Skip to main content
Log in

Analysis of a projection method for the Stokes problem using an \(\varepsilon \)-Stokes approach

  • Original Paper
  • Published:
Japan Journal of Industrial and Applied Mathematics Aims and scope Submit manuscript

Abstract

We generalize pressure boundary conditions of an \(\varepsilon \)-Stokes problem. Our \(\varepsilon \)-Stokes problem connects the classical Stokes problem and the corresponding pressure-Poisson equation using one parameter \(\varepsilon >0\). For the Dirichlet boundary condition, it is proven in Matsui and Muntean (Adv Math Sci Appl, 27:181–191, 2018) that the solution for the \(\varepsilon \)-Stokes problem converges to the one for the Stokes problem as \(\varepsilon \) tends to 0, and to the one for the pressure-Poisson problem as \(\varepsilon \) tends to \(\infty \). Here, we extend these results to the Neumann and mixed boundary conditions. We also establish error estimates in suitable norms between the solutions to the \(\varepsilon \)-Stokes problem, the pressure-Poisson problem and the Stokes problem, respectively. Several numerical examples are provided to show that several such error estimates are optimal in \(\varepsilon \). Our error estimates are improved if one uses the Neumann boundary conditions. In addition, we show that the solution to the \(\varepsilon \)-Stokes problem has a nice asymptotic structure.

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
Fig. 6
Fig. 7
Fig. 8

Similar content being viewed by others

References

  1. Amsden, A.A., Harlow, F.H.: A simplified MAC technique for incompressible fluid flow calculations. J. Comput. Phys. 6, 322–325 (1970)

    Article  MATH  Google Scholar 

  2. Chan, R.K.C., Street, R.L.: A computer study of finite-amplitude water waves. J. Comput. Phys. 6, 68–94 (1970)

    Article  MATH  Google Scholar 

  3. Chorin, A.J.: Numerical solution of the Navier–Stokes equations. Math. Comput. 22(104), 745–762 (1968)

    Article  MathSciNet  MATH  Google Scholar 

  4. Conca, C., Murat, F., Pironneau, O.: The Stokes and Navier–Stokes equations with boundary conditions involving the pressure. Jpn. J. Math. 20(2), 279–318 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  5. Conca, C., Pares, C., Pironneau, O., Thiriet, M.: Navier–Stokes equations with imposed pressure and velocity fluxes. Int. Numer. Methods. Fluids 20, 267–287 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  6. Cummins, S.J., Rudman, M.: An SPH projection method. J. Comput. Phys. 152, 584–607 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  7. Douglas, J., Wang, J.: An absolutely stabilized finite element method for the Stokes problem. Math. Comput. 52(186), 495–508 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  8. Duvaut, G., Lions, J.L.: Inequalities in Mechanics and Physics. Springer, Berlin (1976)

    Book  MATH  Google Scholar 

  9. Girault, V., Raviart, P.A.: Finite Element Methods for Navier–Stokes Equations. Springer, Berlin (1986)

    Book  MATH  Google Scholar 

  10. Glowinski, R.: Finite element methods for incompressible viscous flow. In: Ciarlet, P.G., Lions, J.L. (eds.) Handbook of Numerical Analysis, vol. 9, pp. 3–1176. North-Holland, Amsterdam (2003)

    Google Scholar 

  11. Guermond, J.L., Quartapelle, L.: On stability and convergence of projection methods based on pressure Poisson equation. Int. J. Numer. Methods Fluids 26, 1039–1053 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  12. Harlow, F.H., Welch, J.E.: Numerical calculation of time-dependent viscous incompressible flow of fluid with a free surface. Phys. Fluids 8, 2182–2189 (1965)

    Article  MathSciNet  MATH  Google Scholar 

  13. Hecht, F.: New development in FreeFem++. J. Numer. Math. 20(3–4), 251–265 (2012)

    MathSciNet  MATH  Google Scholar 

  14. Hughes, T., Franca, L., Balestra, M.: A new finite element formulation for computational fluid dynamics: V. Circumventing the Babuška–Brezzi condition: a stable Petrov–Galerkin formulation of the Stokes problem accommodating equal-order interpolations. Comput. Methods Appl. Mech. Eng. 59, 85–99 (1986)

    Article  MATH  Google Scholar 

  15. Kim, J., Moin, P.: Application of a fractional-step method to incompressible Navier–Stokes equations. J. Comput. Phys. 59, 308–323 (1985)

    Article  MathSciNet  MATH  Google Scholar 

  16. Marušić, S.: On the Navier–Stokes system with pressure boundary condition. Ann. Univ. Ferrara 53, 319–331 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  17. Matsui, K., Muntean, A.: Asymptotic analysis of an \(\varepsilon \)-Stokes problem connecting Stokes and pressure-Poisson problems. Adv. Math. Sci. Appl. 27, 181–191 (2018)

