Skip to main content
Log in

Penalty Method for the Stationary Navier–Stokes Problems Under the Slip Boundary Condition

  • Published:
Journal of Scientific Computing Aims and scope Submit manuscript

Abstract

We consider the penalty method for the stationary Navier–Stokes equations with the slip boundary condition. The well-posedness and the regularity theorem of the penalty problem are investigated, and we obtain the optimal error estimate \(O(\epsilon )\) in \(H^k\)-norm, where \(\epsilon \) is the penalty parameter. We are concerned with the finite element approximation with the P1b / P1 element to the penalty problem. The well-posedness of discrete problem is proved. We obtain the error estimate \(O(h+\sqrt{\epsilon }+h/\sqrt{\epsilon })\) for the non-reduced-integration scheme with \(d=2,3\), and the reduced-integration scheme with \(d=3\), where h is the discretization parameter and d is the spatial dimension. For the reduced-integration scheme with \(d=2\), we prove the convergence order \(O(h+\sqrt{\epsilon }+h^2/\sqrt{\epsilon })\). The theoretical results are verified by numerical experiments.

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
Fig. 9

Similar content being viewed by others

References

  1. Amrouche, C., Girault, V.: On the existence and regularity of the solution of the Stokes problem in arbitrary dimension. Proc. Jpn. Acad. 67, 171–175 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  2. Babuška, I.: The finite element method with penalty. Math. Comput. 27, 221–228 (1973)

    Article  MathSciNet  MATH  Google Scholar 

  3. Bänsch, E., Deckelnick, K.: Optimal error estimates for the Stokes and Navier–Stokes equations with slip boundary condition. M2AN 33, 923–938 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  4. Bothe, D., Köhne, M., Prüss, J.: On a class of energy preserving boundary conditions for incompressible Newtonian flows. SIAM J. Math. Anal. 45, 3768–3822 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  5. Boyer, F., Fabrie, P.: Mathematical Tools for the Study of the Incompressible Navier–Stokes Equations on Related Models. Springer, New York (2012)

    MATH  Google Scholar 

  6. Brenner, S.C., Scott, L.R.: The Mathematical Theory of Finite Element Methods. Springer, New York (2002)

    Book  MATH  Google Scholar 

  7. Caglar, A., Liakos, A.: Weak imposition of boundary conditions for the Navier–Stokes equations by a penalty method. Int. J. Numer. Methods Fluids 61, 411–431 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  8. Cattabriga, L.: Su un problema al contorno relativo al sistema di equazioni di Stokes. Rend. Sem. Mat. Univ. Padova 31, 1–33 (1961)

    MathSciNet  MATH  Google Scholar 

  9. Delfour, M.C., Zolésio, J.P.: Shapes and Geometries-Metrics, Analysis, Differential Calculus, and Optimization, 2nd edn. SIAM, Philadelphia (2011)

    MATH  Google Scholar 

  10. Dione, I., Urquiza, J.M.: Penalty: finite element approximation of Stokes equations with slip boundary conditions. Numer. Math. 129, 587–610 (2015)

    Article  MathSciNet  MATH  Google Scholar 

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

    Book  MATH  Google Scholar 

  12. Gilbarg, D., Trudinger, N.S.: Elliptic Partial Differential Equations of Second Order. Springer, New York (1998)

    MATH  Google Scholar 

  13. Grisvard, P.: Elliptic Problems in Nonsmooth Domains, Monographs and Studies in Mathematics, 24. Pitman, New York (1985)

    MATH  Google Scholar 

  14. Hecht, F., Pironneau, O., Le Hyaric, F., Ohtsuka, K.: Freefem++, available online at www.freefem.org

  15. Knobloch, P.: Variational crimes in a finite element discretization of 3D stokes equations with nonstandard boundary conditions, east–west. J. Numer. Math. 7, 133–158 (1999)

    MathSciNet  MATH  Google Scholar 

  16. Kashiwabara, T., Oikawa, I., Zhou, G.: Penalty method with P1/P1 finite element approximation for the Stokes equations under slip boundary condition, arXiv:1505.06540

  17. Lions, J.L., Magenes, E.: Non-Homogeneous Boundary Value Problems and Applications, vol. I. Springer, New York (1972)

    Book  MATH  Google Scholar 

  18. Logg, A., Mardal, K.-A., Well, G.N., et al.: Automated Solution of Differential Equations by the Finite Element Method. Springer, New York (2012)

    Book  Google Scholar 

  19. Saito, N.: On the stokes equation with the leak and slip boundary conditions of friction type: regularity of solutions. Publ. RIMS, Kyoto Univ. 40, 345–383 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  20. Shibata, Y., Shimizu, S.: On a generalized resolvent estimate for the Stokes system with Robin boundary condition. J. Math. Soc. Jpn. 59, 1–34 (2007)

    Article  MathSciNet  Google Scholar 

  21. Tabata, M.: Finite element approximation to infinite Prandtl number Boussinesq equations with temperature-dependent coefficients-thermal convection problems in a spherical shell. Future Gener. Comput. Syst. 22, 521–531 (2006)

    Article  Google Scholar 

  22. Tabata, M., Suzuki, A.: A stabilized finite element method foe the Rayleigh–Bénard equations with infinite Prandtl number in a spherical shell. Comput. Methods Appl. Mech. Eng. 190, 387–402 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  23. Verfürth, R.: Finite element approximation of incompressible Navier–Stokes equations with slip boundary condition. Numer. Math. 50, 697–721 (1987)

    Article  MathSciNet  MATH  Google Scholar 

  24. Wloka, J.: Partial Differential Equations. Cambridge University Press, Cambridge (1987)

    Book  MATH  Google Scholar 

  25. Ženíšek, A.: Nonlinear Elliptic and Evolution Problems and Their Finite Element Approximations. Academic Press, Waltham (1990)

    MATH  Google Scholar 

  26. Zhang, S.: Analysis of finite element domain embedding methods for curved domains using uniform grids. SIAM J. Numer. Anal. 46, 2843–2866 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  27. Zhou, G., Saito, N.: Analysis of the fictitious domain method with penalty for elliptic problems. Jpn. J. Ind. Appl. Math. 31, 57–85 (2014)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgments

The authors would like to thank the anonymous reviewers for their valuable comments and suggestions to improve the quality of the paper.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Guanyu Zhou.

