Skip to main content
Log in

On the Rigid-Lid Approximation for Two Shallow Layers of Immiscible Fluids with Small Density Contrast

  • Published:
Journal of Nonlinear Science Aims and scope Submit manuscript

Abstract

The rigid-lid approximation is a commonly used simplification in the study of density-stratified fluids in oceanography. Roughly speaking, one assumes that the displacements of the surface are negligible compared with interface displacements. In this paper, we offer a rigorous justification of this approximation in the case of two shallow layers of immiscible fluids with constant and quasi-equal mass density. More precisely, we control the difference between the solutions of the Cauchy problem predicted by the shallow-water (Saint-Venant) system in the rigid-lid and free-surface configuration. We show that in the limit of a small density contrast, the flow may be accurately described as the superposition of a baroclinic (or slow) mode, which is well predicted by the rigid-lid approximation, and a barotropic (or fast) mode, whose initial smallness persists for large time. We also describe explicitly the first-order behavior of the deformation of the surface and discuss the case of a nonsmall initial barotropic mode.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

Notes

  1. The models presented in these works are not limited to a flat bottom or horizontal dimension \(d=1\). They present different constants in the velocity equations. This is due to a different choice of scaling in the nondimensionalizing step. We chose our scaling in order to set the typical velocity of the internal wave (obtained by solving explicitly the linear system, i.e., setting \(\alpha =\epsilon =0\)) as \(c_0=\pm 1\), consistently with the rigid-lid system (1.2).

  2. The Saint-Venant model is usually derived using the so-called hydrostatic approximation. Equivalently, one may assume that the horizontal scale is large compared with the vertical scale, so that the horizontal velocity field is accurately described as constant throughout the depth of each layer of fluid.

  3. The justification provided in Bona et al. (2008)—as well as in Duchêne (2010) in the free-surface configuration—is in the sense of consistency: sufficiently smooth solutions of the full Euler system satisfy the equations of (1.2) up to small, i.e., \(\mathcal {O}(\mu ^2)\), remainder terms. The rigorous, full justification follows from the well-posedness of both the full Euler system and the shallow-water model, as well as a stability result, which make it possible to compare the solutions of both systems with corresponding initial data on the relevant time scale. In the rigid-lid situation, Lannes (2013) recently solved the difficult problem of the well-posedness of the full Euler system, consequently completing the full justification of (1.2); see Lannes (2013), Theorem 7. No such result is available in the bifluidic, free-surface configuration.

  4. Of course a fourth vector—second linearly independent element of \(\mathrm{ker}(L_{(0)})\)—could be defined, but this is not necessary in our analysis.

References

  • Abgrall, R., Karni, S.: Two-layer shallow water system: a relaxation approach. SIAM J. Sci. Comput. 31(3), 1603–1627 (2009)

    Article  MATH  MathSciNet  Google Scholar 

  • Alinhac, S., Gérard, P.: Opérateurs pseudo-différentiels et théorème de Nash-Moser. Savoirs Actuels (1991)

  • Barros, R., Gavrilyuk, S.L., Teshukov, V.M.: Dispersive nonlinear waves in two-layer flows with free surface. I. Model derivation and general properties. Stud. Appl. Math. 119(3), 191–211 (2007)

    Article  MathSciNet  Google Scholar 

  • Benjamin, T.B.: Internal waves of finite amplitude and permanent form. J. Fluid Mech. 25(2), 241–270 (1966)

    Article  MATH  MathSciNet  Google Scholar 

  • Bona, J.L., Colin, T., Lannes, D.: Long wave approximations for water waves. Arch. Ration. Mech. Anal. 178(3), 373–410 (2005)

    Article  MATH  MathSciNet  Google Scholar 

  • Bona, J.L., Lannes, D., Saut, J.-C.: Asymptotic models for internal waves. J. Math. Pures Appl. (9) 89(6), 538–566 (2008)

    Google Scholar 

  • Bresch, D., Renardy, M.: Well-posedness of two-layer shallow water flow between two horizontal rigid plates. Nonlinearity 24(4), 1081–1088 (2011)

    Article  MATH  MathSciNet  Google Scholar 

  • Browning, G., Kreiss, H.-O.: Problems with different time scales for nonlinear partial differential equations. SIAM J. Appl. Math. 42(4), 704–718 (1982)

    Article  MATH  MathSciNet  Google Scholar 

  • Castro-Díaz, M.J., Fernández-Nieto, E.D., González-Vida, J.M., Parés-Madroñal, C.: Numerical treatment of the loss of hyperbolicity of the two-layer shallow-water system. J. Sci. Comput. 48(1–3), 16–40 (2011)

    Article  MATH  MathSciNet  Google Scholar 

  • Choi, W., Camassa, R.: Weakly nonlinear internal waves in a two-fluid system. J. Fluid Mech. 313, 83–103 (1996)

