Skip to main content
Log in

Strange Attractors for Oberbeck–Boussinesq Model

  • Published:
Journal of Dynamics and Differential Equations Aims and scope Submit manuscript

Abstract

In this paper, we consider dynamics defined by the Navier–Stokes equations in the Oberbeck–Boussinesq approximation in a two dimensional domain. This model of fluid dynamics involves fundamental physical effects: convection, and diffusion. The main result is as follows: local semiflows, induced by this problem, can generate all possible structurally stable dynamics defined by \(C^1\) smooth vector fields on compact smooth manifolds (up to an orbital topological equivalency). To generate a prescribed dynamics, it is sufficient to adjust some parameters in the equations, namely, the viscosity coefficient, an external heat source, some parameters in boundary conditions and the small perturbation of the gravitational force.

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.

Similar content being viewed by others

References

  1. Arnold, V.I., Afrajmovich, V.S., Il’yashenko, Y.S., Shil’nikov, L.P.: Dynamical Systems V Bifurcation Theory and Catastrophe Theory. Encyclopedia of Mathematics. Springer, Berlin (1994)

    Google Scholar 

  2. Bates, P.W., Lu, K., Zeng, C.: Existence and persistence of invariant manifolds for semiflows in Banach space. Mem. Am. Math. Soc. 645, 1–129 (1998)

    MathSciNet  MATH  Google Scholar 

  3. Chow, S.N., Lu, K.: Invariant manifolds for flows in Banach spaces. J. Differ. Equ. 74, 285–317 (1998)

    Article  MathSciNet  Google Scholar 

  4. Chorin, A.J., Marsden, J.E.: A Mathematical Introduction to Fluid Mechanics, 2nd edn. Springer, New York (1984)

    MATH  Google Scholar 

  5. Dancer, E.N., Poláčik, P.: Realization of vector fields and dynamics of spatially homogeneous parabolic equations. Mem. Am. Math. Soc. 140(668), 121942708 (1999). https://doi.org/10.1090/MEMO/0668

  6. Drazin, P.G., Reid, W.H.: Hydrodynamical Stability. Cambridge University Press, Cambridge (1981)

    Google Scholar 

  7. Feireisl, E., Novotný, A.: The Oberbeck–Boussinesq approximation as a singular limit of the full Navier–Stokes Fourier system. J. Math. Fluid Mech. 11, 274–302 (2009)

    Article  MathSciNet  Google Scholar 

  8. Foias, C., Manley, O., Temam, R.: Attractors for the Bénard problem: existence and physical bounds on their fractal dimension. Nonlinear Anal TMA 11, 939–967 (1987)

    Article  Google Scholar 

  9. Fujita, H., Kato, T.: On the Navier–Stokes initial value problem I. Arch. Ration. Mech. Anal. 16, 269–315 (1964)

    Article  MathSciNet  Google Scholar 

  10. Fujita, H., Morimoto, H.: On fractional powers of the Stokes operator. Proc. Jpn. Acad. 46, 1141–1143 (1970)

    Article  MathSciNet  Google Scholar 

  11. Gershuni, G.Z., Zhukhovickij, E.M.: Convective Stability of Incompressible Fluids. Nauka Publishers, Moscow (1972). (in Russian)

    Google Scholar 

  12. Ghidaglia, J.-M.: On the fractal dimension of attractors for viscous incompressible fluid flows. SIAM J. Math. Anal. 17(5), 1139–1157 (1986)

    Article  MathSciNet  Google Scholar 

  13. Giga, Y.: Analyticity of the semigroup generated by the Stokes operator in \(L_r\) spaces. Math. Z. 178(3), 297–329 (1981)

    Article  MathSciNet  Google Scholar 

  14. Guckenheimer, J., Holmes, P.: Nonlinear Osscillations, Dynamical Systems, and Bifurcations of Vector Fields. Springer, New York (1981)

    Google Scholar 

  15. Henry, D.: Geometric Theory of Semilinear Parabolic Equations. Lecture Notes in Mathematics, vol. 840. Springer, Berlin (1981)

    Book  Google Scholar 

  16. Kato, T.: Perturbation Theory of Linear Operators. Spinger, Berlin (1980)

    MATH  Google Scholar 

  17. Katok, A.B., Hasselblatt, B.: Introduction to the Modern Theory of Dynamical Systems. Encyclopedia of Mathematics and its Applications, vol. 54. Cambridge University Press, Cambridge (1995)

