocean (h = .5 ft., 2a = .33 ft., L = 1.211 ft., w = 4.067 sec-1 , L = 31.17 ft.). Top: harbor closed. Middle: harbor open. Bottom: the radiatei wave. A1 = .015. I 5 CO .0 .5 2r '-r 4w trr Fig. 4.4b. Experiment () vs. theory (-) for the wave amplitude in the ocean (h = .5 ft., 2a = .33 ft., L = 1.211 ft., w = 4.067 sec~ , L = 31.17 ft.). Top: harbor closed. Middle: harbor open. Bottom: the radiatd wave. A1 = .027. ofo O.. CO Fig. 4.4c. Experiment () vs. theory (-) for the wave amplitude in the ocean (h = .5 ft, 2a = .33 ft., L = 1.211 ft., w = 4.067 sec-, L = 31.17 ft.). Top: harbor closed. Middle: harbor open. Bottom: the radiat~d wave. A1 = .040. harbor mode (having L = 1.211 ft.) is excited; the solid curve is (4.2.16). The bottom graph shows the component of the total wave field which is the radiated wave, nR(x,o) = QH l)(kx). Since the agreement is fairly good, it confirms experimentally that (i) the ocean can be treated linearly and (ii) reflection from the wavemaker is not signi- ficant. 4.3 Real Fluid Effects In the following two sections, all variables and coordinates are dimensional. 4.3.1 Viscous Dissipation in Wall Boundary Layers In a laboratory setting, the damping of surface waves in a slightly viscous fluid is a well-known problem. For infinitesimal waves of frequency w, propagating in a fluid of viscousity v, the fluid motion is essentially irrotational except near boundaries, where viscous boundary layers of thickness 6v = /2v/w(4.3.1.1) are formed. Viscous energy dissipation occurs in (a) the boundary layers near solid walls, (b) the boundary layer near the free surface, and (c) the main body of the fluid. Ursell (1952) has shown that the solid wall boundary layers account for most of the dissipated energy. For oscillatory boundary layers, the transition from laminar to turbulent flow on a smooth flat surface occurs at the critical - 89 - Reynolds number (see Phillips (1966), p. 43) R - 6 -~160 c v (4.3.1.2) where U is the particle velocity. In shallow water, U9 = A-/iY (4.3.1.3) where A is the wave amplitude, and c vw (4.3.1.4) In the experiments, h = .5 ft., gh = 16 ft2/sec2, w = 4.067 sec" 1, v = l0-5 ft2/sec, and R = (g)-887 (4.3.1 .5) so that the transition to turbulence occurs when (c) c = .18 (4.3.1.6) In all but one of the harbor resonance experiments, the value of A/h was below this critical value; therefore, laminar flows prevailed. Batchelor (1976, pp. 353-358) derives the following expression for the time-averaged power loss per unit area in an oscillatory laminar boundary layer: dP vP 1 2 S - p IUTI v (4.3.1.7) where Re(UTe-ut ) is the local tangential velocity at the boundary according to inviscid theory and p is the fluid density. - 90 - To estimate the viscous losses in the far field of the long, narrow bay, we shall use the linear expressionsfor the wave field, n(x,y,t) = Re(T cos k(x + L)e-iwt) (4.3.1.8) tI(x,y,t) = Re(i gkT sin k(x + L)e-lwt)x + Q(y2) where 0(p2) terms include the vertical component of velocity and the depth-dependent terms of the horizontal velocity. The expression for T in terms of the incident wave amplitude is found in Equation (1.3.6). From (4.3.1.7), the power loss per unit area on all surfaces parallel to the x-axis is dP vv 1 PVgkT12 sin2 k(x + L) (4.3.1.9)dA 2 v6 xW Lv Integrating over the side walls and bottom of the channel, we obtain for the total power loss P = (2a + 2h) f()dx -L =I pv |w2(2a + 2h) -L(kL - } sin 2kL) ~-jgkT|2(a + h)L (4.3.1.10) v Let us compare P v with the energy loss due to radiation damping at the harbor entrance. The time-averaged radiated power is given by 27T/c P= (2a-h).- dt pgn(ot)u(o,t) 0 = (ah)pg(g9)Tj2 sin 2kL - 91 - !~ah)pg(9K) Tj 2 (4.3.1.11) and, from (4.3.1.10), P /P L(k) a +h L yRR a a = (a2h)(kL)(v) (4.3.1.11) In the harbor resonance experiments, a = 2 in., h = 6 in., W = 4.067 sec , v =10- ft2/sec and - =/- 2.22 x 10-3ft. kL (n + )Tr n = 0,1,2 (4.3.1.12) so that P RR=~(n + )-(2.79 x 102 (4.3.1.13) Thus, the effect of laminar viscous damping is relatively unimportant for the first few modes. It should be noted, however, that a small amount of surface roughness can lead to the formation of turbulent boundary layers, for which v is effectively much larger than its laminar value. For this reason, care was taken to make the walls of the harbor as smooth as possible. In a large scale harbor, surface roughness and the Reynolds number in (4.3.1.4) are much greater than in the laboratory, and so, the boundary layers can be turbulent. From the experiments of Jonsson and Carlsen (1976), the effective viscosity ve is typically of 0(100v). - 92 - Taking the following typical values for a large scale harbor h = 20 m 2a = 100 m L = 1 km Ve = 100V = 10r4 m2/sec kL(1/4 wave mode) = 1.41 w(1/4 wave mode) = .02 sec 1 we obtain = .1 m P RP~R(.7)(1.41)(.005) ~ .005 so that the power loss in the turbulent boundary layers is again negligible. 