nonlinear adjustment of a front over escarpment · 2010. 10. 8. · adjustment problem, as the...

12
Nonlinear adjustment of a front over escarpment F. Bouchut, 1 E. Scherer, 2 and V. Zeitlin 2,a 1 DMA, Ecole Normale Supérieure, 45, rue d’Ulm, 75230 Paris Cedex 05, France 2 LMD, Ecole Normale Supérieure, 24, rue Lhomond, 75231 Paris Cedex 05, France Received 25 June 2007; accepted 20 December 2007; published online 29 January 2008 We present the results of fully nonlinear numerical simulations of the geostrophic adjustment of a pressure front over topography, represented by an escarpment with a linear slope. The results of earlier simulations in the linear regime are confirmed and new essentially nonlinear effects are found. Topography influences both fast and slow components of motion. The fast unbalanced motion corresponds to inertia-gravity waves IGW. The IGW emitted during initial stages of adjustment break and form the localized dissipation zones. Due to topography, the IGW activity is enhanced in certain directions. The slow balanced motion corresponds to topographic Rossby waves propagating along the escarpment. As shown, at large enough nonlinearities they may trap fluid/ tracer and carry it on. There are indications that nonlinear topographic waves form a soliton train during the adjustment process. If the coastal line is added to the escarpment at the shallow side continental shelf, secondary fronts related to the propagation of the coastal Kelvin waves appear. © 2008 American Institute of Physics. DOI: 10.1063/1.2834731 I. INTRODUCTION The relaxation of a pressure front over topography which is represented by an escarpment in a simplest, but practically important situation, see below in a rotating fluid is a classical geophysical fluid dynamics problem. In the case of abrupt step-function escarpment it was treated analyti- cally, in the linear approximation, and by means of labora- tory experiments in the pioneering paper by Gill, Davey, Johnson, and Linden. 1 Complete linear analysis of the prob- lem was achieved by Johnson and Davey 2 for a step-function escarpment of small amplitude in the unbounded domain and in the presence of the coast. Numerical analysis of the linear relaxation problem in the case of escarpment with a linear slope was performed by Allen. 3 The standard framework for studying the relaxation of the front is the rotating shallow water RSW model with topography. In the general perspec- tive, the problem belongs to the “dam-break” class, well known in hydraulics. A thorough study of such a process in the presence of rotation, but without topography, was per- formed by Kuo and Polvani. 4 In the geophysical fluid dy- namics perspective, the relaxation of a front is a geostrophic adjustment problem, as the velocity and the pressure fields should arrive to the geostrophically balanced state by inertia- gravity wave emission. It is, thus, a natural generalization of a classical Rossby problem of the adjustment of a jet. 5 The geostrophic adjustment of jets and fronts in RSW was exten- sively studied recently both analytically 6 and numerically. 7 The use of the new generation of high-resolution finite- volume codes for shallow-water equations allowed us to make substantial progress. 4,7 In the present paper, we undertake a numerical study of the problem of fully nonlinear geostrophic adjustment of a pressure front over escarpment with a linear slope. Both the cases of an escarpment far from the coast and close to the coast will be considered. Unlike the previous theoretical and numerical studies, we are able to treat arbitrary nonlineari- ties. We use in our study the high-resolution, shock-capturing finite-volume method of Bouchut 8,9 see Ref. 7 for a brief description. The advantage of the code is that it is well balanced, i.e., it maintains the geostrophic equilibria in the presence of topography. The essence of the geostrophic adjustment process is that the flow starting from any localized initial condition is split into balanced vortex, jets and unbalanced inertia-gravity waves IGW parts. 10 The IGW carry the excess of energy away from the location of the initial perturbation while the balanced part slowly evolves, remaining close to the geo- strophic equilibrium. In the case of initially plane-parallel jets/fronts, the balanced state is just stationary. 5,6 The new phenomena expected in the process of adjustment of the pressure front over escarpment are, first, related to the fact that topography destroys the one-dimensionality of the prob- lem escarpment is perpendicular to the pressure gradient, and thus the emitted inertia-gravity waves have different speeds at the shallow and deep sides of the escarpment, and therefore the wave-fronts will be curved. Besides, the attain- ability of the fully adjusted stationary state with a constant fluid depth all along the escarpment 1 is questionable. Pre- sumably, the nonstationary intermediate balanced states should arise. Second, the presence of topographic Rossby or double-Kelvin waves introduces a completely new element in the problem, because these waves can move only in the prescribed along-escarpment direction. 11 They are balanced, and thus may be part of the just mentioned nonstationary balanced state. The wave-trapping phenomenon should modify the standard scenario of adjustment, given in Ref. 10 and 6. In this respect, the adjustment of the front over topog- raphy bears a resemblance to the equatorial adjustment, cf. Ref. 12. Both equator and topography act as waveguides for a Author to whom correspondence should be addressed. Electronic mail: [email protected]. PHYSICS OF FLUIDS 20, 016602 2008 1070-6631/2008/201/016602/12/$23.00 © 2008 American Institute of Physics 20, 016602-1 Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp

Upload: others

Post on 20-Jan-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Nonlinear adjustment of a front over escarpment · 2010. 10. 8. · adjustment problem, as the velocity and the pressure fields should arrive to the geostrophically balanced state

Nonlinear adjustment of a front over escarpmentF. Bouchut,1 E. Scherer,2 and V. Zeitlin2,a�

1DMA, Ecole Normale Supérieure, 45, rue d’Ulm, 75230 Paris Cedex 05, France2LMD, Ecole Normale Supérieure, 24, rue Lhomond, 75231 Paris Cedex 05, France

�Received 25 June 2007; accepted 20 December 2007; published online 29 January 2008�

We present the results of fully nonlinear numerical simulations of the geostrophic adjustment of apressure front over topography, represented by an escarpment with a linear slope. The results ofearlier simulations in the linear regime are confirmed and new essentially nonlinear effects arefound. Topography influences both fast and slow components of motion. The fast unbalancedmotion corresponds to inertia-gravity waves �IGW�. The IGW emitted during initial stages ofadjustment break and form the localized dissipation zones. Due to topography, the IGW activity isenhanced in certain directions. The slow balanced motion corresponds to topographic Rossby wavespropagating along the escarpment. As shown, at large enough nonlinearities they may trap fluid/tracer and carry it on. There are indications that nonlinear topographic waves form a soliton trainduring the adjustment process. If the coastal line is added to the escarpment at the shallow side�continental shelf�, secondary fronts related to the propagation of the coastal Kelvin wavesappear. © 2008 American Institute of Physics. �DOI: 10.1063/1.2834731�

I. INTRODUCTION

The relaxation of a pressure front over topography�which is represented by an escarpment in a simplest, butpractically important situation, see below� in a rotating fluidis a classical geophysical fluid dynamics problem. In the caseof abrupt �step-function� escarpment it was treated analyti-cally, in the linear approximation, and by means of labora-tory experiments in the pioneering paper by Gill, Davey,Johnson, and Linden.1 Complete linear analysis of the prob-lem was achieved by Johnson and Davey2 for a step-functionescarpment of small amplitude in the unbounded domain andin the presence of the coast. Numerical analysis of the linearrelaxation problem in the case of escarpment with a linearslope was performed by Allen.3 The standard framework forstudying the relaxation of the front is the rotating shallowwater �RSW� model with topography. In the general perspec-tive, the problem belongs to the “dam-break” class, wellknown in hydraulics. A thorough study of such a process inthe presence of rotation, but without topography, was per-formed by Kuo and Polvani.4 In the geophysical fluid dy-namics perspective, the relaxation of a front is a geostrophicadjustment problem, as the velocity and the pressure fieldsshould arrive to the geostrophically balanced state by inertia-gravity wave emission. It is, thus, a natural generalization ofa classical Rossby problem of the adjustment of a jet.5 Thegeostrophic adjustment of jets and fronts in RSW was exten-sively studied recently both analytically6 and numerically.7