    Book  Google Scholar 

  18. Korzuhin, M.D.: Ph.D. thesis, Moscow (1969)

  19. Mitrea, M., Monniaux, S.: The regularity of the Stokes operator and the Fujita–Kato approach to the Navier–Stokes initial value problem in Lipschitz domains. J. Funct. Anal. 254, 1522–1574 (2008)

    Article  MathSciNet  Google Scholar 

  20. Newhouse, R., Ruelle, D., Takens, F.: Occurence of strange axiom A attractors from quasi periodic flows. Commun. Math. Phys. 64, 35–40 (1971)

    Article  Google Scholar 

  21. Poláčik, P.: Realization of any finite jet in a scalar semilinear parabolic equation on the ball in \(R^2\). Annali Scuola Norm Pisa 17, 83–102 (1991)

    MATH  Google Scholar 

  22. Poláčik, P.: Complicated dynamics in scalar semilinear parabolic equations. Higher space dimensions. J. Differ. Equ. 89, 244–271 (1991)

    Article  MathSciNet  Google Scholar 

  23. Ruelle, D.: Elements of Differentiable Dynamics and Bifurcation Theory. Academic Press, Boston (1989)

    MATH  Google Scholar 

  24. Ruelle, D., Takens, F.: On the nature of turbulence. Commun. Math. Phys. 20(3), 167–192 (1971)

    Article  MathSciNet  Google Scholar 

  25. Shil’nikov, A.L., Shil’nikov, L.P., Turaev, D.V.: Normal forms and Lorenz attractor. Int. J. Bifurc. Chaos 3, 1123–1139 (1993)

    Article  MathSciNet  Google Scholar 

  26. Tang, D., Rankin III, S.M.: Peristaltic transport of a heat-conducting viscous fluid as an application of abstract differential equations and semigroup of operators. J. Math. Anal. Appl. 169, 391–407 (1992)

    Article  MathSciNet  Google Scholar 

  27. Temam, R.: Infinite Dimensional Dynamical Systems in Mechanics and Physics. Springer, New York (1988)

    Book  Google Scholar 

  28. Temam, R.: Navier–Stokes Equations Theory and Numerical Analysis, 3rd edn. North-Holland Publishing Company, Amsterdam (1984)

    MATH  Google Scholar 

  29. Temam, R.: Navier–Stokes Equations and Non-linear Functional Analysis. Society for Industrial and Applied Mathematics, vol. 66, 2nd edn (1987)

  30. Vakulenko, S.A.: Complex attractors and patterns in reaction–diffusion systems. J. Dyn. Differ. Equ. 30, 175–207 (2018)

    Article  MathSciNet  Google Scholar 

  31. Vakulenko, S.A., Sudakov, I.: Complex bifurcations in Bénard–Marangoni convection. J. Phys. A Math. Theor. 49, N42 (2016)

    Article  Google Scholar 

  32. Vakulenko, S., Grigoriev, D., Weber, A.: Reduction methods and chaos for quadratic systems of differential equations. Stud. Appl. Math. 135, 225–247 (2015)

    Article  MathSciNet  Google Scholar 

  33. Zhabotinsky, A.M.: Konzentrazionnie Avtokolebania (Oscillations of Concentrations). Nauka, Moscow (1974). (in Russian)

    Google Scholar 

Download references

Acknowledgements

The author is grateful to Referee for useful remarks that improved that paper. I am thankful to Prof. D. Ruelle, who many years ago encouraged me. This work was supported by the Ministry of Science and Higher Education of the Russian Federation by the Agreement 075-15-2020-933 dated November 13, 2020 on the provision of a grant in the form of subsidies from the federal budget for the implementation of state support for the establishment and development of the world-class scientific center Pavlov center Integrative physiology for medicine, hightech healthcare, and stress-resilience technologies.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Sergei Vakulenko.

Additional information

I dedicate this paper to the memory of my friends Prof. Vladimir Shelkovich, Prof. Andreas Weber, and Prof. A. Yu. Kazakov.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendices

1.1 Appendix 1

Lemma 13.1

Under the conditions (5.13) the form \(((, ))_2\) is positive definite on the space \(W^{1,2}(\Omega )\).