    MathSciNet  Google Scholar 

  18. McKee, S., Tomé, M.F., Cuminato, J.A., Castelo, A., Ferreira, V.G.: Recent advances in the marker and cell method. Arch. Comput. Methods Eng. 2, 107–142 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  19. Nečas, J.: Direct Methods in the Theory of Elliptic Equations. Springer, Berlin (1967). (Translated in 2012 by Kufner, A. and Tronel, G)

    MATH  Google Scholar 

  20. Perot, J.B.: An analysis of the fractional step method. J. Comput. Phys. 108, 51–58 (1993)

    Article  MathSciNet  MATH  Google Scholar 

  21. Temam, R.: Navier–Stokes Equations. North Holland, Amsterdam (1979)

    MATH  Google Scholar 

  22. Viecelli, J.A.: A computing method for incompressible flows bounded by moving walls. J. Comput. Phys. 8, 119–143 (1971)

    Article  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Kazunori Matsui.

Additional information

Publisher's Note

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

Appendix

Appendix

Theorems 2.74.24.3, and Proposition 2.8 are generalizations of severaltheorems stated in [17]. In this appendix, however, we give their proofs for the readers’convenience. We define a continuous coercive bilinear form depending on \(\varepsilon \) and prove Theorem 2.7 by the Lax–Milgram Theorem.

Proof of Theorem 2.7

We take arbitrary \(u_1\in {H^1({\varOmega })}^n\) with \(\gamma _0 u_1=u_b\). Since \({{\text {div}}}:{H^1_0({\varOmega })}^n\rightarrow {L^2({\varOmega })}/{{\mathbb {R}}}\) is surjective [9, Corollary 2.4, 2\({}^\circ \)], there exists \(u_2\in {H^1_0({\varOmega })}^n\) such that \({{\text {div}}}u_2={{\text {div}}}u_1\). We put

$$\begin{aligned} u_0:=u_1-u_2, \end{aligned}$$
(A.1)

and note that \(\gamma _0 u_0=u_b\) and \({{\text {div}}}u_0=0\). To simplify the notation, we set \(u:=u_\varepsilon -u_0\in {H^1_0({\varOmega })}^n,p:=p_\varepsilon -p_0\in Q\), and define \(f\in {H^{-1}({\varOmega })}^n\) and \(g\in Q^*\) by

$$\begin{aligned} \begin{array}{rll} \langle f,v\rangle &{}{\displaystyle :=\int _{\varOmega }Fv-\int _{\varOmega }\nabla u_0:\nabla v-\int _{\varOmega }(\nabla p_0)\cdot v} &{}\text{ for } \text{ all } \;v\in {H^1_0({\varOmega })}^n,\\ \langle g,q\rangle &{}{\displaystyle :=\langle G,q\rangle -\int _{\varOmega }\nabla p_0\cdot \nabla q} &{}\text{ for } \text{ all } \;q\in Q. \end{array} \end{aligned}$$
(A.2)

Then, \((u_\varepsilon ,p_\varepsilon )\) satisfies (ES’) if and only if (up) satisfies

$$\begin{aligned} \left\{ \begin{array}{ll} {\displaystyle \int _{\varOmega }\nabla u:\nabla \varphi +\int _{\varOmega }(\nabla p)\cdot \varphi =\langle f,\varphi \rangle } &{} {\text{ for } \text{ all } } \;\varphi \in {H^1_0({\varOmega })}^n,\\ {\displaystyle \varepsilon \int _{\varOmega }\nabla p\cdot \nabla \psi +\int _{\varOmega }({{\text {div}}}u)\psi =\varepsilon \langle g,\psi \rangle } &{} {\text{ for } \text{ all } } \;\psi \in Q. \end{array}\right. \end{aligned}$$
(A.3)

Adding the equations in (A.3), we get

$$\begin{aligned} \begin{aligned} \left( \left( \begin{array}{c} u\\ p\end{array}\right) , \left( \begin{array}{c} \varphi \\ \psi \end{array}\right) \right) _\varepsilon&:=\int _{\varOmega }\nabla u:\nabla \varphi + \varepsilon \int _{\varOmega }\nabla p\cdot \nabla \psi +\int _{\varOmega }(\nabla p)\cdot \varphi + \int _{\varOmega }({{\text {div}}}u)\psi \\&= \langle f,\varphi \rangle +\varepsilon \langle g,\psi \rangle . \end{aligned} \end{aligned}$$