The use of the new generation of high-resolution finite-volume codes for shallow-water equations allowed us tomake substantial progress.4,7

In the present paper, we undertake a numerical study ofthe problem of fully nonlinear geostrophic adjustment of apressure front over escarpment with a linear slope. Both the

cases of an escarpment far from the coast and close to thecoast will be considered. Unlike the previous theoretical andnumerical studies, we are able to treat arbitrary nonlineari-ties. We use in our study the high-resolution, shock-capturingfinite-volume method of Bouchut8,9 �see Ref. 7 for a briefdescription�. The advantage of the code is that it is wellbalanced, i.e., it maintains the geostrophic equilibria in thepresence of topography.

The essence of the geostrophic adjustment process is thatthe flow starting from any localized initial condition is splitinto balanced �vortex, jets� and unbalanced �inertia-gravitywaves �IGW�� parts.10 The IGW carry the excess of energyaway from the location of the initial perturbation while thebalanced part slowly evolves, remaining close to the geo-strophic equilibrium. In the case of initially plane-paralleljets/fronts, the balanced state is just stationary.5,6 The newphenomena expected in the process of adjustment of thepressure front over escarpment are, first, related to the factthat topography destroys the one-dimensionality of the prob-lem �escarpment is perpendicular to the pressure gradient�,and thus the emitted inertia-gravity waves have differentspeeds at the shallow and deep sides of the escarpment, andtherefore the wave-fronts will be curved. Besides, the attain-ability of the fully adjusted stationary state with a constantfluid depth all along the escarpment1 is questionable. Pre-sumably, the nonstationary intermediate balanced statesshould arise. Second, the presence of topographic Rossby �ordouble-Kelvin� waves introduces a completely new elementin the problem, because these waves can move only in theprescribed along-escarpment direction.11 They are balanced,and thus may be part of the just mentioned nonstationarybalanced state. The wave-trapping phenomenon shouldmodify the standard scenario of adjustment, given in Ref. 10and 6. In this respect, the adjustment of the front over topog-raphy bears a resemblance to the equatorial adjustment, cf.Ref. 12. Both equator and topography act as waveguides for

a�Author to whom correspondence should be addressed. Electronic mail:[email protected].

PHYSICS OF FLUIDS 20, 016602 �2008�

1070-6631/2008/20�1�/016602/12/$23.00 © 2008 American Institute of Physics20, 016602-1

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp

Page 2: Nonlinear adjustment of a front over escarpment · 2010. 10. 8. · adjustment problem, as the velocity and the pressure fields should arrive to the geostrophically balanced state

some types of waves. It was demonstrated that nonlinearequatorial Rossby waves can provide anomalous transport oftracers.13 One can ask whether a similar phenomenon maytake place in the topographic waveguide. We will show thatthis is indeed the case for nonlinear topographic waves thatform soliton-like structures with recirculation zones at strongenough nonlinearities.

We present below the results of numerical simulations ofthe nonlinear adjustment of a pressure front over topographyfor small and strong nonlinearities with and without an adja-cent coast line, with special emphasis on the mixing andtransport properties of the adjustment process. We also dis-cuss the effects of the steepness of the slope. The Appendixcontains a demonstration of the existence of the solitons oftopographically trapped waves in a certain regime of param-eters.

II. THE MODEL AND THE NUMERICAL SETUP

We work in the framework of nonlinear rotating shallowwater model on the f plane in the presence of topography.The equations of the model are

ut + uux + vuy − fv = − g�x, �1�

vt + uvx + vvy + fu = − g�y , �2�

�t + ��H + ��u�x + ��H + ��v�y = 0, �3�

where u�x ,y , t� is the zonal velocity, v�x ,y , t� is the meridi-onal velocity, H�x� is the fluid depth at rest, ��x ,y , t� is thefree-surface elevation, and f is the Coriolis parameter, whichwe assume to be constant and positive. We consider a simplebathymetry consisting of a linear slope of width 2W connect-ing two flat regions with a depth difference along the x axis,H−−H+=�H; cf. Fig. 1.

Using the maximum depth of the fluid at rest H+ as avertical scale, the Rossby deformation radius Rd=�gH+ / f2

as a typical horizontal scale, the inverse Coriolis parameterf−1 as a time scale, and introducing a typical velocity scale Uand a typical value of the perturbation of the free surface ��,we nondimensionalize the dependent and independent vari-ables as follows:

x = xRd, y = yRd, W = WRd, H = HH+,

�4��H = �HH+, � = ��� ,

u = uU, v = vU, T = Tf−1,

�5�

Hx = Hx�H

W= Hx

�H

W

H+

Rd,

where the tilde variables are the nondimensionalized ones.The parameters governing the problem are as follows:

the Rossby �or Froude� number, �= U / fRd , the relative el-evation of the free surface, �= �� / H+ , and the ratio of therelative depth change due to the escarpment to the half-width

of the slope, �= �H / W . Under the assumption ��� �whichis necessary in order to get the geostrophic equilibrium in theleading order in ��, the nondimensional equations become,omitting the tildes,

ut + ��uux + vuy� − fv = − �x, �6�

vt + ��uvx + vvy� + fu = − �y , �7�

��t + �H�ux + vy� + ��Hxu + ����u��x + �v��y� = 0. �8�

This nondimensional form of the equations is used in thenumerical simulations.

As the initial condition, we take a perturbation of thefree surface in a form of step function along a straight lineperpendicular to the escarpment. The perturbation is positivein the negative y domain and null elsewhere. This initialconfiguration is purposely chosen to be the same as the oneused by Allen,3 in order to benchmark our simulations. Forfully nonlinear numerical simulations of this problem we ap-ply the high-resolution shock-capturing finite-volume tech-nique proposed in Ref. 8. The method was first used andextensively tested in the front adjustment problem in Ref. 7.For a complete description of the procedure and discussionof its properties, see Ref. 9.

We briefly review the main ingredients of the method.The shallow water equations in the flux-form are discretizedon a regular grid within the framework of the finite-volumeapproach. The finite-volume scheme is then prescribed bythe choice of the numerical flux function and the treatment ofthe remaining source terms associated with the Coriolisforce. At each time step and in each direction, the Coriolisterms are reformulated following the apparent topographymethod first introduced in Ref. 7. The numerical flux func-tion is associated with a relaxation solver adapted to treattopography as proposed in Ref. 14. The choice of the nu-merical flux function allows us to compute solutions of theshallow water equations even in the case of vanishing depth.The numerical scheme is inviscid, the numerical dissipationbeing extremely small except for the shock locations. Thevariable time step is used in order to verify the Courant–Friedrichs–Lewy �CFL� condition. The numerical simula-tions discussed below are obtained with a typical resolution�x ,�y�0.1Rd. The computations take a few hours on a per-sonal computer at this resolution. For steep topography, theresolution was correspondingly increased. In order to avoidthe returns of IGW emitted during adjustment, the spongelayers are used at the boundaries �except for the coast, ifpresent; see below�.