Additional information

The first author was supported by JST, CREST and by JSPS KAKENHI Grant Number 23340023. The second author was supported by JST, CREST The third author was supported by JSPS KAKENHI Grant Numbers 24224004, 26800089.

The Regularity for the Penalty Problem of the Stokes Equations

The Regularity for the Penalty Problem of the Stokes Equations

In this appendix we prove a regularity result for the Stokes equations subject to a penalized slip boundary condition. This is useful when we treat the nonlinearity in the Navier–Stokes equations as a perturbation term. We emphasize that the regularity estimate [see (6.2) below] is independent of the penalty coefficient \(\epsilon ^{-1}\).

Theorem 6.1

For \(k \in \mathbb {N} \cup \{0\}\), we assume \(\varOmega \) is \(C^{k+4}\) smooth and \(f \in H^k(\varOmega )\), \(g \in H^{k+ \frac{1}{2}}(\varGamma )\). Let \((u_\epsilon ,p_\epsilon ) \in V \times Q\) be the solution of the Stokes problem with penalty, denoted by \(\mathbf{(S_\epsilon )}\):

$$\begin{aligned}&-\nu \varDelta u_\epsilon + \nabla p_\epsilon = f, \quad \nabla \cdot u_\epsilon = 0, \quad \text { in } \varOmega ,\end{aligned}$$
(6.1a)
$$\begin{aligned}&u_\epsilon |_C = 0, \quad \tau (u_\epsilon ,p_\epsilon ) + \epsilon ^{-1} u_{\epsilon n}n = g \quad \text { on } \varGamma . \end{aligned}$$
(6.1b)

Then we have

$$\begin{aligned} \Vert u_\epsilon \Vert _{H^{k+2}(\varOmega )} + \Vert p_\epsilon \Vert _{H^{k+1}(\varOmega )} \le C, \end{aligned}$$
(6.2)

where C is independent of penalty coefficient \(\epsilon ^{-1}\).

The proof of Theorem 6.1 is based on the induction method. Firstly, we show the existence of the weak solution of (6.1). The weak form of (6.1) reads as

$$\begin{aligned} a(u_\epsilon ,v) + b(v,p_\epsilon ) + \frac{1}{\epsilon }\int _\varGamma u_{\epsilon n} v_n~d\varGamma&= (f,v) + \int _\varGamma g \cdot v~d\varGamma ,&\forall&v \in V, \end{aligned}$$
(6.3a)
$$\begin{aligned} b(u_\epsilon ,q)&= 0,&\forall&q \in Q. \end{aligned}$$
(6.3b)

Lemma 6.1

Given \(f \in V'\) and \(g \in H^{-\frac{1}{2}}(\varGamma )\), there exists a unique solution \((u_\epsilon .p_\epsilon ) \in V \times Q\) of (6.3), with

$$\begin{aligned} \Vert u_\epsilon \Vert _{H^1(\varOmega )} + \Vert p_\epsilon \Vert _{L^2(\varOmega )} + \epsilon ^{-\frac{1}{2}}\Vert u_{\epsilon n}\Vert _{L^2(\varGamma )} \le C\left( \Vert f\Vert _{V'} + \Vert g\Vert _{H^{-\frac{1}{2}}(\varGamma )}\right) . \end{aligned}$$
(6.4)

Proof

By Korn’s inequality and the Lax-Milgram theorem, there exists a unique solution \(u_\epsilon \) of

$$\begin{aligned} a(u_\epsilon ,v) + \frac{1}{\epsilon }\int _\varGamma u_{\epsilon n} v_n~d\varGamma = (f,v) + \int _\varGamma g \cdot v~d\varGamma , \quad \forall v \in V^\sigma . \end{aligned}$$
(6.5)

with the estimate

$$\begin{aligned} \Vert u_\epsilon \Vert _{H^1(\varOmega )} + \epsilon ^{-\frac{1}{2}}\Vert u_{\epsilon n}\Vert _{L^2(\varGamma )} \le C\left( \Vert f\Vert _{V'} + \Vert g\Vert _{H^{-\frac{1}{2}}(\varGamma )}\right) \end{aligned}$$
(6.6)

By the inf-sup condition of b, there exists a unique \(\mathring{p}_\epsilon \in \mathring{Q}\) satisfying

$$\begin{aligned}&a(u_\epsilon ,v) + b(v,\mathring{p}_\epsilon ) = (f,v), \quad \forall v \in H_0^1(\varOmega )^d, \end{aligned}$$
(6.7)
$$\begin{aligned}&\Vert \mathring{p}_\epsilon \Vert _{L^2(\varOmega )} \le C(\Vert u_\epsilon \Vert _{H^1(\varOmega )} + \Vert f\Vert _{V'}). \end{aligned}$$
(6.8)

For all \(\phi \in C^\infty (\overline{\varOmega })^d \cap V\) with \(\int _\varGamma \phi _nd\varGamma = 1\), we set

$$\begin{aligned} k_\epsilon = a(u_\epsilon ,\phi ) + b(\phi ,\mathring{p}_\epsilon ) + \epsilon ^{-1} \int _\varGamma u_{\epsilon n} \phi _n~d\varGamma - (f,\phi ) - \int _\varGamma g\cdot \phi ~d\varGamma . \end{aligned}$$
(6.9)

With a similar argument to Proposition 2.2, we see that \(k_\epsilon \) is independent of \(\phi \), and \((u_\epsilon ,p_\epsilon )\) with \(p_\epsilon = \mathring{p}_\epsilon + k_\epsilon \) is a solution of (6.3).