    Article  MATH  MathSciNet  Google Scholar 

  • Craig, W., Guyenne, P., Kalisch, H.: Hamiltonian long-wave expansions for free surfaces and interfaces. Commun. Pure Appl. Math. 58(12), 1587–1641 (2005)

    Article  MATH  MathSciNet  Google Scholar 

  • Craig, W., Guyenne, P., Sulem, C.: Coupling between internal and surface waves. Nat. Hazards 57(3), 617–642 (2010)

    Article  Google Scholar 

  • de Saint-Venant, B.: Théorie du mouvement non-permanent des eaux, avec application aux crues des rivières et à l’introduction des marées dans leur lit. C. R. Acad. Sci. Paris 73:147–154 (1871)

  • Duchêne, V.: 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)

    Article  MATH  MathSciNet  Google Scholar 

  • Duchêne, V.: Boussinesq/Boussinesq systems for internal waves with a free surface, and the KdV approximation. M2AN. Math. Model. Numer. Anal. 46, 145–185 (2011)

    Article  Google Scholar 

  • Duchêne, V.: Decoupled and unidirectional asymptotic models for the propagation of internal waves. M3AS. Math. Models Methods Appl. Sci. 24(01) (2014)

  • Gill, A.E.: Atmosphere-Ocean Dynamics. International Geophysics Series, vol. 30. Academic Press (1982)

  • Grimshaw, R., Pelinovsky, E., Poloukhina, O.: Higher-order korteweg-de vries models for internal solitary waves in a stratified shear flow with a free surface. Nonlinear Processes Geophys. 9, 221–235 (2002)

    Article  Google Scholar 

  • Guyenne, P., Lannes, D., Saut, J.-C.: Well-posedness of the Cauchy problem for models of large amplitude internal waves. Nonlinearity 23(2), 237–275 (2010)

    Article  MATH  MathSciNet  Google Scholar 

  • Helfrich, K.R., Melville, W.K.: Long nonlinear internal waves. Annu. Rev. Fluid Mech. 38, 395–425 (2006)

    Google Scholar 

  • Jackson, C.R.: An atlas of internal solitary-like waves and their properties (2004). Accessible at url URL http://www.internalwaveatlas.com/Atlas2_index.html

  • Kato, T.: The Cauchy problem for quasi-linear symmetric hyperbolic systems. Arch. Ration. Mech. Anal. 58(3), 181–205 (1975)

    Article  MATH  Google Scholar 

  • Kato, T.: Perturbation theory for linear operators. Classics in Mathematics. Springer, Berlin (1995). Reprint of the 1980 edition

  • Kato, T., Ponce, G.: Commutator estimates and the Euler and Navier-Stokes equations. Commun. Pure Appl. Math. 41(7), 891–907 (1988)

    Article  MATH  MathSciNet  Google Scholar 

  • Klainerman, S., Majda, A.: Singular limits of quasilinear hyperbolic systems with large parameters and the incompressible limit of compressible fluids. Commun. Pure Appl. Math. 34(4), 481–524 (1981)

    Article  MATH  MathSciNet  Google Scholar 

  • Lannes, D.: Secular growth estimates for hyperbolic systems. J. Differ. Equ. 190(2), 466–503 (2003)

    Article  MATH  MathSciNet  Google Scholar 

  • Lannes, D.: Sharp estimates for pseudo-differential operators with symbols of limited smoothness and commutators. J. Funct. Anal. 232(2), 495–539 (2006)

    Article  MATH  MathSciNet  Google Scholar 

  • Lannes, D.: A stability criterion for two-fluid interfaces and applications. Arch. Ration. Mech. Anal. 208(2), 481–567 (2013)

    Article  MATH  MathSciNet  Google Scholar 

  • Lannes, D.: The water waves problem. Mathematical Surveys and Monographs, vol. 188. American Mathematical Society (2013)

  • Leonardi, D.: Internal and Surface Waves in a Two-Layer Fluid. PhD thesis, University of Illinois (2011)

  • Long, R.R.: On the Boussinesq approximation and its role in the theory of internal waves. Tellus 17(1), 46–52 (1965)

    Article  MathSciNet  Google Scholar 

  • Métivier, G.: Para-differential calculus and applications to the Cauchy problem for nonlinear systems. Centro di Ricerca Matematica Ennio De Giorgi (CRM) Series, vol. 5 (2008)

  • Schneider, G., Wayne, C.E.: The long-wave limit for the water wave problem. I. The case of zero surface tension. Commun. Pure Appl. Math. 53(12), 1475–1535 (2000)

    Article  MATH  MathSciNet  Google Scholar 

  • Shampine, L.F., Reichelt, M.W.: The MATLAB ODE suite. SIAM J. Sci. Comput. 18(1), 1–22 (1997)

    Article  MATH  MathSciNet  Google Scholar 

  • Stewart, A.L., Dellar, P.J.: Multilayer shallow water equations with complete coriolis force. Part 3. Hyperbolicity and stability under shear. J. Fluid Mech. 723, 289–317, 5 (2013)

    Google Scholar 

  • Taylor, M.E.: Partial differential equations. III Nonlinear equations. Applied Mathematical Sciences, vol. 117. Springer (1997)

  • Trefethen, L.N.: Spectral methods in MATLAB. Software, Environments, and Tools. Society for Industrial and Applied Mathematics (SIAM), vol. 10 (2000).