FIG. 1. The side view of the escarpment. The nonperturbed free surface isindicated by the dashed line.

016602-2 Bouchut, Scherer, and Zeitlin Phys. Fluids 20, 016602 �2008�

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp

Page 3: Nonlinear adjustment of a front over escarpment · 2010. 10. 8. · adjustment problem, as the velocity and the pressure fields should arrive to the geostrophically balanced state

III. EXPECTED SCENARIO OF THE ADJUSTMENTOF THE PRESSURE FRONT AND FURTHERQUESTIONS

From the knowledge of the previous results on nonlinearadjustment of the pressure front without topography,4 of thenumerical results on linear adjustment with topography,3 andof the experimental and theoretical results with step-functionescarpments,1,2 we can expect the following processes totake place: on the shortest time scales, the IGW are to beemitted, to propagate away from the initial discontinuity, andto break at sufficient nonlinearities. Their propagation andbreaking will be modified, with respect to the flat bottomsituation, due to the fact that the phase speed of IGW varieswith varying depth of the fluid. A jet will form along theinitial discontinuity, in order to balance the pressure gradient.On longer time scales, topographically trapped waves �calledeither double Kelvin waves or topographic Rossby waves�should arise at the intersection of the initial front with theslope. The trapped waves propagate along the escarpmentwith shallower water on their right in the northernhemisphere.15 The long trapped waves will form a tonguepropagating in this way, which will distort the jet and make itfollow the contour of the nose of the tongue. The groupvelocities of short and long topographic waves are opposite�see below�. A packet of short waves will also be emitted andwill propagate in the opposite direction with respect to thetongue, but with much slower velocity and, presumably, car-rying not much energy, as in the simulations by Allen.3

Once these predictions for initial stages of adjustmentare checked, the main questions that the fully nonlinear nu-merical simulation should answer are the following:

• How is the propagation and breaking of IGW modified bytopography?

• What is the long time evolution of the trapped waves?• What are the mixing and transport properties of the adjust-

ment process?

IV. THE RESULTS OF NUMERICAL SIMULATIONSFOR THE ADJUSTMENT OF THE FRONTOVER TOPOGRAPHY FAR FROM THE COAST

A. Small nonlinearity

In order to benchmark our simulations and comparethem with those in Ref. 3, we start with a small nonlinearity��= �� / H+ =0.1�.16 We choose the width of the slope equalto five Rossby deformation radii for the simulations in thisand the following sections, and H+=2H−. We will discuss theeffects of varying the parameters of the slope in Sec. IV C. Inall of the figures below, the spatial coordinates are measuredin Rd, so that W=2.5Rd, and the time is measured in f−1.Figure 2 gives the time evolution of the free-surface eleva-tion, the initial state being at rest, u�x ,y ,0�=v�x ,y ,0�=0. Azoom of the last frame in Fig. 2 �at t=150� with superim-posed velocity field �arrows�, and the corresponding three-dimensional snapshot of the free-surface elevation, are givenin Fig. 3. Emission of the IGW takes place at the initialstages of the evolution. The IGW fronts originating from the

initial discontinuity start bending �cf. the second frame ofFig. 2�. The wave packets of IGW, apparently emitted at thelocation of maximum curvature of the primary IGW fronts,are observed �cf. the third frame of Fig. 2�. The IGW frontswill be discussed below in Sec. IV D 2. The topographicRossby waves in the form of localized height extrema alongthe slope that form a tongue appear at the later stages. Thefigure shows clearly that the topographic waves are close tothe geostrophic balance �flow parallel to the isobars�. Theyare, thus, a part of the slow balanced motion, like the “ordi-nary” Rossby waves on the beta-plane. The dispersion rela-tion for the topographic Rossby waves17 is also similar tothat of ordinary Rossby waves;11 see Fig. 4. Our simulationsare, thus, in a good qualitative agreement with the results ofRef. 3: we confirm the formation of an anticyclonic vorticityregion at the slope, the propagation of a tongue of fluid alongthe escarpment keeping the shallow water on its right andtending to separate the flows on the two sides of the slope.Apart from the jet going around the tongue, only a verysmall amount of fluid is crossing the escarpment at the loca-tion of the initial discontinuity.

In order to better visualize the behavior of the topo-graphic wave packet, we present the time evolution of thefluid depth at the center of the slope �x=0� as a function of yin Fig. 5. The amplitude of the topographic waves growsduring approximately ten of inertial periods and then satu-rates, while secondary extrema are formed. It should be no-ticed that the amplitude of the first maximum grows higherthan the initial perturbation �the maximum height is about140% of the initial height�. The topographic waves propagatewithout visible distortion, as seen in the figure.

To push the comparison further, we measured the speedof propagation along the escarpment of both the leading edgeof the tongue and of the principal height maximum. Theformer is obtained, as in Ref. 3, by following the level athalf-height of the initial perturbation, i.e., at h=1.05 in Fig.5. We thus get

t=0.0

y

−10 0 10−10

0

10

20t=4.0

−10 0 10−10

0

10

20t=18.0

−10 0 10−10

0

10

20

t=60.0

y

x−10 0 10

−10

0

10

20t=80.0

x−10 0 10

−10

0

10

20t=150.0

x−10 0 10

−10

0

10

20

FIG. 2. The time evolution of the free-surface elevation in the �x ,y� plane.The initial perturbation is equal to +�� in the lower part of the domain�y�0� and zero in the upper part �y�0�. The edges of the slope are markedwith dashed lines, �=0.1; the difference between the adjacent isolines of thefree-surface height �isobars� is 0.02. The size of the calculation domain inthe y direction is much bigger than the displayed domain.

016602-3 Nonlinear adjustment of a front over escarpment Phys. Fluids 20, 016602 �2008�

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp

Page 4: Nonlinear adjustment of a front over escarpment · 2010. 10. 8. · adjustment problem, as the velocity and the pressure fields should arrive to the geostrophically balanced state

vedge � 0.08, vpeak � 0.06 �9�

in terms of �gH+. The obtained value of vedge is substantiallylarger than the value following from Fig. 8 in Ref. 3 for thesame escarpment.

The difference between vedge and vpeak means that thetopographic wave packet is subject to dispersion. The typicalwavelength of the topographic waves formed in the adjust-ment process may be extracted from the simulations. It al-lows us to estimate the phase and group velocity of the cor-responding linear wave with the help of Fig. 4. For instance,an estimate of the wavelength as a distance between the twomaxima of the height field, cf. Fig. 5, gives k�0.92,cg�0.019, and c��0.056. However, these figures changesignificantly if twice the width of the negative curvature re-gion of the principal height peak is taken as typical wave-length: k�0.56, cg�0.048, and c��0.072. Thus, the iden-tification of the measured propagation speed with the phaseor group velocity of the linear waves cannot be done in aclearcut way. We therefore measured the dispersive effectsdirectly by following the widening of the principal peak in h

during the evolution. We thus obtain a widening of about10% in 100f−1, which is very low comparing with the ex-pected widening due to the difference between the phase andthe group velocity for both estimates given above.