Substituting \(\phi = k_\epsilon \tilde{n}\) into (6.3), where \(\tilde{n}\) is a smooth extension of \(n \in C^3(\varGamma )\) to \(\varOmega \), and noticing that \(\int _\varGamma u_{\epsilon n} k_\epsilon n \cdot n d\varGamma = 0\), we obtain

$$\begin{aligned} \begin{aligned} |k_\epsilon |^2|\varGamma |&= -b(k_\epsilon \tilde{n},k_\epsilon ) \\&= a(u_\epsilon ,\phi ) + b(\phi ,\mathring{p}_\epsilon ) - (f,\phi ) - \int _\varGamma g\cdot \phi ~d\varGamma \\&\le C \left( \Vert u_\epsilon \Vert _{H^1(\varOmega )}+\Vert \mathring{p}_\epsilon \Vert _{L^2(\varOmega )} + \Vert f\Vert _{V'}+\Vert g\Vert _{H^{-\frac{1}{2}}(\varGamma )}\right) . \end{aligned} \end{aligned}$$
(6.10)

Combining (6.6), (6.8) and (6.10), we get (6.4). \(\square \)

Proof of Theorem 6.1

For any interior sub-domain \(\omega \subset \varOmega \) or \(\omega \) near the boundary \(\gamma \), we have (cf. [8])

$$\begin{aligned} \Vert u_\epsilon \Vert _{H^{k+2}(\omega )} + \Vert p_\epsilon \Vert _{H^{k+1}(\omega )} \le C\Vert f\Vert _{H^k(\varOmega )}. \end{aligned}$$

We then consider the regularity near \(\varGamma \). There exist \(\{W_i\}_{i=1}^N \subset \mathbb {R}^d\) covering \(\varGamma \), and \(\{\theta _i\}_{i=1}^N\) with \(\theta _i \subset C_0^\infty (W_i)\), \(\theta _i \ge 0\), \(\sum _{i=1}^N \theta _i = 1\), and \(\text {supp }\theta _i \subsetneq W_i\). We will prove the \(H^{k+2}\)-regularity of \(\theta _i u_\epsilon \) for every \(W_i\), which implies (6.2). In the following, we omit the subscript i of \(W_i\) and \(\theta _i\). Setting \((\bar{u},\bar{p}) = (\theta u_\epsilon , \theta p_\epsilon )\) and \(e_{ij}(u) = \frac{1}{2}(\frac{\partial u_i}{\partial x_j} + \frac{\partial u_j}{\partial x_i})\), we see that

$$\begin{aligned} e_{ij}(\bar{u}) = \theta e_{ij}(u_\epsilon ) + \frac{1}{2}\left( \frac{\partial \theta }{\partial x_j} u_{\epsilon i}+ \frac{\partial \theta }{\partial x_i}u_{\epsilon j}\right) , \quad \nabla \cdot (\theta v) = v \cdot \nabla \theta + \theta \nabla \cdot v. \end{aligned}$$

We re-set \(V = H^1(W \cap \varOmega )^d \cap \{v|_{\partial (W \cap \varOmega ) \backslash \varGamma } = 0\}\), \(Q = L^2(W \cap \varOmega )\), and re-define the bilinear forms \(a(\cdot ,\cdot )\), \(b(\cdot ,\cdot )\) in the domain \(\varOmega \cap W\). From (6.3), we have (here and hereafter the summation convention is employed),

$$\begin{aligned}&\begin{aligned}&a(\bar{u},v) + b(v,\bar{p}) + \frac{1}{\epsilon }\int _{\varGamma \cap W} \bar{u}_n v_n d \varGamma \\&\quad = a(u_\epsilon ,\theta v) + b(\theta v,p_\epsilon ) + \frac{1}{\epsilon }\int _{\varGamma \cap W} u_{\epsilon n} \theta v_n d \varGamma + \nu \int _{\varOmega \cap W} o_{ij}(\theta ,u_\epsilon ) e_{ij}(v)dx \\&\qquad - \nu \int _{\varOmega \cap W} o_{ij}(\theta ,v) e_{ij}(u_\epsilon )dx + \int _{\varOmega \cap W} v \cdot \nabla \theta p_\epsilon ~dx\\&\quad = (\theta f + \nabla \theta p_\epsilon , v)_{W \cap \varOmega } + \nu (o_{ij}(\theta ,u_\epsilon ), e_{ij}(v))_{W \cap \varOmega } \\&\qquad + \nu (o_{ij}(\theta ,v), e_{ij}(u_\epsilon ))_{W \cap \varOmega } + \int _{\varGamma \cap W} \theta g \cdot v~d\varGamma , \quad \forall v \in V, \end{aligned} \end{aligned}$$
(6.11a)
$$\begin{aligned}&b(\bar{u},q) = -(u_\epsilon \cdot \nabla \theta ,q)_{\varOmega \cap W}, \quad \forall q \in L^2(W \cap \varOmega ), \end{aligned}$$
(6.11b)

where \(o_{ij}(\theta ,u) = \frac{\partial \theta }{\partial x_j} u_i + \frac{\partial \theta }{\partial x_i} u_j\). With integration by parts, we obtain