We check that \((\cdot ,\cdot )_\varepsilon \) is a continuous coercive bilinear form on \({H^1_0({\varOmega })}^n\times Q\). The bilinearity and continuity of \((\cdot ,\cdot )_\varepsilon \) are obvious. The coercivity of \((\cdot ,\cdot )_\varepsilon \) is obtained in the following way. Take \((v,q)^T\in {H^1_0({\varOmega })}^n\times Q\). We have the following sequence of inequalities:

$$\begin{aligned} \begin{aligned} \left( \left( \begin{array}{c} v\\ q\end{array}\right) , \left( \begin{array}{c} v\\ q\end{array}\right) \right) _\varepsilon&={\displaystyle \int _{\varOmega }\nabla v:\nabla v + \varepsilon \int _{\varOmega }\nabla q\cdot \nabla q +\int _{\varOmega }v\cdot \nabla q+\int _{\varOmega }({{\text {div}}}v)q}\\&= \Vert \nabla v\Vert ^2_{{L^2({\varOmega })}}+\varepsilon \Vert \nabla q\Vert ^2_{{L^2({\varOmega })}}\\&\ge {\displaystyle \mathrm{min}\{ 1,\varepsilon \}\left( \Vert \nabla v\Vert ^2_{{L^2({\varOmega })}}+\Vert \nabla q\Vert ^2_{{L^2({\varOmega })}}\right) }\\&\ge {\displaystyle c~\mathrm{min}\{ 1,\varepsilon \}\left( \Vert v\Vert ^2_{{H^1({\varOmega })}^n}+\Vert q\Vert ^2_{{H^1({\varOmega })}}\right) }. \end{aligned} \end{aligned}$$

Summarizing, \((\cdot ,\cdot )_\varepsilon \) is a continuous coercive bilinear form and \({H^1_0({\varOmega })}^n\times Q\) is a Hilbert space. Therefore, the conclusion of Theorem 2.7 follows from the Lax–MilgramTheorem. \(\square \)

Let \((u_S,p_S),(u_PP,p_PP)\) and \((u_\varepsilon ,p_\varepsilon )\) be the solutions of (S’), (PP’) and (ES’), respectively, as guaranteed by Theorems 2.52.6 and 2.7. We show that the subtract \(p_S-p_PP\) satisfies

$$\begin{aligned} {\varDelta }(p_S-p_PP)=0 \end{aligned}$$

in distributions sense. The weak harmonicity is the key ingredient to provingProposition 2.8.

Proof of Proposition 2.8

First, we prove that there exists a constant \(c>0\) independent of \(\varepsilon \) such that \( \Vert u_S-u_PP\Vert _{{H^1({\varOmega })}^n} \le c\Vert \gamma _0 p_S-\gamma _0 p_PP\Vert _{H^{1/2}({\varGamma })}, \) and if \(\gamma _0(p_S-p_PP)=0\), then \(p_PP=p_S\). Taking the divergence of the first equation of (S’), we obtain

$$\begin{aligned} {{\text {div}}}F={{\text {div}}}(-{\varDelta }u_S+\nabla p_S)=-{\varDelta }({{\text {div}}}u_S)+{\varDelta }p_S={\varDelta }p_S. \end{aligned}$$

in distributions sense. Since \(p_S\in {H^1({\varOmega })}\) and \(C^\infty _0({\varOmega })\) is dense in \({H^1_0({\varOmega })}\), it follows that

$$\begin{aligned} \int _{\varOmega }\nabla p_S\cdot \nabla \psi =-\int _{\varOmega }({{\text {div}}}F)\psi \end{aligned}$$

for all \(\psi \in {H^1_0({\varOmega })}\). Together with (S’), (PP’) and \({H^1_0({\varOmega })}\subset Q\), we obtain

$$\begin{aligned} \left\{ \begin{array}{ll}{\displaystyle \int _{\varOmega }\nabla (u_S-u_PP):\nabla \varphi =-\int _{\varOmega }(\nabla (p_S-p_PP))\cdot \varphi } &{} {\text{ for } \text{ all } } \;\varphi \in {H^1_0({\varOmega })}^n, \\ {\displaystyle \int _{\varOmega }\nabla (p_S-p_PP)\cdot \nabla \psi =0} &{}{\text{ for } \text{ all } } \psi \in {H^1_0({\varOmega })}\end{array}\right. \end{aligned}$$
(A.4)

from the assumption \(\langle G,\psi \rangle =\int _{\varOmega }\nabla F\cdot \psi \). Putting \(\varphi :=u_S-u_PP\in {H^1_0({\varOmega })}^n\) in (A.4), we get

$$\begin{aligned} \begin{aligned} \Vert \nabla (u_S-u_PP)\Vert ^2_{{L^2({\varOmega })}^{n\times n}}&={\displaystyle -\int _{\varOmega }(\nabla (p_S-p_PP))\cdot (u_S-u_PP)}\\&\le \Vert \nabla (p_S-p_PP)\Vert _{{L^2({\varOmega })}^n}\Vert u_S-u_PP\Vert _{{L^2({\varOmega })}^n}. \end{aligned} \end{aligned}$$