The across-slope profile of the first maximum of thefluid height is close to the profile of the fundamental modegiven by the linear theory, but some differences appear whilesuperimposing them, as illustrated in Fig. 6. The linearmodes shown in Fig. 6, as well as the dispersion curvesabove, are obtained by a straightforward numerical reso-lution of the linear eigenvalue problem for the trapped modesby the shooting method. Figure 4 agrees with the dispersionobtained in Ref. 3 for the same escarpment, e.g., zero cg is atk=1.3. Note that in agreement with Ref. 11, the maximum ofthe curves is shifted toward the shallower side of the slope.

The above observations allow us to suggest that the ef-fects of nonlinearity manifest themselves already at small �and play a role in the coherence of the observed wave struc-ture. We show in the Appendix that modulated topographicwaves, at small nonlinearities, obey the Korteweg–de Vriesequation for a given across-slope mode, and thus can formsolitons. Although our initial perturbation does not project

−10 0 10 20

1

1.05

1.1

1.15t=0.0

y

−10 0 10 20

1

1.05

1.1

1.15t=4.0

−10 0 10 20

1

1.05

1.1

1.15t=18.0

−10 0 10 20

1

1.05

1.1

1.15t=60.0

y

x−10 0 10 20

1

1.05

1.1

1.15t=80.0

x−10 0 10 20

1

1.05

1.1

1.15t=150.0

x

FIG. 5. Time evolution of the fluid height profile at the center of the slope�x=0� along the y axis. Same experiment as in Fig. 2, �=0.1.

−5

0 x

5−10

0y 10

2030

1

1.05

1.1

1.15

x

y

−6 −4 −2 0 2 4 6

−2

0

2

4

6

8

10

12

14

16

FIG. 3. Zoom of the late �t=150� stage of the evolution of the free-surface elevation with superimposed velocity field �arrows� �left panel�; and the 3D viewof the height field at the same time �right panel�. Same experiment as in Fig. 2.

0 1 2 3 4 5 6−0.01

0

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

k

ω

cg

FIG. 4. Dispersion curve, phase �c� and group �cg� velocities of the topo-graphic Rossby waves, obtained by a straightforward numerical resolutionof the linear eigenproblem for the topographic Rossby waves by the shoot-ing method �cf. Ref. 11�. The along-slope wavenumber is nondimensional-ized by Rd, and the frequency is nondimensionalized by the inertial fre-quency f .

016602-4 Bouchut, Scherer, and Zeitlin Phys. Fluids 20, 016602 �2008�

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp

Page 5: Nonlinear adjustment of a front over escarpment · 2010. 10. 8. · adjustment problem, as the velocity and the pressure fields should arrive to the geostrophically balanced state

onto a single mode, the observed behavior of the subsequentpressure maxima is qualitatively consistent with that of asoliton train.

B. Strong nonlinearity

We now focus on the same process but at a larger initialnonlinearity �=0.3. Figure 7 shows the nonlinear evolutionof the free-surface elevation and Fig. 8 gives a zoom of thelast frame at t=150 with superimposed velocity field �ar-rows�. The scenario of the adjustment is qualitatively thesame as at �=0.1. The topographic waves propagate withoutdistortion, and the amplitude of the first maximum againgrows higher than the initial height of the perturbation�maximum height is about 130% the initial height; seeFig. 9�. The measured speeds of the edge of the tongue andthat of the principal height maximum are close to those atsmaller nonlinearity,

vedge � 0.08, vpeak � 0.05. �10�

The left panel of Fig. 8 shows that the flow is again close tothe geostrophic balance but it is less equilibrated on the shal-low side.

There are, however, some differences between the twocases. The emission of inertia-gravity waves takes place dur-ing a longer period of time �about 30f−1�. The secondaryextrema of the height field are formed later: at t=200 thereare three maxima and two minima in the profile of h at smallnonlinearity, and only two maxima and two minima at strongnonlinearity �not shown�.

The estimates of the typical wavenumber, group, andphase velocities based on the same criteria as in thepreceding section give, respectively, k�0.98, cg�0.015,c��0.054 and k�0.73, cg�0.033, c��0.064. The widen-ing of the principal pressure peak during 100f−1 is again �.1,and is much slower than that given by the linear dispersion.The comparison of the profiles of linear modes with the mea-

−10 −8 −6 −4 −2 0 2 4 6 8 100

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

x

FIG. 6. The nondimensional profiles of the fundamental modes of the lineartheory corresponding to k=0.92 �dashed line� and to k=0.56 �dash-dottedline� vs the section of the first pressure maximum in the numerical simula-tion of Fig. 2 �solid line� at time t=150 as functions of nondimensional x.

t=0.0

y

−10 0 10−10

0

10

20t=4.0

−10 0 10−10

0

10

20t=18.0

−10 0 10−10

0

10

20

t=60.0

y

x−10 0 10

−10

0

10

20t=80.0

x−10 0 10

−10

0

10

20t=150.0

x−10 0 10

−10

0

10

20

FIG. 7. Same as in Fig. 2, but at �=0.3. The difference between the adjacentisolines of the free-surface height �isobars� is 0.05.

x

y

−6 −4 −2 0 2 4 6

0

5

10

15

FIG. 8. Same as in the left panel of Fig. 3 but at �=0.3. The isobars fromh=1.02 up at step 0.05 are displayed.

−10 0 10 201

1.1

1.2

1.3

1.4

1.5t=0.0

y

−10 0 10 201

1.1

1.2

1.3

1.4

1.5t=4.0

−10 0 10 201

1.1

1.2

1.3

1.4

1.5t=18.0

−10 0 10 201

1.1

1.2

1.3

1.4

1.5t=60.0

y

x−10 0 10 201

1.1

1.2

1.3

1.4

1.5t=80.0

x−10 0 10 201

1.1

1.2

1.3

1.4

1.5t=150.0

x

FIG. 9. Same as in Fig. 5 but at �=0.3.

016602-5 Nonlinear adjustment of a front over escarpment Phys. Fluids 20, 016602 �2008�

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp

Page 6: Nonlinear adjustment of a front over escarpment · 2010. 10. 8. · adjustment problem, as the velocity and the pressure fields should arrive to the geostrophically balanced state

sured profile of the main pressure peak presented in Fig. 10shows some differences, e.g., the widening of the nonlinearprofile.

Globally, the simulations at �=0.3 confirm the coherent-structure character of the topographic wave pattern formedduring adjustment. This character is strengthened even moreby its stirring properties displayed in Sec. IV D.

C. Steep topography

In this section, we briefly describe how the previous re-sults are modified for the steep topography. One of the mo-tivations is the comparison with the theoretical and experi-mental results obtained in Refs. 1 and 2 for an “infinitelysteep,” i.e., step-function profile of the escarpment. The nu-merical code we use, by construction �cf. the explicit expres-sions for numerical fluxes in Ref. 7�, cannot treat the step-function topography. Hence, the only way to model theabrupt topography is to take a steep escarpment. This means,in turn, that spatial resolution across escarpment should beincreased. We thus consider a slope of a half-width,W=0.25Rd, i.e., ten times steeper than in the previous sub-sections, with the same ratio of depths H−=0.5H+. We startfrom a weakly nonlinear pressure front with �=0.1. The evo-lution of the free surface is presented in Fig. 11. We stoppedthe simulation at t=40 as it is getting increasingly costly togo further in time because of the increased resolution. Theoverall scenario of adjustment is qualitatively the same asin the case of gentle topography, with the following maindifferences:

�1� The velocity of the topographic waves is much slower.�2� The tongue is less confined to the escarpment, i.e., it

spreads outside the slope.�3� The peak is shifted toward the top of the slope.