$$\begin{aligned}&-\nabla \cdot \sigma (\bar{u}, \bar{p}) = \bar{f}, \quad \nabla \cdot \bar{u} = \bar{h} , \quad \text { in } W \cap \varOmega , \end{aligned}$$
(6.12a)
$$\begin{aligned}&\bar{u} = 0, \quad \text { on } \partial (W \cap \varOmega )\backslash \overline{\varGamma }, \quad \tau (\bar{u}, \bar{p}) + \epsilon ^{-1}\bar{u}_n n = \bar{g} \quad \text { on } \partial (W \cap \varOmega ) \cap \varGamma , \end{aligned}$$
(6.12b)

where

$$\begin{aligned}&\bar{f} = \theta f + \nabla \theta p_\epsilon - \nu ((u_\epsilon \cdot \nabla )\nabla \theta + 2 (\nabla \theta \cdot \nabla )u_\epsilon + \varDelta \theta u_\epsilon + \nabla u_\epsilon \nabla \theta ),\\&\bar{g} = \theta g - \nu ( u_{\epsilon n} \nabla \theta + (\nabla \theta \cdot n)u_\epsilon ), \quad \bar{h} = u_\epsilon \cdot \nabla \theta . \end{aligned}$$

There exists a \(C^{k+3}\)-diffeomorphism \(\varvec{\Phi }\) (cf. [19, Proof of Lemma 4.1], [24] ), such that \(0 < c \le |\text {Jac }\varvec{\Phi }| \le C\), \(|\text {Jac }\varvec{\Phi }||\text {Jac }\varvec{\Phi }^{-1}| = 1\), and

  1. (1)

    \(\varvec{\Phi }(W \cap \varOmega ) = Q_R := \{y = (y',y_d) \in \mathbb {R}^{d-1} \times \mathbb {R} : |y'|<R, \ 0 < y_d < R\}\);

  2. (2)

    \(\varvec{\Phi }(W \cap \varGamma ) = S_R := \{y = (y',y_d) \in \mathbb {R}^{d-1} \times \mathbb {R} : |y'|<R, \ y_d =0\}\);

  3. (3)

    \( \frac{\partial \varPhi _d}{\partial x_j} = \frac{\partial \varPhi _j}{\partial x_d} = 0, \quad \frac{\partial \varPhi _d}{\partial x_d} = -1, \text { on } W \cap \varGamma \ (j = 1,\ldots ,d-1)\);

  4. (4)

    \(\varvec{\Phi } : n(x) \mapsto \tilde{n}(y) = (0,\ldots ,0,-1) \text { for } x \in W \cap \varGamma .\)

Here, the \(C^{k+4}\)-smoothness of \(\varGamma \) is sufficient to obtain (3) and (4) (cf. [24, §1.2.4, Theorem 2.12]).

We set \(y = \varvec{\Phi }(x) = (\varPhi _1(x), \ldots , \varPhi _d(x))\), and

$$\begin{aligned}&\tilde{U}(y) = \bar{u}(x), \quad \tilde{P}(y) = \bar{p}(x),\\&\tilde{F}(y) = \bar{f}(x) |\text {Jac }\varvec{\Phi }|, \quad \tilde{G}(y) = (\theta g)(x) \sqrt{\mathrm {det}\,A} , \quad \tilde{H}(y) = \bar{h}(x) |\text {Jac }\varvec{\Phi }|,\\&\tilde{u}_\epsilon (y) = u_\epsilon (x), \quad \tilde{p}_\epsilon (y) = p_\epsilon (x), \quad \tilde{\theta }(y) = \theta (x), \end{aligned}$$

where A is a \((d-1)\times (d-1)\) matrix, whose components are given by \(A_{ij} = \frac{\partial \varvec{\Phi }^{-1}}{\partial y_i}\cdot \frac{\partial \varvec{\Phi }^{-1}}{\partial y_j}.\) Under the assumption of \(k=0\) and (6.4), we see that

$$\begin{aligned} \Vert \tilde{F}\Vert _{L^2(Q_R)} + \Vert \tilde{G}\Vert _{H^\frac{1}{2}(S_R)} + \Vert \tilde{H}\Vert _{H^1(Q_R)} \le C. \end{aligned}$$

We introduce \(K(Q_R) = \{ \varphi \in H^1(Q_R)^d : \varphi (y) = 0 \text { for } |y'| = R, \ y_d = R\}.\) Then \((\tilde{U},\tilde{P})\) satisfies

$$\begin{aligned}&\begin{aligned} \tilde{a}(\tilde{U},\tilde{\varphi }) +&\tilde{b}(\tilde{\varphi },\tilde{P}) + \frac{1}{\epsilon }\int _{S_R} \tilde{U}_d \tilde{\varphi }_d \sqrt{\mathrm {det}\,A} ~dy' \\ =&(\tilde{F}, \tilde{\varphi })_{Q_R} + \int _{S_R} \tilde{G}\cdot \tilde{\varphi } ~dy', \quad \forall \tilde{\varphi } \in K(Q_R), \end{aligned} \end{aligned}$$
(6.13a)
$$\begin{aligned}&\tilde{b}(\tilde{U},\tilde{q}) = (\tilde{H}, \tilde{q})_{Q_R}, \quad \forall \tilde{q} \in L^2(Q_R). \end{aligned}$$
(6.13b)

Here we have put

$$\begin{aligned}&\tilde{a}(\tilde{U},\tilde{\varphi }) = 2\nu \int _{Q_R} \tilde{e}_{ij}(\tilde{U}) \tilde{e}_{ij}(\tilde{\varphi })|\text {Jac } \varvec{\Phi } |~dy, \quad \tilde{e}_{ij}(\tilde{U}) = \frac{1}{2}\left( \frac{\partial \varPhi _k}{\partial x_j} \frac{\partial \tilde{U}_i}{\partial y_k} + \frac{\partial \varPhi _k}{\partial x_i} \frac{\partial \tilde{U}_j}{\partial y_k}\right) ;\nonumber \\ \end{aligned}$$
(6.14)
$$\begin{aligned}&\tilde{b}(\tilde{U},\tilde{q}) = - \int _{Q_R} \tilde{q} \frac{\partial \varPhi _k}{\partial x_j} \frac{\partial \tilde{U}_j}{\partial y_k} |\text {Jac } \varvec{\Phi } |~dy. \end{aligned}$$
(6.15)

