Relaxing the sharp density stratification and columnar motion assumptions in layered shallow water systems
Abstract
We rigorously justify the bilayer shallow-water system as an approximation to the hydrostatic Euler equations in situations where the flow is density-stratified with close-to-piecewise constant density profiles, and close-to-columnar velocity profiles. Our theory accommodates with continuous stratification, so that admissible deviations from bilayer profiles are not pointwise small. This leads us to define refined approximate solutions that are able to describe at first order the flow in the pycnocline. Because the hydrostatic Euler equations are not known to enjoy suitable stability estimates, we rely on thickness-diffusivity contributions proposed by Gent and McWilliams. Our strategy also applies to one-layer and multilayer frameworks.
1 Introduction
Motivation
The bilayer shallow water system is a standard model for the description of internal waves in density-stratified flows in situations where the density distribution is such that the fluid can be approximately described as two layers with almost-constant densities separated by a thin a pycnocline; see e.g. [27, Chap. 6]. In addition to this sharp stratification assumption, the formal derivation of the bilayer shallow water system relies on two additional ingredients. Firstly, the internal pressure is assumed to be hydrostatic, that is the pressure-gradient force balances the external force due to gravity. Secondly, the flow velocity is assumed to be columnar, that is the horizontal velocity of fluid particles is constant with respect to the vertical variable within each layer. Of course the validity of bilayer shallow water system relies on the expectation that, if originally approximately satisfied, these three assumptions remain accurate on a relevant timescale as the flow evolves.
The rigorous justification of the hydrostatic assumption in the shallow water regime —that is when the typical horizontal wavelength of the flow is large with respect to the vertical depth of the layer— has been rigorously analyzed either in situations of homogeneous density [5, 36, 23, 37], with smooth density distributions [42, 41], or in the bilayer framework [11, 21, 20]. In the bilayer framework and assuming that the pressure is hydrostatic, the columnar assumption is propagated exactly by the flow. By this we mean that the bilayer shallow water system produces exact solutions to the hydrostatic (incompressible) Euler equations with a density and horizontal velocity distributions which are piecewise constant with respect to the vertical variable. This statement is made explicit below.
In this work we investigate solutions to the hydrostatic Euler equations in the vicinity of such solutions, that is relaxing the sharp stratification as well as columnar motion assumptions. We prove that for initial data suitably close to the bilayer framework, the emerging solutions to the hydrostatic Euler equations remain close to the solution predicted by the bilayer shallow water system on a relevant timescale.
This task is made difficult in part because we lack good stability estimates for the hydrostatic Euler equations in the presence of density stratification. For that matter, as in our previous works [9, 1], we rely on the regularizing properties of thickness-diffusivity terms proposed by Gent and McWilliams [25] so as to model the effective contributions of geostrophic eddy correlations in non-eddy-resolving systems.
Description of our results
Specifically, the hydrostatic Euler equations we consider take the form
(1.1) | ||||
where the Montgomery potential is given by
(1.2) |
Here, the equations are formulated using isopycnal coordinates (in particular we assume that the fluid is stratified in the sense that the two-dimensional fluid domain is foliated through lines of equal density, namely isopycnals). The variable represents the deviation of the infinitesimal depth of isopycnals from the reference value , and is the deviation of the horizontal velocity of the fluid particles from the reference value , and their density. The unknowns depend on the time , the horizontal space , and the variable referring to the isopycnal line at stake, while and are given and depend only on . The derivation of these equations from the more standard formulation in Eulerian coordinates is described for instance in [9].111Let us point out that in [9] we choose to label isopycnal lines using the value of the density of fluid particles: . Here we use a different convention, so as to set the reference infinitesimal depth of isopycnals at value . Notice that the change of variable is bijective in the stably stratified situation, i.e. when is strictly decreasing, but that our choice in this work allows to consider stratifications that are not strictly monotonic such as homogeneous and layered configurations. Incidentally, let us mention that in (1.1) we have set the gravity acceleration to and the total depth of the fluid domain at rest to through suitable rescaling. Finally is the thickness diffusivity coefficient and is often referred to as the “bolus velocity”.
The corresponding bilayer shallow water system reads
(1.3) |
Here, (resp. ) represents the deviation of the depth of the upper (resp. lower) layer from the reference constant value (resp. ), and (resp. ) is the deviation of the horizontal velocity within the upper (resp. lower) layer from the reference constant value (resp. ). Notice depend only on the time and horizontal space variables, bringing about a relative ease of use of the bilayer model. We denote (resp. ) the constant density of fluid particles in the upper (resp. lower) layers. Finally, is again the thickness diffusivity coefficient.
As mentioned above, solutions to (1.3) provide exact solutions to (1.1). Specifically, if we denote
(1.4) | ||||
where is a solution to (1.3), then is a solution to (1.1)-(1.2). In this work we shall compare these solutions with the ones emerging from profiles satisfying (1.4) only approximately. Specifically, our results are twofold.
-
i.
We prove that strong solutions to the bilayer shallow water system (1.3) emerge from sufficiently regular initial data satisfying some hyperbolicity conditions.
- ii.