4.3.2 Separation Loss at the Entrance Except at very low velocities, flow separation occurs at the sharp corners of the harbor entrance, and energy is expended in the production of eddies. The mechanism of flow separation is related to the adverse pressure gradient caused by the sharp curvature of the boundary. If the flow in the entrance is assumed to be quasi-static, then the pressure drop due to flow separation can be described by the hydraulic formula for steady flow n(0~,t) - (0+,t) = |%Iu(0,t)Iu(0,t) (4.3.2.1) - 93 - where 0 is on the harbor side of the entrance and 0+ is on the ocean side. The constant f is an empirical friction factor whose value depends on the Reynolds number and the geometry of the entrance. For an asymmetrical entrance, f may take on different values depending on whether u(0,t) is positive or negative. An expression similar to (4.3.2.1) was used by Ingard and Ising (1967) and Ingard (1970) in studying nonlinear sound transmission through an orifice, with application to the absorption of sound at high pressure levels. Experimental measurements of both the energy loss and the distortion of the frequency spectrum of the transmitted sound were found to agree with the hydraulic, steady-flow formula. In particular, for a monochromatic wave incident on a symmetrical orifice, only odd harmonics are generated by the nonlinear term fuful; whereas, for an asymmetrical orifice, both even and odd harmonics are generated. A reduction in the mass reactance, or inertia, of the orifice was also observed by Ingard and Ising (1967), but this is not accounted for in (4.3.2.1). The magnitude of the mass reactance depends on the details of the velocity field, which is difficult to calculate when the flow is separated. Since the inertia of the orifice is associated primarily with the irrotational part of the flow, it is expected to be smaller when flow separation occurs. At high sound pressure levels, the mass reactance is observed to be one half its unseparated value, which suggests that, at any given time, the flow field on one side of the orifice is fully separated. In water wave problems, separation loss has been studied in connection with the construction of perforated breakwaters; see - 94 - Jarlan (1965); Terrett, Osorio, and Lean (1968); and Mei, Liu, and Ippen (1974). The separation loss at a narrow harbor entrance has been studied numerically by Ito (1970) and analytically by Unluata and Mei (1975). The latter show that the higher harmonics generated by the nonlinear term -!-Ju(O,t)|u(0,t) in (4.3.2.1) are not large under usual2g circumstances; rather, the primary effect of separation is to reduce the amplitude of the fundamental harmonic. We shall therefore present a simplified treatment of flow separation based on a monochromatic entrance velocity, u(0,t). To allow for an asymmetrical harbor entrance, we shall approximate f by f 1 when u(0,t) > 0 f = { (4.3.2.2.) f 2 when u(0,t) < 0 This is a mathematical simplification since f is more likely to be a continuous function of time. We then have f-uluj = uju + +fu2 (4.3.2.3) where I f +f2 (1 f2) (4.3.2.4) 2 1 Next, we Fourier analyze each term of (4.3.2.3): -u | = UI LY eint2g lul -2g n - 95 - fU2= nXf - int (4.3.2.5)2g 2g _0n If we assume that u(0,t) has the simple form u 'Ue-it + U*eit) = jul cos(t-t) (4.3.2.6)2 where is the phase of U, viz. U =uele then the Fourier coefficients in (4.3.2.5) are found to be fi f ,T u Iueint dt 0 4 sin(nir/2) fUJ2 ein 'r n(4-nt ) 2 2 . (4.3.2.7) f = -L f u2 e'nt dt (o1/2 for n = 0 = fjUl 2 ein . 1/4 for n = +2 0 for all other n Observe that fn is zero for even values of n; thus, we retrieve the known result that, for a symmetric entrance (f=O), only odd harmonics are generated by separation. For an asymmetrical entrance, the second and zeroth harmonics are also generated; higher even harmonics would also be generated if u(O,t) were more complicated than in (4.3.2.6). For steady flow through the entrance of Fig. 1.1, the values of f and f2 are found to be 1.0 and .4, respectively, according to Daily and Harleman (1966), pp. 314-319. For oscillatory flows, the separation zone on either side of x = 0 has only a half wave period during which to develop. It may therefore be expected that the effective friction factors are smaller for oscillatory flow. Moreover, in the experimental - 96 - harbor model, the sharp corners at x = 0 are slightly rounded, which further reduces flow separation. Thus, the average friction factor 1(f1 + f2) in the experiments is expected to be smaller than .7. We shall see in section 5.1 that the value of f in the experiments is roughly half this value. Taking f1 = .5 and f2 = .2 in the experiments, we have f = .35, f = .15 and f = TIU|2 1 - .15|UU1 3'T .j f3 = 4.f|Ui2e3i4 = .03 U3/JUJ3 157T (4.3.2.8) f 0 = 2f UI = .075jUI f2 = If IU 2e2 ic1 = .0375 U2 Since U is 0(s), the above coefficients are all 0(62). Note that the main effect of separation is to