Proof

It is sufficient to verify that

$$\begin{aligned} ||\nabla u||^2 + I_{\beta }(u, u) \ge c ||\nabla u||^2, \end{aligned}$$
(13.1)

for a constant \(c >0\) and for each \(u \in H^1(\Omega )=W^{2,1}(\Omega )\). Here \(I_{\beta }(u, u)\) is defined by (5.12). To estimate the term \(I_{\beta }\) we use the inequality

$$\begin{aligned} |u(x,h)-u(x,0)|=\int _0^h u_y dy \le \sqrt{h} \Big (\int _0^h u_y(x,y)^2 dy\Big )^{1/2}. \end{aligned}$$

Consequently,

$$\begin{aligned} u^2(x,h) \le 2 \big (u^2(x,0) + h \int _0^h u_y(x,y)^2 dy\big ) \end{aligned}$$

and

$$\begin{aligned} I_{\beta }(u, u)\ge (\beta -2|\beta _1|) \int _0^{\pi } u^2(x,0) dx - 2h |\beta _1| \ || u_y||^2. \end{aligned}$$
(13.2)

The last inequality and (5.13) imply (13.1). \(\square \)

Proof of Proposition 5.1

I Conditions (5.3) and (5.4) are fulfilled for \(\theta _1=0\) (see [27], Ch. III, Sect. 3.5.1, page 136). Let us check (5.5). Let us denote \(z=(\mathbf{v}, u)\) and \(U(z)=a(z,z) +(Rz, z)\). In our case U(z) has the form

$$\begin{aligned} U(z)= \nu || \nabla \mathbf{v}||^2 + || \nabla u ||^2 + I_{\beta } (u) + \kappa \langle \mathbf{e} (1 +\gamma g_1 u), \mathbf{v} \rangle , \end{aligned}$$

where

$$\begin{aligned} \langle \mathbf{a}, \mathbf{b}\rangle =\int \int _{\Omega }( a_1(x,y) b_1(x,y) + a_2(x,y) b_2(x,y)) dx dy, \quad || \mathbf{v} ||^2 =\langle \mathbf{v}, \mathbf{v} \rangle . \end{aligned}$$

We should verify that

$$\begin{aligned} U(z) \ge c_0 ||| z|||^2 =c_0 (||\nabla \mathbf{v}||^2+ ||\nabla u||^2) \end{aligned}$$
(13.3)

for a constant \(c_0 >0\). The function \(g_1\) is bounded thus

$$\begin{aligned} U(z) \ge \nu ||\nabla \mathbf{v}||^2+ ||\nabla u||^2 + I_{\beta } (u) - C |\kappa | || u|| ||\mathbf{v}||, \end{aligned}$$

where \(C>0\) is a constant. For each \(\rho >0\) by the Poincaré inequality one has

$$\begin{aligned} U(z) \ge \nu ||\nabla \mathbf{v}||^2 + || \nabla u||^2 + I_{\beta } (u) - \frac{ C |\kappa |}{2} \Big (\rho c_1|| \nabla u||^2 + \rho ^{-1} ||\mathbf{v}||^2 \Big ), \quad c_1 >0.\qquad \end{aligned}$$
(13.4)

Lemma 13.1 and estimate (13.4) proves (13.3) for sufficiently small \(|\kappa |\).

As it was pointed out by Referee, the general case of arbitrary \(\kappa \) can be reduced to that case of small \(\kappa \) by a temperature scaling \(u =\epsilon {\bar{u}}\) with a small \(\epsilon >0\). Indeed, then for new variables \((\mathbf{v}, {\bar{u}})\) we have a small \(|\kappa |\). It completes the proof of (13.3).

II Checking condition (5.6) Let us check that under our choice of operators \(A_i\) and quadratic forms \(B_i\) the key condition (5.6) is fulfilled (all other conditions on B can be checked as in [27], Ch. III, Sect. 3.5.1). Let \(v=(\mathbf{v}, u)\), \(w=({ \mathbf{w}}, {\tilde{u}})\), where all \(\mathbf{v}, \mathbf{w}, u, {\tilde{u}}\) are sufficiently smooth. One has

$$\begin{aligned} (B (v, w), w)=\nu I_1(\mathbf{v} , \mathbf{w}) + I_2(\mathbf{v}, {\tilde{u}}), \end{aligned}$$

where

$$\begin{aligned} I_1(\mathbf{v} , \mathbf{w})= & {} \int \int _{\Omega } [\mathbf{v} \cdot \nabla \mathbf{w}] \mathbf{w} dxdy,\\ I_2(\mathbf{v} , {\tilde{u}})= & {} \int \int _{\Omega } [\mathbf{v} \cdot \nabla {\tilde{u}}] {\tilde{u}} dxdy. \end{aligned}$$

