Skip to main content
Log in

Penalty method with P1/P1 finite element approximation for the Stokes equations under the slip boundary condition

  • Published:
Numerische Mathematik Aims and scope Submit manuscript

Abstract

We consider the P1/P1 or P1b/P1 finite element approximations to the Stokes equations in a bounded smooth domain subject to the slip boundary condition. A penalty method is applied to address the essential boundary condition \(u\cdot n = g\) on \(\partial \Omega \), which avoids a variational crime and simultaneously facilitates the numerical implementation. We give \(O(h^{1/2} + \epsilon ^{1/2} + h/\epsilon ^{1/2})\)-error estimate for velocity and pressure in the energy norm, where h and \(\epsilon \) denote the discretization parameter and the penalty parameter, respectively. In the two-dimensional case, it is improved to \(O(h + \epsilon ^{1/2} + h^2/\epsilon ^{1/2})\) by applying reduced-order numerical integration to the penalty term. The theoretical results are confirmed 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

Similar content being viewed by others

References

  1. Bänsch, E., Deckelnick, K.: Optimal error estimates for the Stokes and Navier–Stokes equations with slip-boundary condition. Math. Model. Numer. Anal. 33, 923–938 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  2. Bänsch, E., Höhn, B.: Numerical treatment of the Navier–Stokes equations with slip boundary condition. SIAM J. Sci. Comput. 21, 2144–2162 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  3. da Beirão da Veiga, H.B.: Regularity for Stokes and generalized Stokes systems under nonhomogeneous slip-type boundary conditions. Adv. Differ. Equ. 9, 1079–1114 (2004)

    MathSciNet  MATH  Google Scholar 

  4. Brenner, S.C., Scott, L.R.: The Mathematical Theory of Finite Element Methods, 3rd edn. Springer, Berlin (2007)

    MATH  Google Scholar 

  5. Castillo, P.: Performance of discontinuous Galerkin methods for elliptic PDEs. SIAM J. Sci. Comput. 24, 524–547 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  6. Çağlar, A.: Weak imposition of boundary conditions for the Navier–Stokes equations. Appl. Math. Comput. 149, 119–145 (2004)

    MathSciNet  MATH  Google Scholar 

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

    MATH  Google Scholar 

  9. Dione, I., Tibirna, C., Urquiza, J.: Stokes equations with penalised slip boundary conditions. Int. J. Comput. Fluid Dyn. 27, 283–296 (2013)

    Article  MathSciNet  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  11. Dziuk, G.: Finite elements for the Beltrami operator on arbitrary surfaces. Lect. Notes Math. 1357, 142–155 (1988)

    Article  MathSciNet  MATH  Google Scholar 

  12. Geuzaine, C., Remacle, J.F.: Gmsh: a three-dimensional finite element mesh generator with built-in pre- and post-processing facilities. Int. J. Numer. Methods Eng. 79, 1309–1331 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  13. Gilbarg, D., Trudinger, N.S.: Elliptic Partial Differential Equations of Second Order. Springer, Berlin (1998)

    MATH  Google Scholar 

  14. Gresho, P.M., Sani, R.L.: Incompressible Flow and the Finite Element Method. Wiley, New York (2000)

    MATH  Google Scholar 

  15. Hecht, F., Pironneau, O., Hyaric, F.L., Ohtsuka, K.: FreeFem++. www.freefem.org

  16. John, V.: Slip with friction and penetration with resistance boundary conditions for the Navier–Stokes equations–numerical tests and aspects of the implementation. J. Comput. Appl. Math. 147, 287–300 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  17. Knobloch, P.: Discrete Friedrichs’ and Korn’s inequalities in two and three dimensions. East-West J. Numer. Math. 4, 35–51 (1996)

    MathSciNet  MATH  Google Scholar 

  18. 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 

  19. Knobloch, P.: A finite element convergence analysis for 3D Stokes equations in case of variational crimes. Appl. Math. 45, 99–129 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  20. Krantz, S.G., Parks, H.R.: The Implicit Function Theorem-History, Theory, and Applications. Birkhäuser, London (2002)

    MATH  Google Scholar 

  21. Layton, W.: Weak imposition of “no-slip” conditions in finite element methods. Comput. Math. Appl. 38, 129–142 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  22. Lenoir, M.: Optimal isoparametric finite elements and error estimates for domains involving curved boundaries. SIAM J. Numer. Anal. 21, 562–580 (1986)