alter the first harmonic, through the coefficient f1. The term fi will lead to a slight increase in the mean set-up, n0(x). Taking the first harmonic component of (5.3.2.1) and using the impedance condition, n1O(0+) = A1 + Z1U, on the ocean side, we obtain YO(0) = A1 + Z1U + T/g (4.3.2.9) A I+ (Z1 + c )U where c, = $ TU|/g (4.3.2.10) - 97 - The constant ce is an effective damping constant due to separation. Eq. (4.3.2.9) is an effective harbor boundary condition which includes (i) the forcing of the incident wave, A.; (ii) the inertia of the har- bor entrance, Im(Z); (iii) the damping due to the radiated wave, Re(Z1); and (iv) the damping due to separation loss at the entrance, ce. Using (4.3.1.8) for the wave field in the harbor, we have ri (0~) = T cos kL (4.3.2.11) U = i T sin kL (4.3.2.12) and, from (4.3.2.9), T = A[cos kL - i9kS(Z + c ) sin kL] co 1 e = A[(cos kL + 9k Im(Z1) sin kL) - i9 (Re(Z ) + 4 k T|T|) sin kL 1-~W 1 W 1 37r w (4.3.2.13) Resonance occurs, as before, when cos kL + Im( Z1) sin kL = 0. At resonance, ITI = IAl[ (Re(Z 1) + - k K |TI) Isin kL|]~I = ITJ|[l + aT/|Tfol] (at resonance) (4.3.2.14) where T| =AJ(9 t Re(Z 1)|sin kL)~ = ITI when f = 0 (no separation) S-ky |T/Re(Z) (4.3.2.15)37JTJwRe - 98 - Physically, a measures the amount of energy loss due to separation as compared with that due to radiation damping. When a >> 1, separation losses dominate radiation losses. Solving (4.3.2.14) for ITI/ITJ|, we find |Tj/jTo| = 1 1 + 4cc ~ a-1/2 when a >> 1/4 (4.3.2.16) A plot of fT/IT vs. a is shown in Fig. 4.5. Suppose we allow Re(Z) to approach zero while maintaining f and A constant (and nonzero). This corresponds to the well-known harbor paradox, in which Re(ZI) is made smaller by decreasing the width of the harbor entrance and JT0j is observed to increase without limit. From (4.3.2.16), we see that |TI/IT0! + 0 and that, specifically, IT! jT a-1/2 = (4 f)-/2 (Re(Z)fTf)1/2 = (I 4k -/2 (JJAl 1/2 (4.3.2.17) 37 r w kisin kLj Instead of increasing without limit, ITI approaches a finite value proportional to (IAI/f)l/2. Thus, the presence of separation losses effectively removes the harbor paradox. - 99 - 00 I /.I/- V So /oo Fig. 4.5. Entrance loss: Normalized first harmonic amplitude, JT/T0j, vs. normalized friction factor, a. CHAPTER V EXPERIMENTAL AND THEORETICAL RESULTS 5.1 Comparison of Theory with Experiment The experimental set-up is shown in Fig. 4.1, where the coordinates are defined. Five length scales are involved in the study of the long, narrow bay: harbor width, 2a; mean water depth, h; harbor length, L; wave- length of the first harmonic of the incident wave, approximately vg"hi-2ir/w; and incident wave amplitude, |A1 Jh. In the experiments, these had the values 2a = 1/3 ft. h = 1/2 ft. L = 1.211, 4.173, and 7.136 ft. /gh.(2w/w) = (4.0125 ft/sec) (1.545 sec) = 6,200 ft. A1 -h = .0075, .0135, and .020 ft. Three harbor lengths and three incident wave amplitudes were tested. The corresponding dimensionless parameters are 6 E a-(w//gh) = .169 9 E L.(w/vg-h) = 1.227, 4.230, and 7.233 2 2P = W h/g = .257 A1J = .015, .027, and .040 The values of z were chosen so as to produce the first three linearly - 101 - resonant harbor modes. Each resonant mode was studied for three different values of 1A1 I; so, altogether, nine resonant harbor oscillations were examined. The incident waves produced by the wave generator are not per- fectly monochromatic. With the harbor entrance closed, the standing wave amplitude in the ocean was measured at x = y = 0 and was found to have the harmonic composition given in Table 5.1. The phases, Pn, are defined by An E aAnt n and are given in radians. Table 5.1 |A1 1 .015 .027 .040 A21 .001 .003 .012 A31 .000 (2) .001 .003 2 - 21 6,210 5.646 5.506 3 - 3 .671 1.795 5.314 These are used as inputs in the numerical calculations. Table 5.2 compares the wavenumbers kn calculated by the approximate Boussinesq dispersion relation (knh)2 =n2 2(1 +TI n 2) + 0(np)6n3 2 = ,2 where p = W h/g = .257, with those calculated by the exact dispersion relation n2 2 = (knh) tanh (knh) - 102 - which is valid for arbitrary depth, e.g. arbitrary (np). Table 5.2 n Exact kn (ft:.) Boussinesq kn (ft:1 ) 1 1.060 1.056 2 2.448 2.350 3 4.710 4.048 4 8.228 6.244 5 12.850 8.986 Note that the error in the Boussinesq values for kn is quite large for the higher harmonics. The theoretical assumption that the far field of the harbor is one- dimensional is only valid when transverse modes of oscillation, or "cross modes," are not present. The minimum wavenumber, kc, for which cross modes are possible, is given by (kC).