Integrating by parts, we obtain

$$\begin{aligned} I_1(\mathbf{v} , \mathbf{w})= S_1 + S_2, \end{aligned}$$
(13.5)

where

$$\begin{aligned} S_1= & {} \frac{1}{2} \int _0^h v_1(x,y) \big (w_1^2+ w_2^2\big )(x,y) dy \Big \vert _{x=0}^{x=\pi },\\ S_2= & {} \frac{1}{2} \int _0^{\pi } v_2(x,y) \big (w_1^2+ w_2^2\big )(x,y) dx \Big \vert _{y=0}^{y=h}. \end{aligned}$$

The term \(S_1\) equals zero because \(v_1(x, y)\equiv 0\) at \(x=0\) and \(x=\pi \) due to the boundary condition (2.7). The second term \(S_2\) equals zero according to the boundary condition (2.8), so we conclude that \(I_1=0\). Similarly, one can show that \(I_2=0\) and therefore (5.6) holds. \(\square \)

1.2 Appendix 2

Proof of Lemma 8.1

This assertion is, for example, a consequence of Theorem 6.1.7 [15] (however, we could also use results [2] or [3]). In the variables \({\hat{z}}=(\hat{\mathbf{v}}, {\hat{w}})^{tr}\), X system (8.12), (8.13), (8.14) can be rewritten as

$$\begin{aligned} X_t= & {} \gamma {\hat{F}}(X, {\hat{z}}), \end{aligned}$$
(13.6)
$$\begin{aligned} {\hat{z}}_t= & {} L {\hat{z}} + {\hat{G}}(X, {\hat{z}}). \end{aligned}$$
(13.7)

Using the standard truncation trick we modify Eq. (13.6) as follows:

$$\begin{aligned} X_t= \gamma {\hat{F}}(x, {\tilde{w}})\chi _{R_0}(X), \end{aligned}$$
(13.8)

where \(\chi _{R_{0}}\) is a smooth function such that \(\chi _{R_0}(X) =1 \) for \(|X| < R_0\) and \(\chi _{R_0}(X) =0 \) for \(|X| > 2R_0\). As a result of that modification, X-trajectories of (13.8) are defined for all \(t \in (-\infty , +\infty )\) (as in Theorem 6.1.7 [15]). Then an invariant manifold for the semiflow defined by system (13.8), (13.7) is a locally invariant one for the semiflow generated by (13.6), (13.7).

For \(\alpha \in [0,1]\) let us denote by \(H_{\alpha }\) the fractional spaces (see [10] for details) associated with Stokes operator

$$\begin{aligned} H_{\alpha }= \{ \mathbf{v} \in D( (-\Delta _D)^{\alpha }\}), \quad div \mathbf{v}=0, \quad \mathbf{v}\cdot \mathbf{n} \big \vert _{\partial \Omega }=0 \}, \end{aligned}$$
(13.9)

where D(A) denotes the domain of the operator A, \(\Delta _D\) is the Laplace operator with the domain corresponding to our boundary conditions for \(\mathbf{v}\) and \(\mathbf{n}\) is a normal vector to the boundary \(\partial \Omega \). Let \({\tilde{H}}_{\alpha }\) be another fractional space associated with \(L_2(\Omega )\):

$$\begin{aligned} {\tilde{H}}_{\alpha }= \{ u \in L_2(\Omega ): ||u||_{\alpha } =||(I-\Delta _N)^{\alpha } u|| < \infty \}, \end{aligned}$$
(13.10)

where \(\Delta _N\) is the Laplace operator with the domain corresponding to the boundary conditions (2.5), (2.6).

We use the Sobolev embeddings

$$\begin{aligned} H_{\alpha } \subset L_{\infty } (\Omega ), \quad \alpha >1/2, \end{aligned}$$
(13.11)

and analogous embeddings for \({\tilde{H}}_{\alpha }\). The Sobolev embeddings and estimates for the Stokes operator in the Lipshitz domains \(\Omega \) were considered, for example, in [19]. In our case of boundary conditions (2.7) and (2.8) the spectral problem for the Stokes operator can be resolved by the stream function and variable separation (as well as for periodical boundary conditions), and embedding (13.11) for \(H_{\alpha }\) can be proved in a standard way.