Notice that to accommodate our aim of comparing solutions with a piecewise constant density distribution with a solution with a continuous density distribution, we need to consider deviations that can be pointwise large in small regions, that is the pycnocline. This demands to weaken the topology measuring the size of deviations in (ii). Yet a control in the strong topology associated with a Banach algebra turns out to be necessary to secure suitable convergence estimates. Our strategy then relies on the construction of a refined approximate solution, which is proved to be close to the corresponding solution to the hydrostatic Euler equation in a strong topology, and close to the bilayer solution in a weaker topology. Hence this refined approximate solution improves the description of the exact solution within the pycnocline.
A second important remark is that while the contribution of thickness diffusivity is essential to our stability estimates, we wish to control and compare solutions on a time interval which is uniform with respect to the thickness diffusivity parameter, . The dependency on the thickness parameter will appear only as a restriction on the size of admissible deviations. Concerning (i) this is made possible by the well-known fact that the bilayer shallow-water system is well-posed (under some hyperbolicity conditions) when (see [40]). Yet obtaining the corresponding result for is not straightforward and demands to use finely the structure of the thickness diffusivity parameters, following the “two-velocity” strategy developed in the context of the BD-entropy (see e.g. [16]). Concerning (ii) we use the existence of the bilayer solution and consequently the existence of the refined approximate solution to bootstrap the control of sufficiently close solutions to the hydrostatic Euler equations on the relevant timescale. For that purpose we strongly use the regularizing effect of thickness diffusivity contributions, but any non-uniformity with respect to the diffusivity parameter, , can be balanced through the smallness of the deviations.
Let us point out that the result (ii) applies to any given (sufficiently regular with respect to the time and horizontal space variables) solution to the hydrostatic Euler equations. Hence our work provides the same stability estimates around other solutions, constructed for instance in the framework of multiple layers and/or simple waves.
Related literature
Several existing works discuss the matter of modeling thin pycnoclines through the bilayer framework. Let us first recall that in the bilayer framework, non-hydrostatic pressure contributions trigger strong Kelvin–Helmholtz instabilities that in particular prevent the well-posedness of the initial-value problem in the absence of any additional regularizing ingredients; see [32]. Such regularizing ingredients include interfacial tension as proved in [34] but this is not expected to be te physically relevant mechanism. In [10], Bogucki and Garrett describe and model a scenario of interface-thickening due to mixing triggered by shear instabilities, up to a situation where the Richardson number in the interface becomes compatible with the celebrated Miles [39] and Howard [30] stability ion. Recall however that the bilayer shallow water system does not suffer from shear-induced instabilities when shear velocities are sufficiently small. Consistently, the authors in [19] discuss the simultaneous limits of sharp stratification together with shallow water (which can be considered as a low-frequency or hydrostatic pressure limit). In our work we impose the hydrostatic pressure assumption thus taming shear instabilities, although as pointed out previously the stability of continuously stratified hydrostatic Euler equations is poorly understood in the presence of density variations (see [43, 12, 28, 38] in the homogeneous framework). Notice also that shear-induced instabilities disappear when restricting the framework to purely traveling waves. In this framework, James [31] (improving upon [44, 4]) was able to rigorously justify the sharp stratification limit: writing the bilayer and continuously stratified problems in a unified formulation, James proved the existence of internal traveling waves associated with density stratifications in a small neighborhood (according to the topology) of the bilayer framework which converge towards the bilayer solution in the limit of sharp stratification. Our results are in the same spirit: system (1.1) is our unified formulation, and the topology controlling the limit of sharp stratification in our work is the topology. However, our results admit non-trivial dynamics thanks to the hydrostatic assumption and the presence of thickness diffusivity.
The propagation of internal waves with thin pycnoclines in relation with bilayer models was also investigated through experiments. In particular Grue et al. [29] set up precise experiments generating large-amplitude solitary waves and reported the dynamical development of rolls on the trailing side of the largest considered waves in accordance with the mechanism promoted by Bogucki and Garrett, while bilayer models provide very accurate predictions otherwise. Almgren, Camassa and Tiron [3] investigate thoroughly this matter through careful numerical simulations and analytical results of asymptotic bilayer models, analyzing the triggering of shear-induced instabilities in the region of maximal displacement as well as their advection into stable regions of the flow. White and Helfrich [46] consider internal bores generated by a dam-break, and compare continuously stratified and bilayer models with numerical experiments. From their findings they suggest an improvement on existing bilayer theories. In [18], Camassa and Tiron optimize bilayer models (specifically calibrating the top and bottom densities and the position of the sharp interface) and compare the analytical predictions of the optimized bilayer models with respect to the numerically computed continuously stratified solutions —considering infinitesimally small waves, internal bores and solitary waves— showing excellent agreement even in situations of relatively thick pycnoclines. Furthermore they propose a new asymptotic model taking into account thin pycnoclines in view of reconstructing analytically local properties of traveling waves within the pycnocline, which is similar in spirit with the “refined approximate solution” that we introduce in this work. Notice the authors consider that “fully time-dependent models governing the evolution of the pycnocline thickness probably constitute one of the most relevant extensions of the model [they] have introduced”, and we believe that our work provides a partial answer in that respect.