(2a) = T so that, with 2a = 1/3 ft., kc = 9.42 ft:1 From Table 5.2, we see that cross modes can exist for the fifth harmonic or higher. In the experiments, the amplitudes of these harmonics were negligible, and the far field of the harbor was observed to be approx- imately one-dimensional. - 103 - The wave amplitude in the harbor was Fourier analyzed at intervals of two inches along the centerline of the harbor channel. The zeroth harmonic, or mean set-up, was removed by the Fourier analysis. The vari- ation of the first three harmonic amplitudes with distance from the back wall of the harbor is shown, for each of the nine experimenal cases, in Figs. 5.1(a) through 5.1(i). The horizontal dotted line indicates the maximum amplitude of the first harmonic predicted by linearized theory. The solid lines are theoretical curves based on the inviscid nonlinear theory of Chapter 3. These solutions were computed by the numerical procedure of Section 3.5, with the incident wave composition given in Table 5.1 and the Boussinesq values of kn given in Table 5.2. Convergence of the iteration scheme was achieved in all cases with five harmonics or less. For the shortest harbor (L = 1.211 ft.) the higher harmonics are small, owing to the short distance over which nonlinear interactions take place, and the inviscid nonlinear theory agrees closely with linearized theory. The reduction of the first harmonic amplitude is due primarily to separation loss at the entrance, rather than to nonlinearity, and will be discussed later. In the two longer harbors (L = 4.173 ft. and 7.136 ft.), there is considerable harmonic generation; in some instances, the second harmonic is as large as fifty percent of the first. In such cases, the second harmonic siphons an appreciable amount of energy from the first harmonic, and therefore, the maximum amplitude of the first harmonic is observed to be significantly less than that predicted by linearized theory. For the longest harbor (L = 7.136 ft.), the agreement between experiment and - 104 - and inviscid theory is fairly good. In particular, the significant reduction of the first harmonic and the presence of higher harmonics are well accounted for by the nonlinear theory. However, for the second harbor (L = 4.173 ft.), we note the qualitative feature that the ob- served second harmonic is higher than calculated, and that the observed reduction of the first harmonic is larger than that calculated. These discrepancies cannot be attributed to harmonic generation by flow separa- tion at the entrance, as this was shown to be quite small, in Section 4.3.2. We believe that part of the discrepancy between theoretical and experimental values of the higher harmonics is most likely due to errors in the Boussinesq values for kn, cf. Table 5.2. Errors in the values of k n lead to corresponding errors in the positions of nodes and anti- nodes of the harmonic amplitudes. These errors increase with increasing distance from the back wall of the harbor. Furthermore, because the harbor response is large at resonant values of knL, a small error in k n can significantly alter the absolute magnitude of the n-th harmonic. We suspect that this happens for the second harbor (L = 4,173 ft.). In this case, the Boussinesq value of k2L is 9.807 which does not lie close to a resonant mode; whereas the exact value of k2L is 10.216, which 7 coincides with the fourth resonant mode, having kL = f + Im(Z). Because of this, the second harmonic predicted by Boussinesq theiry is expected to be smaller than that observed in the laboratory. The energy conservation theorem of Section 3.4 implies that the first har-- monic must also be affected. Figures 5.2(a), (b), and (c) show the results of a numerical - 105 - calculation in which the exact values of kn are used instead of the Boussinesq values, for the second harbor. Note that the second har- monic is now larger and the first harmonic smaller than that obtained previously (cf. Figs. 5.l(d),(e),(f)). This improves the agreement between theory and experiment, though the quantitative agreement is still not as good as for the long harbor (L = 7.136 ft.) case. Of course, this ad hoc correction of the values of kn must await rigorous justifica- tion by a nonlinear theory which is valid for arbitrary depth. We have used the shortest harbor to determine the average friction factor, ? = (f + f2), which appears in the entrance loss theory of2 1 Section 4.3.2. The value f = .35, which is half the upper limit expect- ed on the basis of steady flow values for f1 and f2, was found to give good agreement with the experimental data on the shortest harbor. In Figs. 5.3(a) through 5.3(i), the experimental data for all nine harbor resonances is compared with theoretical curves computed using the entrance loss theory of Section 4.3.2, with ? = .35. Since harmonic generation by separation has been shown to be small, only the main effect of separation, e.g. reduction of the first harmonic, has been included in the numerical computation. The inclusion of entrance loss does not significantly improve the agreement between experiment and theory for the second and third harbor lengths. In particular, the quantitative agreement is still better for the longest harbor than for