Download references

Acknowledgments

The author is grateful to Christophe Cheverry, Jean-François Coulombel, and Frédéric Rousset for helpful advice and stimulating discussions. This work was partially supported by Project ANR-13-BS01-0003-01 DYFICOLTI.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Vincent Duchêne.

Additional information

Communicated by P. Newton.

Appendix: Proof of Proposition 2.2

Appendix: Proof of Proposition 2.2

In this section, we detail the proof of Proposition 2.2, which follows the classical theory concerning Friedrichs-symmetrizable quasilinear systems. The proof is based on a priori energy estimates, for which the key ingredients are product and commutator estimates in Sobolev spaces. We first recall such results and refer the reader to, e.g., Alinhac and Gérard (1991), Lannes (2013) for the proof of Lemmata 5.1 and 5.3.

Lemma 5.1

(Product estimates) Let \(s\ge 0\). For all \(f,g\in H^s(\mathbb {R})\bigcap L^\infty (\mathbb {R})\), one has

$$\begin{aligned} \big \vert \ f \ g\ \big \vert _{H^s} \ \lesssim \ \big \vert \ f \ \big \vert _{L^\infty } \big \vert \ g\ \big \vert _{H^s}+\big \vert \ f \ \big \vert _{H^s} \big \vert \ g\ \big \vert _{L^\infty }. \end{aligned}$$

If \(s\ge s_0>1/2\), then one deduces, thanks to a continuous embedding of Sobolev spaces,

$$\begin{aligned} \big \vert \ f \ g\ \big \vert _{H^s} \ \lesssim \ \big \vert \ f \ \big \vert _{H^s} \big \vert \ g\ \big \vert _{H^s}. \end{aligned}$$

Let \(F\in C^\infty (\mathbb {R})\) such that \(F(0)=0\). If \(g\in H^s(\mathbb {R})\bigcap L^\infty (\mathbb {R})\) with \(s\ge 0\), then one has \(F(g)\in H^s(\mathbb {R})\) and

$$\begin{aligned} \big \vert \ F(g) \ \big \vert _{H^s} \ \le \ C(\big \vert \ g\ \big \vert _{L^\infty },\big \vert \ F\ \big \vert _{C^\infty })\big \vert \ g \ \big \vert _{H^s}. \end{aligned}$$

Throughout the paper, we repeatedly make use of the following corollary.

Corollary 5.2

Let \(f,\zeta \in L^\infty \bigcap H^{ s}\), with \(s\ge 0\) and \(h(\zeta )\equiv 1-\zeta \), with \(h(\zeta )\ge h_0>0\) for any \(x\in \mathbb {R}\). Then one has

$$\begin{aligned} \big \vert \frac{1}{h(\zeta )} f \big \vert _{H^{s}} \&\le \ C(h_0^{-1},\big \vert \zeta \big \vert _{L^{\infty }})\big ( \big \vert f\big \vert _{H^{s}}+ \big \vert \zeta \big \vert _{H^{s}}\big \vert f\big \vert _{L^\infty }\big )\\ \big \vert f-\frac{1}{h(\zeta )} f \big \vert _{H^{s}} \&\le \ \ C(h_0^{-1},\big \vert \zeta \big \vert _{L^{\infty }})\big ( \big \vert \zeta \big \vert _{L^{\infty }}\big \vert f\big \vert _{H^{s}}+ \big \vert \zeta \big \vert _{H^{s}}\big \vert f\big \vert _{L^\infty }\big ). \end{aligned}$$

Proof

We will use the identity

$$\begin{aligned} \frac{1}{h(\zeta )} f \ = \ \frac{1}{1-\zeta }f \ = \ f \ + \ \frac{\zeta }{1-\zeta }f. \end{aligned}$$

By Lemma 5.1, we deduce

$$\begin{aligned} \big \vert \frac{1}{h(\zeta )} f \big \vert _{H^{s}} \le \ \big \vert f \big \vert _{H^{s}} \ + \ \big \vert \frac{\zeta }{1-\zeta } f \big \vert _{H^{s}}\\&\lesssim \ \big \vert f \big \vert _{H^{s}} \ + \ \big \vert \frac{\zeta }{1-\zeta } \big \vert _{L^\infty }\big \vert f \big \vert _{H^{s}} \ + \ \big \vert \frac{\zeta }{1-\zeta }\big \vert _{H^{s}}\big \vert f \big \vert _{L^\infty }. \end{aligned}$$