Outline
This manuscript is structured as follows. In Section 2 we study the bilayer shallow water systems. We first recall some known results without thickness diffusivity, and then consider the system with thickness diffusivity contributions. The main result is Proposition 2.7 which provides for any sufficiently regular initial data satisfying some hyperbolicity criterion the existence and control of strong solutions to the bilayer system on a time interval which is uniform with respect to , while Proposition 2.8 states the strong convergence as . In Section 3 we study the hydrostatic Euler equations. We provide first some stability estimates with respect to perturbations of the equations and of the data using suitable distances. As a second step we introduce refined approximate solutions associated with some given reference exact solution and close-by profiles. Building upon these approximate solutions, we prove Proposition 3.8 which controls the difference between the reference solution and exact solutions to the hydrostatic Euler equations emerging from close-by profiles. Together, Proposition 2.7 and Proposition 3.8 provide the announced result that for profiles suitably close to the bilayer framework and satisfying some hyperbolicity criterion the emerging solutions to the hydrostatic Euler equations remain close to the solution predicted by the bilayer shallow water system on a relevant timescale. The rigorous statement is displayed in Section 4, completed with a discussion on analogous statements in the one-layer and multilayer frameworks.
Notations
Let us introduce some notations for functional spaces used in this work.
-
•
The spaces are the standard Lebesgue spaces endowed with the usual norms denoted .
-
•
The spaces for are the -based Sobolev spaces endowed with the usual norms denoted .
-
•
The spaces for are the -based Sobolev spaces endowed with the usual norms denoted .
-
•
Given a real interval and a Banach space, (respectively ) the space of -integrable (respectively -continuously differentiable) -valued functions, endowed with their usual norms.
-
•
When useful, we provide insights on the variables at stake in aforementioned functional spaces by means of subscripts. For instance for we may denote
-
•
We sometimes also use subscripts to provide information on the interval at stake in functional spaces. For instance for and we may denote
2 The bilayer shallow water system
In this section we analyze the bilayer shallow water system (1.3). We first consider the case without diffusivity () and recall the hyperbolicity analysis due to Ovsjannikov [40]. We complete it by exhibiting explicit symmetrizers of the system of conservation laws. The standard theory for quasilinear systems then provides the local well-posedness of the initial-value problem, that we state in Proposition 2.4.
Extending such result for the system with diffusivity () uniformly with respect to is not as obvious as one could naively think, because the aforementioned symmetrizer behaves poorly with respect to diffusivity contributions. In order to deal with this issue, we exhibit regularization effects stemming from the diffusivity contributions that apply to the total velocity (that is adding the bolus velocities to the velocity unknowns). This is in the spirit of the BD entropy that arose in the context of the barotropic Euler equations with degenerate viscosities (see [16]). We infer a stability result on the linearized system, Lemma 2.6, which eventually yields the “large-time” —that is uniform with respect to — control of solutions stated in Proposition 2.7, and their strong convergence towards corresponding solutions to the non-diffusive system as stated in Proposition 2.8.
2.1 The system without thickness diffusivity
We consider the system
(2.1) |
We shall also always assume and . Through rescaling and Galilean invariance we can assume without loss of generality that and .
In compact form, the system reads
with , and where we introduce the matrix-valued function
(2.2) |
The following Lemma concerning the hyperbolicity domain of the bilayer shallow water system is proved in [40, 7, 45].
Lemma 2.1.
Let and be such that that . There exist two values such that the following holds:
-
1.
If , then there exist four distinct real eigenvalues of the matrix .
-
2.
If , then there exist two distinct real eigenvalues of the matrix and two distinct complex conjugate eigenvalues.
-
3.
If , then there exist four distinct real eigenvalues of the matrix .
Moreover, and depend only and smoothly on and .
Remark 2.2.
Ovsjannikov [40] —revisited by Barros and Choi [7] and then by Viríssimo and Milewski [45]— provided a nice geometrical approach to the critical values and . The characteristic polynomial associated to is
Notice that is a real root of if and only if satisfies the following identities:
(2.3) |
The first equality describes a fourth-order curve parametrized by having four axes of symmetry and consisting of an inner closed curve and four hyperbolic branches and the second equality describes the straight line with slope and intercept . In this geometrical approach, and (and their opposite) are the intercepts of the tangents to the fourth-order curves with slope .
Figure 1 reproduces the aforementioned curves and straight lines for several parameter values.



In this work we restrict our analysis to the hyperbolic domain described by Lemma 2.1(1). While standard theory for strictly hyperbolic systems guarantees the existence of a symmetrizer to (2.1) by using spectral projections [8], the following Lemma provides an (almost) explicit expression for such a symmetrizer.
Lemma 2.3.
There exists such that, denoting for ) the matrix
satisfies (i) is symmetric; and (ii) is symmetric, definite positive.
Moreover, can be chosen so that (for ) depends only and smoothly on , , and .
Proof.
It is straightforward to check that and are symmetric (and real-valued) for any value of . In order to prove that is definite positive for a suitable choice of , we rely on Sylvester’s criterion. We obtain the requirements
and
The last inequality is equivalent to where is the aforementioned characteristic polynomial. By Lemma 2.1, under the condition (2.4) there are four distinct real roots to , which we can denote and for any . Moreover for all , belongs to the domain delimited by the inner closed curve, and in particular we have . Hence we find that for all all principal minors are positive, and hence is definite positive.