the medium -ength harbor. By using T = .35 and an ad hoc application of the exact dispersion relation, as explained previously, we obtain the theoretical curves shown in Figs. 5.4(a),(b),(c) for the medium length harbor. - 106 - *10 .4L, Fig. 5.la. Experiment vs. inviscid nonlinear t'heory (f = 0). (A,+,x): measured first, second, and third harmonic amplitudes. (-): nonlinear theory. (----): linear theory for first harmonic amplitude. L = 1.211 ft., A1 = .015. C 00 2& L 1/. Fig. 5.1b. L = 1.211 ft., A = .027 (See Fig. 5.la). .301 .151 10 Fig. 5.lc. L = 1.211 ft., A1 = .040 (See Fig. 5.la). to. ;t0 9 '4. '4 '4 '4 '4 Sb & .05 & a & & U, * , hi: 'a:'. *: L = 4.173 ft., A1 = .015 (See Fig. 5.la).Fig. 5.1d. &4 5 & & & & 4 42K32: Fig. 5.le. L = 4.173, A1 = .027 (See Fig. 5.la). .1-1 r%33t Fig. 5.1f. L = 4.173 ft., A1 = .040 (See Fig. 5.la). IPM am Oft 4e OR .. o - e/m1%9I}\\1/ *os X 10 L = 7.136 ft., A1 = .015 (See Fig. 5.la).Fi g. 5.l1g. *1\ X+ Fig. 5.1h. L = 7.136 ft., A3 .027 (See Fig. 5.1a). .30 / \ C- 39 I\ / / \ 01 - +4*.+ I\t L = 7.136 ft., A = .040 (See Fig. 5.la).Fig. 5.1i. .lc - WAD +2 3 Fig. 5.2a. Experiment vs. inviscid nonlinear theory (f = 0) with exact dispersion relation. (A,+,x): measured first, second, and third harmonic amplitudes. (-): nonlinear theory. (----): linear theory for first harmonic amplitude. L = 4.173 ft., A1 = .015. .et foo am\u a x 3c 33 Fig. 5.?b. 1 = 4.173 ft., A1 = .027 (See Fig. 5.2a). 930 wo 3cIr-r+-93 itt rf Fig. 5.2c. L = 4.173, A = .040 (See Fig. 5.2a). )42 Fig. 5.3a. Experiment vs. nonlinear theory with separation loss ( = .35). (A,+,x): measured first, second, and third harmonic amplitudes. (-): nonlinear theory. (---): linear theory for first harmonic amplitude. L = 1.211 ft., A1 = .015. sawoh 46 Fig. 5.3b. L = 1.211 ft., A1 = .027 (See Fig. 5.3a). -4 13'4 I Fig. 5.3c. L = 1.211 ft., A1 = .040 (See Fig. 5.3a). ,30 .a5S- -J A6 /.2 w6a It .10~ a &3 Fig. 5.3d. L = 4.173 ft. , A1 = .015 (See Fig. 5.3a). 4m I %% &a & / 393 Fig. 5.3e. L = 4.173, A = .027 (See Fig. 5.3a). .30 - /9x39 31/ i i i i : : i/ .153 Fig. 5.3f. L = 4.173 ft., A1 = .040 (See Fig. 5.3a). .0 am J112 01P/ l o% /1~4 & + L = 7. 136 ft., A1 = .015 (See Fig. 5.3a).Fig. 5.3g. e4t* ql ~ % /-9 / 3tX 3/3 /C L = 7.136 ft., A1 = .027 (See Fig. 5.3a). I 4e we %* %. Fig. 5.3h. / \ / \ '*+Is+. . ./+/ & x L = 7.136 ft., AI = .040 (See Fig. 5.3a). .30 w7 Fig. 5. 3i . \ / \ /4 ama Fig. 5.4a. Experiment vs. nonlinear theory with separation loss (F = .35) and exact dispersion relation. (A,+,x): measured first, second, and third harmonic amplitudes. (-): nonlinear theory. (----): linear theory for first harmonic amplitude. L = 4.173 ft., A1 = .015. 5lo W, &m/Gr 4m &~ 4D. at s 3a ax / a3 9 Fig. 5.4b. L = 4.173 ft., A1 = .027 (See Fig. 5.4a). .30 r -4.. r r C r '4 / '4, / '4 / "S .13- 4 & & 03C a / a '6 5.4c. L = 4.173 ft. ,A1 = .040 (See Fig. 5.4a) . 5.2 Some Theoretical Predictions for a Large Scale Harbor Having obtained some experimental confirmation of the nonlinear theory, we shall investigate theoretically the response of a harbor having the practical dimensions 2a = 100 m, h = 20 m, and L = 1000 m. The first three lineary resonant modes occur for z = wL//g-i = 1.41, 4,36, and 7.36 corresponding to wave periods of 5.305, 1.716, and 1.016 minutes, respectively. Table 5.3 summarizes the linear results per- taining to these harbor modes. Table 5.3 _ p_ Amplification (.n1C-z)/A1 ) 1.41 7.95 x 10-4 .0705 14.35 4.36 7.60 x 10-3 .218 4.83 7.36 2.17 x 10-2 .368 2.99 Note that each value of P 2 is much smaller than the corresponding value of 6; this is typical for large scale harbors. In practice then, the assumption 6 << 1 is likely to be more restrictive than the assumption P2 << 1. p2 Consider the response of the harbor to monochromatic incident waves, with normalized amplitude A1 = .03, A2 = A3 = A4... = 0. Linearized theory predicts that, for the first resonant mode (z = 1.41), the normalized wave amplitude at the back wall of the harbor will be ,nl(-k) = .430. This corresponds to a peak-to-trough wave height of .86 times the depth, which is greater than the value at which waves break in - 131 - shallow water. The predictions of inviscid nonlinear theory are shown in Figs. 5.5(a) through 5.5(d), for the case A = .03. Fig. 5.5(a) shows the number of harmonics required in the numerical solution, N*, versus the normalized incident wave frequency, z = wL/vgW. As expected, the greatest number of harmonics are required at values of z corresponding to resonant harbor modes, for which the wave amplitude in the harbor is large. Fig. 5.5(b) shows the variation of the first harmonic amplitude at the back wall, (rI(-9J), with incident wave frequency, w. The solid line is the prediction of linearized theory (and regular perturbation theory, also); the points are nonlinear results. According to nonlinear theory, the first resonant mode occurs at z = 1.37 and has a peak height