Hence,

$$\begin{aligned} \Vert u_S-u_PP\Vert _{{H^1({\varOmega })}^n}\le c_1\Vert \nabla (p_S-p_PP)\Vert _{{L^2({\varOmega })}^n} \end{aligned}$$
(A.5)

holds. From the second equation of (A.4), we obtain

$$\begin{aligned}&\Vert \nabla (p_S-p_PP-\psi )\Vert ^2_{{L^2({\varOmega })}^n}\\&\quad ={\displaystyle \Vert \nabla (p_S-p_PP)\Vert ^2_{{L^2({\varOmega })}^n}+\Vert \nabla \psi \Vert ^2_{{L^2({\varOmega })}^n} -2\int _{\varOmega }\nabla (p_S-p_PP)\cdot \nabla \psi }\\&\quad =\Vert \nabla (p_S-p_PP)\Vert ^2_{{L^2({\varOmega })}^n}+\Vert \nabla \psi \Vert ^2_{{L^2({\varOmega })}^n}\\&\quad \ge \Vert \nabla (p_S-p_PP)\Vert ^2_{{L^2({\varOmega })}^n} \end{aligned}$$

for all \(\psi \in {H^1_0({\varOmega })}\). Thus we find

$$\begin{aligned} \Vert \nabla (p_S-p_PP)\Vert _{{L^2({\varOmega })}^n} \le \inf _{\psi \in {H^1_0({\varOmega })}^n}\left( \Vert \nabla (p_S-p_PP-\psi )\Vert _{{L^2({\varOmega })}^n}\right) . \end{aligned}$$

Since \(\gamma _0\) is surjective and the space \(\text{ Ker }(\gamma _0)={H^1_0({\varOmega })}\), \({H^1({\varOmega })}/{H^1_0({\varOmega })}\) and \(H^{1/2}({\varGamma })\) are isomorphic, there exists a constant \(c_2>0\) such that \(\Vert q\Vert _{{H^1({\varOmega })}/{H^1_0({\varOmega })}}\le c_2\Vert \gamma _0 q\Vert _{H^{1/2}({\varGamma })}\) for all \(q\in {H^1({\varOmega })}\). Hence, we obtain

$$\begin{aligned} \begin{aligned} \Vert \nabla (p_S-p_PP)\Vert _{{L^2({\varOmega })}^n}&\le {\displaystyle \inf _{\psi \in {H^1_0({\varOmega })}^n}\Vert \nabla (p_S-p_PP-\psi )\Vert _{{L^2({\varOmega })}^n}}\\&\le {\displaystyle \inf _{\psi \in {H^1_0({\varOmega })}^n}\Vert p_S-p_PP-\psi \Vert _{H^1({\varOmega })}}\\&=\Vert p_S-p_PP\Vert _{{H^1({\varOmega })}/{H^1_0({\varOmega })}}\\&\le c_2\Vert \gamma _0 p_S-\gamma _0 p_PP\Vert _{H^{1/2}({\varGamma })}. \end{aligned} \end{aligned}$$
(A.6)

Together with (A.5), we obtain \( \Vert u_S-u_PP\Vert _{{H^1({\varOmega })}^n} \le c_1c_2\Vert \gamma _0 p_S-\gamma _0 p_PP\Vert _{H^{1/2}({\varGamma })}. \) Moreover, if \(\gamma _0(p_S-p_PP)=0\), then \(p_PP=p_S\).

Next, we prove that there exists a constant \(c>0\) independent of \(\varepsilon \) such that \( \Vert u_S-u_\varepsilon \Vert _{{H^1({\varOmega })}^n} \le c\Vert \gamma _0 p_S-\gamma _0 p_\varepsilon \Vert _{H^{1/2}({\varGamma })}, \) and if \(\gamma _0(p_S-p_PP)=0\), then \(p_PP=p_\varepsilon \). Let \(w_\varepsilon :=u_S-u_\varepsilon \in {H^1_0({\varOmega })}^n\) and \(r_\varepsilon :=p_PP-p_\varepsilon \in Q\). By (S’), (PP’) and (ES’), we obtain

$$\begin{aligned} \left\{ \begin{array}{ll} {\displaystyle \int _{\varOmega }\nabla w_\varepsilon :\nabla \varphi +\int _{\varOmega }(\nabla r_\varepsilon )\cdot \varphi =-\int _{\varOmega }(\nabla (p_S-p_PP))\cdot \varphi }&{} {\text{ for } \text{ all } } \;\varphi \in {H^1_0({\varOmega })}^n, \\ {\displaystyle \varepsilon \int _{\varOmega }\nabla r_\varepsilon \cdot \nabla \psi +\int _{\varOmega }({{\text {div}}}w_\varepsilon )\psi =0} &{}{\text{ for } \text{ all } } \;\psi \in Q. \end{array}\right. \end{aligned}$$
(A.7)