A zoom of the last frame in Fig. 11 with superimposed ve-locity vectors is shown in Fig. 12, with the three-dimensionalview of this same stage. We see that the peak height exceeds

the initial maximum height and that the flow is approxi-mately in geostrophic balance at the deep side, but at theshallow side we notice significant departures from the geo-strophic balance, probably due to relatively short integrationtime. Note also the presence of the small-amplitude IGW inFig. 12, which are deforming the isobars. The speed of thepropagation of the tongue is much faster, which is consistentwith the increase of the phase velocity of the trapped waveswith increasing slope; cf. Ref. 11. Nevertheless, the velocityof the main pressure peak is of the same order of magnitudeas in the case of gentle topography,

vedge � 0.2, vpeak � 0.08. �11�

Qualitatively, these results are consistent with the linearanalysis of Ref. 2 with a notable exception of the observeddissymmetry with respect to the axis of the slope.

D. Transport, stirring, and mixingduring the adjustment of the pressure front

1. Anomalous transport due to nonlineartopographic waves

In order to study the transport and stirring properties ofthe adjustment process, we add a new variable, the density ofa passive tracer, to the previous runs, and follow itsevolution.18 We use the same method, well adapted for thefinite-volume code, and the same mode of visualization ofthe tracer as in Ref. 13, i.e., we follow the evolution of theisopycnals of the tracer distribution initially linear in one ofthe coordinates. We show in the figures below the simulta-neous evolution of the pair of tracers, distributed initiallylinearly in x and in y, respectively. Thus, each intersection ofthe isopycnals represents a fluid particle.

As was already mentioned, we are motivated by similarstudies for the equatorial adjustment,12,13 where it was shownthat the waveguide modes at sufficient nonlinearities mayform coherent structures �modons of equatorial Rossby

−10 −8 −6 −4 −2 0 2 4 6 8 100

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

x

FIG. 10. The nondimensional profiles of the fundamental modes of thelinear theory corresponding to k=0.98 �dashed line� and to k=0.73 �dash-dotted line� vs the section of the first pressure maximum in the numericalsimulation of Fig. 7 �solid line� at time t=150, as functions of nondimen-sional x.

t=0.0

y

−4 −2 0 2 4−10

0

10

20t=1.0

−4 −2 0 2 4−10

0

10

20t=4.0

−4 −2 0 2 4−10

0

10

20

t=9.0

y

x−4 −2 0 2 4

−10

0

10

20t=18.0

x−4 −2 0 2 4

−10

0

10

20t=40.0

x−4 −2 0 2 4

−10

0

10

20

FIG. 11. The time evolution of the free-surface elevation in the �x ,y� plane.The initial perturbation is equal to +�� in the lower part of the domain�y�0� and zero in the upper part �y�0�. The edges of the slope are markedwith dashed lines �W=0.25�, �=0.1, and the height difference between theadjacent isolines of the free-surface height is 0.02. The length of the calcu-lation domain in the y direction is much bigger than the displayed domain.

016602-6 Bouchut, Scherer, and Zeitlin Phys. Fluids 20, 016602 �2008�

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp

Page 7: Nonlinear adjustment of a front over escarpment · 2010. 10. 8. · adjustment problem, as the velocity and the pressure fields should arrive to the geostrophically balanced state

waves in that case� trapping fluid and carrying it along thewaveguide. We show here that although the spatial structureof the waveguide modes in the present case is not the same�they are monopolar, and not dipolar, as at the equator�, thesame phenomenon takes place.

Figure 13 gives the spatio-temporal evolution of thetracer field during the adjustment of the front with �=0.1over the gentle topography W=2.5Rd. In the last frame�t=150�, we observe a formation of a recirculation pattern,with subsequent trapping of the fluid. This is a “dispersivebreaking” phenomenon, as it was called in Ref. 13. At largernonlinearity, trapping appears earlier, as we can see in Fig.14, where the nonlinearity is �=0.3. The topographic wavesare again trapping fluid and carrying it along the escarpment.

2. Wave-breaking as an indicator of mixing

As was already mentioned in the Introduction, our nu-merical model is essentially nondissipative except for the

shock locations.7 The total energy is conserved during thesimulations, apart from the events of wave-breaking. Weshow in Fig. 15 the time evolution of the total energy loss inthe flow in the frontal adjustment process. The energy loss ispronounced during the initial stages of evolution �severalinertial periods� where the strong gradients �due to initialconditions, and to the secondary shock formation; see below�are present. The secondary peak of the energy loss at t�6 iscorrelated with the appearance of the strongly curved regionat the emitted wave-front; see Fig. 16. The energy loss isvery low during the rest of the simulation, which amounts tothe overall loss of energy of the order of one per thousand.

Due to the difference in the phase speeds of waves onthe left and on the right of the escarpment �IGW propagatingfaster in deeper fluid�, the wave-fronts of the IGW emitted atthe initial stage of adjustment should bend. This is a novelaspect of the dam-break problem in the presence of topogra-phy, which should affect the wave-breaking accompanyingthe propagation of such wave-fronts.4,7 By monitoring thelocal dissipation rates, we can identify the wave-breaking

−4

−2

0

2

4

−100

1020

30

1

1.05

1.1

1.15

x

y

x−2 −1.5 −1 −0.5 0 0.5 1 1.5 2 2.5

y

−2

0

2

4

6

8

10

12

14

FIG. 12. The late �t=40� stage of the evolution of the free-surface elevation with superimposed velocity field �arrows� �left panel�; and the corresponding 3Dview of the height field �right panel�. Same experiment as in Fig. 11. Note that the x scale here is different with respect to Figs. 2 and 7.

−10 0 10

−10

0

10

20

30

y

t=0.0

−10 0 10

−10

0

10

20

30t=4.0

−10 0 10

−10

0

10

20

30t=18.0

−10 0 10

−10

0

10

20

30

y

x

t=60.0

−10 0 10

−10

0

10

20

30

x

t=80.0

−10 0 10

−10

0

10

20

30

x

t=150.0

FIG. 13. Time evolution of the tracer field at �=0.1 and W=2.5Rd. Theisopycnals of a pair of passive tracers with initial distributions linear in xand y, respectively, are shown in subsequent times. Tracers are advected bythe velocity field corresponding to the simulation of Fig. 2.

−10 0 10

−10

0

10

20

30

y

t=0.0

−10 0 10

−10

0

10

20

30t=4.0

−10 0 10

−10

0

10

20

30t=18.0

−10 0 10

−10

0

10

20

30

y

x

t=60.0

−10 0 10

−10

0

10

20

30

x

t=80.0

−10 0 10

−10

0

10

20

30

x

t=150.0

FIG. 14. Time evolution of the tracer field at nonlinearity �=0.3 andW=2.5Rd. See the caption of Fig. 13.