By the standard perturbation theory [33], for and depend smoothly on , and . What is more, we can always choose (smoothly) so that and when , or and when , which corresponds to enforcing . This concludes the proof. ∎
The following proposition follows from the standard theory on strictly hyperbolic systems (see e.g. [8]). For convenience, we define for a compact subset of the domain of strict hyperbolicity as
(2.5) |
where is defined in Lemma 2.1.
Proposition 2.4 (Well-posedness).
Let , and . There exist and such that the following holds.
For all such that and and such that
and
there exists a unique maximal-in-time (classical) solution to (2.1) emerging from the initial data .
Moreover, one has and for any one has
and
Moreover, the maximal existence time (resp. the emerging solution in ) is a lower semi-continuous (resp. continuous) function of the initial data in and if then
2.2 The system with diffusivity
We now consider the system
(2.6) |
Proposition 2.5 (Small time well-posedness).
Let , , and . There exists such that the following holds.
For all , for all such that and and for all such that
and
there exists a unique maximal-in-time strong solution to (2.6) emerging from the initial data .
Moreover, and for any one has
and
Moreover, the maximal existence time (resp. the emerging solution in ) is a lower semi-continuous (resp. continuous) function of the initial data in and if then
Proof.
The proof has been given in [1], but we sketch it here for convenience. We view (2.6) as a system of two transport-diffusion equations and two transport equations, coupled only through order-zero source terms:
The standard theory on transport and transport-diffusion equations (see [6]) allows to bootstrap the standard fixed-point strategy through Picard iterates
which defines a sequence satisfying the following estimates (where is a non-essential constant depending on )
and converging in provided where is chosen sufficiently small.
The proof of the continuity of the flow map can be obtained along the same lines, using the continuity with respect to the initial data and Lipschitz-continuity with respect to source terms of the transport-diffusion and transport equations. ∎
The terminology “Small time well-posedness” in Proposition 2.5 refers to the fact that the time of existence and control of solutions of the above result is limited to , and in particular may vanish as . Notice that, differently from the statement of Proposition 2.4, we do not assume that the flow is stably stratified, namely . Assuming additionally that the flow is stably stratified, the first author improved this result in situations with small shear velocities and small deviations from the shear equilibrium, and obtains in [1] the existence and uniform control of solution up to times .
In the following results, we complete the picture by showing that, in the situation where the shear velocity is small enough to guarantee that the flow is in the hyperbolic domain of the non-diffusive equation, then the time of existence is uniform with respect to . In fact we shall prove the expected property that solutions to the diffusive system (2.6) converge as towards corresponding solutions to the non-diffusive system (2.1) as long as the non-diffusive solution is bounded.
In order to obtain stability estimates that are uniform with respect to , we rely on two main ideas. Firstly, we shall use energy estimates using the explicit symmetrizer adapted to the non-diffusive system introduced in Lemma 2.3 (while the strategy in [1] used only its block-diagonal component). Because the non-diagonal components of the symmetrizer behave poorly with respect to the diffusive contributions, we need another ingredient. Specifically, we notice that the total velocities () associated to solutions to (2.6) satisfy the system
(2.7) |
We observe that diffusive terms act as effective viscosity contributions on the total velocities. The last two equations read equivalently
and we recognize the shallow-water equations with degenerate viscosity contributions which were advocated by Gent in [24] and derived from the Navier–Stokes equations in [26, 17]. In their analysis of such systems (and generalizations thereof), Bresch and Desjardins introduced the so-called BD entropy in [15, 13, 14] (see also [16] for a refined analysis), which is based precisely in the reformulation of (2.7) as (2.6) (in dimension ).
In the same spirit, we combine the regularizing effects of the effective diffusivity and viscosity terms with aforementioned energy estimates, which allows us to obtain suitable stability estimates presented in Lemma 2.6, below.
Applying Lemma 2.6 to (the derivatives of) solutions to (2.6)-(2.7), we find a time of existence which is uniform with respect to . We state the result in forthcoming Proposition 2.7.
Applying Lemma 2.6 to (the derivatives of) the difference between solutions to (2.6)-(2.7) and corresponding solutions to the non-diffusive system (2.1) () yields the aforementioned convergence of the former towards the latter as , on a time interval defined by the solutions without diffusivity. We state the result in forthcoming Proposition 2.8.
Lemma 2.6 (Stability).
Let and . There exists depending only on and depending also on such that the following holds.
Let and be sufficiently regular solutions to the linearized equations with remainders
where we denote
and
Moreover, denote such that
(2.8) |
Then, for any , one has the estimate
Proof.
Denote
where (for ) with provided in Lemma 2.3. Using that and are symmetric, and integration by parts, we have the energy identities
By means of Cauchy–Schwarz inequality we find that
where denotes a multiplicative constant depending only on and , and which may change from line to line. We now focus on the remaining terms. We first notice that defects of symmetry in arise only in the first two rows, and that the first two components of equal the first two components of . Hence using integration by parts and Cauchy–Schwarz inequality we infer
Then, again making use of the identity we infer that
After integration by parts, and since , we find that
Then, using from (2.8) that (where ), we obtain the identities
where we used integration by parts in the last line. We infer
Combining all these estimate and denoting
one has
(2.9) |
By using that is definite positive, we find that there exists depending only on such that
Hence we find (using the Peter Paul inequality and augmenting ) that
and the result follows by Gronwall’s Lemma. ∎
Proposition 2.7 (Large-time well-posedness).