Putting \(\varphi :=w_\varepsilon \) and \(\psi :=r_\varepsilon \) and adding the two equations of (A.7), we get

$$\begin{aligned} \Vert \nabla w_\varepsilon \Vert ^2_{{L^2({\varOmega })}^{n\times n}}+\varepsilon \Vert \nabla r_\varepsilon \Vert ^2_{{L^2({\varOmega })}^n} \le \Vert \nabla (p_S-p_PP)\Vert _{{L^2({\varOmega })}^n}\Vert w_\varepsilon \Vert _{{L^2({\varOmega })}^n} \end{aligned}$$
(A.8)

from \(\int _{\varOmega }(\nabla r_\varepsilon )\cdot w_\varepsilon =-\int _{\varOmega }({{\text {div}}}w_\varepsilon )r_\varepsilon \). Thus we find

$$\begin{aligned} \Vert w_\varepsilon \Vert _{{H^1({\varOmega })}^n}\le c_3\Vert \nabla (p_S-p_PP)\Vert _{{L^2({\varOmega })}^n}. \end{aligned}$$

Together with (A.6), we obtain

$$\begin{aligned} \Vert u_S-u_\varepsilon \Vert _{{H^1({\varOmega })}^n} =\Vert w_\varepsilon \Vert _{{H^1({\varOmega })}^n} \le c_2c_3\Vert \gamma _0 p_S-\gamma _0 p_PP\Vert _{H^{1/2}({\varGamma })}. \end{aligned}$$

Moreover, by (A.8), we obtain

$$\begin{aligned} \varepsilon \Vert p_PP-p_\varepsilon \Vert ^2_{L^2({\varOmega })}=\varepsilon \Vert r_\varepsilon \Vert ^2_{L^2({\varOmega })}\le c_4\Vert \nabla (p_S-p_PP)\Vert _{{L^2({\varOmega })}^n}\Vert w_\varepsilon \Vert _{{L^2({\varOmega })}^n}. \end{aligned}$$

Hence, if \(\gamma _0(p_S-p_PP)=0\), then \(p_PP=p_\varepsilon \). \(\square \)

We show that the sequence \(((u_\varepsilon ,p_\varepsilon ))_{\varepsilon >0}\) is bounded in \({H^1({\varOmega })}^n\times ({L^2({\varOmega })}/{{\mathbb {R}}})\). By the reflexivity of \({H^1({\varOmega })}^n\times ({L^2({\varOmega })}/{{\mathbb {R}}})\), the sequence \(((u_\varepsilon ,p_\varepsilon ))_{\varepsilon >0}\) has a subsequence converging weakly to somewhere in \({H^1({\varOmega })}^n\times ({L^2({\varOmega })}/{{\mathbb {R}}})\). It is sufficient to check that the limit satisfies (S’). Since the solution of (S’) is unique, the sequence \(((u_\varepsilon ,p_\varepsilon ))_{\varepsilon >0}\) converges weakly.

Proof of Theorem 4.2

We take \(u_b\in {H^1({\varOmega })}^n,f\in {H^{-1}({\varOmega })}^n\) and \(g\in Q^*\) as (A.1) and (A.2) in the proof of Theorem 2.7. We put \({{\tilde{u}}}_\varepsilon :=u_\varepsilon -u_b\in {H^1_0({\varOmega })}^n,{\tilde{p}}_\varepsilon :=p_\varepsilon -p_0\in Q\). Then we obtain

$$\begin{aligned} \left\{ \begin{array}{ll} {\displaystyle \int _{\varOmega }\nabla {\tilde{u}}_\varepsilon :\nabla \varphi +\int _{\varOmega }(\nabla {\tilde{p}}_\varepsilon )\cdot \varphi =\langle f,\varphi \rangle } &{} {\text{ for } \text{ all } } \;\varphi \in {H^1_0({\varOmega })}^n,\\ {\displaystyle \varepsilon \int _{\varOmega }\nabla {\tilde{p}}_\varepsilon \cdot \nabla \psi +\int _{\varOmega }({{\text {div}}}\tilde{u}_\varepsilon )\psi =\varepsilon \langle g,\psi \rangle } &{} {\text{ for } \text{ all } } \;\psi \in Q. \end{array}\right. \end{aligned}$$
(A.9)

Putting \(\varphi :={\tilde{u}}_\varepsilon ,\psi :={\tilde{p}}_\varepsilon \) and adding the two equations of (A.9), we get

$$\begin{aligned} \Vert \nabla {\tilde{u}}_\varepsilon \Vert ^2_{{L^2({\varOmega })}^{n\times n}}+\varepsilon \Vert \nabla \tilde{p}_\varepsilon \Vert ^2_{{L^2({\varOmega })}^n}\le & {} \Vert f\Vert _{{H^{-1}({\varOmega })}^n}\Vert \nabla \tilde{u}_\varepsilon \Vert _{{L^2({\varOmega })}^{n\times n}} \\&+\varepsilon \Vert g\Vert _{Q^*}\Vert \nabla \tilde{p}_\varepsilon \Vert _{{L^2({\varOmega })}^n} \end{aligned}$$

since \(\int _{\varOmega }(\nabla {\tilde{p}}_\varepsilon )\cdot \tilde{u}_\varepsilon =-\int _{\varOmega }({{\text {div}}}{\tilde{u}}_\varepsilon ){\tilde{p}}_\varepsilon \). Hence,

$$\begin{aligned} (\Vert {\tilde{u}}_\varepsilon \Vert _{{H^1({\varOmega })}^n})_{0<\varepsilon<1} \text{ and } (\Vert \sqrt{\varepsilon }{\tilde{p}}_\varepsilon \Vert _{H^1({\varOmega })})_{0<\varepsilon <1} \text{ are } \text{ bounded. } \end{aligned}$$
(A.10)

Moreover, by Lemma 4.1, we obtain

$$\begin{aligned} \Vert {\tilde{p}}_\varepsilon \Vert _{{L^2({\varOmega })}/{{\mathbb {R}}}}\le {c}(\Vert \nabla \tilde{u}_\varepsilon \Vert _{{L^2({\varOmega })}^{n\times n}}+\Vert f\Vert _{{H^{-1}({\varOmega })}^n}), \end{aligned}$$

i.e., \((\Vert {\tilde{p}}_\varepsilon \Vert _{{L^2({\varOmega })}/{{\mathbb {R}}}})_{0<\varepsilon <1}\) is bounded. By Theorem 3.2, \((\Vert u_\varepsilon \Vert _{{H^1({\varOmega })}^n})_{\varepsilon \ge 1}\) and \((\Vert {\tilde{p}}_\varepsilon \Vert _{{L^2({\varOmega })}/{{\mathbb {R}}}})_{\varepsilon \ge 1}\) are bounded, and thus \((\Vert u_\varepsilon \Vert _{{H^1({\varOmega })}^n})_{\varepsilon >0}\) and \((\Vert \tilde{p}_\varepsilon \Vert _{{L^2({\varOmega })}/{{\mathbb {R}}}})_{\varepsilon >0}\) are bounded.

Since \({H^1({\varOmega })}^n\times ({L^2({\varOmega })}/{{\mathbb {R}}})\) is reflexive and \(({\tilde{u}}_\varepsilon ,[\tilde{p}_\varepsilon ])_{0<\varepsilon <1}\) is bounded in \({H^1({\varOmega })}^n\times ({L^2({\varOmega })}/{{\mathbb {R}}})\), there exist \((u,p)\in {H^1({\varOmega })}^n\times ({L^2({\varOmega })}/{{\mathbb {R}}})\) and a subsequence of pairs \(({\tilde{u}}_{\varepsilon _k}, {\tilde{p}}_{\varepsilon _k})_{k\in {{\mathbb {N}}}}\subset {H^1_0({\varOmega })}^n\times Q\) such that

$$\begin{aligned} {\tilde{u}}_{\varepsilon _k}\rightharpoonup u~\mathrm{weakly~\text{ in } }{H^1({\varOmega })}^n,~ [{\tilde{p}}_{\varepsilon _k}]\rightharpoonup p~\mathrm{weakly~\text{ in } }{L^2({\varOmega })}/{{\mathbb {R}}}\quad \mathrm{as~}k\rightarrow \infty . \end{aligned}$$

Hence, from (A.9) with \(\varepsilon :=\varepsilon _k\), taking \(k\rightarrow \infty \), we obtain

$$\begin{aligned} \left\{ \begin{array}{ll}{\displaystyle \int _{\varOmega }\nabla u:\nabla \varphi +\langle \nabla p,\varphi \rangle =\langle f,\varphi \rangle } &{} {\text{ for } \text{ all } } \;\varphi \in {H^1_0({\varOmega })}^n\\ {\displaystyle \int _{\varOmega }({{\text {div}}}u)\psi =0}&{} {\text{ for } \text{ all } } \;\psi \in Q, \end{array}\right. \end{aligned}$$
(A.11)

where we have used that

$$\begin{aligned} |\varepsilon _k\int _{\varOmega }\nabla {\tilde{p}}_{\varepsilon _k}\cdot \nabla \psi |\le & {} \sqrt{\varepsilon _k}\Vert \sqrt{\varepsilon }{\tilde{p}}_\varepsilon \Vert _{H^1({\varOmega })}\Vert \psi \Vert _{H^1({\varOmega })}\rightarrow 0, \\ \int _{\varOmega }\nabla {\tilde{p}}_{\varepsilon _k}\cdot \varphi= & {} -\int _{\varOmega }[\tilde{p}_{\varepsilon _k}]{{\text {div}}}\varphi \rightarrow -\int _{\varOmega }p{{\text {div}}}\varphi =\langle \nabla p,\varphi \rangle \end{aligned}$$