Now we consider the case of \(k=0\). According the smoothness assumption on the data, we have \(\Vert \tilde{F}\Vert _{L^2(Q_R)} + \Vert \tilde{G}\Vert _{H^\frac{1}{2}(S_R)} \le C\). Substituting \(\tilde{\varphi } = \tilde{U}\) into (6.13) and using Korn’s inequality \(\tilde{a}(\cdot ,\cdot ) \ge \tilde{\alpha }_1 \Vert \cdot \Vert _{H^1(Q_R)}^2\), we obtain

$$\begin{aligned} \Vert \tilde{U}\Vert _{H^1(Q_R)}^2 + \epsilon ^{-1}\Vert \tilde{U}_d\Vert _{L^2(S_R)}^2 \le C. \end{aligned}$$
(6.16)

For \( \zeta > 0\), and \(v \in K(Q_R)\), we introduce a difference quotient operator \(D_\zeta ^i\) by

$$\begin{aligned} v(x_i+\zeta ) = v(x_1,\cdots ,x_i+\zeta ,\cdots ,x_d), \quad D_\zeta ^i v = (v(x_i+\zeta )-v(x))/\zeta . \end{aligned}$$

The following facts are well known:

$$\begin{aligned}&(w,D_{-\zeta }^i v) = (D_\zeta ^i w,v), \quad \Vert D_\zeta ^i w\Vert _{L^p(Q_R)} \le C \Vert \nabla _{y_i} w \Vert _{L^p(Q_R)}, \quad p \in (1,\infty ), \end{aligned}$$
(6.17)
$$\begin{aligned}&D_\zeta ^i(w(x)v(x)) = (D_\zeta ^i w) v(x_i+\zeta ) + w(x) (D_\zeta ^i v) ,\quad \forall w,v \in K(Q_R). \end{aligned}$$
(6.18)

Substituting \(\tilde{\varphi } = D_{-\zeta }^i D_\zeta ^i \tilde{U}\) (\(i = 1, \ldots ,d-1\)) into (6.13) and using the above facts, together with Sobolev’s and Poincaré’s inequalities, we have the following five estimates:

$$\begin{aligned}&\begin{aligned} |(\tilde{F} ,D_{-\zeta }^i D_\zeta ^i \tilde{U})_{Q_R} | \le C\Vert D_{-\zeta }^i D_\zeta ^i \tilde{U}\Vert _{L^2(Q_R)}, \end{aligned} \end{aligned}$$
(6.19)
$$\begin{aligned}&\begin{aligned} \left| \int _{S_R} \tilde{G}\cdot \left( D_{-\zeta }^i D_\zeta ^i \tilde{U}\right) ~dy'\right|&\le C\left( \Vert D_\zeta ^i \tilde{G} \Vert _{H^{-\frac{1}{2}}(S_R)} + \Vert \tilde{G} \Vert _{H^{-\frac{1}{2}}(S_R)}\right) \Vert D_\zeta ^i \tilde{U}\Vert _{H^\frac{1}{2}(S_R)}\\&\le C \Vert \tilde{G} \Vert _{H^\frac{1}{2}(S_R)} \Vert D_\zeta ^i \tilde{U}\Vert _{H^1(Q_R)} \quad (\text {cf. [19, Lemma 6.1]})\\&\le C \Vert \nabla D_\zeta ^i \tilde{U} \Vert _{L^2(Q_R)}, \\ \end{aligned} \end{aligned}$$
(6.20)
$$\begin{aligned}&\begin{aligned}&\tilde{a}\left( \tilde{U}, D_{-\zeta }^i D_\zeta ^i \tilde{U}\right) \ge \tilde{a}\left( D_\zeta ^i \tilde{U},D_\zeta ^i \tilde{U}\right) - C_2\Vert \tilde{u}_\epsilon \Vert _{H^1(Q_R)} \Vert \nabla D_\zeta ^i \tilde{U}\Vert _{L^2(Q_R)} \\&\qquad \ge C_1 \Vert \nabla D_\zeta ^i \tilde{U}\Vert _{L^2(Q_R)} ^2 - C_2\Vert \tilde{u}_\epsilon \Vert _{H^1(Q_R)} \Vert \nabla D_\zeta ^i \tilde{U}\Vert _{L^2(Q_R)} \quad (\text {cf. [19, (4.17)]}), \end{aligned} \end{aligned}$$
(6.21)
$$\begin{aligned}&\begin{aligned} \frac{1}{\epsilon }\int _{S_R} \tilde{U}_d \left( D_{-\zeta }^i D_\zeta ^i \tilde{U}_d\right) \sqrt{\mathrm {det}\,A}~dy'&\ge C_3\frac{1}{\epsilon }\int _{S_R} |D_\zeta ^i \tilde{U}_{d}|^2 ~dy' - C_4 \frac{1}{\epsilon } \int _{S_R} |D_\zeta ^i \tilde{U}_{d}| |\tilde{U}_d|~dy'\\&\ge C_5\epsilon ^{-1}\Vert D_\zeta ^i \tilde{U}_{d}\Vert ^2_{L^2(S_R)} - C \quad (\text {by (6.16)}), \end{aligned} \end{aligned}$$
(6.22)
$$\begin{aligned}&\begin{aligned}&|\tilde{b}\left( D_{-\zeta }^i D_\zeta ^i \tilde{U},\tilde{P}\right) | = \left| \left( \frac{\partial \tilde{U}_j}{\partial y_k}, D_{\zeta }^i D_{-\zeta }^i \left( \tilde{P} \frac{\partial \varPhi _k}{\partial x_j} |\text {Jac } \varvec{\Phi } | \right) \right) _{Q_R} \right| \\&\qquad = \left| \left( \tilde{H}, \left( \frac{\partial \varPhi _k}{\partial x_j} |\text {Jac } \varvec{\Phi } |\right) ^{-1} D_{\zeta }^i D_{-\zeta }^i \left( \tilde{P} \frac{\partial \varPhi _k}{\partial x_j} |\text {Jac } \varvec{\Phi } | \right) \right) _{Q_R} \right| \quad (\text {by (6.13b)}) \\&\qquad \le C\left( \Vert \nabla D_\zeta ^i \tilde{H}\Vert _{L^2(Q_R)} + \Vert \tilde{H}\Vert _{H^1(Q_R)}\right) \Vert \tilde{P}\Vert _{L^2(Q_R)} \\&\qquad \le C \Vert \nabla D_\zeta ^i \tilde{U}\Vert _{L^2(Q_R)} \quad (\text {by } \tilde{H}(y) = (u_\epsilon \cdot \nabla \theta )(x)). \end{aligned} \end{aligned}$$
(6.23)