Let , and . There exists and such that the following holds.
Let , such that and , and let such that the hyperbolicity condition holds:
where is defined in (2.5), and
Denote the maximal-in-time solution to (2.6) emerging from the initial data as defined in Proposition 2.5.
One has and for any ,
and
Proof.
We assume that the initial data is smooth, so that are smooth on their domain of existence. The general case is obtained by regularizing the initial data and passing to the limit, thanks to the persistence of regularity and continuity of the flow map stated in Proposition 2.5.
Denote (), and
Applying the operator to (2.6) and (2.7), we obtain
and
with remainders and defined as
Moreover, applying the operator to the identity () yields
with
Standard commutator estimates and composition estimates in Sobolev spaces; see e.g. [35, Appendix B] yield
where is a positive constant depending only on , and .
Moreover, notice that by using the equations (2.6), (2.7) and the identity for , we have
where the multiplicative constant depends on and (for ).
We may thus apply Lemma 2.6, and infer that we can set depending on , and depending only on , so that as long as
(2.10) |
and
(2.11) |
one has (using the standard continuous embedding and the fact that and )
Applying Gronwall’s Lemma, we find that if is smaller than a universal constant, then
What is more we have (augmenting if necessary)
Hence lowering further , we infer that for all . By the usual continuity argument we infer that the assumptions (2.10) and (2.11) do hold for with depending on . This yields the lower bound on the maximal time of existence and the claimed upper bound on the solution follows from the above estimates replacing with . ∎
Combined together, Proposition 2.5 and Proposition 2.7 yield a time of existence for solutions to (2.6) emerging from sufficiently regular initial data which is independent of . The following result describes the behavior of these solutions as .
Proposition 2.8 (Convergence).
Let , and .
Let , such that and , and be such that the hyperbolicity condition holds
where is defined in (2.5), and
Denote
- •
- •
-
•
and such that for all ,
and
Then there exists and , both depending only on , such that for all , is well-defined for all and satisfies the hyperbolicity condition
and the upper bound
Moreover, one has for all
Proof.
Denote and (), , and
Substracting (2.6), (2.7) and (2.1) we obtain
and
where, for any , we denote
Moreover, one has by definition
where
Standard commutator estimates and composition estimates in Sobolev spaces; see e.g. [35, Appendix B] yield
where is a positive constant depending only on , , and for any .
Moreover notice that by the equations (2.6) and (2.7) and using the identity we have
where the multiplicative constant depends on and for .
We may thus apply Lemma 2.6, and infer that we can set depending on , and depending only on , so that as long as
(2.12) |
and
(2.13) |
one has, using that ,
Now, we use in the integrand the triangle inequality
The first contribution may be absorbed by the left-hand side if is sufficiently small (depending on ), and the second contribution is estimated, as other terms, using the assumption
Applying then Gronwall’s Lemma, we find that
(2.14) |
where depends only on .
3 The hydrostatic Euler equations
In this section we study the stability of the hydrostatic Euler equations for stratified flows:
(3.1) | ||||
where the Montgomery potential is given by
(3.2) |
We recall that our stability results must accommodate with solutions generated by the bilayer system that are piecewise constant and allow the comparison with continuously stratified flows. Hence we must allow for deviations that can be large pointwise, while smallness stems from integration (with respect to the -variable). In practice we shall manipulate simultaneously the pointwise as well as the topologies depending on the need. This is the case for instance in the following Lemma, measuring the Lipschitz continuity with respect to the density variable of the Montgomery operator defined (3.2).
Lemma 3.1.
Let . There exists such that for any () such that
and for any , one has for almost any ,
Proof.
∎
Note also that we are seeking stability estimates for the system (3.1)-(3.2) with respect to perturbations of the equations —in particular through and — and with respect to perturbations of the initial data.
In Section 3.1, we first state a local well-posedness result associated with the initial-value problem for the system (3.1)-(3.2), and then provide some stability estimates. These two results by themselves are not sufficient to bootstrap in a standard manner the strong convergence of solutions as the size of deviations shrink, because the topology involved in the first result, namely , is stronger than the topology used in the second result, which is roughly speaking . For that matter we introduce and study in Section 3.2 a refined approximate solution which, compared with the original reference solution, improves the description of solutions associated with nearby profiles. Specifically, the refined approximate solution satisfies the following three properties:
-
i.
it is well-defined and controlled on a time interval which is uniform with respect to ;
-
ii.
the difference with respect to the nearby solution is controlled for the strong norm associated with ;
-
iii.
the difference with respect to the reference solution is controlled for the weak norm associated with .
The resulting convergence result, Proposition 3.8, is stated and proved in Section 3.3.
3.1 Stability estimates
Proposition 3.2 (Well-posedness).
Let , , and . There exists such that the following holds.
For all , all such that
and all such that for almost all ,
and
there exists a unique maximal-in-time strong solution to (3.1)–(3.2) emerging from the initial data .