The only nontrivial term to estimate is now \(\big \vert \frac{\zeta }{1-\zeta }\big \vert _{H^{s}}\). Using that \(h(\zeta )=1-\zeta \ge h_0>0\), we introduce a function \(F\in C^\infty (\mathbb {R})\) such that

$$\begin{aligned} F(X) \ = \ {\left\{ \begin{array}{ll} \frac{X}{1-X} &{} \text { if } 1-X\ge h>0,\\ 0 &{} \text { if } 1-X\le 0. \end{array}\right. } \end{aligned}$$

The function \(F\) satisfies the hypotheses of Lemma 5.1, and we have

$$\begin{aligned} \big \vert \frac{\zeta }{1-\zeta }\big \vert _{H^{s}} \ = \ \big \vert F(\zeta )\big \vert _{H^{s}} \ \le \ C(\big \vert \zeta \big \vert _{L^\infty },h_0^{-1})\big \vert \zeta \big \vert _{H^{s}}. \end{aligned}$$

The first estimate of the lemma is proved. The second estimate is obtained in the same way using

$$\begin{aligned} f-\frac{1}{h(\zeta )} f \ =\ - \frac{\zeta }{1-\zeta }f. \end{aligned}$$

The corollary is proved. \(\square \)

The following lemma presents a generalization of the Kato–Ponce (Kato and Ponce 1988) commutator estimates due to Lannes Lannes (2006) (one has \(\big \vert f\big \vert _{H^{s}}\) instead of \(\big \vert \partial _x f\big \vert _{H^{s-1}}\) in the standard Kato–Ponce estimate).

Lemma 5.3

(Commutator estimates) For any \(s\ge 0\) and \(\partial _x f, g\in L^\infty (\mathbb {R})\bigcap H^{s-1}(\mathbb {R})\) we have

$$\begin{aligned} \big \vert \ [\Lambda ^s,f]g\ \big \vert _{L^2}\ \lesssim \ \big \vert \ \partial _x f\ \big \vert _{H^{s-1}}\big \vert \ g\ \big \vert _{L^\infty }+\big \vert \ \partial _x f\ \big \vert _{L^\infty }\big \vert \ g\ \big \vert _{H^{s-1}}. \end{aligned}$$

Thanks to the continuous embedding of Sobolev spaces, we have for \(s\ge s_0+1, \ s_0>\frac{1}{2}\),

$$\begin{aligned} \big \vert \ [\Lambda ^s,f]g\ \big \vert _{L^2}\ \lesssim \ \big \vert \ \partial _x f\ \big \vert _{H^{s-1}}\big \vert \ g\ \big \vert _{H^{s-1}}. \end{aligned}$$

Let us now continue with the proof of Proposition 2.2. System (1.1) is quasilinear. In what follows we prove that it is Friedrichs-symmetrizable under conditions (2.2). We present below the symmetrizer of the system and compute the necessary energy estimates in Lemmata 5.5 and 5.6.

System symmetrizer. Recall that (1.1) reads \(\partial _t U \ + \ A[U]\partial _x U \ = \ 0\), with

$$\begin{aligned} A[U] \ \equiv \ \begin{pmatrix} u_1 &{}\quad \frac{ u_2-u_1}{\varrho } &{}\quad \frac{h_1}{\varrho } &{}\quad \frac{h_2}{\varrho } \\ 0 &{}\quad u_2 &{}\quad 0 &{}\quad h_2 \\ \frac{1}{\varrho } &{}\quad 0 &{}\quad u_1 &{}\quad 0 \\ \frac{\gamma }{\varrho } &{}\quad \delta +\gamma &{}\quad 0 &{}\quad u_2 \end{pmatrix}, \end{aligned}$$
(5.1)

where we use the notation \(h_1\equiv 1+\varrho \zeta _1-\zeta _2\) and \(h_2\equiv \delta ^{-1}+\zeta _2\). Define

$$\begin{aligned} S[U]\equiv \begin{pmatrix} \gamma &{}\quad 0&{}\quad 0 &{}\quad 0 \\ 0&{}\quad \gamma +\delta &{}\quad 0 &{}\quad u_2-u_1\\ 0 &{}\quad 0 &{}\quad \gamma h_1 &{}\quad 0 \\ 0&{}\quad u_2-u_1 &{}\quad 0 &{}\quad h_2 \end{pmatrix}. \end{aligned}$$
(5.2)