as \(k\rightarrow \infty \). By (A.2), the first equation of (A.11) implies that

$$\begin{aligned} \int _{\varOmega }\nabla (u+u_b):\nabla \varphi +\langle \nabla (p+p_0),\varphi \rangle =\int _{\varOmega }F\cdot \varphi \end{aligned}$$

for all \(\varphi \in {H^1_0({\varOmega })}^n\). From the second equation of (A.11) and \(C^\infty _0({\varOmega })\subset Q\), \({{\text {div}}}(u+u_b)=0\) follows. Hence, we obtain that \((u+u_b,p+[p_0])\) satisfies (S’), i.e., \(u_S=u+u_b\) and \(p_S=p+[p_0]\). Then we have

$$\begin{aligned} u_{\varepsilon _k}-u_S= & {} u_{\varepsilon _k}-u-u_b=\tilde{u}_{\varepsilon _k}-u_S\rightharpoonup 0 \text{ weakly } \text{ in } {H^1({\varOmega })}^n, \\ {[}p_{\varepsilon _k}{]}-p_S= & {} [p_{\varepsilon _k}-p-p_0]=[\tilde{p}_{\varepsilon _k}]-p\rightharpoonup 0 \text{ weakly } \text{ in } {L^2({\varOmega })}/{{\mathbb {R}}}\end{aligned}$$

as \(k\rightarrow \infty \). Since any arbitrarily chosen subsequence of \(((u_\varepsilon ,[p_\varepsilon ]))_{0<\varepsilon <1}\) has a subsequence which converges to \((u_S,p_S)\), we can conclude the proof. \(\square \)

Using Theorem 4.2 and the Rellich–Kondrachov Theorem, it is easy to proveTheorem 4.3.

Proof of Theorem 4.3

We have from (ES’) and (S’) that

$$\begin{aligned} \left\{ \begin{array}{ll} {\displaystyle \int _{\varOmega }\nabla (u_\varepsilon -u_S):\nabla \varphi +\int _{\varOmega }(\nabla (p_\varepsilon -p_S))\cdot \varphi =0 } &{} {\text{ for } \text{ all } } \;\varphi \in {H^1_0({\varOmega })}^n,\\ {\displaystyle \varepsilon \int _{\varOmega }\nabla p_\varepsilon \cdot \nabla \psi +\int _{\varOmega }({{\text {div}}}u_\varepsilon )\psi =\varepsilon \langle G,\psi \rangle } &{} {\text{ for } \text{ all } } \;\psi \in Q. \end{array} \right. \end{aligned}$$

Putting \(\varphi :=u_\varepsilon -u_S\in {H^1_0({\varOmega })}^n\) and \(\tilde{p}_S:=p_S-p_0\in {H^1({\varOmega })}\), we get

$$\begin{aligned} \begin{aligned} \Vert \nabla (u_\varepsilon -u_S)\Vert ^2_{{L^2({\varOmega })}^{n\times n}}&={\displaystyle -\int _{\varOmega }(\nabla (p_\varepsilon -p_S))\cdot (u_\varepsilon -u_S)}\\&=\displaystyle -\int _{\varOmega }(\nabla (p_\varepsilon -p_0))\cdot (u_\varepsilon -u_S) \\&\quad +\int _{\varOmega }(\nabla (p_S-p_0))\cdot (u_\varepsilon -u_S) \\&={\displaystyle \int _{\varOmega }(p_\varepsilon -p_0){{\text {div}}}(u_\varepsilon -u_S) +\int _{\varOmega }(\nabla {\tilde{p}}_S)\cdot (u_\varepsilon -u_S)}\\&={\displaystyle \int _{\varOmega }(p_\varepsilon -p_0){{\text {div}}}u_\varepsilon +\int _{\varOmega }(\nabla \tilde{p}_S)\cdot (u_\varepsilon -u_S),} \end{aligned} \end{aligned}$$

since \(-\int _{\varOmega }(\nabla (p_\varepsilon -p_0)) \cdot (u_\varepsilon -u_S)= \int _{\varOmega }(p_\varepsilon -p_0){{\text {div}}}(u_\varepsilon -u_S)\) and \({{\text {div}}}u_S=0\). Thus,

$$\begin{aligned} \Vert \nabla (u_\varepsilon -u_S)\Vert ^2_{{L^2({\varOmega })}^{n\times n}} =\int _{\varOmega }(p_\varepsilon -p_0){{\text {div}}}u_\varepsilon +\int _{\varOmega }(\nabla \tilde{p}_S)\cdot (u_\varepsilon -u_S). \end{aligned}$$
(A.12)