Moreover, and for any and almost all one has
and
Moreover, the maximal existence time (resp. the emerging solution in ) is a lower semi-continuous (resp. continuous) function of the initial data in and if then
Proof.
The proof is very similar to the proof of Proposition 2.5, using estimates for transport and transport-diffusion equations pointwisely with respect to the variable . The essential arguments are that is a Banach algebra and that differentiation with respect to the space variable as well as the Fourier multiplier commute with the operator , and that the linear operator is bounded for any . ∎
Proposition 3.3 (Stability).
Let , and . There exists such that the following holds.
Proof.
Let us denote , , and . We have on the time interval
where we denote ,
and
We can now use standard estimates and transport-diffusion and transport equations ([6]) to infer that there exists depending only on such that for almost any , one has
and
Using that is a Banach algebra for all , we find that for any
and, making additionally use of standard composition estimates in Sobolev spaces ([35, Appendix B]) and Lemma 3.1,
Finally, using product and composition estimates on , we have
Collecting these estimates we infer that there exists depending only on such that
Hence there exists , depending only on such that for any sufficiently small so that one has , and one has
(3.5) |
Integrating this inequality with respect to the variable and using Minkowski’s inequality, we infer the first stability estimate, (3.3). Plugging (3.3) in the right-hand side of (3.5), the second stability estimate, (3.4), follows immediately. ∎
Remark 3.4.
The restriction in Propositions 3.2 and 3.3 is quite stringent. In [9], some improved stability estimates concerning the system (3.1)-(3.2) were derived by the authors. The latter estimates exploit a partial symmetric structure of the equations, and demand some extra regularity with respect to the variable . Because we cannot afford such regularity since our stability estimates will be used with piecewise constant functions, we use in the proof of Proposition 3.3 in a stronger way the parabolic regularization of thickness diffusivity.
In order to obtain a final result on a timescale which is independent of the parameter we shall exploit some a priori control on a reference solution and restrict to initial data as well as shear velocity and density distributions that are close to the reference data.
3.2 Refined approximation
In this section we consider a given reference solution to the hydrostatic Euler equation (3.1)-(3.2) and build from it a refined approximate solution associated with nearby profiles. Specifically, let be a solution to
(3.6) | ||||
where we recall that the operator is defined in (3.2). Considering profiles which are in some sense close to we construct approximate solutions as the solutions to
(3.7) | ||||
Notice first that satisfies approximately the hydrostatic Euler equations associated with profiles .
Proposition 3.5.
For any the refined approximate solution satisfies
(3.8) | ||||
with
Proof.
We have
and the result follows since . ∎
That the above remainder term is small is a consequence of the subsequent Proposition 3.7. We first prove that for any initial data , the emerging solution is well-defined and controlled on a time interval uniform with respect to .
Proposition 3.6.
Let , , and . There exists and such that the following holds.
Let be such that
For all , all such that
and all such that for almost all
and
there exists a unique maximal solution to (3.7) emerging from the initial data . Moreover one has and for any and almost any one has the upper bound
Proof.
The existence and uniqueness of maximal solution to (3.7) is obtained as in Proposition 3.2. We set . By standard estimates on transport and transport-diffusion equations ([6]) applied to (3.7) we have
and
where we denote , , , and the constant depends only on .
Now we notice that for almost any ,
From this we infer that there exists , depending only on such that
By the standard continuity argument, we find that
for almost all and for all such that
where depends only on . This concludes the proof. ∎
We conclude this section by investigating the difference between the reference solution and the refined approximate solution.
Proposition 3.7.
Let , and . There exists such that the following holds.
Proof.
Let us denote , , and . We have on the time interval for which both function are well-defined
where ,
and
We can proceed as in the proof of Proposition 3.3 with some straightforward adjustments as for the contributions of since the contribution is nonexistent.
We infer that there exists depending only on such that for all and such that , one has for almost any ,
Iterating this control on , we find that
which yields the claimed estimate. ∎
3.3 Convergence
We now conclude our analysis with the following stability result for solutions to the hydrostatic Euler equations.
Proposition 3.8 (Convergence).
Let , , and . Then there exists independent of and (depending on ) such that the following holds.
Proof.
We first recall that Proposition 3.7 provides an estimate on the difference between the reference solution and the corresponding approximate solution defined as the solution to the system (3.7) with initial data , , whose existence and control on the time interval (lessening if necessary) is provided by Proposition 3.6. Specifically we have for almost any
where depends only on and .
We then consider the difference between the exact solution —whose existence is provided by Proposition 3.2— and the approximate solution. By means of the consistency result, Proposition 3.5, we can adapt the proof of Proposition 3.3 and we find that under the assumptions that
(3.11) |
there exists depending only on such that for all and such that and for almost any one has
Iterating the stability estimate on we deduce that (augmenting if necessary)
Since by construction and using the above control on we infer (augmenting if necessary)
Notice that this estimate implies by triangle inequality an upper bound on which (choosing sufficiently large and sufficiently small) enforces strictly the condition (3.11), so that by the standard continuity argument the above holds without restriction on . By the persistence of regularity stated in Proposition 3.2 we infer that and the above holds for any . The estimate (3.10) then immediately follows from the triangle inequality. ∎
4 Conclusion
Thanks to Proposition 2.7 (concerning the well-posedness and control of solutions to the bilayer system) on one hand and Proposition 3.8 (concerning the control of the deviations of nearby solutions to some given reference solutions) on the other hand, one infers immediately the announced rigorous justification of the propagation in time of the columnar motion and sharp stratification assumptions in near-bilayer situations (within the hydrostatic framework). Specifically, we have the following result.