We can easily check that \(S[U]A[U]\equiv \Sigma [U]\) and \(S[U]\) are symmetric. More precisely, we have

$$\begin{aligned}&\Sigma [U]\nonumber \\&\quad \equiv \begin{pmatrix} \gamma u_1 &{}\quad \frac{\gamma (u_2-u_1)}{\varrho }&{}\quad \frac{\gamma h_1}{ \varrho }&{}\quad \frac{\gamma h_2}{ \varrho } \\ \frac{\gamma (u_2-u_1)}{\varrho } &{}\quad 2 (\gamma +\delta ) (2u_2-u_1)&{}\quad 0 &{}\quad (\gamma +\delta )h_2+ u_2(u_2-u_1)\\ \frac{\gamma h_1}{ \varrho }&{}\quad 0 &{}\quad \gamma h_1u_1 &{}\quad 0 \\ \frac{\gamma h_2}{\varrho } &{}\quad (\gamma +\delta )h_2+ u_2(u_2-u_1)&{}\quad 0 &{}\quad h_2 (2u_2-u_1) \end{pmatrix}.\nonumber \\ \end{aligned}$$
(5.3)

We can easily check that \(S[U]\) is positive definite provided that the following holds:

$$\begin{aligned} \gamma >0 \quad ; \quad \gamma +\delta >0\quad ; \quad h_1 > 0 \quad ; \quad h_2-\frac{|u_2-u_1|^2}{\gamma +\delta }> 0, \end{aligned}$$

which is guaranteed by condition (2.2).

Energy of our system. The natural energy of our system is

$$\begin{aligned} E^s(U)&\equiv \big ( S[\underline{U}] \Lambda ^s U,\Lambda ^s U\big )\nonumber \\&= \gamma \big \vert \zeta _1\big \vert _{H^s}^2 + (\gamma +\delta )\big \vert \zeta _2\big \vert _{H^s}^2+\gamma \int _\mathbb {R}\underline{h}_1 \big \vert \Lambda ^s u_1\big \vert ^2\nonumber \\&+\int _\mathbb {R}\underline{h}_2 \big \vert \Lambda ^s u_2\big \vert ^2 +2\int _\mathbb {R}(\underline{u}_2- \underline{u}_1) \big \{\Lambda ^su_2\big \}\big \{\Lambda ^s\zeta _2\big \}, \end{aligned}$$
(5.4)

with \(\underline{h}_1\equiv 1+ \varrho \underline{\zeta }_1-\underline{\zeta }_2\) and \(\underline{h}_2\equiv \delta ^{-1}+\underline{\zeta }_2 \).

In what follows, we specify the equivalence between our energy and the norm \(X^s\) offered by the well-posedness of the symmetrizer. Recall that \(X^s\) denotes the space \(H^s(\mathbb {R})^4\), endowed with the following norm:

$$\begin{aligned} \big \vert U\big \vert _{X^s}^2 \ = \ \gamma \big \vert \zeta _1\big \vert _{H^s}^2 \ + \ \big \vert \zeta _2\big \vert _{H^s}^2+\gamma \big \vert u_1\big \vert _{H^s}^2+\big \vert u_2\big \vert _{H^s}^2. \end{aligned}$$

Lemma 5.4

Let \(s\ge 0\) and \( \underline{\zeta }\in L^{\infty }(\mathbb {R})\), satisfying (2.2). Then \(E^s(U)\) is uniformly equivalent to the \(\vert \cdot \vert _{X^s}\)-norm. More precisely, there exists positive constants \(C_2=C(h_{0}^{-1},\delta _{\min }^{-1})>0\), and \(C_1=C(\big \vert \underline{h}_1\big \vert _{L^\infty },\big \vert \underline{h}_2\big \vert _{L^\infty },\delta _{\max })>0\) such that

$$\begin{aligned} \frac{1}{C_1}E^s(U) \ \le \ \big \vert U \big \vert _{X^s}^2 \ \le \ C_2 E^s(U). \end{aligned}$$

Proof

The fact that \(E^s(U) \ \le \ C_1 \big \vert U \big \vert _{X^s}\) is a simple consequence of the Cauchy–Schwarz inequality, applied to (5.4), where we use that (2.2) yields \(\vert \underline{u}_2-\underline{u}_1 \vert ^2 < (\gamma +\delta ) \underline{h}_2\).

The other inequality follows directly from (2.2). More precisely, we have

$$\begin{aligned} E^s(U) \ge&\,\, \gamma \big \vert \zeta _1\big \vert _{H^s}^2+\gamma h_0 \int _\mathbb {R}\big \vert \Lambda ^s u_1\big \vert ^2 \\&+ (\gamma +\delta )\big \vert \zeta _2\big \vert _{H^s}^2\!+\!\int _\mathbb {R}\underline{h}_2 \big \vert \Lambda ^s u_2\big \vert ^2 \!-2\int _\mathbb {R}\sqrt{(\underline{h}_2-h_0)(\gamma +\delta )} \big \{\Lambda ^su_2\big \}\big \{\Lambda ^s\zeta _2\big \}, \end{aligned}$$

and the result is now clear. Lemma 5.4 is proved. \(\square \)

We now highlight energy estimates with respect to the linearized system from (1.1), namely

$$\begin{aligned} \partial _t U \ + \ A[\underline{U}]\partial _x U \ = \ \mathcal {R}\ , \end{aligned}$$
(5.5)

with given \(\underline{U},\mathcal {R}\).