of .382; as compared with z = 1.41 and a peak height of .430 in linear- ized theory. For the second and third resonant modes, the resonant &mplification is smaller, and the discrepancy between nonlinear and linear theory is therefore smaller. Fig. 5.5(c) shows second harmonic generation due to nonlinearity. The solid line is calculated from the regular perturbation theory of section 3.3; the points are numerically-generated results. From Table 5.3, we see that, in the range of z covering the first three 2 resonant modes, z << 1/B2. Therefore, we expect that regular perturbation will adequately describe second harmonic generation so long as the wave amplitudes are not too large. Comparing 5.5(c) and 5.5(a), we see that the major discrepancies between regular perturbation and numerical theory occur where the numerical solution required many more - 132 - than two harmonics. The various peaks in second harmonic generation have been discussed in Section 3.3; here we simply recall that the three central peaks correspond to linear resonance of the first har- monic, whereas the two small peaks flanking each central peak correspond to linear resonance of the second harmonic. The effect of dispersion increases with increasing k and tends to shift the small peaks to the right relative to the central peak. At k = 7.36, one of the small peaks has almostoverlapped its adjacent central peak; this corresponds to a situation in which both the first and second harmonics are reson- ated simultaneously. As a result of this phenomenon, the peaks in Fig. 5.5(c) do not diminish with z as fast as the first harmonic peaks of Fig. 5.5(b). The variation of J|l 3 (-z)J with z is shown in Fig. 5.5(d). Note that the largest peak in Ir3(-R)| occurs at z = 1.41, corresponding to the first resonant harbor mode, and has a magnitude of .023, which is comparable to the magnitude of n2(-0). This is contrary to the usual situation in which the higher harmonics are progressively smaller. The reason for this can be seen from Table 5.3. Since the second resonant harbor mode (z = 4.36) occurs at a frequency very nearly three times that of the first resonant harbor mode (z = 1.41), we see that, when z = 1.41, both the first and third harmonics are resonated simultaneously. This is quite generally true for the long, narrow bay, because its resonant modes appear at roughly odd multiples of '/2. This resonance mechanism for the third harmonic is similar to the one occurring at z = 7.36 for the second harmonic, though the underlying reasons are different for the two cases. - 133 - Fig. 5.5(e) shows the mean set-up at the back wall, n(-z), vs. Z. The solid line is the prediction of regular perturbation theory, cf. Fig. 3.1(c); the points are calculated from the numerical solutions using (3.2.17). Note that the peaks occur at values of z for which the first harmonic is linearly resonated. Since no depends primarily on the square of nl, the peaks of no diminish in height more rapidly than thoseof T,. The numerically-predicted spatial dependence of the harbor response is shown in Figs. 5.6(a)-(c), 5.7(a)-(c), and 5.8(a)-(c) for each of the first three resonant modes. These figures may be compared with the predictions of regular perturbation theory, shown in Figs. 3.2, 3.3, and 3.4, Note that the higher harmonics are progressively smaller; this is the a posteriori justification for truncating the numerical solution at a finite number of harmonics. The differences between the numerical solutions and regular perturbation theory are mainly due to the fact that the latter overestimates the first harmonic amplitude. For incident waves with amplitudes greater than A1 = .03, the numerical schemefrequern1y requires more than ten harmonics to achieve convergence. This is not entirely satisfactory because the value of 6 associated with the tenth harmonic, at say k = 4.36, is 2.18. This exceeds the value 6 =7T/2 = 1.5708 at which cross modes first appear in the bay. Thus, at larger amplitudes, the assumption of one-dimension- ality in the far field of the harbor is no longer tenable. Further- more, the peak-to-trough wave heights predicted by the nonlinear theory exceed half the depth. At such large amplitudes, the waves are no longer correctly described by the Boussinesq equations. - 134 - Thus far, we have not accounted for entrance loss due to separation, whose primary effect is to reduce the resonant amplitude of the first harmonic. The magnitude of this reduction depends upon the average friction factor f, which in turn depends strongly upon the particular shape of the harbor entrance. As argued in Section 4.3.2, we expect ? to be less than .7 for the long narrow bay. To estimate the influence of separation, we shall use the conservative value f = .35, which coincides with the value chosen for the laboratory experiments. Of practical interest is the relative importance of nonlinearity and separation in