Theorem 4.1.
Let , ,, , and . Then there exist independent of and (depending on ) such that the following holds.
Remark 4.2.
Proposition 3.8 is not limited to the bilayer framework and applies to any suitably regular reference solution. Hence it can be combined with results analogous to Proposition 2.7 to provide results analogous to Theorem 4.1 in the one-layer and multilayer frameworks.
The result analogous to Proposition 2.7 in the one-layer framework (that is associated with the standard shallow water equation with thickness diffusivity that was discussed for instance in [24]) is stated and proved in [2, Sect. 2.2]. Notice it requires neither the discussion on the hyperbolicity domain of the non-diffusive equations (since the standard non-cavitation assumption guarantees hyperbolicity) nor the discussion on the parabolic regularization of the total velocity (since the natural symmetrizer behaves well with diffusivity contributions).
A result analogous to Proposition 2.7 in the multilayer framework follows from combining the result of [22] with the analysis of Section 2.2. In the former, it is proved that assuming sufficiently small shear velocities is a sufficient condition for the (strict) hyperbolicity of the multilayer system in the stably stratified situation. Notice however that this smallness condition is implicit, and not uniform with respect to the number of layers.
Acknowledgements
The authors thank the Institute for applied mathematics “Mauro Picone” (IAC) for its hospitality when this work was initiated, and Laboratoire Ypatia des Sciences Mathématiques LYSM CNRS-INdAM International Research Laboratory as well as Univ. Rennes funding for its financial support. MA and VD thank the program Centre Henri Lebesgue ANR-11-LABX-0020-0, for fostering an attractive mathematical environment. RB acknowledges fundings from the Italian Ministry of University and Research, PRIN 2022HSSYPN (Teseo).
References
- [1] M. Adim. Approximating a continuously stratified hydrostatic system by the multi-layer shallow water system. arXiv preprint: 2307.11426.
- [2] M. Adim. Modèles continument stratifiés et systèmes multi-couches pour les écoulements géophysiques. PhD thesis, Univ. Rennes, 2024.
- [3] A. Almgren, R. Camassa, and R. Tiron. Shear instability of internal solitary waves in Euler fluids with thin pycnoclines. J. Fluid Mech., 710:324–361, 2012.
- [4] C. J. Amick and R. E. L. Turner. A global theory of internal solitary waves in two-fluid systems. Trans. Amer. Math. Soc., 298(2):431–484, 1986.
- [5] P. Azérad and F. Guillén. Mathematical justification of the hydrostatic approximation in the primitive equations of geophysical fluid dynamics. SIAM J. Math. Anal., 33(4):847–859, 2001.
- [6] H. Bahouri, J.-Y. Chemin, and R. Danchin. Fourier analysis and nonlinear partial differential equations, volume 343. Springer, 2011.
- [7] R. Barros and W. Choi. On the hyperbolicity of two-layer flows. In Frontiers of applied and computational mathematics, pages 95–103. World Sci. Publ., Hackensack, NJ, 2008.
- [8] S. Benzoni-Gavage and D. Serre. Multidimensional hyperbolic partial differential equations. First-order systems and applications. Oxford Mathematical Monographs. The Clarendon Press Oxford University Press, Oxford, 2007.
- [9] R. Bianchini and V. Duchêne. On the hydrostatic limit of stably stratified fluids with isopycnal diffusivity. To appear in Comm. Partial Differential Equations. arXiv preprint: 2206.01058.
- [10] D. Bogucki and C. Garrett. A simple model for the shear-induced decay of an internal solitary wave. J. Phys. Oceanogr., 23(8):1767–1776, 1993.
- [11] J. L. Bona, D. Lannes, and J.-C. Saut. Asymptotic models for internal waves. J. Math. Pures Appl. (9), 89(6):538–566, 2008.
- [12] Y. Brenier. Homogeneous hydrostatic flows with convex velocity profiles. Nonlinearity, 12(3):495–512, 1999.
- [13] D. Bresch and B. Desjardins. Existence of global weak solutions for a 2D viscous shallow water equations and convergence to the quasi-geostrophic model. Comm. Math. Phys., 238(1-2):211–223, 2003.
- [14] D. Bresch and B. Desjardins. Some diffusive capillary models of Korteweg type. C. R., Méc., Acad. Sci. Paris, 332(11):881–886, 2004.
- [15] D. Bresch, B. Desjardins, and C.-K. Lin. On some compressible fluid models: Korteweg, lubrication, and shallow water systems. Comm. Partial Differential Equations, 28(3-4):843–868, 2003.
- [16] D. Bresch, B. Desjardins, and E. Zatorska. Two-velocity hydrodynamics in fluid mechanics: Part II. Existence of global -entropy solutions to the compressible Navier-Stokes systems with degenerate viscosities. J. Math. Pures Appl. (9), 104(4):801–836, 2015.