016602-7 Nonlinear adjustment of a front over escarpment Phys. Fluids 20, 016602 �2008�

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp

Page 8: Nonlinear adjustment of a front over escarpment · 2010. 10. 8. · adjustment problem, as the velocity and the pressure fields should arrive to the geostrophically balanced state

zones. The evolution of the local dissipation rate in the cal-culation domain at the initial stages of the simulation with�=0.3 is shown in Fig. 16. The dissipation rate is calculatedas the deviation from the energy conservation in each com-putation cell. The zones where the dissipation is prominentare located first at the two wave-fronts propagating from theinitial discontinuity in both along-slope directions. After sev-eral of the inertial periods, the maximum of dissipation islocalized at the specific zones of the IGW wave-fronts propa-gating at the shallow side of the escarpment. As seen fromthe figures, the IGW wave-front intensity is decreased, ex-cept for these precise locations that continue to propagate aslocalized wave-packets at an angle that is the same in simu-lations with different front intensities.

Obviously the highly idealized model cannot pretend tocorrectly represent the dissipative phenomena in the realflow �recall that the typical resolution is 0.1 Rd in the simu-

lations above�. However, it gives a good first guess aboutlocations of enhanced dissipation and hence mixing in thereal flow in similar configurations.

V. THE ROLE OF THE COAST

In the previous simulations, the escarpment was situatedin the “open ocean.” Let us briefly comment on the neweffects introduced by an adjacent coast line. The effects ofthe coast perpendicular to the escarpment were discussed byGill, Davey, Johnson, and Linden1 and Allen.3 We repeatedour simulations with W=2.5Rd by placing a rigid wall paral-lel to the escarpment at x=10 at the shallow side in order tomodel a “continental shelf” configuration. The results for theevolution of the height field, the dissipation, and the profileof the height field along the x=9 section are given in Figs.17–19 , respectively.

The main new element introduced by the coast is thepropagation of the boundary Kelvin wave front along it. Dueto the fact that the boundary Kelvin waves are dispersionless,they break �cf. Ref. 19� and form the mixing zones. The

0 20 40 60 80 100 120 140 160 180 200−0.08

−0.07

−0.06

−0.05

−0.04

−0.03

−0.02

−0.01

0

t

Rel

ativ

een

ergy

loss

FIG. 15. Time evolution of the energy loss due to numerical dissipation ofthe flow at �=0.3. The relative energy loss per unit time �E / E0�t is cal-culated as deviation from the energy balance in each calculation cell, inte-grated over the whole calculation domain and normalized by the total initialenergy.

t=0.0

y

−5 0 5

−10

0

10

20

30t=2.0

−5 0 5

−10

0

10

20

30t=6.0

−5 0 5

−10

0

10

20

30

t=8.0

y

x−5 0 5

−10

0

10

20

30t=14.0

x−5 0 5

−10

0

10

20

30t=16.0

x−5 0 5

−10

0

10

20

30

FIG. 16. Time evolution of the dissipation rate �black contours correspond-ing to the threshold of 2�10−5 in nondimensionalized energy flux unit�superimposed on the surface elevation �gray contours at �h=0.05 interval�in the �x ,y� plane, at �=0.3.

t=0.0

y

−5 0 5

−10

0

10

20

30t=4.0

−5 0 5

−10

0

10

20

30t=18.0

−5 0 5

−10

0

10

20

30

t=60.0

x

y

−5 0 5

−10

0

10

20

30t=80.0

x−5 0 5

−10

0

10

20

30t=150.0

x−5 0 5

−10

0

10

20

30

FIG. 17. Time evolution of the free-surface elevation in the �x ,y� plane, at�=0.3. The solid wall is on the right side.

t=0.0

y

−5 0 5

−10

0

10

20

30t=2.0

−5 0 5

−10

0

10

20

30t=6.0

−5 0 5

−10

0

10

20

30

t=8.0

y

x−5 0 5

−10

0

10

20

30t=12.0

x−5 0 5

−10

0

10

20

30t=18.0

x−5 0 5

−10

0

10

20

30

FIG. 18. Joint evolution of the dissipation rate �black contours� and thefree-surface elevation �gray contours� in the �x ,y� plane, at �=0.3.

016602-8 Bouchut, Scherer, and Zeitlin Phys. Fluids 20, 016602 �2008�

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp

Page 9: Nonlinear adjustment of a front over escarpment · 2010. 10. 8. · adjustment problem, as the velocity and the pressure fields should arrive to the geostrophically balanced state

details of the Kelvin front formation are given in Fig. 19.Another phenomenon related to the Kelvin front propagationis a secondary wave emission behind the front well-distinguishable in the pressure contours in Fig. 18. It is simi-lar to that observed for equatorial Kelvin fronts20,12 and hasthe same explanation. Namely, the nonlinear Kelvin waveshaving phase speed higher than the linear one may be in adirect resonance with the IGW.

VI. CONCLUSIONS AND DISCUSSION

We have performed detailed fully nonlinear numericalsimulations of the classical problem of the adjustment of apressure front over topography. We confirmed the results ofearlier investigations that were performed previously in thelinear regime and found new essentially nonlinear phenom-ena. The most important of them, in what concerns dynam-ics, is excitation of finite-amplitude topographic wavespropagating essentially without change of form for longtimes, and resembling the soliton trains known from thetheory of the KdV equation. Nevertheless, the quantitativecheck of the soliton hypothesis is yet to be done. In whatconcerns transport and mixing, the most characteristic fea-tures are �i� formation of recirculation regions in the cores oftopographic “solitons” with subsequent anomalous tracertransport along the topography, and �ii� appearance of spe-cific mixing zones related to the propagation of IGW, or ofthe boundary Kelvin waves in the presence of the coast,emitted during the early stages of adjustment.

Therefore, the geostrophic adjustment in the presence ofthe topographic waveguide is specific and resembles that atthe equator, as part of the initial perturbation projects ontothe waveguide modes. These waveguide modes form coher-ent structures providing the anomalous transport along thewaveguide. The waveguide modes belong to the balancedpart of the motion, as we confirm by direct calculations.They evolve much slower than unbalanced IGW, whichevacuate the excess of energy at the early stages of the ad-

justment process. The IGW are, however, also influenced bytopography, which leads to the enhancement of dissipation/mixing in specific locations.

It should be mentioned that fully nonlinear numericalsimulations of the interaction of jets with topography usingthe rigid-lid configuration are known in the literature.21 Therigid-lid approximation corresponds formally to Rd→�,which physically means that all typical scales, including thescale of the slope, are much smaller than the deformationradius. The steep topography case in Sec. IV C above is theonly one approaching this regime in the present study �suchscales are largely subgrid in the gentle slope case�. However,our simulations reveal the formation of the coherent struc-tures of the scale still much larger than the width of theslope, which was not the case in the corresponding cases inRef. 21. The dipolar structures corresponding to the highertopographic modes were typically observed in Ref. 21, whilein our simulations the monopolar ones are dominating, al-though higher modes can appear on the late stages as sug-gested by the gentle-slope simulations �cf., e.g., Fig. 8�. Itprobably means that even with steep topographies, the rigid-lid regime is hard to reach in the full rotating shallow watermodel. This is also true for the adjustment with the reversedjet �initial pressure anomaly of opposite sign�, which wasalso studied in Ref. 21. We present in Fig. 20 the late stage ofthe evolution for strong nonlinearity �=0.3 in this case, to becompared with Fig. 8. The topology of the isobars is globallythe same, although it corresponds to a trough in Fig. 20 andto an elevation in Fig. 8, with respect to the nonperturbedfluid surface at rest. Some asymmetry, however, is observedin what concerns the intensity of the resulting balanced vor-tex �more intense in Fig. 20� and its position �at the shal-lower side of the slope for the cyclonic vortex, Fig. 20, andat the deeper side for the anticyclonic vortex, Fig. 8�. Beingmore intense, the vortex/pressure extremum propagatesfaster in Fig. 20, although the tip of the tongue has practi-cally the same velocity in both cases.