Combining (6.19)–(6.23), we conclude, for all \(i = 1,\ldots ,d-1\),

$$\begin{aligned} \Vert \nabla D_\zeta ^i \tilde{U}\Vert _{L^2(Q_R)}^2 + \epsilon ^{-1} \Vert D_\zeta ^i \tilde{U}_{d}\Vert ^2_{L^2(S_R)} \le C, \end{aligned}$$
(6.24)

which implies (passing to the limit \(\zeta \rightarrow 0\)), for all \(j = 1,\ldots ,d\) and \(i = 1,\ldots ,d-1\),

$$\begin{aligned} \Vert \nabla _{y_j} \nabla _{y_i} \tilde{U}\Vert _{L^2(Q_R)}^2 + \epsilon ^{-1} \Vert \nabla _{y_i} \tilde{U}_{d}\Vert ^2_{L^2(S_R)} \le C. \end{aligned}$$
(6.25)

From (6.25) and the strong from of (6.13) (which can be obtained by the integration by parts), we have

$$\begin{aligned} \Vert \nabla _{y_d} \nabla _{y_d} \tilde{U}\Vert _{L^2(Q_R)} \le C\Vert \tilde{F}\Vert _{L^2(Q_R)} + C \sum _{i+j \le 2d-1}\Vert \nabla _{y_j} \nabla _{y_i} \tilde{U}\Vert _{L^2(Q_R)} \le C . \end{aligned}$$
(6.26)

Hence, we have proved the case of \(k=0\). As a result of (6.12b), we have

$$\begin{aligned} \Vert u_{\epsilon n}\Vert _{H^\frac{1}{2}(\varGamma )} \le C\epsilon . \end{aligned}$$
(6.27)

Let us show the case of \(k=1\), which is equivalent to prove \(\Vert \nabla _{y_l} \tilde{U}\Vert _{H^2}\le C\), for \(l = 1,\ldots ,d\). First, we calculate the equation of \(\nabla _{y_l} \tilde{U}\), for \(l = 1,\ldots ,d-1\). Since \(\text {supp } \theta \subsetneq W\), compact, we have

$$\begin{aligned} \text {supp } \tilde{U} \subsetneq W, \text { compact }; \quad \text {supp } D_\zeta ^l \tilde{U} \subsetneq W, \text { compact, for all } l = 1,\ldots , d-1. \end{aligned}$$

Therefore, for any \( \tilde{v} \in K(Q_R)\), \(\tilde{q}_1 \in L^2(Q_R)\), we can substitute \(\tilde{\varphi } = D_{-\zeta }^l \tilde{v}\), \(\tilde{q} = D_{-\zeta }^l \tilde{q}_1\) into (6.13) (here, \( \tilde{q}_1\) and \(\tilde{v}\) are extended to \(\mathbb {R}^{d-1} \times (0,R)\) by zero). Then, applying (6.17), (6.18), and passing to the limit \(\zeta \rightarrow 0\), we obtain

$$\begin{aligned}&\begin{aligned}&\tilde{a}\left( \nabla _{y_l} \tilde{U},\tilde{v}\right) + \tilde{b}\left( \tilde{v},\nabla _{y_l} \tilde{P}\right) + \frac{1}{\epsilon }\int _{S_R} \left( \nabla _{y_l} \tilde{U}_d\right) \tilde{v}_d \sqrt{\mathrm {det}\,A} ~dy' \\&\quad = \left( \tilde{J}_l,\tilde{v}\right) _{Q_R} + \int _{S_R} \tilde{L}_l \cdot \tilde{v}~dy' - \frac{1}{\epsilon }\int _{S_R} \tilde{U}_d \tilde{v}_d \nabla _{y_l} \sqrt{\mathrm {det}\,A} ~dy', \quad \forall \tilde{v} \in K(Q_R), \end{aligned} \end{aligned}$$
(6.28a)
$$\begin{aligned}&\tilde{b}(\nabla _{y_l} \tilde{U},\tilde{q}_1) = (\tilde{H}_l ,\tilde{q}_1)_{Q_R}, \quad \forall \tilde{q}_1 \in L^2(Q_R), \end{aligned}$$
(6.28b)

where

$$\begin{aligned}&\tilde{J}_l = d_{1l} \frac{\partial \tilde{F} }{\partial y_l} + d_{2l} \frac{\partial \tilde{P} }{\partial y_l} + d_{3lik} \frac{\partial ^2 \tilde{U}_{i} }{\partial y_l \partial y_k} + d_{4li}\frac{\partial \tilde{U}_{i}}{\partial y_l} + d_{5l} \tilde{P} + d_{6l}\tilde{F}, \end{aligned}$$
(6.29)
$$\begin{aligned}&\tilde{L}_l = d_{7l} \frac{\partial \tilde{G} }{\partial y_l} + d_{8l} \tilde{G}, \end{aligned}$$
(6.30)
$$\begin{aligned}&\tilde{H}_l = d_{9li} \frac{\partial \tilde{U}_i }{\partial y_l} + d_{10l} \tilde{U}_{l}. \end{aligned}$$
(6.31)