- [17] D. Bresch and P. Noble. Mathematical justification of a shallow water model. Methods Appl. Anal., 14(2):87–117, 2007.
- [18] R. Camassa and R. Tiron. Optimal two-layer approximation for continuous density stratification. J. Fluid Mech., 669:32–54, 2011.
- [19] B. Desjardins, D. Lannes, and J.-C. Saut. Normal mode decomposition and dispersive and nonlinear mixing in stratified fluids. Water Waves, 3(1):153–192, 2021.
- [20] V. Duchêne. Many Models for Water Waves. Open Math Notes, OMN:202109.111309, 2021.
- [21] V. Duchêne. Asymptotic shallow water models for internal waves in a two-fluid system with a free surface. SIAM J. Math. Anal., 42(5):2229–2260, 2010.
- [22] V. Duchêne. A note on the well-posedness of the one-dimensional multilayer shallow water model. HaL preprint: 00922045, 2013.
- [23] K. Furukawa, Y. Giga, M. Hieber, A. Hussein, T. Kashiwabara, and M. Wrona. Rigorous justification of the hydrostatic approximation for the primitive equations by scaled Navier-Stokes equations. Nonlinearity, 33(12):6502–6516, 2020.
- [24] P. R. Gent. The energetically consistent shallow-water equations. Journal of the atmospheric sciences, 50(9):1323–1325, 1993.
- [25] P. R. Gent and J. C. McWilliams. Isopycnal mixing in ocean circulation models. J. Phys. Oceanogr., 20(1):150–155, 1990.
- [26] J.-F. Gerbeau and B. Perthame. Derivation of viscous Saint-Venant system for laminar shallow water; numerical validation. Discrete Contin. Dyn. Syst. Ser. B, 1(1):89–102, 2001.
- [27] A. E. Gill. Atmosphere-ocean dynamics, volume 30 of International geophysics series. Academic Press, 1982.
- [28] E. Grenier. On the derivation of homogeneous hydrostatic equations. M2AN Math. Model. Numer. Anal., 33(5):965–970, 1999.
- [29] J. Grue, A. Jensen, P.-O. Rusås, and J. K. Sveen. Properties of large-amplitude internal waves. J. Fluid Mech., 380:257–278, 1999.
- [30] L. N. Howard. Note on a paper of John W. Miles. J. Fluid Mech., 10:509–512, 1961.
- [31] G. James. Internal travelling waves in the limit of a discontinuously stratified fluid. Arch. Ration. Mech. Anal., 160(1):41–90, 2001.
- [32] V. Kamotski and G. Lebeau. On 2D Rayleigh-Taylor instabilities. Asymptot. Anal., 42(1-2):1–27, 2005.
- [33] T. Kato. Perturbation theory for linear operators. Classics in Mathematics. Springer-Verlag, Berlin, 1995. Reprint of the 1980 edition.
- [34] D. Lannes. A Stability Criterion for Two-Fluid Interfaces and Applications. Arch. Ration. Mech. Anal., 208(2):481–567, 2013.
- [35] D. Lannes. The water waves problem, volume 188 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2013. Mathematical analysis and asymptotics.
- [36] J. Li and E. S. Titi. The primitive equations as the small aspect ratio limit of the Navier-Stokes equations: rigorous justification of the hydrostatic approximation. J. Math. Pures Appl. (9), 124:30–58, 2019.
- [37] J. Li, E. S. Titi, and G. Yuan. The primitive equations approximation of the anisotropic horizontally viscous 3 Navier-Stokes equations. J. Differential Equations, 306:492–524, 2022.
- [38] N. Masmoudi and T. K. Wong. On the theory of hydrostatic Euler equations. Arch. Ration. Mech. Anal., 204(1):231–271, 2012.
- [39] J. W. Miles. On the stability of heterogeneous shear flows. J. Fluid Mech., 10:496–508, 1961.
- [40] L. V. Ovsjannikov. Models of two-layered “shallow water”. Zh. Prikl. Mekh. i Tekhn. Fiz., (2):3–14, 180, 1979.
- [41] X. Pu and W. Zhou. On the rigorous mathematical derivation for the viscous primitive equations with density stratification. Acta Math. Sci. Ser. B (Engl. Ed.), 43(3):1081–1104, 2023.
- [42] X. Pu and W. Zhou. Rigorous derivation of the full primitive equations by the scaled Boussinesq equations with rotation. Bull. Malays. Math. Sci. Soc., 46(3):Paper No. 88, 23, 2023.
- [43] V. M. Teshukov. On Cauchy problem for long wave equations. In Free boundary problems in continuum mechanics (Novosibirsk, 1991), volume 106 of Internat. Ser. Numer. Math., pages 331–338. Birkhäuser, Basel, 1992.
- [44] R. E. L. Turner. Internal waves in fluids with rapidly varying density. Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4), 8(4):513–573, 1981.
- [45] F. d. M. Viríssimo and P. A. Milewski. Nonlinear stability of two-layer shallow water flows with a free surface. Proc. A., 476(2236):20190594, 20, 2020.
- [46] B. L. White and K. R. Helfrich. A model for internal bores in continuous stratification. J. Fluid Mech., 761:282–304, 2014.