By (13.11) we observe that

$$\begin{aligned} || (\mathbf{v} \cdot \nabla ) \mathbf{v}|| \le |\mathbf{v}|_{\infty } ||\mathbf{v}||_{\alpha } \end{aligned}$$
(13.12)

where \(\alpha \in (1/2, 1)\)

$$\begin{aligned} |\mathbf{v}|_{\infty }=||v_1||_{L_{\infty }(\Omega )} + ||v_2||_{L_{\infty }(\Omega )}. \end{aligned}$$

Estimate (13.12) and an analogous estimate for \(|| (\mathbf{v} \cdot \nabla ) u||\) show that for \(\alpha >1/2\) the map \({\hat{G}}\) is a bounded \(C^{\infty }\)- map from a bounded domain in \({\mathcal H}_{\alpha }=H_{\alpha } \times {\tilde{H}}_{\alpha }\) to \({\mathcal {H}}\) [15]. It allows us to apply results on existence of smooth invariant manifolds [2, 3, 15]. We will check conditions of Theorem 6.1.7 from [15].

Let us consider the semigroup \(\exp (Lt)\). We have estimates (7.128), (7.129), where \(M, {\bar{M}}, \rho >0\) do not depend on \(\gamma \). Moreover,

$$\begin{aligned} M_0= & {} \gamma \sup _{(X, {\hat{z}}) \in {{{\mathcal {D}}}_{\gamma , 2R_0, C_1, C_2, \alpha }}} || {\hat{F}} \chi _{R_0} || < c_2\gamma , \end{aligned}$$
(13.13)
$$\begin{aligned} \lambda= & {} \gamma \sup _{(X, {\hat{z}}) \in {{{\mathcal {D}}}_{\gamma , 2R_0, C_1, C_2, \alpha }}} ||D_X {\hat{F}} \chi _{R_0}|| + ||D_{{\hat{z}}} {\hat{F}} \chi _{R_0}|| < c_3\gamma , \end{aligned}$$
(13.14)
$$\begin{aligned} M_2= & {} \gamma \sup _{(X, {\tilde{w}}) \in {{{\mathcal {D}}}_{\gamma , 2R_0, C_1, C_2, \alpha }}} ||D_{{\hat{z}}} {\hat{G}}|| < c_4\gamma , \end{aligned}$$
(13.15)

We set \(\mu _0=\kappa /4\). Then for small \(\gamma \)

$$\begin{aligned} M_3=\gamma \sup _{(X, {\hat{z}}) \in {{{\mathcal {D}}}_{\gamma , 2R_0, C_1, C_2, \alpha }} } ||D_{X} {\hat{G}}|| < c_5\gamma . \end{aligned}$$
(13.16)

We set \(\delta _1=2\theta _1\), where

$$\begin{aligned} \theta _p =\lambda M_0 \int _0^{\infty } u^{-\alpha } \exp (-(\kappa - p\mu ^{\prime }) u) du, \quad 1 \le p \le 1+\delta , \end{aligned}$$
(13.17)

and \(\mu ^{\prime }= \mu _0 + \delta _1 M_2\). For sufficiently small \(\gamma \) one has \(\mu ^{\prime } <\rho /2\), therefore, the integral in the right hand side of (13.17) converges and, according to (13.15), one obtains \(\theta < c_6 \gamma \) (since M is independent of \(\gamma \)). We notice then that for sufficiently small \(\gamma \) the following estimates

$$\begin{aligned}&(1+\delta ) \mu ^{\prime }< \rho /2,\\&\theta _1< \delta _1(1 + \delta _1)^{-1}< 1, \quad \theta _1(1+\delta _1)M_2{\mu ^{\prime }}^{-1} < 1, \end{aligned}$$

and

$$\begin{aligned} \theta _p(1 + \frac{(1+\delta _1)M_2}{r \mu ^{\prime }}) < 1 \end{aligned}$$

hold. Those estimates show that all conditions of Theorem 6.1.7 [15] are satisfied, and Lemma 8.1 is proved. \(\square \)

1.3 Appendix 3

Proof of Lemma 7.3

Let us prove that

$$\begin{aligned} |G_{k}(y, y_0) \!-\! {\bar{G}}_{ k}(y, y_0)| \!<\! C_0|k|^{-3}\!\!( \exp (- c_0 k (|h-y| \!+\! |h-y_0|)) \!\!+\!\! \exp (- c_0 (k (y \!+\! |h\!-\!y_0|) ),\nonumber \\ \end{aligned}$$
(13.18)

where \(C_0, c_0, C_1 >0\) are constants and \({\bar{G}}_{ k}\) is defined by

$$\begin{aligned} {\bar{G}}_k = k^{-3} \frac{1}{2\tau (1-\tau ^2)} \big (\tau \exp (-k|y-y_0|) - \exp (-\tau k |y-y_0|)\big ). \end{aligned}$$
(13.19)