Lemma 5.5

(\(L^2\) energy estimate) Set \(T,M>0\). Let \(U\in L^\infty ( [0,T];X^0)\) satisfy (5.5), with given \(\mathcal {R}\in L^1([0,T];X^0)\), and \(\underline{U}\) satisfying (2.2), with \(h_0>0\) (for any \(t\in [0,T]\)) as well as

$$\begin{aligned} \big \Vert \underline{U} \big \Vert _{L^\infty ([0,T]\times \mathbb {R})^4}+\big \Vert \partial _x\underline{U} \big \Vert _{L^\infty ([0,T]\times \mathbb {R})^4}+\varrho \big \Vert \partial _t \underline{U} \big \Vert _{L^\infty ([0,T]\times \mathbb {R})^4} \ \le \ M. \end{aligned}$$

Then there exists \(C_0\equiv C(M,h_0^{-1},\delta _{\min }^{-1},\delta _{\max })\) such that \(\forall t\in [0,T],\)

$$\begin{aligned} E^0(U)(t)\le e^{C_0M\varrho ^{-1} t}E^0(U\left| _{\scriptstyle t=0}\right. )+\,\, C_0 \int ^{t}_{0} e^{C_0M\varrho ^{-1}( t-t')}\big \vert \mathcal {R}(t',\cdot )\big \vert _{X^s}\ dt'. \end{aligned}$$
(5.6)

Proof

Let us consider the \(L^2\) inner product of (5.5) and \( S[\underline{U}] U\):

$$\begin{aligned} \big (\partial _t U,S[\underline{U}] U\big ) \ + \ \big (A[\underline{U}]\partial _x U,S[\underline{U}] U\big ) \ = \ \big ( \mathcal {R},S[\underline{U}] U\big ). \end{aligned}$$

From the symmetry property of \(S[\underline{U}],\Sigma [\underline{U}]\), and using the definition of \(E^0(U)\), we deduce

$$\begin{aligned} \frac{1}{2} \frac{d}{dt}E^0(U) = \frac{1}{2}\big ( U,\big [\partial _t, S[\underline{U}]\big ] U\big )-\big (\Sigma [\underline{U}]\partial _xU, U\big ) + \big (\mathcal {R},S[\underline{U}] U\big )\nonumber \\&= \frac{1}{2}\big ( U,\big [\partial _t, S[\underline{U}]\big ] U\big )+\frac{1}{2}\big (\big [\partial _x,\Sigma [\underline{U}]\big ] U, U\big ) + \big (\mathcal {R},S[\underline{U}] U\big ). \end{aligned}$$
(5.7)

We now estimate each of the terms on the right-hand side of (5.7).

Estimate of \(\big ( U,\big [\partial _t, S[\underline{U}]\big ] U\big )\). We have \( \big ( U,\big [\partial _t, S[\underline{U}]\big ] U\big ) \ = \ \big ( U, \mathrm{d}S[\partial _t \underline{U}] U\big )\), with

$$\begin{aligned} \mathrm{d}S[\partial _t\underline{U}] \equiv \begin{pmatrix} 0 &{}0&{} 0 &{} 0 \\ 0&{} 0&{} 0 &{} \partial _t(\underline{u}_2-\underline{u}_1)\\ 0 &{} 0 &{}\gamma \partial _t (\varrho \underline{\zeta }_1-\underline{\zeta }_2) &{} 0 \\ 0&{} \partial _t(\underline{u}_2-\underline{u}_1) &{} 0 &{}\partial _t \underline{\zeta }_2 \end{pmatrix}. \end{aligned}$$

Using the Cauchy–Schwarz inequality and Lemma 5.4 we have straightforwardly

$$\begin{aligned} \big | \big ( U,\big [\partial _t, S[\underline{U}]\big ] U\big ) \big | \ \le \ C_0\big \vert \partial _t \underline{U}\big \vert _{L^\infty } C_2^{-1}\ \big \vert U \big \vert _{X^0}^2 \ \le \ C_0\ M\ \varrho ^{-1}\ E^0(U) , \end{aligned}$$
(5.8)

with \(C_0=C(h_0^{-1},\delta _{\min }^{-1},\delta _{\max })\).