−10 0 10 20 30

1

1.1

1.2

1.3

1.4

1.5t=0.0

y−10 0 10 20 30

1

1.1

1.2

1.3

1.4

1.5t=2.0

y−10 0 10 20 30

1

1.1

1.2

1.3

1.4

1.5t=6.0

y

−10 0 10 20 30

1

1.1

1.2

1.3

1.4

1.5t=8.0

y−10 0 10 20 30

1

1.1

1.2

1.3

1.4

1.5t=12.0

y−10 0 10 20 30

1

1.1

1.2

1.3

1.4

1.5t=18.0

y

FIG. 19. Time evolution of the free-surface elevation at x=9 �close to thewall�, at �=0.3.

y

x−6 −4 −2 0 2 4 6

0

5

10

15

FIG. 20. Same as in Fig. 8 but for the reversed initial anomaly. Isobars fromh=0.98 down at step 0.05 are displayed.

016602-9 Nonlinear adjustment of a front over escarpment Phys. Fluids 20, 016602 �2008�

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp

Page 10: Nonlinear adjustment of a front over escarpment · 2010. 10. 8. · adjustment problem, as the velocity and the pressure fields should arrive to the geostrophically balanced state

APPENDIX: KdV SOLITONS OF MODULATEDTOPOGRAPHIC WAVES

Consider the shallow-water equations with topography�1�–�3� on the f plane.

We use a slightly different nondimensionalization withrespect to Eqs. �4� and �5� for the horizontal scale, the verti-cal scale, and the velocity, because we want to allow fordifferent scales in the two horizontal directions. Using thehalf-width of the slope W as a typical horizontal scale in thex direction, L, a typical horizontal scale in the y directionwith the corresponding aspect ratio parameter =W /Lsupposed to be small, the mean depth of the fluid at rest

H= H++H− / 2 as a vertical scale, and U and U for the zonaland meridional velocity, respectively, we nondimensionalizethe dependent and independent variables as follows:

x = xW, y = yL, H = HH, �H = �HH, � = ��� ,

�A1�

u = u U, v = vU,

�A2�

T = Tf−1, Hx = Hx�H

W= Hx�H

H

W,

where the tilde variables are the nondimensionalized ones.The parameters governing the problem, thus, are as follows:the Rossby number �= U / fW , the relative elevation of the

free surface �= �� / H , the Burger number �= �Rd / W �2,where the Rossby deformation radius is defined as

Rd=�gH / f2 , and the relative depth change due to the es-

carpment, �=�H.We get the following nondimensionalized set of equa-

tions:

ut + � 2�uux + vuy� − v = −��

��x, �A3�

vt + � �uvx + vvy� + u = − ��

��y , �A4�

��t + � H�ux + vy� + � �Hxu + �� ��u��x + �v��y� = 0.

�A5�

We are interested in slow motions close to the geostrophicequilibrium. The latter is achieved at the leading order by thefollowing choice of parameters: �� / � �1, which is standardin the quasigeostrophic dynamics. The filtering of the fastinertia-gravity waves is achieved by using the hierarchy ofthe slow time scales, tn= nt, and supposing that all the fieldsare slow, i.e., do not depend on the fast time scale t0. Finally,the nonlinearity, as well as the Rossby number, are supposedto be small and we choose the following relation among thesmall parameters: ���� 2, ��1, meaning that the widthof the escarpment is of the order of the Rossby radius. Wethus get, omitting the numerical factors of order 1,

2�ut1+ ut2

+ 2ut3+ ¯ � + 4�uux + vuy� − v = − �x,

�A6�

�vt1+ vt2

+ 2vt3+ ¯ � + 2�uvx + vvy� + u = − �y ,

�A7�

��t1+ �t2

+ 2�t3+ ¯ � + Hux + Hvy + �Hxu

+ 2��u��x + �v��y� = 0. �A8�

The last hypothesis consists in supposing the order 1 depthchange across the escarpment, ��1. All fields are then ex-panded in asymptotic series in ,

u = u�0� + u�1� + 2u�2� + ¯ ,

v = v�0� + v�1� + 2v�2� + ¯ ,

� = ��0� + ��1� + 2��2� + ¯ .

At the lowest order we get

v�0� = �x�0�, �A9�

vt1�0� + u�0� = − �y

�0�, �A10�

�t1�0� + Hux

�0� + Hvy�0� + Hxu

�0� = 0, �A11�

which is equivalent to a single equation for ��0�,

�t1�H�xx

�0� + Hx�x�0� − ��0�� + Hx�y

�0� = 0. �A12�

Looking for a propagating in y solution with a given struc-ture in x,

��0� = F�0��x��0��ky − t1,t2,t3, . . . � �A13�

we get the following equation for the across-escarpmentstructure functions,

�HFx�0��x − F�0��1 +

k

Hx = 0. �A14�

The solutions of this equation in the case of linear topogra-phy are expressed in terms of the modified Bessel function ofthe first kind and of the second kind.

Due to the assumed scaling, this equation is the long-wave low-frequency limit of the equation for the linearmodes obtained by the straightforward linearization aboutthe rest state of Eqs. �1�–�3�, and used earlier to get Fig. 4. Itis easy to see that in order to get nontrivial eigenmodesFn

�0��x� with eigenfrequencies n, it is necessary that�1+ k / Hx� changes sign. Thus, for monotonous escarpment-like topography, the well-known unique direction of propa-gation of the topographic waves which keep shallower wateron their right in the Northern hemisphere15 follows.

Proceeding to the higher orders in will give, as usual inthe asymptotic expansions method, a series of inhomoge-neous problems for higher corrections ��i� , i=1,2 , .. with theright-hand side �r.h.s.� depending on the previous-orderfields. Again, according to the standard rules, the solutions ofthese problems may be obtained by removing the resonanceson the r.h.s., i.e., imposing the orthogonality condition and

016602-10 Bouchut, Scherer, and Zeitlin Phys. Fluids 20, 016602 �2008�

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp

Page 11: Nonlinear adjustment of a front over escarpment · 2010. 10. 8. · adjustment problem, as the velocity and the pressure fields should arrive to the geostrophically balanced state

introducing dependence on slower times. This gives the non-linear evolution equation for ��0�. Proceeding in this way, itis easy to see that it is sufficient to require that there is no t2

dependence in the zero-order fields to get rid of the reso-nances at the first order in . At the order 2, we have