Let us make substitution \(z=ky\). Then \(G_k= k^{-3} H(ky, ky_0)\), where \(H(z, z_0)\) is the Green function defined by the following boundary value problem:

$$\begin{aligned}&{[(D_z^2-1)^2 - {\bar{\lambda }} (D_z^2 -1)]} H = \delta (z-z_0), \quad z, z_0 \in [0, kh], \end{aligned}$$
(13.20)
$$\begin{aligned}&H (0, z_0)=H(kh, z_0)=0, \quad \frac{\partial H(z,z_0)}{\partial z} \Big \vert _{z=0, kh}=0, \end{aligned}$$
(13.21)

where \(D_z=\frac{\partial }{\partial z}\).

Let us define the auxiliary function \({\bar{H}}\) as follows. Let \(\tau =\sqrt{1 + {\bar{\lambda }}}\) and \(Re \ \tau \ge 0\). Note that then for \(Re \lambda > -1/2\) one has \(Re \ \tau > 1/2\) (since \(\nu>>1\) and \(k \ge 1\)). One has the following relations:

$$\begin{aligned} {\bar{H}}(z, z_0)=\frac{1}{2\tau (1-\tau ^2)} \big (\tau \exp (-|z-z_0|) - \exp (-\tau |z-z_0|)\big ) \end{aligned}$$

for \(\tau \ne 1\), and

$$\begin{aligned} {\bar{H}}(z, z_0) = \frac{1}{4} \exp (-|z-z_0|) ( 1 + |z-z_0|) \end{aligned}$$

for \(\tau =1\). For all \(\tau \) such that \(Re \ \tau >1/2\) the function H satisfies Eq. (13.20) and H is smooth in \(\tau \). We are seeking for H in the form

$$\begin{aligned} H= {\tilde{H}}_1(z)+ \xi {\tilde{H}}_2(z) + {\bar{H}}(z, z_0), \quad \xi =(\tau -1)^{-1}, \end{aligned}$$

where

$$\begin{aligned}&{\tilde{H}}_1=C_1 \exp (-z) + C_3 \exp (z-kh),\\&{\tilde{H}}_2= C_2 (\exp (-\tau z) -\exp (-z)) + C_4 (\exp (\tau (z- kh)) - \exp (z-kh)), \end{aligned}$$

and \(C_i\) are unknown coefficients. Note that \({\tilde{H}}\) is a smooth in z and \(\tau \). Then for \(C=(C_1, C_2, C_3, C_4)^{tr}\) one obtains the linear system

$$\begin{aligned} D C= B, \end{aligned}$$
(13.22)

where D is a \(4\times 4\) matrix and \(B=({\bar{H}}(0,z_0), \bar{H}_z(0, z_0), {\bar{H}}(kh,z_0), {\bar{H}}_z(kh, z_0))^{tr}\). For large h the matrix D is an exponentially small perturbation of a block diagonal matrix

$$\begin{aligned} D= \begin{bmatrix} 1 &{}\quad 0 &{}\quad \exp (-kh) &{}\quad \xi (\exp (- \tau kh) - \exp (-kh)) \\ -1 &{}\quad -1 &{}\quad -\exp (-kh) &{}\quad \xi (\exp (- kh) - \tau \exp (-\tau kh)) \\ \exp (-kh) &{}\quad \xi (\exp (- \tau kh) - \exp (-kh)) &{}\quad 1 &{}\quad 0 \\ \exp (- kh) &{}\quad \xi (\tau \exp (- \tau kh) - \exp (-\tau kh)) &{}\quad 1 &{}\quad 1 \end{bmatrix}. \end{aligned}$$

For \(h>>1\) this matrix has a non-zero determinant since \(|Det \ D + 1|= kh \big (O(\exp -kh)) + O(\exp (-\tau kh)) \big )\), and coefficients C can be found by iterations that immediately gives us (13.18). Returning to the variable y we obtain the conclusion of Lemma 7.3. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Vakulenko, S. Strange Attractors for Oberbeck–Boussinesq Model. J Dyn Diff Equat 33, 303–343 (2021). https://doi.org/10.1007/s10884-020-09939-z

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10884-020-09939-z

Keywords

Navigation