Estimate of \(\big (\big [\partial _x,\Sigma [\underline{U}]\big ] U, U\big ) \). We have \(\big (\big [\partial _x,\Sigma [\underline{U}]\big ] U, U\big ) =\big ( U, \mathrm{d}\Sigma [\underline{U}] U\big )\), with

$$\begin{aligned} \mathrm{d}\Sigma [\underline{U}]\equiv \begin{pmatrix} \gamma \partial _x \underline{u}_1 &{}\frac{\gamma \partial _x(\underline{u}_2-\underline{u}_1)}{\varrho }&{} \frac{\gamma \partial _x(\varrho \underline{\zeta }_1-\underline{\zeta }_2)}{ \varrho }&{} \frac{\gamma \partial _x \underline{\zeta }_2}{ \varrho } \\ \frac{\gamma \partial _x(\underline{u}_2-\underline{u}_1)}{\varrho } &{} (\gamma +\delta )\partial _x(2 \underline{u}_2-\underline{u}_1)&{} 0 &{}\partial _x \big ((\gamma +\delta ) \underline{\zeta }_2+\underline{u}_2(\underline{u}_2-\underline{u}_1)\big )\\ \frac{\gamma \partial _x(\varrho \underline{\zeta }_1-\underline{\zeta }_2)}{ \varrho } &{} 0 &{} \gamma \partial _x (\underline{h}_1\underline{u}_1) &{} 0 \\ \frac{\gamma \partial _x \underline{\zeta }_2}{ \varrho } &{}\partial _x \big ((\gamma +\delta ) \underline{\zeta }_2+\underline{u}_2(\underline{u}_2-\underline{u}_1)\big ) &{} 0 &{}2\partial _x\big (\underline{h}_2(2 \underline{u}_2-\underline{u}_1)\big ) \end{pmatrix}. \end{aligned}$$

As previously, the Cauchy–Schwarz inequality and Lemmata 5.1 and 5.4 yield

$$\begin{aligned} \big |\big (\Sigma [\underline{U}]\partial _xU, U\big )\big | \ \le \ C_0\ M\ \varrho ^{-1}\ E^0(U) , \end{aligned}$$
(5.9)

with \(C_0=C(M,h_0^{-1},\delta _{\min }^{-1},\delta _{\max })\).

Estimate of \(\big (\mathcal {R},S[\underline{U}] U\big )\). By the Cauchy–Schwarz inequality and Lemmata 5.1 and 5.4,

$$\begin{aligned} \big | \big (\mathcal {R},S[\underline{U}] U\big ) \big | \ \le \ C_0\ \big \vert U \big \vert _{X^s} \big \vert \mathcal {R}\big \vert _{X^s} \ \le \ C_0' E^s(U)^{1/2} \big \vert \mathcal {R}\big \vert _{X^s} \ , \end{aligned}$$
(5.10)

with \(C_0,C_0'=C(M,h_0^{-1},\delta _{\min }^{-1},\delta _{\max })\).

Estimate (5.6) is now a consequence of the Gronwall–Bihari inequality applied to the differential inequality obtained when plugging (5.8), (5.9), (5.10) into (5.7). \(\square \)

Lemma 5.6

(\(H^s\) energy estimate) Set \(M,T>0\) and \(s\ge s_0+1, s_0>1/2\). Let \(U\in L^\infty ([0,T];X^s)\) satisfy (5.5), with \(\mathcal {R}\in L^1([0,T];X^s)\), and \(\underline{U}\in L^\infty ([0,T];X^s)\) satisfying (2.2) as well as

$$\begin{aligned} \big \Vert \underline{U} \big \Vert _{L^\infty ([0,T];X^s)}+\varrho \big \Vert \partial _t \underline{U} \big \Vert _{L^\infty ([0,T];X^{s-1} )} \ \le \ M. \end{aligned}$$

Then there exists \(C_0\equiv C(M,h_0^{-1},\delta _{\min }^{-1},\delta _{\max })\) such that, for all \(t \in [0, T]\),

$$\begin{aligned}&E^s(U)(t)\le e^{C_0M\varrho ^{-1}}E^s(U\left| _{\scriptstyle t=0}\right. ) + C_0 \int ^{t}_{0} e^{C_0M\varrho ^{-1}( t-t')}\big \vert \mathcal {R}(t',\cdot )\big \vert _{X^s}\ dt'.\quad \quad \quad \end{aligned}$$
(5.11)