reducing the resonant amplitudes of the first harmonic. Fig. 5.9 shows the variation of 1n1(-z| with increasing incident wave amplitudes, A1 I) for each of the first three resonant harbor modes. The dashed lines represent linearized theory; their slopes are just the amplification factors of Table 5.3. The symbols "A" and "x" indicate the values of Inj-z)I according to inviscid nonlinear theory (f = 0) and nonlinear theory with separation (f = .35), respectively. Note that, for the first resonance mode (t = 1.41),, the combined reduction due to both separation and nonlinearity is roughly twice that due to nonlinearity alone; thus both effects are of roughly equal importance. For the second and third harbor modes (z = 4.37 and 7.36), the effect of nonlinearity dominates the effect of separation. This is reasonable since nonlinearity increases with increasing z, whereas separation does not. The spatial dependence of the harbor response according to numerical theory with f = .35 and |A1 | = .03 is shown in Figs. 5.10(a)-(c), 5.11(a)-(c), and 5.12(a)-(c), for each of the first three resonant - 135 - modes. These figures may be compared with the predictions of inviscid nonlinear theory, shown in Figs. 5.6, 5.7, and 5.8. As expected, the inclusion of separation losses leads to a reduction in the predicted wave heights. For the first resonant mode, this reduction is sub- stantial; for the second and third resonant modes, the viscous theory agrees fairly closely with the inviscid theory. When the incident wave is not monochromatic, nonlinear interactions generate a great many new frequencies, some of which may be resonated by the linear mechanism. Consider, for example, an incident wave having A1 = 0 but A2 and A3 nonzero, as in Fig. 5.13(a). The incident wave spectrum is dominated by two frequencies, 2w and 3w, where w coincides with the first resonant harbor mode, viz. z= wL//gh = 1.41. The harbor response according to linearized theory is simply an amplification of A3, corresponding to the resonant mode with k = 4.36, and is shown in Fig. 5.13(b). However, nonlinear theory allows for the generation of additional harmonics; among them the first and fifth harmonics, both of which coincide with resonant harbor modes. Fig. 5.13(c) shows the harbor response at x = -z predicted by inviscid nonlinear theory. Nine harmonics are present in the numerical solution. This means that the wave spectrum of an incident wave can be significantly changed by non- linearity. This possibility, somewhat similar to subharmonic resonance in simpler physical systems, can be important to the design of mooring systems for ships in the harbor. - 136 - -I .2 I 9. I Ct) Fig. 5.5&-f Inviscid nonlinear theory 2a = 100 m., L = 1000 m., numerical solutions. = 0): A1 = .03). Frequency response of large-scale harbor (h 20 m., Number of harmonics, N*, required in the N * s -- ) I 10 mmim | | | iI I .I.I. I I I I I1 .& &SJ J A A~.I.4..bj.A i..~sriikvAa UI I U9 a O I 'pi 00 . L Af WLt/r Fig. 5.5b. First harmonic amplitude at the back wall according to nonlinear theory (A) and regular perturbation theory (-). (See Fig. 5.5a). I 4%: 8./0 Fig. 5.5c. Second harmonic amplitude at the back wall according to nonlinear theory (+) and regular perturbation theory (- ). (See Fig. 5.5a). I.01- a it a 3sv% x x"x orLa a a ax a KW I aaII 1 1 02 I. 2 P/0 Fig. 5.5d. Third harmonic amplitude at the back wall according to nonlinear theory (x). (See Fig. 5.5a). .1 '7~ (-#) . 09 I (41 L/v37 Fig. 5.5e. Zeroth harmonic at the back wall according to nonlinear theory (A) and regular perturbation theory (-). (See Fig. 5.5a). I I I I I I I SF I I l0 I .5" .4'-- .3- I) .1I Fig. 5.6a. 'ja .I I AF .8 -F -. 2lia T Inviscid nonlinear theory (~ = 0): First resonant mode of large-scale harbor (h = 20 m., 2a = 100 m., L = 1000 m., A = .03, z = 1.41). Spatial variation of the lower harmonics. *vac 3 1131 .016-~ * .oIC .0105 -I .os. -X'S Fig. 5.6b. Spatial variation of the higher harmonics. (See Fig. 5.6a). .5. -. 5 am /. Fig. 5.6c. Surface profiles at times t = m7/4, m = 0,1,2...7. W4e S21 R R 3 t Fig. 5.7a. Inviscid nonlinear theory C? = 0): Second resonant mode of large-scale harbor (h = 20 m., 2a = 100 m., L = 1000 m., A1 = .03, z = 4.36). Spatial variation of the lower harmonics. C)010 I3 .oloS I 1 t 3dt- Fig. 5.7b. Spatial variation of the higher harmonics. (See Fig. 5.7a). evesoo oo 0. COEM 0' 0..*.gof f. o P w -e 4::m I01103600 Fig. 5.7c. Surface profiles at times t = mfr/4, m = 0,l,2...7. (See Fig. 5.7a). 4t.ob 00 C2 -7 Fig. 5.8a. Inviscid nonlinear theory (~ = 0): Third resonant mode of large-scale harbor (h = 20 m., 2a = 100 m., L = 1000 m., A1 = .03, t =1.36). Spatial variation of the lower harmonics. .ol - - 13 .ooC / a 3 7 Fig. 5.8b. Spatial variation of the higher harmonics. (See Fig. 5.8a). .II C, 00g +aow 3 4L- 7 Fig. 5.8c. Surface profiles at times t = mir/4, m = 0,1,2...7. (See Fig. 5,8a). 