ut1�0� − v�2� = − �x

�2�, �A15�

vt1�2� + vt3

�0� + u�0�vx�0� + v�0�vy

�0� + u�2� = − �y�2�, �A16�

�t1�2� + �t3

�0� + H�ux�2� + vy

�2�� + u�2�Hx + �u�0���0��x

+ �v�0���0��y = 0. �A17�

By expressing the velocity field in terms of the free-surfaceelevation,

v�2� = �x�2� − �yt1

�0� − �xt1t1�0� , �A18�

u�2� = − �y�2� − �xt1

�2� + �yt1t1�0� + �xt1t1t1

�0� − �xt3�0� − �x

�0��xy�0�

+ ��y�0� + �xt1

�0���xx�0�, �A19�

we get

− H�xxt1�2� − Hx��xt1

�2� + �y�2�� + �t1

�2�

= − �t3�0� − H��xyt1t1

�0� + �xxt1t1t1�0� − �xxt3

�0� − �x�0��xxy

�0�

+ �xxt1�0� �xx

�0� + �y�0��xxx

�0� + �xt1�0��xxx

�0� − �yyt1�0� − �xyt1t1

�0� �

− Hx− �xt3�0� + �yt1t1

�0� + �xt1t1t1�0� − �x

�0��xy�0�

+ ��y�0� + �xt1

�0���xx�0�� + ��0��xxt1

�0� + �x�0��xt1

�0�. �A20�

We look for a solution of this equation in the form��2�=F�2��x��2��ky−t1 , t3 , . . . �. The t1 derivatives shouldbe replaced in Eq. �A20� by the y ones according to the rule�y =−k / �t1

consistent with the unidirectional propagationof the topographic waves. We thus get the following equationfor F�2��x�:

t1�2��F�2� + Hx� k

F�2� − Fx

�2� − HFxx�2� = r.h.s� , �A21�

where r.h.s.� stands for the r.h.s. of Eq. �A20�. By multiply-ing by F�0��x� and integrating by x, we get the Korteweg–deVries equation for �0�,

at3�0� + byyy

�0� + c�0�y�0� = 0, �A22�

where the coefficients are given by

a = �−�

+�

F�0��− F�0� + HxFx�0� + HFxx

�0��dx , �A23�

b = �−�

+�

F�0��− Hx�2

k2 F�0� −3

k3 Fx�0�

− H�−3

k3 Fxx�0� +

kF�0� dx , �A24�

c = �−�

+�

F�0��− Hx�− Fx�0�Fx

�0� + F�0�Fxx�0� −

kFx

�0�Fxx�0�

− H��Fx�0�Fxx

�0� + F�0�Fxxx�0� −

k�Fxx

�0�Fxx�0� + Fx

�0�Fxxx�0� �

k�Fx

�0�Fx�0� + F�0�Fxx

�0�� dx . �A25�

The numerical values of a, b, and c may be easily found forany linear escarpment.

The appearance of the KdV equation for slow evolutionof weakly nonlinear long topographic waves could be ex-pected. Indeed, the dispersion of long unidirectionally propa-gating topographic waves is weak �cf. the dispersion curveFig. 4� and may be compensated by weak nonlinearity togive solutions preserving their form, a typical situation inwhich the KdV equation arises.

Surprisingly, soliton-like structures in topographic wavesarise at strong nonlinearities, as our numerical simulationshows. In this respect, again, the situation is similar to thatobserved in simulations of the equatorial adjustment.12

1A. E. Gill, M. K. Davey, E. R. Johnson, and P. F. Linden, “Rossby adjust-ment over a step,” J. Mar. Res. 44, 713 �1986�.

2E. R. Johnson and M. K. Davey, “Free-surface adjustment and topo-graphic waves in coastal currents,” J. Fluid Mech. 219, 273 �1990�.

3S. E. Allen, “Rossby adjustment over a slope in a homogeneous fluid,” J.Phys. Oceanogr. 26, 1646 �1996�.

4A. C. Kuo and L. M. Polvani, “Time-dependent fully nonlinear geo-strophic adjustment,” J. Phys. Oceanogr. 27, 1614 �1997�.

5C.-G. Rossby, “On the mutual adjustment of pressure and velocity distri-butions in certain simple current systems, ii,” J. Mar. Res. 1, 239 �1937�.

6V. Zeitlin, S. B. Medvedev, and R. Plougonven, “Frontal geostrophic ad-justment, slow manifold and nonlinear wave phenomena in one-dimensional rotating shallow water. Part 1: Theory,” J. Fluid Mech. 481,269 �2003�.

7F. Bouchut, J. LeSommer, and V. Zeitlin, “Frontal geostrophic adjustmentand non-linear wave phenomena in one-dimensional rotating shallow wa-ter. Part 2: High resolution numerical simulations,” J. Fluid Mech. 514, 35�2004�.

8F. Bouchut, Nonlinear Stability of Finite Volume Methods for HyperbolicConservation Laws, and Well-balanced Schemes for Sources, Frontiers inMathematics Series �Birkhaüser, Basel, 2004�.

9F. Bouchut, “Efficient numerical finite-volume schemes for shallow watermodels” in Nonlinear Dynamics of Rotating Shallow Water: Methods andAdvances, edited by V. Zeitlin, Edited Series on Advances in NonlinearScience and Complexity �Elsevier, Amsterdam, 2007�, pp. 191–258.

10G. M. Reznik, V. Zeitlin, and M. Ben Jelloul, “Nonlinear theory of geo-strophic adjustment. Part 1: Rotating shallow water model,” J. Fluid Mech.445, 93 �2001�.

11M. S. Longuet-Higgins, “Double Kelvin waves with continuous depth pro-file,” J. Fluid Mech. 34, 49 �1968�.

12J. LeSommer, G. M. Reznik, and V. Zeitlin, “Nonlinear geostrophic ad-justment of long-wave disturbances in the shallow water model on theequatorial beta-plane,” J. Fluid Mech. 515, 135 �2004�.

13F. Bouchut, J. Le Sommer, and V. Zeitlin, “Breaking of balanced andunbalanced equatorial waves,” Chaos 15, 013503 �2005�.

14E. Audusse, F. Bouchut, M.-O. Bristeau, R. Klein, and B. Perthame, “Afast and stable well-balanced scheme with hydrostatic reconstruction forshallow water flows,” SIAM J. Sci. Comput. �USA� 25, 2050 �2004�.

15P. H. LeBlond and L. A. Mysak, Waves in the Ocean, Elsevier Oceanog-raphy Series �Elsevier, Amsterdam, 1978�.

16Note, however, that the exact value of the amplitude of the perturbation isnot given in Ref. 3.

17Due to their above-mentioned balanced character, we find this name moreappropriate than the double Kelvin waves.

016602-11 Nonlinear adjustment of a front over escarpment Phys. Fluids 20, 016602 �2008�

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp

Page 12: Nonlinear adjustment of a front over escarpment · 2010. 10. 8. · adjustment problem, as the velocity and the pressure fields should arrive to the geostrophically balanced state

18The diagnostic of transport with the help of potential vorticity frequentlyused in literature is not well suited for the present study because initialpotential vorticity takes four different �constant� values out of the slopeplus a continuous spectrum of values in the region of the slope.

19A. V. Fedorov and W. K. Melville, “Propagation and breaking of nonlinearKelvin waves,” J. Phys. Oceanogr. 25, 2518 �1995�.

20A. V. Fedorov and W. K. Melville, “Kelvin fronts on the equatorial ther-mocline,” J. Phys. Oceanogr. 30, 1692 �2000�.

21G. F. Carnevale, S. G. Llewelyn-Smith, F. Crisciani, R. Purini, and R.Serraval, “Bifurcation of a coastal current at an excarpment,” J. Phys.Oceanogr. 29, 969 �1999�.

016602-12 Bouchut, Scherer, and Zeitlin Phys. Fluids 20, 016602 �2008�

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp