Skip to main content
Log in

Strong \(L^2\) convergence of time Euler schemes for stochastic 3D Brinkman–Forchheimer–Navier–Stokes equations

  • Published:
Stochastics and Partial Differential Equations: Analysis and Computations Aims and scope Submit manuscript

Abstract

We prove that some time Euler schemes for the 3D Navier–Stokes equations modified by adding a Brinkman–Forchheimer term and a random perturbation converge in \(L^2(\varOmega )\). This extends previous results concerning the strong rate of convergence of some time discretization schemes for the 2D Navier Stokes equations. Unlike the 2D case, our proposed 3D model with the Brinkman–Forchheimer term allows for a strong rate of convergence of order almost 1/2, that is independent of the viscosity parameter.

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.

We’re sorry, something doesn't seem to be working properly.

Please try refreshing the page. If that doesn't work, please contact support so we can address the problem.

Availability of data and materials

Data sharing not applicable to this article as no datasets were generated or analysed during the current study.

References

  1. Adams, R.A.: Sobolev Spaces. Academic Press, New York (1975)

    MATH  Google Scholar 

  2. Barret, J.W., Liu, W.B.: Finite elements approximations for the parabolic \(p\)-Laplacian. SIAM J. Numer. Anal. 31, 413–428 (1994)

    Article  MathSciNet  Google Scholar 

  3. Bensoussan A.: Some existence results for stochastic partial differential equations. Pitman Res. Notes Math. Ser., vol. 268, Longman Sci. Tech., Harlow, Trento, pp. 37–53 (1990)

  4. Bensoussan, A., Glowinski, R., Rascanu, A.: Approximation of some stochastic differential equations by splitting up method. Appl. Math. Optim. 25, 81–106 (1992)

    Article  MathSciNet  Google Scholar 

  5. Bessaih, H., Brzeźniak, Z., Millet, A.: Splitting up method for the 2D stochastic Navier–Stokes equations. Stoch. PDE Anal. Comput. 2–4, 433–470 (2014)

    Article  MathSciNet  Google Scholar 

  6. Bessaih, H., Millet, A.: On stochastic modified 3D Navier–Stokes equations with anisotropic viscosity. J. Math. Anal. Appl. 462, 915–956 (2018)

    Article  MathSciNet  Google Scholar 

  7. Bessaih, H., Millet, A.: Strong \(L^2\) convergence of time numerical schemes for the stochastic two-dimensional Navier–Stokes equations. IMA J. Numer. Anal. 39–4, 2135–2167 (2019)

    Article  Google Scholar 

  8. Bessaih, H., Millet, A.: Space–time Euler discretization schemes for the stochastic 2D Navier–Stokes equations. Stoch. PDE Anal. Comput. (2021). https://doi.org/10.1007/s40072-021-00217-7

  9. Bessaih, H., Millet, A.: Strong rates of convergence of space-time discretization schemes for the 2D Navier–Stokes equations with additive noise. Stoch. Dyn. (2022). https://doi.org/10.1142/S0219493722400056

  10. Bessaih, H., Trabelsi, S., Zorgati, H.: Existence and uniqueness of global solutions for the modified anisotropic 3D Navier–Stokes equations. M2AN 50, 1817–1823 (2016)

    Article  MathSciNet  Google Scholar 

  11. Breckner, H.: Galerkin approximation and the strong solution of the Navier–Stokes equation. J. Appl. Math. Stochastic Anal. 13–3, 239–259 (2000)

    Article  MathSciNet  Google Scholar 

  12. Brzeźniak, Z., Carelli, E., Prohl, A.: Finite element base discretizations of the incompressible Navier–Stokes equations with multiplicative random forcing. IMA J. Numer. Anal. 33–3, 771–824 (2013)

    Article  Google Scholar 

  13. Carelli, E., Prohl, A.: Rates of convergence for discretizations of the stochastic incompressible Navier–Stokes equations. SIAM J. Numer. Anal. 50–5, 2467–2496 (2012)

    Article  MathSciNet  Google Scholar 

  14. Chemin, J.-Y., Desjardin, B.,Gallagher, I., Grenier, E.: Mathematical Geophysics: An Introduction to Rotating Fluids and the Navier–Stokes Equations. Oxford Lecture Series in Mathematics and its Applications, p. 32 (2006)

  15. Chueshov, I., Millet, A.: Stochastic 2D hydrodynamical type systems: Well posedness and large deviations. Appl. Math. Optim. 61–3, 379–420 (2010)

    Article  MathSciNet  Google Scholar 

  16. Da Prato, G., Zabczyk, J.: Stochastic Equations in infinite Dimensions. Cambridge University Press, Cambridge (1992)

  17. Dörsek, P.: Semigroup splitting and cubature approximations for the stochastic Navier–Stokes Equations. SIAM J. Numer. Anal. 50–2, 729–746 (2012)

    Article  MathSciNet  Google Scholar 

  18. Flandoli, F.: A stochastic view over the open problem of well-posedness for the 3D Navier–Stokes equations. Stochastic analysis: a series of lectures. Progr. Probab., vol. 68. Birkhäuser/Springer, Basel, pp. 221–246 (2015)

  19. Flandoli, F., Gatarek, D.: Martingale and stationary solutions for stochastic Navier–Stokes equations. Probab. Theory Relat. Fields 102, 367–391 (1995)

    Article  MathSciNet  Google Scholar 

  20. Flandoli, F., Mahalov, A.: Stochastic three-dimensional rotating Navier–Stokes equations: averaging, convergence and regularity. Arch. Ration. Mech. Anal. 205–1, 195–237 (2012)

    Article  MathSciNet  Google Scholar 

  21. Giga, Y., Miyakawa, T.: Solutions in \(L_r\) of the Navier–Stokes initial value problem. Arch. Ration. Mech. Anal. 89–3, 267–281 (1985)

    Article  Google Scholar 

  22. Girault, V., Raviart, P.A.: Finite element method for Navier–Stokes equations: theory and algorithms. Springer, Berlin (1981)

    MATH  Google Scholar 

  23. Hutzenthaler, M., Jentzen, A.: Numerical approximations of stochastic differential equations with non-globally Lipschitz continuous coefficients. Mem. Amer. Math. Soc. 236, 1112 (2015)

  24. Kunita, H.: Stochastic Flows and Stochastic Differential Equations. Cambridge University Press, Cambridge (1990)

    MATH  Google Scholar 

  25. Mohan, M.: Stochastic convective Brinkman–Forchheimer equations. arXiv:2007.09376 (2020)

  26. Printems, J.: On the discretization in time of parabolic stochastic partial differential equations. M2AN Math. Model. Numer. Anal. 35–6, 1055–1078 (2001)

    Article  MathSciNet  Google Scholar 

  27. Temam, R.: Navier–Stokes Equations. Theory and Numerical Analysis. Studies in Mathematics and Its Applications, vol. 2. North-Holland, Amsterdam (1979)

  28. Temam, R.: Navier–Stokes Equations and Nonlinear Functional Analysis, CBMS-NSF Regional Conference Series in Applied Mathematics. Society for Industrial and Applied Mathematics (SIAM), Philadelphia (1995)

    Google Scholar 

Download references

Acknowledgements

The authors want to thank an anonymous referee for a very careful reading and valuable comments, which enabled them to improve the presentation of the paper. Annie Millet’s research has been conducted within the FP2M federation (CNRS FR 2036). This research was started while both authors stayed at the Mathematisches Forschung Institute Oberwolfach during a Research in Pairs program. They want to thank the MFO for the financial support and excellent working conditions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Annie Millet.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Additional information

This article is dedicated to István Gyöngy on the occasion of his 70th birthday.

Publisher's Note

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

Hakima Bessaih was partially supported by Simons Foundation Grant: 582264 and NSF Grant DMS: 2147189.

Appendix

Appendix

In this section, we provide the proof of the well-posedness result stated in Sect. 3.

1.1 Proofs of preliminary estimates

The following results gather some estimates of the bilinear term, and more generally of the non linear part in (1.1). They are deduced from the Brinkman–Forchheimer smoothing term. The proofs are somewhat similar to the corresponding ones in [6] in a different functional setting.

The next lemma gathers further properties of B.

Lemma 3

Suppose that \(\alpha \in [1,+\infty )\).

  1. (i)

    Let \(u\in L^\infty (0,T;H)\cap L^{2\alpha +2}([0,T]\times D;{{\mathbb {R}}}^3)\), \(v\in X_0\). Then

    $$\begin{aligned}&\int _0^T \! \big | \langle B(u(t), u(t)) , v(t)\rangle \big | dt \nonumber \\&\quad \le \Vert \nabla v\Vert _{L^2(0,T;{{\mathbb {L}}}^2)} \; \underset{{t\in [0,T]}}{\mathrm{ess \,sup }} \Vert u(t)\Vert _{{{\mathbb {L}}}^2}^{\frac{\alpha -1}{\alpha }}\; \Vert u\Vert _{L^{2\alpha +2}([0,T]\times D;{{\mathbb {R}}}^3)}^{\frac{\alpha +1}{\alpha }} \; T^{\frac{\alpha -1}{2\alpha }}. \end{aligned}$$
    (7.1)
    $$\begin{aligned}&\int _0^T\! \big | \langle B(u(t), u(t)) - B(v(t),v(t)) , u(t)- v(t)\rangle \big | dt \le \Vert \nabla v\Vert _{L^2(0,T;H)} \nonumber \\&\quad \times \; \underset{{t\in [0,T]}}{\mathrm{ess\, sup }} \Vert (u-v)(t)\Vert _{H}^{\frac{\alpha -1}{\alpha }}\; \Vert u-v\Vert _{L^{2\alpha +2}([0,T]\times D;{{\mathbb {R}}}^3)}^{\frac{\alpha +1}{\alpha }} \; T^{\frac{\alpha -1}{2\alpha }}. \end{aligned}$$
    (7.2)
  2. (ii)

    Let \(u\in L^4(\varOmega ; L^\infty (0,T;H))\cap L^{2\alpha +2}(\varOmega _T\times D;{{\mathbb {R}}}^3)\) and \(v\in {{\mathcal {X}}}_0\). Then

    $$\begin{aligned}&{{\mathbb {E}}}\int _0^T \! \big | \langle B(u(t), u(t)) , v(t)\rangle \big | dt \le \Big \{ {{\mathbb {E}}}\Big | \int _0^T \Vert \nabla v(t)\Vert _{{{\mathbb {L}}}^2}^2 dt \Big |^2 \Big \}^{\frac{1}{4}} \nonumber \\&\quad \times \Big \{ {{\mathbb {E}}}\Big ( \underset{{t\in [0,T]}}{\mathrm{ess\, sup }} \Vert u(t)\Vert _{H}^{4}\Big ) \Big \}^{ \frac{\alpha -1}{4\alpha } } \Big \{ {{\mathbb {E}}}\int _0^T \!\! dt \int _{D} \! |u(t,x)|^{2\alpha +2} dx \Big \}^{\frac{1}{2\alpha }} \; T^{\frac{\alpha -1}{2\alpha }}, \end{aligned}$$
    (7.3)
    $$\begin{aligned}&{{\mathbb {E}}}\int _0^T\! \big | \langle B(u(t), u(t)) - B(v(t),v(t)) , u(t)- v(t)\rangle \big | dt \nonumber \\&\quad \le T^{\frac{\alpha -1}{2\alpha }} \, \Big \{ {{\mathbb {E}}}\Big | \int _0^T \! \Vert \nabla v(t)\Vert _{{{\mathbb {L}}}^2}^2 dt \Big |^2 \Big \}^{\frac{1}{4}} \Big \{ {{\mathbb {E}}}\Big ( \underset{{t\in [0,T]}}{\mathrm{ess\, sup }} \Vert (u-v)(t)\Vert _{H}^{4}\Big ) \Big \}^{ \frac{\alpha -1}{4\alpha } } \nonumber \\&\qquad \times \; \Big \{ {{\mathbb {E}}}\! \int _0^T \!\! dt\! \int _{D} \! |(u-v)(t,x)|^{2\alpha +2} dx \Big \}^{\frac{1}{2\alpha }} . \end{aligned}$$
    (7.4)

Proof

  1. (i)

    Suppose \(\alpha >1\). Using (2.4) with \(h=\partial _i v_j\), \(f=u_i\) and \(g=u_j\), we deduce

    $$\begin{aligned} |\langle B(u,u),v\rangle |&= |-\langle B(u,v),u\rangle | \le \sum _{i,j=1}^3 \int _{D} |u_i(x) \partial _i v_j(x) u_j(x)| dx \\&\le \big \Vert |u| \, |u|^{\frac{1}{\alpha }}\big \Vert _{{{\mathbb {L}}}^{2\alpha }} \; \big \Vert |u|^{1-\frac{1}{\alpha }} \big \Vert _{{{\mathbb {L}}}^{\frac{2\alpha }{\alpha -1}}}\; \Vert \nabla v\Vert _{{{\mathbb {L}}}^2}. \end{aligned}$$

    Integrating on the time interval [0, T] and using the Cauchy–Schwarz inequality, we obtain

    $$\begin{aligned} \int _0^T \! \big | \langle B(u(t), u(t)) , v(t)\rangle \big | dt&\le \; \underset{{t\in [0,T]}}{\mathrm{ess\, sup }} \Vert u(t)\Vert _{H}^{\frac{\alpha -1}{\alpha }}\; \Big ( \int _0^T\!\! \Vert u(t)\Vert _{{{\mathbb {L}}}^{2\alpha +2}}^{\frac{2\alpha +2}{\alpha }} dt \Big )^{\frac{1}{2}} \\&\quad \; \times \Big ( \int _0^T \!\! \Vert \nabla v(t)\Vert _{{{\mathbb {L}}}^2}^2 dt \Big )^{\frac{1}{2}}. \end{aligned}$$

    Hölder’s inequality implies

    $$\begin{aligned} \int _0^T\!\! \Vert u(t)\Vert _{{{\mathbb {L}}}^{2\alpha +2}}^{\frac{2\alpha +2}{\alpha }} dt \le \Vert u\Vert _{L^{2\alpha +2}([0,T]\times D;{{\mathbb {R}}}^3)}^{\frac{2\alpha +2}{\alpha }}\; T^{\frac{\alpha -1}{\alpha }}. \end{aligned}$$

    This completes the proof of (7.1) for \(\alpha >1\).

    If \(\alpha =1\), since \( |\langle B(u,u),v\rangle | \le \big \Vert u\big \Vert _{{{\mathbb {L}}}^4}^2 \; \Vert \nabla v\Vert _{{{\mathbb {L}}}^2}\), a straightforward computation implies (7.1).

    Since \(\langle B(u,u)-B(v,v) \, , \, u-v\rangle = \langle B(u-v,v)\, , \, u-v\rangle \), using the antisymmetry (2.1) it is easy to see that the upper estimate (7.1) implies (7.2).

  2. (ii)

    For \(\alpha>1>\frac{2}{3}\), we have \(\frac{4\alpha }{3\alpha -2}>1\). Using Hölder’s inequality for the expected value with exponents 4, \(\frac{4\alpha }{3\alpha -2}\) and \(2\alpha \) in (7.1), we deduce

    $$\begin{aligned} {{\mathbb {E}}}\int _0^T \! \big | \langle B(u(t), u(t)) , v(t)\rangle \big | dt&\le \Big \{ {{\mathbb {E}}}\Big ( \Vert \nabla v\Vert _{L^2(0,T;{{\mathbb {L}}}^2)}^4 \Big )\Big \} ^{\frac{1}{4 }} \\&\quad \times \Big \{ {{\mathbb {E}}}\Big ( \underset{{t\in [0,T]}}{\mathrm{ess sup }} \Vert u(t)\Vert _{{{\mathbb {L}}}^2}^{\frac{4(\alpha -1)}{3\alpha -2}} \Big ) \Big \}^{\frac{3\alpha -2}{4\alpha } } \\&\quad \times \Big \{ E \int _0^T dt \int _{D} |u(t,x)|^{2\alpha +2} \, dx \Big \}^{\frac{1}{2\alpha }}\; T^{\frac{\alpha -1}{2\alpha }} . \end{aligned}$$

    Since \(\alpha >\frac{1}{2}\) we have \(\frac{4(\alpha -1)}{3\alpha -2} < 4\); this completes the proof of (7.3) for \(\alpha >1\).

For \(\alpha =1\), using the antisymmetry (2.1), and twice the Cauchy–Schwarz inequality, we deduce

$$\begin{aligned}&{{\mathbb {E}}}\int _0^T \! \big | \langle B(u(t), u(t)) , v(t)\rangle \big | dt \le \Big \{ {{\mathbb {E}}}\int _0^T \Vert \nabla v(t)\Vert _{{{\mathbb {L}}}^2}^2dt \Big \}^{\frac{1}{2}} \Big \{ {{\mathbb {E}}}\int _0^T \Vert u(t)\Vert _{{{\mathbb {L}}}^4}^4dt \Big \}^{\frac{1}{2}} \\&\quad \le \Big \{ {{\mathbb {E}}}\Big | \int _0^T \Vert \nabla v(t)\Vert _{{{\mathbb {L}}}^2}^2 dt \Big |^2 \Big \}^{\frac{1}{4}} \Big \{ {{\mathbb {E}}}\int _0^T \Vert u(t)\Vert _{{{\mathbb {L}}}^4}^4dt \Big \}^{\frac{1}{2}}. \end{aligned}$$

This completes the proof of (7.3).

A similar argument based on the identity \(\langle B(u,u)-B(v,v) \, , \, u-v\rangle = \langle B(u-v,v)\, , \, u-v\rangle \) shows (7.4). \(\square \)

We next prove upper estimates for the gradient of the bilinear term.

Lemma 4

  1. (i)

    There exists a positive constant C such that for \(\alpha \in (1,\infty )\), some constant \(C_\alpha >0\), any constants \(\varepsilon _0, \varepsilon _1 >0\) we have for \(u\in X_1\),

    $$\begin{aligned} |\langle A^{1/2} B(u,u)\, , \, A^{1/2} u\rangle |\le & {} C \Big [ \varepsilon _0 \Vert Au\Vert _{{{\mathbb {L}}}^2}^2 + \frac{\varepsilon _1}{4\varepsilon _0} \big \Vert |u|^\alpha \nabla u \big \Vert _{{{\mathbb {L}}}^2}^2 \nonumber \\&+ \frac{C_\alpha }{\varepsilon _0 \varepsilon _1^{\frac{1}{\alpha -1}}} \Vert \nabla u\Vert _{{{\mathbb {L}}}^2}^2 \Big ]. \end{aligned}$$
    (7.5)
  2. (ii)

    Let \(\alpha =1\); for every \(\epsilon >0\), we have for some constant \(C>0\) and any \(u\in X_1\)

    $$\begin{aligned} |\langle A^{1/2} B(u,u)\, , \,A^{1/2} u\rangle | \le \varepsilon \Vert Au\Vert _{{{\mathbb {L}}}^2}^2 + \frac{1}{4\varepsilon } \big \Vert |u| \nabla u \big \Vert _{{{\mathbb {L}}}^2}^2 . \end{aligned}$$
    (7.6)

Proof

  1. (i)

    Let \(\alpha >1\) and \(u\in X_1\). Then

    $$\begin{aligned} \langle A^{1/2} B(u,u)\, , \,A^{1/2} u\rangle = \sum _{i,j,k=1}^3 \int _{D} \partial _k\big [ u_i \, \partial _i u_j \big ]\, \partial _k u_j dx = T_1+T_2, \end{aligned}$$

    where, using the antisymmetry property (2.1), we get

    $$\begin{aligned} T_1&= \sum _{i,j,k=1}^3 \int _{D} \partial _k u_i \; \partial _i u_j \; \partial _k u_j dx, \\ T_2&= \sum _{i,j,k=1}^3 \int _{D} u_i \; \partial _k \partial _i u_j \; \partial _k u_j dx = \sum _{k=1}^3 \langle B(u,\partial _k u)\, , \, \partial _k u\rangle = 0. \end{aligned}$$

    Using integration by parts, we deduce \(T_1=T_{1,1}+T_{1,2}\), where since \(\mathrm{div\,} u=0\)

    $$\begin{aligned} T_{1,1}&=- \sum _{j,k=1}^3 \int _{D} \partial _k \Big ( \sum _{i=1}^3 \partial _i u_i\Big )\, u_j \, \partial _k u_j dx =0, \\ T_{1,2}&=- \sum _{i,j,k=1}^3 \int _{D} \partial _k u_i \, u_j \, \partial _i\partial _k u_j dx. \end{aligned}$$

    The inequality (2.5) applied with \(f=u_j\), \(g=\partial _k u_i\) and \(h=\partial _i \partial _k u_j\) implies

    $$\begin{aligned} |T_{1,2}|&\le \; \sum _{i,j,k=1}^3 \varepsilon _0 \Vert \partial _i\partial _k u_j\Vert _{L^2}^2 + \sum _{i,j,k=1}^3 \frac{\varepsilon _1}{4\varepsilon _0} \big \Vert |u_j|^\alpha \partial _k u_i \Vert _{L^2}^2 \\&\quad \; + \sum _{i,j,k=1}^3 \frac{C_\alpha }{\varepsilon _0 \varepsilon _1^{\frac{1}{\alpha -1}}} \Vert \partial _k u_i\Vert _{L^2}^2. \end{aligned}$$

    This completes the proof of (7.5).

  2. (ii)

    Let \(\alpha =1\) and \(u\in X_1\). Then an integration by parts implies

    $$\begin{aligned} \langle A^{1/2} B(u,u)\, , \,A^{1/2} u\rangle&=\; \sum _{i,j,k=1}^3 \int _{D} \partial _k\big [ u_i \, \partial _i u_j \big ]\, \partial _k u_j dx \\&=\; - \sum _{i,j=1}^3 \int _{D} u_i \, \partial _i u_j \; \varDelta u_j \, dx. \end{aligned}$$

    The Cauchy–Schwarz and Young inequalities imply (7.6).

\(\square \)

For \(\varphi \in X_0\), set

$$\begin{aligned} F(\varphi )= -\nu A\varphi - B(\varphi , \varphi ) -a \varPi |\varphi |^{2\alpha } \varphi . \end{aligned}$$
(7.7)

Lemma 2.2 page 415 in [2] provides upper and lower bounds of the non linear Brinkman–Forchheimer term. Let \(\alpha \in [1,\infty )\); there exist positive constants C and \(\kappa \) such that for \(u,v\in {{\mathbb {R}}}^3\)

$$\begin{aligned} \big | |u|^{2\alpha } u - |v|^{2\alpha } v \big |&\le C |u-v|\, \big ( |u|^{2\alpha } + |v|^{2\alpha } \big ), \end{aligned}$$
(7.8)
$$\begin{aligned} \big ( |u|^{2\alpha } u - |v|^{2\alpha } v \big ) \cdot (u-v)&\ge \kappa |u-v|^2 \big ( |u|+|v|\big )^{2\alpha }. \end{aligned}$$
(7.9)

The following lemma gives upper bounds of F for any \(\alpha \in [1,\infty )\).

Lemma 5

Let \(\alpha \in [1,+\infty )\).

  1. (i)

    Let \(u\in X_0\), \(v\in L^2(0,T;V) \cap L^{2\alpha +2}([0,T]\times D;{{\mathbb {R}}}^3)\). Then

    $$\begin{aligned}&\int _0^T\!\! | \langle F(u(t)), v(t)\rangle | dt \le C\big [ \Vert v\Vert _{L^2(0,T;V)} \Vert u\Vert _{L^2(0,T;V)} \nonumber \\&\quad + \Vert v\Vert _{L^{2\alpha +2}([0,T]\times D;{{\mathbb {R}}}^3)} \Vert u\Vert _{L^{2\alpha +2}([0,T]\times D;{{\mathbb {R}}}^3)}^{2\alpha +1} \nonumber \\&\quad + \Vert v\Vert _{L^2(0,T;V)}\; \underset{{t\in [0,T]}}{\mathrm{ess sup }} \Vert u(t)\Vert _{H}^{\frac{\alpha -1}{\alpha }} \Vert u\Vert _{L^{2\alpha +2}([0,T]\times D ; {{\mathbb {R}}}^3)}^{\frac{\alpha +1}{\alpha }} T^{\frac{\alpha -1}{2\alpha }}\big ] \end{aligned}$$
    (7.10)

    for some positive constant C.

  2. (ii)

    Let \(u\in {{\mathcal {X}}}_0\), \(v\in L^4(\varOmega ;L^2(0,T;V))\cap L^{2\alpha +2}(\varOmega _T\times D;{{\mathbb {R}}}^3)\). Then

    $$\begin{aligned}&{{\mathbb {E}}}\int _0^T \! |\langle F(u(t)),v(t)\rangle | dt \le C\Big [ \Vert v\Vert _{L^2(\varOmega _T;V)} \Vert u\Vert _{L^2(\varOmega _T;V)} \nonumber \\&\quad + \Vert v\Vert _{L^{2\alpha +2}(\varOmega _T \times D;{{\mathbb {R}}}^3)} \Vert u\Vert _{L^{2\alpha +2}(\varOmega _T\times D;{{\mathbb {R}}}^3)}^{2\alpha +1} \nonumber \\&\quad + \Vert v\Vert _{L^4(\varOmega ; L^2(0,T;V))} \Big \{ {{\mathbb {E}}}\Big ( \underset{{t\in [0,T]}}{\mathrm{ess\, sup }} \Vert u(t)\Vert _{H}^4\Big )\Big \}^{\frac{\alpha -1}{\alpha }} \Vert u\Vert _{L^{2\alpha +2}(\varOmega _T\times D ; {{\mathbb {R}}}^3)}^{\frac{\alpha +1}{\alpha }} T^{\frac{\alpha -1}{2\alpha }}\Big ] \end{aligned}$$
    (7.11)

    for some positive constant C.

Proof

Integration by parts and the Cauchy–Schwarz inequality imply

$$\begin{aligned} \nu \int _0^T \! \! | \langle A u(t)\, , \, v(t)\rangle | dt&=\; \int _0^T\! \Big | - \nu \int _{D} \! A^{\frac{1}{2}} u(t,x) A^{\frac{1}{2}} v((t,x) dx \Big | dt \\&\le \; \; \nu \Vert u\Vert _{L^2(0,T;V)} \Vert v\Vert _{L^2(0,T;V)}. \end{aligned}$$

Furthermore, Hölder’s inequality with conjugate exponents \(2\alpha +2\) and \(\frac{2\alpha +2}{2\alpha +1}\) yields

$$\begin{aligned} \int _0^T \Big | \int _{D} |u(t,x)|^{2\alpha } u(t,x) v(t,x) dx \Big | dt&\le \; \big \Vert |u|^{2\alpha } u\Vert _{L^{\frac{2\alpha +2}{2\alpha +1}}([0,T]\times D;{{\mathbb {R}}}^3)} \\&\quad \times \Vert v\Vert _{L^{2\alpha +2}([0,T]\times D;{{\mathbb {R}}}^3)}. \end{aligned}$$

Using the above upper estimates with the inequality (7.1) concludes the proof of (7.10).

(ii) The upper estimate (7.11) is a straightforward consequence of the upper estimates (7.3), (7.10), the Cauchy–Schwarz and Hölder inequalities. \(\square \)

The next lemma provides estimates of the gradient of F(u) for \(\alpha \in [1,+\infty )\). Note that when \(\alpha =1\), this requires that the coefficient a in front of the Brinkman–Forchheimer smoothing term is “not too smal” compared to the viscosity \(\nu \).

Lemma 6

  1. (i)

    Let \(\alpha >1\). For \(\eta \in (0,\nu )\), \({\tilde{a}}\in (0,a)\), there exists a positive constant \(C:=C(\alpha , \eta , {\tilde{a}})\) such that for \(u\in X_1\) and \(t\in [0,T]\),

    $$\begin{aligned}&\int _0^t \!\! \langle A^{1/2} F(u(s)) , A^{1/2}u(s) \rangle ds \nonumber \\&\quad \le -\eta \! \int _0^t\!\! \Vert A u(s)\Vert _{{{\mathbb {L}}}^2}^2 ds - {\tilde{a}}\! \int _0^t \!\! \big \Vert |u(s)|^{\alpha } \nabla u(s) \big \Vert _{{{\mathbb {L}}}^2}^2 ds + C\! \int _0^t \!\! \Vert \nabla u(s)\Vert _{{{\mathbb {L}}}^2}^2 ds. \end{aligned}$$
    (7.12)
  2. (ii)

    Let \(\alpha =1\) and suppose \(4\nu a >1\). Then for \(\eta \in \big ( 0, \nu - \frac{1}{4a}\big )\) and \({\tilde{a}} = a-\frac{1}{4(\nu -\eta )}\) we have

    $$\begin{aligned} \int _0^t \!\! \langle A^{1/2} F(u(s)) , A^{1/2}u(s)\rangle ds\le & {} -\eta \! \int _0^t\!\! \Vert A u(s)\Vert _{{{\mathbb {L}}}^2}^2 ds \nonumber \\&- {\tilde{a}}\! \int _0^t \!\! \big \Vert |u(s)|^{\alpha } \nabla u(s) \big \Vert _{{{\mathbb {L}}}^2}^2 ds . \end{aligned}$$
    (7.13)

Proof

  1. (i)

    Let \(\alpha \in (1,\infty )\). For \(u\in X_1\), integration by parts implies for a.e. \(s\in [0,t]\),

    $$\begin{aligned} \nu \langle A^{\frac{1}{2}} \varDelta u(s), A^{\frac{1}{2}} u(s)\rangle = -\nu \Vert A u(s)\Vert _{{{\mathbb {L}}}^2}^2.\end{aligned}$$

    Furthermore,

    $$\begin{aligned}&\int _{D} \nabla \big ( |u(s)|^{2\alpha } u(s)\big ) \cdot \nabla u(s) dx \nonumber \\&\quad = \int _{D}\big [ |u(s)|^{2\alpha } \nabla u(s) \cdot \nabla u(s) + 2\alpha |u(s)|^{2(\alpha -1)} \big ( u(s)\cdot \nabla u(s) \big )^2 \big ] dx \nonumber \\&\quad \ge \int _{D} |u(s)|^{2\alpha } \nabla u(s) \cdot \nabla u(s) dx = \big \Vert |u(s)|^\alpha \nabla u(s) \big \Vert _{{{\mathbb {L}}}^2}^2. \end{aligned}$$
    (7.14)

    Hence, using (7.5) with \(C\, \varepsilon _0\in (0, \nu -\eta )\), then \(\varepsilon _1\) such that \(C\, \frac{\varepsilon _1}{4\varepsilon _0} \in (0, a-{\tilde{a}})\), we deduce that for a.e. \(s\in [0,T]\),

    $$\begin{aligned} \langle A^{1/2} F(u(s)) , A^{1/2}u(s)\rangle&\le \; -\eta \Vert Au(s)\Vert _{{{\mathbb {L}}}^2}^2 - {\tilde{a}} \big \Vert |u(s)|^\alpha \nabla u(t)\Vert _{{{\mathbb {L}}}^2}^2 \nonumber \\&\quad \; + C(\alpha , \eta , {\tilde{a}}) \Vert \nabla u(s)\Vert _{{{\mathbb {L}}}^2}^2. \end{aligned}$$
    (7.15)

    Integrating this inequality on the time interval [0, t] concludes the proof of (7.12).

  2. (ii)

    Let \(\alpha =1\). Then using (7.6) and (7.14), we deduce for \(\epsilon >0\) and \(s\in [0,T]\)

    $$\begin{aligned} \langle A^{1/2} F(u(s)) , A^{1/2}u(s)\rangle&\le \; -(\nu -\epsilon ) \Vert A u(s)\Vert _{{{\mathbb {L}}}^2}^2 + \frac{1}{4\epsilon } \Vert |u(s)| \nabla u(s)\Vert _{{{\mathbb {L}}}^2}^2 \\&\quad \; -a \Vert |u(s)| \nabla u(s)\Vert _{{{\mathbb {L}}}^2}^2. \end{aligned}$$

    Since \(4a\nu >1\), for \(\eta \in \big ( 0, \nu - \frac{1}{4a}\big )\), \(\epsilon = \nu -\eta \) and \({\tilde{a}} = a-\frac{1}{4(\nu -\eta )}\) we deduce

    $$\begin{aligned} \langle A^{1/2} F(u(s)) , A^{1/2}u(s)\rangle \le -\eta \Vert Au(s)\Vert _{{{\mathbb {L}}}^2}^2 - {\tilde{a}} \big \Vert |u(s)|^\alpha \nabla u(s)\Vert _{{{\mathbb {L}}}^2}^2 . \end{aligned}$$
    (7.16)

    Integrating on the time interval [0, t], we deduce (7.13).

\(\square \)

We finally prove upper estimates of increments \(F(u)-F(v)\) for \(\alpha \in [1,\infty )\).

Lemma 7

There exists a positive constant \(\kappa \) depending on \(\alpha \in [1,+\infty )\), and for \(\eta \in (0,\nu )\) a positive constant \({\bar{C}}(\eta )\), such that for \(u,v\in V\cap L^{2\alpha +2}(D;{{\mathbb {R}}}^3)\),

$$\begin{aligned} \langle F(u)-F(v), u-v\rangle&\le \; - \eta \Vert \nabla (u-v)\Vert _{{{\mathbb {L}}}^2}^2 - a\kappa \big \Vert (|u|+|v|)^\alpha (u-v)\big \Vert _{{{\mathbb {L}}}^2}^2 \nonumber \\&\quad \; + {\bar{C}}(\eta ) \Vert \nabla v\Vert _{{{\mathbb {L}}}^2}^4 \Vert u-v\Vert _{{{\mathbb {L}}}^2}^2. \end{aligned}$$
(7.17)

Proof

Using integration by parts, we obtain

$$\begin{aligned} \nu \langle \varDelta (u-v),u-v\rangle = -\nu \Vert \nabla (u-v)\Vert _{{{\mathbb {L}}}^2}^2. \end{aligned}$$

The monotonicity property (7.9) implies

$$\begin{aligned}&a\int _{D} \big ( |u(x)|^{2\alpha } u(x) - |v(x)|^{2\alpha } v(x)\big ) \cdot \big ( u(x)-v(x)\big ) dx \nonumber \\&\quad \ge a \kappa \big \Vert (|u|+|v])^\alpha (u-v)\big \Vert _{{{\mathbb {L}}}^2}^2. \end{aligned}$$
(7.18)

Finally, Hölder’s inequality and the Gagliardo–Nirenberg inequality (2.2) for the \({{\mathbb {L}}}^4\) norm imply

$$\begin{aligned} |\langle B(u,u) - B(v,v) , u-v\rangle |&= |\langle B(u-v,v) , u-v\rangle |\\&\le \; \Vert u-v\Vert _{{{\mathbb {L}}}^4}^2 \Vert \nabla v\Vert _{{{\mathbb {L}}}^2} \le {\bar{C}}_4^2 \Vert u-v\Vert _{{{\mathbb {L}}}^2}^{\frac{1}{2}} \Vert \nabla (u-v)\Vert _{{{\mathbb {L}}}^2}^{\frac{3}{2}} \Vert \nabla v\Vert _{{{\mathbb {L}}}^2} \\&\le \; \frac{3}{4} \varepsilon ^{\frac{4}{3}} \Vert \nabla (u-v)\Vert _{{{\mathbb {L}}}^2}^2 + \frac{1}{4} \frac{1}{\varepsilon ^4} {\bar{C}}_4^8 \Vert \nabla v\Vert _{{{\mathbb {L}}}^2}^4 \Vert u-v\Vert _{{{\mathbb {L}}}^2}^2, \end{aligned}$$

where the last inequality holds for any \(\varepsilon >0\) by Young’s inequality. Choosing \(\frac{3}{4} \varepsilon ^{\frac{4}{3}} \in (0, \nu -\eta )\), we conclude the proof of (7.17). \(\square \)

We next prove that (1.1) has a unique strong solution in \({\mathcal {X}}_1\). The outline is quite classical, based on some Galerkin approximation and a priori estimates.

1.2 Galerkin approximation and a priori estimates

Recall that D is periodic domain of \({{\mathbb {R}}}^3\). Let \((e_n, n\ge 1)\) be the orthonormal basis of H defined in Sect. 3.1 (that is made of functions in H which are also orthogonal in V). For every integer \(n\ge 1\) we set \({{\mathcal {K}}}_n:=\mathrm{span} (\zeta _1, \ldots , \zeta _n)\) where \(\{\zeta _j\}_{j\ge 1}\) is an ONB of K mode of eigenfunctions of Q. Let \(\varPi _n\) denote the projection from K onto \(Q^{1/2}({{\mathcal {K}}}_n)\), and let \(W_n(t)=\sum _{j=1}^n \sqrt{q_j} \zeta _j \beta _j(t) = \varPi _n W(t)\).

Recall that if \({{\mathcal {H}}}_n= \mathrm{span} (e_1,\ldots , e_n)\), the orthogonal projection \(P_n\) of H onto \({{\mathcal {H}}}_n\) restricted to V coincides with the orthogonal projection of V onto \({{\mathcal {H}}}_n\).

Fix \(n\ge 1\) and consider the following stochastic ordinary differential equation on the n-dimensional space \({{\mathcal {H}}}_n\) defined by \(u_{n}(0)=P_n u_0\), and for \(t\in [0,T]\) and \(v \in {{\mathcal {H}}}_n\):

$$\begin{aligned} d(u_{n}(t), v)= \big \langle P_n F(u_{n}(t)),v\big \rangle dt + (P_n\, G( u_{n}(t)) \, \varPi _n\, dW(t), v), \quad {{\mathbb {P}}}\, \mathrm{a.s.},\qquad \end{aligned}$$
(7.19)

where F is defined in (7.7). Then for \(k=1, \, \ldots , \, n\) we have for \(t\in [0,T]\):

$$\begin{aligned} d(u_{n}(t), e_k)= \big \langle P_n F(u_{n}(t)),e_k \big \rangle \, dt + \sum _{j=1}^n q_j^{\frac{1}{2}} \big ( P_n\, G(u_{n}(t)) \zeta _j\, ,\, e_k \big ) \, d\beta _j(t), \quad {{\mathbb {P}}}\, \text{ a.s. } \end{aligned}$$

Note that for \(v\in {{\mathcal {H}}}_n\) the map \( u\in {{\mathcal {H}}}_n \mapsto \langle F(u) \, ,\, v\rangle \) is locally Lipschitz. Indeed, \({{\mathbb {H}}}^2\subset {{\mathbb {L}}}^{2\alpha +2}\) and there exists some constant C(n) such that \(\Vert v\Vert _{{{\mathbb {H}}}^2} \le C(n) \Vert v\Vert _{{{\mathbb {L}}}^2}\) for \(v\in {{\mathcal {H}}}_n\). Let \(\varphi , \psi , v\in {{\mathcal {H}}}_n\); integration by parts implies that

$$\begin{aligned} |\langle \varDelta \varphi - \varDelta \psi ,v\rangle | \le \Vert \varphi - \psi \Vert _V \, \Vert v\Vert _V \le C(n)^2 \Vert \varphi - \psi \Vert _{{{\mathbb {L}}}^2}\, \Vert v\Vert _{{{\mathbb {L}}}^2}. \end{aligned}$$

In the polynomial nonlinear term, the upper estimate (7.8), the Hölder inequality with exponents \(\frac{\alpha +1}{\alpha }\), \(2\alpha +2\), and \(2\alpha +2\), and the Sobolev embedding \({{\mathbb {H}}}^2\subset {{\mathbb {L}}}^{2\alpha +2}\) imply

$$\begin{aligned}&\Big | \int _{D}\! \big ( |\varphi (x) |^{2\alpha }\varphi (x)- |\psi (x)|^{2\alpha } \psi (x) \big ) v(x) dx \Big | \\&\quad \le C\, \big ( \Vert \varphi \Vert _{{{\mathbb {L}}}^{2\alpha +2}}^{2\alpha } + \Vert \psi \Vert _{{{\mathbb {L}}}^{2\alpha +2}}^{2\alpha } \big ) \, \Vert \varphi - \psi \Vert _{{{\mathbb {L}}}^{2\alpha +2}}\, \Vert v\Vert _{{{\mathbb {L}}}^{2\alpha +2}} \\&\quad \le C \, C(n)^{2(\alpha +1)} \big ( \Vert \varphi \Vert _{{{\mathbb {L}}}^2}^{2\alpha } +\Vert \psi \Vert _{{{\mathbb {L}}}^2}^{2\alpha }\big ) \, \Vert \varphi - \psi \Vert _{{{\mathbb {L}}}^2}\, \Vert v\Vert _{{{\mathbb {L}}}^2} . \end{aligned}$$

Finally, using integration by parts, the Hölder and Gagliardo–Nirenberg inequalities, we deduce:

$$\begin{aligned} |\langle B(\varphi ,\varphi ) - B(\psi ,\psi ), v\rangle |&=\big | -\langle B(\varphi -\psi , v)\, , \, \varphi \rangle - \langle B(\psi , v)\, , \, \varphi -\psi \rangle \big | \\&\le C\, \Vert \varphi - \psi \Vert _{{{\mathbb {L}}}^4} \big ( \Vert \varphi \Vert _{{{\mathbb {L}}}^4}+ \Vert \psi \Vert _{{{\mathbb {L}}}^4} \big ) \Vert \nabla v\Vert _{{{\mathbb {L}}}^2} \\&\le C C(n)^3 \Vert \varphi - \psi \Vert _{{{\mathbb {L}}}^2} \big ( \Vert \varphi \Vert _{{{\mathbb {L}}}^2}+\Vert \psi \Vert _{{{\mathbb {L}}}^2}\big ) \Vert v\Vert _{{{\mathbb {L}}}^2}. \end{aligned}$$

Condition (G) implies that the map \(u\in {{\mathcal {H}}}_n \mapsto \big ( \sqrt{q_j}\, \big (G(u) \zeta _j\, ,\, e_k\big ) : 1\le j,k\le n \big )\) satisfies the classical global linear growth and Lipschitz conditions from \({{\mathcal {H}}}_n\) to \(n\times n\) matrices uniformly in \(t\in [0,T]\). Hence by a well-known result about existence and uniqueness of solutions to stochastic differential equations (see e.g. [24]), there exists a maximal solution \(u_{n}=\sum _{k=1}^n (u_{n}\, ,\, e_k\big )\, e_k\in {{\mathcal {H}}}_n\) to (7.19), i.e., a stopping time \(\tau _{n}^*\le T\) such that (7.19) holds for \(t< \tau _{n}^*\) and if \(\tau _n^* <T\), \(\Vert u_{n}(t)\Vert _{L^2} \rightarrow \infty \) as \(t \uparrow \tau _{n}^*\).

The following proposition shows that \(\tau _n^*=T\) a.s., and provides a priori estimates on norms of \(u_n\), which do not depend on n.

Proposition 7

Let \(\alpha \in [1,\infty )\), and if \(\alpha =1\), suppose that \(4\nu a >1\).

  1. (i)

    Let \(u_0\) be \({{\mathcal {F}}}_0\)-measurable such that \({{\mathbb {E}}}\big ( \Vert u_0\Vert _H^2\big )<\infty \), \(T>0\) and G satisfy (3.2) and (3.4). Then the evolution equation (7.19) with initial condition \(P_n u_0\) has a unique global solution on [0, T] (i.e., \(\tau _n^*=T\) a.s.) with a modification \(u_n\in C([0,T];{{\mathcal {H}}}_n)\). Furthermore, if \({{\mathbb {E}}}\big ( \Vert u_0\Vert _H^{2p}\big ) <\infty \) for some \(p\in [1,\infty )\), we have \(u_n \in {{\mathcal {X}}}_0\) and

    $$\begin{aligned}&\sup _n {{\mathbb {E}}}\Big ( \sup _{t\in [0,T]} \Vert u_n(t)\Vert _H^{2p} + \int _0^T \big [ \Vert u_n(t)\Vert _V^2 + \Vert u_n(t)\Vert _{{{\mathbb {L}}}^{2\alpha +2}}^{2\alpha +2}\big ] \Vert u_n(t)\Vert _H^{2p-2} \, dt \Big ) \nonumber \\&\quad \le C \big [ 1+{{\mathbb {E}}}(\Vert u_0\Vert _H^{2p}) \big ]. \end{aligned}$$
    (7.20)
  2. (ii)

    If \({{\mathbb {E}}}(\Vert u_0\Vert _V^{2p}) <\infty \) for some \(p\in [1,\infty )\) and G satisfies also (3.3), we have furthermore

    $$\begin{aligned}&\sup _n {{\mathbb {E}}}\Big ( \sup _{t\in [0,T]} \Vert u_n(t)\Vert _V^{2p} \nonumber \\&\qquad + \int _0^T \big [ \Vert A u_n(t)\Vert _{{{\mathbb {L}}}^2}^2 + \Vert |u_n(t)|^\alpha \nabla u_n(t)\Vert _{{{\mathbb {L}}}^2}^2 \big ] \Vert u_n(t)\Vert _V^{2p-2} dt\Big ) \nonumber \\&\quad \le C \big [ 1+{{\mathbb {E}}}(\Vert u_0\Vert _V^{2p}) \big ] . \end{aligned}$$
    (7.21)

Proof

(i) For fixed \(N>0\) set \(\tau _N:= \inf \{ t\ge 0 : \Vert u_n(t)\Vert _H \ge N\} \wedge \tau _n^*\). Itô’s formula and the antisymmetry property of B imply

$$\begin{aligned} \Vert u_n(t\wedge \tau _N)\Vert _H^2&=\; \Vert P_n u_0\Vert _H^2 -2\int _0^{t\wedge \tau _N} \!\! \big [ \nu \Vert \nabla u_n(s)\Vert _{{{\mathbb {L}}}^2}^2 +a \Vert u_n(s)\Vert _{L^{2\alpha +2}}^{2\alpha +2} \big ] ds \nonumber \\&\quad + \sum _{i=1}^2 T_i(t), \end{aligned}$$
(7.22)

where

$$\begin{aligned} T_1(t)&= \; 2 \int _0^{t\wedge \tau _n} \big ( G(u_n(s))\, dW_n(s)\, , \, u_n(s)\big ), \\ T_2(t)&=\; \int _0^{t\wedge \tau _n} \Vert P_n G(u_n(s)) \varPi _n \Vert _{{\mathcal {L}}}^2 \; ds. \end{aligned}$$

Apply once more the Itô formula to \(z\mapsto z^p\) and \(z=\Vert u_n(t\wedge \tau _N)\Vert _H^2\) for \(p\in [2,\infty )\). We obtain

$$\begin{aligned} \Vert u_n(t\wedge \tau _N)\Vert _H^{2p}&= \Vert P_n u_0\Vert _H^{2p} + \sum _{i=1}^3 {\bar{T}}_i(t) \nonumber \\&\quad -2p\int _0^{t\wedge \tau _N} \!\! \big [ \nu \Vert \nabla u_n(s)\Vert _{{{\mathbb {L}}}^2}^2 +a \Vert u_n(s)\Vert _{L^{2\alpha +2}}^{2\alpha +2}\big ] \Vert u_n(s)\Vert _H^{2p-2} ds, \end{aligned}$$
(7.23)

where

$$\begin{aligned} {\bar{T}}_1(t)&= \; 2p \int _0^{t\wedge \tau _N} \big ( P_n G(u_n(s))\, dW_n(s) \, , \, u_n(s)\big ) \, \Vert u_n(s)\Vert _H^{2p-2}, \\ {\bar{T}}_2(t)&=\; p\int _0^{t\wedge \tau _N} \Vert P_n G(u_n(s)) \varPi _n\Vert _{{\mathcal {L}}}^2 \, \Vert u_n(s)\Vert _H^{2p-2}\, ds , \\ {\bar{T}}_3(t)&=\; 2p(p-1) \int _0^{t\wedge \tau _N} \Vert \big ( G(u_n(s))\, \varPi _n \big )^* u_n(s)\big \Vert _{K}^2\, \Vert u_n(s)\Vert _H^{2p-4}\, ds. \end{aligned}$$

The growth condition (3.2) implies

$$\begin{aligned} {\bar{T}}_2(t) + {\bar{T}}_3(t) \le p(2p-1) \int _0^{t} [K_0 + K_1 \Vert u_n(s\wedge \tau _N)\Vert _H^2] \, \Vert u_n(s\wedge \tau _N)\Vert _H^{2p-2}\, {\mathrm{Tr}} Q\, ds. \end{aligned}$$

Using the Davis inequality, the growth condition (3.2) and Young’s inequality, we deduce for \(\beta \in (0,1)\),

$$\begin{aligned} {{\mathbb {E}}}\Big ( \sup _{s\le t\wedge \tau _n} \; {\bar{T}}_1(s)\Big )\le & {} \; 6p \; {{\mathbb {E}}}\Big ( \Big \{ \int _0^{t\wedge \tau _N} \Vert G(u_n(s))\Vert _{{\mathcal {L}}}^2 \; \Vert u_n(s)\Vert _H^{4p-2}\; {\mathrm{Tr}}Q\, ds \Big \}^{\frac{1}{2}} \Big ) \\\le & {} \; \beta \; {{\mathbb {E}}}\Big ( \sup _{s\le t} \Vert u_n(s\wedge \tau _N)\Vert _H^{2p}\Big ) \\&\; + \frac{9p^2}{\beta } {{\mathbb {E}}}\!\! \int _0^t\!\! \big [ K_0+K_1 \Vert u_n(s\wedge \tau _N)\Vert _H^2\big ] \Vert u_n(s\wedge \tau _N)\Vert _H^{2p-2} \, {\mathrm{Tr}} Q\, ds . \end{aligned}$$

Neglecting the first integral in the right hand side of (7.23), using the above upper estimates of \({\bar{T}}_i\) and the Gronwall lemma, we deduce that for \(\beta \in (0,1)\),

$$\begin{aligned} \sup _{n\ge 1} {{\mathbb {E}}}\Big ( \sup _{s\le T} \Vert u_n(s\wedge \tau _N)\Vert _H^{2p}\Big ) \le C(\beta ,p, K_0, K_1, {\mathrm{Tr}}Q) \big [ 1+{{\mathbb {E}}}(\Vert u_0\Vert _H^{2p}\big ].\qquad \end{aligned}$$
(7.24)

As \(N\rightarrow \infty \), the sequence of stopping times \(\tau _N\) increases to \(\tau ^*_n\) and on the set \(\{ \tau ^*_n<T\}\), we have \(\sup _{s\in [0, \tau _N]} \Vert u_n(s)\Vert _H \rightarrow \infty \). Hence (7.24) implies \(P(\tau _n^*<T)=0\) and for almost every \(\omega \), for \(N(\omega )\) large enough we have \(\tau _{N(\omega )}(\omega )=T\). Plugging the upper estimate (7.24) in (7.23), we conclude the proof of (7.20).

Note that the above argument based on (7.22) instead of (7.23) proves that if \({{\mathbb {E}}}(\Vert u_0\Vert _H^2) <\infty \) we have once more \(\tau _{N(\omega )}(\omega )=T\) for \(N(\omega )\) large enough and a.e. \(\omega \), and that (7.20) holds for \(p=1\).

We next prove that \(u_n\in {{\mathcal {X}}}_0\). Plugging the above upper estimate for \(p=1\) in (7.22), taking expected values and using Condition (3.2), we obtain

$$\begin{aligned} {{\mathbb {E}}}\int _0^T \big [ \Vert u_n(s)\Vert _V^2 + \Vert u_n(s)\Vert _{L^{2\alpha +2}}^{2\alpha +2} \big ] ds <\infty . \end{aligned}$$

A similar argument using (7.24) in (7.23) completes the proof of (7.20) when the H-norm of the initial condition has 2p moments.

(ii) Taking the gradient of both hand sides of (7.19), using the Itô formula and (3.1), we deduce for \( {\tilde{\tau }}_N:=\inf \{ s\ge 0\, : \, \Vert u_n(s)\Vert _V\ge N\} \wedge T\),

$$\begin{aligned} \Vert A^{\frac{1}{2}} u_n(t\wedge {\tilde{\tau }}_N)\Vert _{{{\mathbb {L}}}^2}^2&= \Vert A^{\frac{1}{2}} P_n u_0\Vert _{{{\mathbb {L}}}^2}^2 + 2 \int _0^{t\wedge {\tilde{\tau }}_N} \langle A^{\frac{1}{2}} P_n F(u_n(s)) , A^{\frac{1}{2}} u_n(s)\rangle \, ds \\&\quad + 2\! \int _0^{t\wedge {\tilde{\tau }}_N} \!\! \big ( A^{\frac{1}{2}} P_n G(u_n(s)) dW_n(s) , A^{\frac{1}{2}} u_n(s)\big ) \\&\quad + \int _0^{t\wedge {\tilde{\tau }}_N} \!\! \Vert A^{\frac{1}{2}} P_n G(u_n(s)) \varPi _n\Vert _{{\mathcal {L}}}^2 \, ds\\&= \, \Vert A^{\frac{1}{2}} P_n u_0\Vert _{{{\mathbb {L}}}^2}^2 + 2 \int _0^{t\wedge {\tilde{\tau }}_N} \langle A^{\frac{1}{2}} F(u_n(s)) , A^{\frac{1}{2}} u_n(s)\rangle \, ds \\&\quad + 2\! \int _0^{t\wedge {\tilde{\tau }}_N} \!\! \big ( A^{\frac{1}{2}} G(u_n(s)) \varPi _n dW(s) , A^{\frac{1}{2}} u_n(s)\big ) \\&\quad + \int _0^{t\wedge {\tilde{\tau }}_N} \!\! \Vert A^{\frac{1}{2}} P_n G(u_n(s)) \varPi _n\Vert _{{\mathcal {L}}}^2 \, ds. \end{aligned}$$

Indeed, since \(u_n(s)\in V\) for \(s\le t\wedge {\tilde{\tau }}_N\), we deduce \(A^{\frac{1}{2}} u_n(s)\in H\) and \(A^{\frac{1}{2}} G(u_n(s))\in {{\mathcal {L}}}(K,H)\).

Using once more the Itô formula for the function \(z\mapsto z^p\) for \(p\in [2,\infty )\), we obtain

$$\begin{aligned} \Vert A^{\frac{1}{2}} u_n(t\wedge {\tilde{\tau }}_N)\Vert _{{{\mathbb {L}}}^2}^{2p}&\le \; \Vert A^{\frac{1}{2}} u_0\Vert _{{{\mathbb {L}}}^2}^{2p} + \sum _{i=1}^3 {\tilde{T}}_i(t) \nonumber \\&\quad + 2 \int _0^{t\wedge {\tilde{\tau }}_N} \langle A^{\frac{1}{2}} \nabla F(u_n(s)) , A^{\frac{1}{2}} u_n(s)\rangle \, \Vert A^{\frac{1}{2}} u_n(s)\Vert _{{{\mathbb {L}}}^2}^{2(p-1)} ds, \end{aligned}$$
(7.25)

where

$$\begin{aligned} {\tilde{T}}_1(t)&=\; 2p \int _0^{t\wedge {\tilde{\tau }}_N} \big ( A^{\frac{1}{2}} G(u_n(s)) dW_n(s)\, , \, A^{\frac{1}{2}} u_n(s)\big ) \, \Vert A^{\frac{1}{2}} u_n(s)\Vert _{{{\mathbb {L}}}^2}^{2(p-1)} ds,\\ {\tilde{T}}_2(t)&=\; p\int _0^{t\wedge {\tilde{\tau }}_N} \Vert G(u_n(s) ) \varPi _n\Vert _{\tilde{{\mathcal {L}}}}^2 \; \Vert A^{\frac{1}{2}} u_n(s)\Vert _{{{\mathbb {L}}}^2}^{2(p-1)} ds, \\ {\tilde{T}}_3(t)&=\; 2p(p-1)\! \int _0^{t\wedge {\tilde{\tau }}_N} \!\! \Vert \big ( A^{\frac{1}{2}} G(u_n(s)) \, \varPi _n \big )^* A^{\frac{1}{2}} u_n(s)\Vert _{K}^2 \; \Vert A^{\frac{1}{2}} u_n(s)\Vert _{{{\mathbb {L}}}^2}^{2(p-2)}\, ds. \end{aligned}$$

Since

$$\begin{aligned} \Vert \big (A^{\frac{1}{2}} G(u_n(s)) \, \varPi _n \big )^*\Vert _{{{\mathcal {L}}}(H;K)} \le \Vert A^{\frac{1}{2}} G(u_n(s))\Vert _{{{\mathcal {L}}}(K;H)} \le \Vert G(u_n(s))\Vert _{{{\mathcal {L}}}(K;V)}, \end{aligned}$$

the growth condition (3.3) and Young’s inequality imply

$$\begin{aligned} {\tilde{T}}_2(t)+{\tilde{T}}_3(t) \le C(p,T,{\mathrm{Tr}} Q, {\tilde{K}}_0, {\tilde{K}}_1) \Big [ 1+ \! \int _0^{t\wedge {\tilde{\tau }}_N}\!\!\! \big ( \Vert u_n(s)\Vert _H^{2p} + \Vert \nabla u_n(s)\Vert _{{{\mathbb {L}}}^2}^{2p}\big ) ds\Big ]. \end{aligned}$$

The growth condition (3.3), the Gundy and Young inequalities imply that for \({\tilde{\beta }}\in (0,1)\),

$$\begin{aligned} {{\mathbb {E}}}\Big ( \sup _{s\le t} {\tilde{T}}_1(s) \Big )&\le \,C(p) {{\mathbb {E}}}\Big ( \Big \{\! \int _0^{t\wedge {\tilde{\tau }}_N} \!\!\!\big [ {\tilde{K}}_0 + {\tilde{K}}_1 \Vert u_n(s)\Vert _V^2\big ] \Vert \nabla u_n(s)\Vert _{{{\mathbb {L}}}^2}^{4p-2} \, {\mathrm{Tr}}Q ds \Big \}^{\frac{1}{2}} \!\Big ) \\&\le \; {\tilde{\beta }} {{\mathbb {E}}}\Big ( \sup _{s\le t} \Vert u_n(s\wedge {\tilde{\tau }}_N)\Vert _{{{\mathbb {L}}}^2}^{2p} \Big ) + {\tilde{\beta }} {{\mathbb {E}}}\Big ( \sup _{s\le t} \Vert \nabla u_n(s\wedge {\tilde{\tau }}_N)\Vert _{{{\mathbb {L}}}^2}^{2p} \Big ) \\&\quad \; + C({\tilde{\beta }}, {\mathrm{Tr}} Q, {\tilde{K}}_0, {\tilde{K}}_1) \Big [ 1+ {{\mathbb {E}}}\Big ( \int _0^t \Vert \nabla u_n(s\wedge {\tilde{\tau }}_N)\Vert _{{{\mathbb {L}}}^2}^{2p} ds\Big ) \Big ]. \end{aligned}$$

Let \(\rho \in (0,\nu )\) and \({\tilde{a}}\in (0,a)\). Using (7.12) for \(\alpha >1\) and (7.13) for \(\alpha =1\), (7.20) and the Gronwall lemma, we deduce

$$\begin{aligned} {{\mathbb {E}}}\Big ( \sup _{s\le {\tilde{\tau }}_N} \Vert u_n(s)\Vert _V^{2p}\Big ) \le C\big [ 1+ {{\mathbb {E}}}(\Vert u_0\Vert _V^{2p})\big ] \end{aligned}$$

for some positive constant C which does not depend on N and n. For fixed n, letting \(N\rightarrow \infty \) and using the monotone convergence theorem we deduce \(u_n\in L^{2p}(\varOmega ; L^\infty (0,T;V))\). Plugging this in (7.25) and taking expected values, we conclude the proof of (7.21). \(\square \)

1.3 Proof of global well-posedness of the solution

The proof of Theorem 2 is classical and uses the upper estimates (7.2) and (7.4) for the uniqueness; see e.g. [6] for details.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bessaih, H., Millet, A. Strong \(L^2\) convergence of time Euler schemes for stochastic 3D Brinkman–Forchheimer–Navier–Stokes equations. Stoch PDE: Anal Comp 10, 1005–1049 (2022). https://doi.org/10.1007/s40072-022-00255-9

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s40072-022-00255-9

Keywords

Mathematics Subject Classification

Navigation