Here \(d_{sl}, d_{sli}, d_{slik} \in C^{k+1}(Q_R)\) (or \(C^{k+1}(Q_R)^d\)) for any \(s=1,\ldots , 10\) and \(1\le i,k\le d\), which are composed of \(\tilde{\theta }\), \(\frac{\partial \tilde{\theta } }{\partial y_k}\), \(\frac{\partial ^2 \tilde{\theta } }{\partial y_k \partial y_l}\), \(\varvec{\Phi }\), \(|\text {Jac } \varvec{\Phi }|\), \(\frac{\partial \varPhi _k }{\partial x_i}\), \(\frac{\partial ^2 \varPhi _k }{\partial x_i \partial x_j}\), and \(\frac{\partial |\text {Jac } \varvec{\Phi }| }{\partial x_i}\).

With the assumption of \(k=1\) and the regularity results of \(k=0\), we see that

$$\begin{aligned} \Vert f\Vert _{H^1(\varOmega )} + \Vert g\Vert _{H^\frac{3}{2}(\varGamma )} + \Vert u_\epsilon \Vert _{H^2(\varOmega )} + \Vert p_\epsilon \Vert _{H^1(\varOmega )} \le C, \end{aligned}$$

which implies

$$\begin{aligned} \Vert \tilde{J}_l\Vert _{L^2(Q_R)} + \Vert \tilde{L}_l\Vert _{H^\frac{1}{2}(S_R)} + \Vert \tilde{H}_l\Vert _{H^1(Q_R)} \le C. \end{aligned}$$

Now, substituting \(\tilde{v} = \nabla _{y_l} \tilde{U}\) into (6.28) and using \(\epsilon ^{-1} \Vert \tilde{U}_d\Vert _{L^2(\varGamma )} \le C\) obtained from (6.27), we get:

$$\begin{aligned} \left\| \nabla \frac{\partial \tilde{U}}{\partial y_l} \right\| _{L^2(Q_R)}^2 + \epsilon ^{-1} \left\| \frac{\partial \tilde{U}_d}{\partial y_l} \right\| _{L^2(S_R)}^2 \le C, \quad \forall l = 1,\ldots ,d-1. \end{aligned}$$
(6.32)

Substituting \(\tilde{v} = D_{-\zeta }^i D_\zeta ^i \nabla _{y_l}\tilde{U}\) (\(i = 1, \ldots ,d-1\)) into (6.28) and using (6.17), (6.18) together with the Sobolev inequality, we have the following six estimates:

$$\begin{aligned}&\begin{aligned} \left| \left( \tilde{J}_l,D_{-\zeta }^i D_\zeta ^i \nabla _{y_l} \tilde{U}\right) _{Q_R}\right| \le C\Vert D_{-\zeta }^i D_\zeta ^i \nabla _{y_l} \tilde{U}\Vert _{L^2(Q_R)}, \end{aligned} \end{aligned}$$
(6.33)
$$\begin{aligned}&\begin{aligned}&\left| \int _{S_R}\tilde{L}_l \cdot (D_{-\zeta }^i D_\zeta ^i \nabla _{y_l} \tilde{U})~dy'\right| \\&\qquad \le C \Vert D_\zeta ^i \tilde{L}_l \Vert _{H^{-\frac{1}{2}}(S_R)} \Vert D_\zeta ^i \nabla _{y_l} \tilde{U}\Vert _{H^\frac{1}{2}(S_R)} \\&\qquad \le C \Vert \tilde{L}_l \Vert _{H^\frac{1}{2}(S_R)} \Vert \nabla D_\zeta ^i \nabla _{y_l} \tilde{U}\Vert _{L^2(Q_R)} \le C \Vert \nabla D_\zeta ^i \nabla _{y_l} \tilde{U}\Vert _{L^2(Q_R)} , \end{aligned} \end{aligned}$$
(6.34)
$$\begin{aligned}&\begin{aligned}&\tilde{a}(\nabla _{y_l} \tilde{U}_l,D_{-\zeta }^i D_\zeta ^i \nabla _{y_l} \tilde{U}) \\&\qquad \ge C_1 \Vert \nabla D_\zeta ^i \nabla _{y_l} \tilde{U}\Vert _{L^2(Q_R)} ^2 - C_2\Vert \nabla _{y_l} \tilde{u}_\epsilon \Vert _{H^1(Q_R)} \Vert \nabla D_\zeta ^i \nabla _{y_l} \tilde{U}\Vert _{L^2(Q_R)} , \end{aligned} \end{aligned}$$
(6.35)
$$\begin{aligned}&\begin{aligned}&| \tilde{b}(D_{-\zeta }^i D_\zeta ^i \nabla _{y_l} \tilde{U}, \nabla _{y_l} \tilde{P}) | = \left| \left( \frac{\partial (\nabla _{y_l} \tilde{U}_j)}{\partial y_k}, D_{\zeta }^i D_{-\zeta }^i \left( \nabla _{y_l} \tilde{P} \frac{\partial \varPhi _k}{\partial x_j} |\text {Jac } \varvec{\Phi } | \right) \right) _{Q_R} \right| \\&\qquad = \left| \left( \tilde{H}_l, \left( \frac{\partial \varPhi _k}{\partial x_j} |\text {Jac } \varvec{\Phi } |\right) ^{-1} D_{\zeta }^i D_{-\zeta }^i \left( \nabla _{y_l} \tilde{P} \frac{\partial \varPhi _k}{\partial x_j} |\text {Jac } \varvec{\Phi } | \right) \right) _{Q_R} \right| \quad (\text {by (6.28b)} ) \\&\qquad \le C(\Vert \nabla D_\zeta ^i \tilde{H}_l\Vert _{L^2(Q_R)} + \Vert \tilde{H}_l\Vert _{H^1(Q_R)}) \Vert \nabla _{y_l} \tilde{P}\Vert _{L^2(Q_R)} \\&\qquad \le C \Vert \nabla D_\zeta ^i \nabla _{y_l} \tilde{U}\Vert _{L^2(Q_R)} \quad (\text {by } (6.31)), \end{aligned} \end{aligned}$$
(6.36)
$$\begin{aligned}&\begin{aligned}&\frac{1}{\epsilon }\int _{S_R} \nabla _{y_l} \tilde{U}_d (D_{-\zeta }^i D_\zeta ^i \nabla _{y_l} \tilde{U}_d) \sqrt{\mathrm {det}\,A} ~dy' \\&\qquad \ge C_3\frac{1}{\epsilon }\int _{S_R} |D_\zeta ^i \nabla _{y_l} \tilde{U}_{d}|^2 ~dy' - C_4 \frac{1}{\epsilon } \int _{S_R} |D_\zeta ^i \nabla _{y_l} \tilde{U}_{d}| |\nabla _{y_l} \tilde{U}_d|~dy'\\&\qquad \ge C_5\epsilon ^{-1}\Vert D_\zeta ^i \nabla _{y_l} \tilde{U}_{d}\Vert ^2_{L^2(S_R)} - C \quad (\text {by (6.32)}) , \end{aligned} \end{aligned}$$
(6.37)
$$\begin{aligned}&\begin{aligned}&\left| \frac{1}{\epsilon }\int _{S_R} \tilde{U}_d (D_{-\zeta }^i D_\zeta ^i \nabla _{y_l} \tilde{U}_d) \nabla _{y_l} \sqrt{\mathrm {det}\,A} ~dy'\right| \\&\qquad \le C_6 \epsilon ^{-1} (\Vert D_{\zeta }^i \tilde{U}_d \Vert _{L^2(S_R)} + \Vert \tilde{U}_d \Vert _{L^2(S_R)} )\Vert D_\zeta ^i \nabla _{y_l} \tilde{U}_{d}\Vert _{L^2(S_R)} \\&\qquad \le \frac{C_5}{2\epsilon } \Vert D_\zeta ^i \nabla _{y_l} \tilde{U}_{d}\Vert _{L^2(S_R)}^2 + C_7 \quad (\text {by (6.27) and (6.32)}). \end{aligned} \end{aligned}$$
(6.38)

Combining (6.33)–(6.38), we conclude

$$\begin{aligned} \Vert \nabla D_\zeta ^i \nabla _{y_l} \tilde{U}_{d}\Vert _{L^2(Q_R)}^2 + \epsilon ^{-1} \Vert D_\zeta ^i \nabla _{y_l} \tilde{U}_{d}\Vert _{L^2(S_R)}^2 \le C, \quad \forall i = 1,\ldots ,d-1, \end{aligned}$$
(6.39)

which yields (passing to the limit \(\zeta \rightarrow 0\)):

$$\begin{aligned} \Vert \nabla \nabla _{y_i} \nabla _{y_l} \tilde{U}\Vert _{L^2(Q_R)}^2 + \epsilon ^{-1} \Vert \nabla _{y_i} \nabla _{y_l} \tilde{U}_{d}\Vert _{L^2(S_R)}^2 \le C, \quad \forall i = 1,\ldots ,d-1. \end{aligned}$$
(6.40)

From (6.40) and the equation of \((\nabla _{y_l}\tilde{U},\nabla _{y_l}\tilde{P})\) (the strong from of (6.28), which can be obtained by the integration by parts), we obtain

$$\begin{aligned} \Vert \nabla _{y_d} \nabla _{y_d} \nabla _{y_l} \tilde{U}\Vert _{L^2(Q_R)}^2 \le C. \end{aligned}$$
(6.41)

Since we have (6.40) and (6.41) for all \(l = 1,\ldots ,d-1\), from the equation of \((\tilde{U},\tilde{P})\) (the strong from of (6.13), which can be obtained by the integration by parts), we conclude

$$\begin{aligned} \Vert \nabla _{y_d} \nabla _{y_d} \tilde{U}\Vert _{H^1(Q_R)} \le C\Vert \tilde{F}\Vert _{H^1(Q_R)} + C \sum _{j+i \le 2d-1} \Vert \nabla _{y_j} \nabla _{y_i} \tilde{U}\Vert _{H^1(Q_R)} \le C. \end{aligned}$$
(6.42)

Hence, we show the case of \(k=1\) for (6.2). For \(k \ge 2\), it follows from the induction method. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhou, G., Kashiwabara, T. & Oikawa, I. Penalty Method for the Stationary Navier–Stokes Problems Under the Slip Boundary Condition. J Sci Comput 68, 339–374 (2016). https://doi.org/10.1007/s10915-015-0142-0

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10915-015-0142-0

Keywords

Mathematics Subject Classification

Navigation