Proof

As previously, we deduce from (5.5) the identity

$$\begin{aligned} \big (\Lambda ^s\partial _t U,S[\underline{U}] \Lambda ^s U\big ) \ + \ \big (\Lambda ^s A[\underline{U}]\partial _x U,S[\underline{U}] \Lambda ^s U\big ) \ = \ \big ( \Lambda ^s \mathcal {R},S[\underline{U}] \Lambda ^s U\big ) \ , \end{aligned}$$

where we recall the notation \(\Lambda \equiv ({{\mathrm{Id}}}-\partial _x^2)^{1/2}\). It follows that

$$\begin{aligned} & \frac{1}{2} \frac{d}{dt}E^s(U) = \frac{1}{2}\big ( \Lambda ^s U,\big [\partial _t, S[\underline{U}]\big ] \Lambda ^s U\big )-\big (S[\underline{U}]\Lambda ^sA[\underline{U}]\partial _xU, \Lambda ^s U\big ) \nonumber \\&\qquad +\,\, \big ( \Lambda ^s \mathcal {R},S[\underline{U}] \Lambda ^s U\big )\nonumber \\&= \frac{1}{2}\big ( \Lambda ^s U,\big [\partial _t, S[\underline{U}]\big ] \Lambda ^s U\big )+\frac{1}{2}\big (\big [\partial _x,\Sigma [\underline{U}]\big ] \Lambda ^s U, \Lambda ^s U\big )\nonumber \\&\qquad +\,\, \big ( \Lambda ^s \mathcal {R},S[\underline{U}] \Lambda ^s U\big )\nonumber \\&\qquad -\big (S[\underline{U}]\big [\Lambda ^s,A[\underline{U}]\big ]\partial _xU, \Lambda ^s U\big ). \end{aligned}$$
(5.12)

The first three terms are bounded exactly as previously when replacing \(U\) with \(\Lambda ^s U\). The only novelty lies in the use of continuous Sobolev embeddings, so that

$$\begin{aligned} \big \Vert \underline{U} \big \Vert _{L^\infty ([0,T]\times \mathbb {R})^4}+\big \Vert \partial _x\underline{U} \big \Vert _{L^\infty ([0,T]\times \mathbb {R})^4}\ \lesssim \ \big \Vert \underline{U}\big \Vert _{L^\infty ([0,T];X^{s})}. \end{aligned}$$

Similarly, we have

$$\begin{aligned} \varrho \big \Vert \partial _t \underline{U} \big \Vert _{L^\infty ([0,T]\times \mathbb {R})^4} \ \lesssim \ \varrho \big \Vert \partial _t \underline{U}\big \Vert _{L^\infty ([0,T];X^{s-1})}. \end{aligned}$$

The remaining term is estimated as follows. Using the commutator estimate in Lemma 5.3 we have

$$\begin{aligned} \big \vert \big [\Lambda ^s,A[\underline{U}]\big ]\partial _xU\big \vert _{L^2} \ \le \ C \big \vert \partial _x U\big \vert _{H^{s-1}} \big \vert \big [\partial _x, A[\underline{U}]\big ]\big \vert _{H^{s-1}}\ \le \ C_0\ M \ \varrho ^{-1}\ \big \vert U\big \vert _{X^{s}} , \end{aligned}$$

with \(C_0=C(M,h_0^{-1},\delta _{\min }^{-1},\delta _{\max })\). Altogether, we deduce from (5.12)

$$\begin{aligned} \frac{1}{2} \frac{d}{dt}E^s(U) \ \le \ C_0M \varrho ^{-1} E^s(U) \ + \ C_0 E^s(U)^{1/2} \big \vert \mathcal {R}\big \vert _{X^s}. \end{aligned}$$

Estimate (5.11) is now a consequence of the Gronwall–Bihari inequality, and the lemma is proved. \(\square \)

Completion of Proof of Proposition 2.2

The well-posedness of system (1.1) is now a consequence of the energy estimates of Lemmata 5.5 and 5.6, following the standard strategy (we refer the reader to standard textbooks, e.g., Taylor 1997; Alinhac and Gérard 1991; Métivier 2008, for more details). More precisely, we first show that the linearized problem (5.5) is well posed, then the solution of the nonlinear problem (1.1) is obtained as the limit of an iterative scheme:

$$\begin{aligned} \partial _t U^{n+1} \ + \ A[U^n]\partial _x U^{n+1} \ = \ 0. \end{aligned}$$

The restriction on the time scale \(t\in [0, T\varrho ]\) is necessary to guarantee that \((U^n)_{n\in \mathbb {N}}\) is a Cauchy sequence, and in particular that \(U^n\) is uniformly bounded with respect to \(n\), over a time domain which can be chosen independent of \(n\). The desired estimate on \(\big \vert U\big \vert _{X^s} \) follows directly from Lemma 5.6, with \(\underline{U}=U\) and \(R\equiv 0\), and the corresponding estimate on \(\big \vert \partial _t U\big \vert _{X^s} \) is then deduced using (1.1). The uniqueness comes from a similar estimate on the difference between two solutions, and the blow-up criterion as \(t\rightarrow T_{\max }\) if \(T_{\max }<\infty \) follows from standard continuation arguments. This concludes the proof of Proposition 2.2. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Duchêne, V. On the Rigid-Lid Approximation for Two Shallow Layers of Immiscible Fluids with Small Density Contrast. J Nonlinear Sci 24, 579–632 (2014). https://doi.org/10.1007/s00332-014-9200-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00332-014-9200-2

Keywords

Navigation