Putting \(\psi :=p_\varepsilon -p_0\in Q\), we have

$$\begin{aligned} \varepsilon \int _{\varOmega }\nabla p_\varepsilon \cdot \nabla (p_\varepsilon -p_0)+\int _{\varOmega }({{\text {div}}}u_\varepsilon )(p_\varepsilon -p_0) =\varepsilon \langle G,p_\varepsilon -p_0\rangle . \end{aligned}$$

Hence,

$$\begin{aligned} \varepsilon \Vert \nabla (p_\varepsilon -p_0)\Vert ^2_{{L^2({\varOmega })}^n}&=-\varepsilon \int _{\varOmega }\nabla (p_\varepsilon -p_0)\cdot \nabla p_0 \nonumber \\&\quad -\int _{\varOmega }(p_\varepsilon -p_0){{\text {div}}}u_\varepsilon +\varepsilon \langle G,p_\varepsilon -p_0\rangle . \end{aligned}$$
(A.13)

Together with (A.12) and (A.13), we obtain

$$\begin{aligned} \begin{aligned}&\Vert \nabla (u_\varepsilon -u_S)\Vert ^2_{{L^2({\varOmega })}^{n\times n}}+ \varepsilon \Vert \nabla (p_\varepsilon -p_0)\Vert ^2_{{L^2({\varOmega })}^n}\\&\quad ={\displaystyle \int _{\varOmega }\nabla {\tilde{p}}_S\cdot (u_\varepsilon -u_S) -\varepsilon \int _{\varOmega }\nabla (p_\varepsilon -p_0)\cdot \nabla p_0 +\varepsilon \langle G,p_\varepsilon -p_0\rangle }\\&\quad \le \Vert \nabla {\tilde{p}}_S\Vert _{{L^2({\varOmega })}^n}\Vert u_\varepsilon -u_S\Vert _{{L^2({\varOmega })}^n} \\&\qquad +\varepsilon (\Vert \nabla p_0\Vert _{{L^2({\varOmega })}^n}+\Vert G\Vert _{Q^*})\Vert \nabla (p_\varepsilon -p_0)\Vert _{{L^2({\varOmega })}^n}. \end{aligned} \end{aligned}$$

By Theorem 4.2 and the Rellich–Kondrachov Theorem, there exists a sequence \((\varepsilon _k)_{k\in {{\mathbb {N}}}}\subset {{\mathbb {R}}}\) such that

$$\begin{aligned} u_{\varepsilon _k}\rightarrow u_S \text{ strongly } \text{ in } {L^2({\varOmega })}^n \quad \text{ as } k\rightarrow \infty . \end{aligned}$$

Therefore, by (A.10),

$$\begin{aligned} \begin{aligned} \Vert \nabla (u_{\varepsilon _k}-u_S)\Vert ^2_{{L^2({\varOmega })}^{n\times n}}&\le \Vert \nabla \tilde{p}_S\Vert _{{L^2({\varOmega })}^n}\Vert u_{\varepsilon _k}-u_S\Vert _{{L^2({\varOmega })}^n} \\&\quad +\varepsilon _k(\Vert \nabla {\tilde{p}}_0\Vert _{{L^2({\varOmega })}^n}+\Vert G\Vert _{Q^*})\Vert \nabla (p_{\varepsilon _k}-p_0)\Vert _{{L^2({\varOmega })}^n}\\&\rightarrow 0 \end{aligned} \end{aligned}$$

as \(k\rightarrow \infty \). This implies that

$$\begin{aligned} \Vert [p_{\varepsilon _k}]-p_S\Vert _{{L^2({\varOmega })}/{{\mathbb {R}}}}= & {} \Vert p_{\varepsilon _k}-p_S\Vert _{{L^2({\varOmega })}/{{\mathbb {R}}}} \\\le & {} c\Vert \nabla (u_{\varepsilon _k}-u_S)\Vert _{{L^2({\varOmega })}^{n\times n}} \rightarrow 0 \text{ as } k\rightarrow \infty \end{aligned}$$

by Lemma 4.1. Since any arbitrarily chosen subsequence of \(((u_\varepsilon ,[p_\varepsilon ]))_{0<\varepsilon <1}\) has a subsequence which converges to \((u_S,p_S)\), we can conclude the proof. \(\square \)

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kimura, M., Matsui, K., Muntean, A. et al. Analysis of a projection method for the Stokes problem using an \(\varepsilon \)-Stokes approach. Japan J. Indust. Appl. Math. 36, 959–985 (2019). https://doi.org/10.1007/s13160-019-00373-3

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s13160-019-00373-3

Keywords

Mathematics Subject Classification

Navigation