040 -V0 60 #40 4e .3 - - a a a I I =to+H .00 It 044 a mom;2:3 POWe wool,, *40# ~.k7.3L~g- 0 *a - S r 5 I a .*a a I A A1 First harmonic amplitude at the back wall of the large-scale harbor vs. incident wave amplitude. (l----): linear theory. ( ): inviscid nonlinear theory U 0). (x): nonlinear theory with separation loss (U = .35). 04e M" -' I , -p Fig. 5.9. x 4w spow 4opp- Ono" moomd" , s O 5.104.. Nonlinear theory with separation loss (f .35): First resonant mode of large-scale harbor (h = 20 m., 2a = 100 m., L = 1000 m., A1 = .03, 9, = 1.41). Spatial variation of the lower harmonics. ."os 13t pa(. 5l.a. 5.l0b. Spatial variation of the higher harmonics. (See Fig. 5.l0a). x3 6'1 5.1Oc. Surface profiles at times t = m7/4, m = 0,1,2...7. (See Fig. 5.10a). ILIL CY,4 Fig. 5.lla. Nonlinear theory with separation loss (~ = .35): Second resonant mode of large- scale harbor (h = 20 m., 2a = 100 m., L = 1000 m., A1 = .03, z = 4.36). Spatial variation of the lower harmonics. *a em .010 - .00 a 3 4 Spatial variation of the higher harmonics, (See Fig. 5.lla).Fig. 5.11b. 'fx)t) U, R Won %Ira dOo \40. 4MUD owe con" WOW GPM low qpftn W-M 4 moo am" *AND O*ft Goo WOM IWO ago* %ft arm* 4w wom am, 4wb 4,m 40M wow WOW 0#90 -4r-r- - - 0*0 .2 3 Fig. 5.llc. Surface profiles at times t = mfr/4, m = 0,1,2...7. (See Fig. 5.lla). 010 - .00 Fig. 5.12a. Nonlinear theory with separation loss (f = .35): Third resonant mode of large- scale harbor (h = 20 m. , 2a = 100 m., L = 1000 m., A1 = .03, = 7.36). Spatial variation of the lower harmonics. .010- 113 I a 3 9- 4 7 Spatial variation of the higher harmonics. (See Fig. 5.12a).Fig. 5.1 2b. 06 ra r 0-r*ra o2 3+7 Fig. 5.12c. Surface profiles at times t = m-f/4, m = 0,1,2...7. (See Fig. 5.12a). (a) INCIDENT WAVE ols 1Aql *05 (a) HARBOR RESPONSE (NONLINEAR, INVISCID THEORY) .003 1 .OvaI .t0.3 ,so' 3 i Fig. 5.13. Response of large-scale L = 1000 m., A2 = .050, harbor to nonmonochromatic incident wave (h = 20 m.,, 2a = 100 m., A3 = .015, z = 1.41). .o-s (b) HARBOR RESPONSE (LINEAR THEORY) I~ p .o47 I .ol7 I' t 0? s3 + -7 9 ,Oli 96 7 CHAPTER VI CONCLUDING REMARKS 6.1 Validity of the Various Approximations The proposed nonlinear theory for harbor resonance hinges on the smallness of three parameters--E, 2, and 6--corresponding to the assumptions of weak nonlinearity, weak dispersion, and a narrow harbor entrance. The assumption c n p2 << 1 is required by the depth-averaged Boussinesq equations; the smallness of 6 permits a linear treatment of the radiated wave in the ocean and a one-dimensional treatment of the nonlinear response in the long, narrow bay. Two major approximations are made in the present work, namely to decompose the solution into a finite number of harmonics and to use an impedance boundary condition at the harbor entrance. The impedance boundary condition is, however, only valid for the lowest few harmonics, hence the calculated higher harmonics may be quantitatively unreliable even if numerical convergence is achieved by including many harmonics. Nonetheless, the lowest few harmonics are expected to be substantially correct. In a large scale harbor, the depth is usually comparable to or smaller than the entrance width, so that p2= 5 THEN 1040 1030 GOTO 1010 1040 CALL (I,D,F) 1050 IF D <= 5 THEN 1070 1060 GOTO 1040 1070 WAIT (500) 1080 RETURN 3000 REM ******************** FOURIER ANALYSIS ********* 3010 PRINT I 3020 LET F[IJ1=(F(1J1+F[ TI1) /2 3030 LET H=6.28318/T 3040 FOR K= TO K! 3050 LET A=B:0 3060 FOR J=1 TO TI-1 3070 LET A=A+F[J]*COS((J-1)*K*H) 3080 LET B=B+F[JI*SIN(CJ-1)*K*H) 3090 NEXT J 3100 PRINT 2*SQR(A*A+B*B)/T/C, 3110 LET P[K]=ATN(B/A)+(1-SGN(A))*1.5708 3120 NEXT K 3130 PRINT 3140 PRINT " ", 3150 FOR K=2 TO K1 3160 LET P=P[K]-K*P[1] 3170 PRINT P-6.28318*INT(P/6.28318), 3180 NEXT K 3190 PRINT 3200 RETURN 9999 END - 183 - BIOGRAPHY Steven R. Rogers was born on February 28, 1952, in Washington, D. C. . His primary education was received at the Hebrew Academy of Washington and his secondary education at the Talmudical Academy of Baltimore and Montgomery Blair High School of Silver Spring. He was awarded a Bachelor of Science degree in physics by M. I. T. in June, 1972. His bachelor's thesis, "Ising Chain with Random Magnetic Moments", appears in Studie's in Apptied Maathemaica, Vol. 52, No. 2, 1973, pp. 163-173. A fellowship from the Fannie and John Hertz Foundation of Los Angeles, California, has enabled him to pursue graduate studies at M. I. T.. His employment experience includes work at the Technion-- Israel Institute of Technology (1974), the Weizmann Institute of Science (1972), and the X-ray Exposure Control Laboratory (1970 and 1971), where he served as a commissionned officer in the U. S. Public Health Service. The author is a member of the American Physical Society and the American Geophysical Union, and is a licensed amateur radio operator. He is also a member of Phi Beta Kappa and Sigma Xi honorary associations. - 184 -