    Article  MathSciNet  MATH  Google Scholar 

  23. Logg, A., Mardal, K.A., Wells, G.N. (eds.): Automated Solution of Differential Equations by the Finite Element Method. Springer, Berlin (2012)

    MATH  Google Scholar 

  24. Nečas, J.: Direct Methods in the Theory of Elliptic Equations. Springer, Berlin (2012)

    MATH  Google Scholar 

  25. Saito, H., Scriven, L.E.: Study of coating flow by the finite element method. J. Comput. Phys. 42, 53–76 (1981)

    Article  MATH  Google Scholar 

  26. Solonnikov, V.A., Ščadilov, V.E.: On a boundary value problem for a stationary system of Naiver–Stokes equations. Proc. Steklov Inst. Math. 125, 186–199 (1973)

    MATH  Google Scholar 

  27. Stokes, Y., Carey, G.: On generalized penalty approaches for slip, free surface and related boundary conditions in viscous flow simulation. Int. J. Numer. Methods Heat Fluid Flow 21, 668–702 (2011)

    Article  Google Scholar 

  28. Tabata, M.: Uniform solvability of finite element solutions in approximate domains. Jpn. J. Ind. Appl. Math. 18, 567–585 (2001)

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  Google Scholar 

  30. Tabata, M., Suzuki, A.: A stabilized finite element method for 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 

  31. Verfürth, R.: Finite element approximation of steady Navier–Stokes equations with mixed boundary conditions. Math. Model. Numer. Anal. 19, 461–475 (1985)

    MathSciNet  MATH  Google Scholar 

  32. 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 

  33. Verfürth, R.: Finite element approximation of incompressible Navier–Stokes equations with slip boundary condition II. Numer. Math. 59, 615–636 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  34. Wei, H., Chen, L., Huang, Y.: Superconvergence and gradient recovery of linear finite elements for the Laplace–Beltrami operator on general surfaces. SIAM J. Numer. Anal. 48, 1920–1943 (2010)

    Article  MathSciNet  MATH  Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

  36. Zhou, G., Kashiwabara, T., Oikawa, I.: Penalty method for the stationary Navier-Stokes problems under the slip boundary condition. J. Sci. Comput. (2015). doi:10.1007/s10915-015-0142-0

Download references

Acknowledgments

The authors appreciate the suggestions from the anonymous referees that have improved the quality of this paper. They thank Professor Fumio Kikuchi, Professor Norikazu Saito, and Professor Masahisa Tabata for valuable comments to the results of this paper. They also thank Professor Xuefeng Liu for giving them a program which converts matrices described by the PETSc-format into those by the CRS-format.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Takahito Kashiwabara.

Additional information

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

Appendices

Appendix 1: Transformation between \(\varGamma \) and \(\varGamma _h\) and related estimates

Let \(\Omega \) be a bounded domain in \(\mathbb R^N\). Its boundary \(\varGamma \) is assumed to be \(C^{1,1}\)-smooth, namely, there exist a system of local coordinates \(\{(U_r, y_r, \varphi _r)\}_{r=1}^M\) and positive numbers \(\alpha , \beta \) such that: 1) \(\{U_r\}_{r=1}^M\) forms an open covering of \(\varGamma \); 2) \(y_r = (y_{r1}, \dots , y_{rN-1}, y_{rN}) = (y_r', y_{rN})\) is a rotated coordinate of the original one x, that is, \(y_r = A_rx\) for some orthogonal transformation \(A_r\); 3) \(\varphi _r\in C^{1,1}(\varDelta _r)\) gives a graph representation of \(\varGamma \cap U_r\), where \(\varDelta _r := \{y_r' \in \mathbb R^{N-1} :\; |y_r'| < \alpha \}\), that is,

$$\begin{aligned} \varGamma \cap U_r&= \{ y_r \in \mathbb R^N \,:\; y_r' \in \varDelta _r \text { and } y_{rN} = \varphi _r(y_r') \}, \\ \Omega \cap U_r&= \{ y_r \in \mathbb R^N \,:\; y_r' \in \varDelta _r \text { and } \varphi _r(y_r') < y_{rN} < \varphi _r(y_r') + \beta \}, \\ \overline{\Omega }^c \cap U_r&= \{ y_r \in \mathbb R^N \,:\; y_r' \in \varDelta _r \text { and } \varphi _r(y_r') - \beta < y_{rN} < \varphi _r(y_r') \}. \end{aligned}$$

Since \(\varGamma \) is compact, there exists \(h_0>0\) such that for any \(x\in \varGamma \) the open ball \(B(x; h_0)\) is contained in some local coordinate neighborhood \(U_r\). According to the fact that \(C^{1,1}(\varDelta _r) = W^{2,\infty }(\varDelta _r)\), the derivatives of \(\varphi _r\) are bounded up to second order, i.e.,

$$\begin{aligned} \Vert \varphi _r\Vert _{L^\infty (\varDelta _r)} \le C_0, \quad \Vert \nabla '\varphi _r\Vert _{L^\infty (\varDelta _r)} \le C_1, \quad \Vert \nabla '^2\varphi _r\Vert _{L^\infty (\varDelta _r)} \le C_2, \end{aligned}$$

where \(C_0,\,C_1,\,C_2\) are constants independent of r, and \(\nabla '\) means \(\nabla _{y'}\).

The smoothness of \(\varGamma \) is connected with that of the signed distance function d(x) defined by

$$\begin{aligned} d(x) = {\left\{ \begin{array}{ll} -\mathrm {dist}(x, \varGamma ) &{} \text {if }\; x\in \Omega , \\ \mathrm {dist}(x, \varGamma ) &{} \text {if }\; x\in \Omega ^c. \end{array}\right. } \end{aligned}$$

We collect several known properties on d(x) below. For the details, see e.g. [13, Section 14.6] or [8, Section 7.8]. Let \(\varGamma (\delta ) := \{ x\in \mathbb R^N :\; |d(x)| < \delta \}\) be a tubular neighborhood of \(\varGamma \) with width \(2\delta \). Then there exists \(\delta \) depending only on the curvature of \(\varGamma \) such that for arbitrary \(x\in \varGamma (\delta )\) the decomposition

$$\begin{aligned} x = \pi (x) + d(x)\, n(\pi (x)), \quad \pi (x)\in \varGamma , \end{aligned}$$
(8.1)

is uniquely determined. Here, n is the outer unit normal field defined on \(\varGamma \), which coincides with \(\nabla d|_{\varGamma }\). We extend n from \(\varGamma \) to \(\varGamma (\delta )\) by \(n(x) = \nabla d(x)\), which also agrees with \(n(\pi (x))\). The fact that \(\varGamma \) is \(C^{1,1}\)-smooth implies that \(d\in C^{1,1}(\varGamma (\delta ))\), \(n\in C^{0,1}(\varGamma (\delta ))\), and \(\pi \in C^{0,1}(\varGamma (\delta ))\). We call \(\pi : \varGamma (\delta )\rightarrow \varGamma \) the orthogonal projection onto \(\varGamma \), in view of its geometrical meaning. We may assume that

$$\begin{aligned} \Vert d\Vert _{L^\infty (\varGamma (\delta ))} \le C_0, \quad \Vert \nabla d\Vert _{L^\infty (\varGamma (\delta ))} \le C_1, \quad \Vert \nabla ^2d\Vert _{L^\infty (\varGamma (\delta ))} \le C_2, \end{aligned}$$

where we re-choose the constants \(C_0,\,C_1,\,C_2\) if necessary.

Now we introduce a regular family of triangulations \(\{\mathcal T_h\}_{h\downarrow 0}\) of \(\overline{\Omega }\) in the sense of Sect. 3.1. As before, we denote by \(\mathcal S_h\) the boundary mesh inherited from \(\mathcal T_h\), and we set \(\overline{\Omega }_h = \cup _{T\in \mathcal T_h}T\) and \(\varGamma _h = \cup _{S\in \mathcal S_h}S\). In order for \(\varGamma _h\) to be compatible with the local-coordinate system \(\{(U_r, y_r, \varphi _r)\}_{r=1}^M\), we assume the following:

  1. (1)

    the mesh size h is less than \(\min \{h_0, 1\}\);

  2. (2)

    for every \(r=1,\dots ,M\), \(\varGamma _h\cap U_r\) is represented by a graph \(\{(y_r', \varphi _{rh}(y_r'))\in \mathbb R^N :\; y_r'\in \varDelta _r \}\);

  3. (3)

    every vertex of \(S\in \mathcal S_h\) lies on \(\varGamma \).

From these we see that every \(S\in \mathcal S_h\) is contained in some \(U_r\) and that \(\varphi _{rh}\) is a piecewise linear interpolation of \(\varphi _r\). As a result of interpolation error estimates, for arbitrary r we obtain

$$\begin{aligned} \begin{aligned} \Vert \varphi _{rh}\Vert _{L^\infty (\varDelta _r)}&\le C_0, \\ \Vert \nabla '\varphi _{rh}\Vert _{L^\infty (\varDelta _r)}&\le C_1, \end{aligned} \qquad \begin{aligned} \Vert \varphi _r - \varphi _{rh}\Vert _{L^\infty (\varDelta _r)}&\le C_{0E}h^2, \\ \Vert \nabla '(\varphi _r - \varphi _{rh})\Vert _{L^\infty (\varDelta _r)}&\le C_{1E}h, \end{aligned} \end{aligned}$$
(8.2)

where we re-choose \(C_0,\,C_1\) if necessary and the subscript E refers to “error”. In the following, h is made small enough to satisfy \(2C_{0E}h^2< \min \{h_0, \delta \}\), which in particular ensures that \(\pi \) is well-defined on \(\varGamma _h\). We may assume further that \(\pi (S)\) is contained in the same \(U_r\) that contains S.

Based on the observations above, we see that the orthogonal projection \(\pi \) maps \(\varGamma _h\) into \(\varGamma \). It indeed gives a homeomorphism between \(\varGamma _h\) and \(\varGamma \), and is element-wisely a diffeomorphism, as shown below. The representation of a function f(x) in each local coordinate \((U_r, y_r, \varphi _r)\) is defined as \(\tilde{f}(y_r) := f(A_r^{-1}y_r)\). However, with some abuse of notation, we denote it simply by \(f(y_r)\). Then, the local representation of the outer unit normal n associated to \(\varGamma \) is given by

$$\begin{aligned} n(y_r', \varphi _r(y_r')) = \frac{1}{ K_r(y_r') } \begin{pmatrix} \nabla '\varphi _r(y_r') \\ -1 \end{pmatrix}, \qquad y_r'\in \varDelta _r, \end{aligned}$$

where \(K_r(y_r') := \sqrt{1 + |\nabla '\varphi _r(y_r')|^2}\). We are ready to state the following.

Proposition 8.1

If \(h>0\) is sufficiently small, then \(\pi |_{\varGamma _h}\) is a homeomorphism between \(\varGamma _h\) and \(\varGamma \).

Proof

Let us construct an inverse map \(\pi ^*: \varGamma \rightarrow \varGamma _h\) of \(\pi |_{\varGamma _h}\) which is continuous. To this end, we fix arbitrary \(x\in \varGamma \) and choose a local coordinate \((U_r, y_r, \varphi _r)\) of x such that \(B(x; h_0)\subset U_r\). For simplicity, we omit the subscript r in the following. In view of the definition of \(\pi (x)\) by (8.1), we introduce a segment given by

$$\begin{aligned} \begin{pmatrix} y' \\ \varphi (y') \end{pmatrix} + \frac{t}{K(y')} \begin{pmatrix} \nabla '\varphi (y') \\ -1 \end{pmatrix}, \qquad |t|\le 2C_{0E}h^2. \end{aligned}$$

To each point on this segment we associate its height H(t) with respect to the graph of \(\varphi _h\), that is,

$$\begin{aligned} H(t) = \varphi (y') - \frac{t}{K(y')} - \varphi _h\left( y' + \frac{t}{K(y')} \nabla '\varphi (y') \right) . \end{aligned}$$

Then we assert that \(\frac{d}{dt}H(t) < 0\) and that \(H(-2C_{0E}h^2) > 0\), \(H(2C_{0E}h^2) < 0\).

To prove the first assertion, letting

$$\begin{aligned} Y' := y' + \frac{t}{K(y')} \nabla '\varphi (y'), \end{aligned}$$

we have \(\frac{d}{dt}H(t) = -\frac{1}{K(y')}(1 + \nabla '\varphi (y')\cdot \nabla '\varphi _h(Y'))\). One sees that

$$\begin{aligned} \nabla '\varphi _h(Y')&= \nabla '\varphi (y') + [ -\nabla '\varphi (y') + \nabla '\varphi (Y') ] + [-\nabla '\varphi (Y') + \nabla '\varphi _h(Y')] \\&=: \nabla '\varphi (y') + I_1 + I_2, \end{aligned}$$

where \(I_1\) and \(I_2\) satisfy

$$\begin{aligned} |I_1| \le \Vert \nabla '^2\varphi \Vert _{L^\infty (\varDelta )} |Y' - y'| \le C_2\cdot 2C_{0E}h^2, \qquad |I_2| \le C_{1E}h. \end{aligned}$$

Then it follows that

$$\begin{aligned} 1 + \nabla '\varphi (y')\cdot \nabla '\varphi _h(Y') \ge K(y')^2 - C_1(2C_2C_{0E}h^2 + C_{1E}h). \end{aligned}$$

From this we have \(\frac{d}{dt}H(t) < 0\) provided \(C_1(2C_2C_{0E}h^2 + C_{1E}h) < 1/2\). For the second assertion, one finds that

$$\begin{aligned} H(-2C_{0E}h^2)&= \frac{2C_{0E}h^2}{K(y')} + \varphi (y') - \varphi _h(Y') \\&= \frac{2C_{0E}h^2}{K(y')} + [ \varphi (y') - \varphi (Y') ] + [ \varphi (Y') - \varphi _h(Y') ], \end{aligned}$$

where \(Y'\) is \(y' - \frac{2C_{0E}h^2}{K(y')} \nabla '\varphi (y')\). By Taylor’s theorem, there exists some \(\theta \in (0,1)\) such that

$$\begin{aligned} \varphi (Y') - \varphi (y')&= \nabla '\varphi (y')\cdot (Y' - y') + \frac{1}{2} (Y' - y')^T \nabla '^{2}\varphi |_{y' + \theta (Y' - y')} (Y' - y') \\&\le - \frac{2C_{0E}h^2}{K(y')} |\nabla '\varphi (y')|^2 + 2 C_2C_{0E}^2 h^4. \end{aligned}$$

By the definition of K(y) and by \(|\varphi (Y') - \varphi _h(Y')| \le C_{0E}h^2\), we obtain

$$\begin{aligned} H(-2C_{0E}h^2)&\ge 2K(y')C_{0E}h^2 - 2C_2C_{0E}^2 h^4 - C_{0E} h^2 \\&\ge 2C_{0E}h^2 - 2C_2C_{0E}^2 h^4 - C_{0E} h^2 = C_{0E}h^2(1 - 2C_2C_{0E}h^2), \end{aligned}$$

which implies that \(H(-2C_{0E}h^2) > 0\) provided \(C_2C_{0E}h^2 \le 1/4\). In the same way, the last assertion \(H(2C_{0E}h^2) < 0\) can be proved.

From these assertions we deduce that there exists a unique \(t^*(x) \in [-2C_{0E}h^2, 2C_{0E}h^2]\) such that \(H(t^*(x)) = 0\). Consequently, the map \(\pi ^*: \varGamma \rightarrow \varGamma _h; \; x\mapsto x + t^*(x)n(x)\) is well-defined. A direct computation combined with the uniqueness of the decomposition (8.1) shows that \(\pi ^*\) is the inverse of \(\pi |_{\varGamma _h}\). The continuity of \(\pi ^*\), especially that of \(t^*\), follows from an argument similar to the proof of the implicit function theorem (see e.g. [20, Theorem 3.2.1]). \(\square \)

Remark 8.1

Proposition 8.1 enables us to define an exact triangulation of \(\varGamma \) by \(\pi (\mathcal S_h) = \{ \pi (S) \,:\, S\in \mathcal S_h \}\). In particular, we can subdivide \(\varGamma \) into disjoint sets as \(\varGamma = \bigcup _{S\in \mathcal S_h}\pi (S)\). Under assumption (3) this fact is obvious to see for the two-dimensional case, but not for the three-dimensional case. Furthermore, the above construction of an exact triangulation even applies to the situation where the vertices of \(\varGamma _h\) do not exactly belong to \(\varGamma \) but still lie closely to it so that (8.2) is fulfilled. This may be useful information and is left as a side remark.

Next we observe that S and \(\pi (S)\) admit the same domain of parametrization for each \(S\in \mathcal S_h\), which is important in the subsequent analysis. To describe this fact, we choose a local coordinate \((U_r, y_r, \varphi _r)\) such that \(U_r \supset S\cup \pi (S)\), and introduce the projection to the base set \(b_r:\mathbb R^N\rightarrow \mathbb R^{N-1}\) by \(b_r(y_r) = y_r'\). The domain of parametrization is then defined to be \(S' = b_r(\pi (S))\). We see that the mappings

are bijective and that \(\varPhi \) is smooth on \(S'\). If in addition \(\varPhi _h\), especially \(t^*\), is also smooth on \(S'\), then \(\varPhi \) and \(\varPhi _h\) may be employed as smooth parametrizations for \(\pi (S)\) and S respectively. The next proposition verifies that this is indeed the case.

Proposition 8.2

Under the setting above, we have

$$\begin{aligned} \Vert t^*\Vert _{L^\infty (S')} \le \tilde{C}_{0E} h^2, \qquad \Vert \nabla ' t^*\Vert _{L^\infty (S')} \le \tilde{C}_{1E} h, \end{aligned}$$

where \(\tilde{C}_{0E}\) and \(\tilde{C}_{1E}\) are constants depending only on N and \(\varGamma \).

Proof

Since the first relation is already obtained in Proposition 8.1 with \(\tilde{C}_{0E} = 2C_{0E}\), we focus on proving the second one. For notational simplicity, we omit the subscript r and also use the abbreviation \(\partial _i = \frac{\partial }{\partial y_i}\,(i=1,\dots ,N)\). The fact that \(t^*\) is differentiable with respect to \(y'\) can be shown in a way similar to the proof of the implicit function theorem. Thereby it remains to evaluate the supremum norm of \(\nabla ' t\) in \(S'\), which we address in the following.

Recall that \(t^*(y')\) is determined according to the equation

$$\begin{aligned} \tilde{t}(y') = \varphi (y') - \varphi _h(y' + \tilde{t}(y') \nabla '\varphi (y')), \end{aligned}$$
(8.3)

where we have set \(\tilde{t}(y') := t^*(y')/K(y')\). Because \(\nabla 't^* = K\nabla '\tilde{t} + \frac{\nabla '^2\varphi \nabla '\varphi }{K^2}\, t^*\), it follows that

$$\begin{aligned} \Vert \nabla 't^*\Vert _{L^\infty (S')} \le (1 + C_1) \Vert \nabla '\tilde{t}\Vert _{L^\infty (S')} + C_2C_1\tilde{C}_{0E}h^2. \end{aligned}$$

Therefore, it suffices to prove that \(\Vert \nabla '\tilde{t}\Vert _{L^\infty (S')} \le Ch\); here and hereafter C denotes various constants which depends only on N and \(\varGamma \).

Applying \(\nabla '\) to (8.3) gives

$$\begin{aligned} \big ( 1 + \nabla '\varphi (y')\cdot \nabla '\varphi _h(Y') \big ) \nabla '\tilde{t} = \nabla '\varphi (y') - \nabla '\varphi _h(Y') - \tilde{t}(y') \nabla '^2\varphi (y') \nabla '\varphi _h(Y'), \end{aligned}$$
(8.4)

where \(Y' := y' + \tilde{t}(y') \nabla '\varphi (y')\). By the same way as we estimated \(I_1\) and \(I_2\) in the proof of Proposition 8.1, we obtain

$$\begin{aligned} |\nabla '\varphi (y') - \nabla '\varphi _h(Y')|&\le C_2\tilde{C}_{0E}h^2 + C_{1E}h \le Ch, \\ 1 + \nabla '\varphi (y')\cdot \nabla '\varphi _h(Y')&\ge K(y')^2 - C_1(C_2\tilde{C}_{0E}h^2 + C_{1E}h) \ge \frac{1}{2}. \end{aligned}$$

Also we see that

$$\begin{aligned} |\tilde{t}(y') \nabla '^2\varphi (y') \nabla '\varphi _h(Y')| \le \tilde{C}_{0E}h^2\cdot C_2C_1 \le Ch^2. \end{aligned}$$

Combining these observations with (8.4), we deduce the desired estimate \(\Vert \nabla '\tilde{t}\Vert _{L^\infty (S')} \le Ch\). \(\square \)

Remark 8.2

When \(\varGamma \in C^{2,1}\), further differentiation of (8.4) gives us \(t^*\in C^{1,1}(\pi (S))\); in fact we have \(\Vert \nabla '^2 t^*\Vert _{L^\infty (S')} \le \tilde{C}_{2E}\) with some constant \(\tilde{C}_{2E}\). This implies that \(\pi |_{S}\) is a \(C^{1,1}\)-diffeomorphism between S and \(\pi (S)\). However, since \(\nabla '\phi _h\) is smooth only within b(S), \(\pi \) is not globally a diffeomorphism.

Now we give an error estimate for surface integrals on \(\varGamma \) and \(\varGamma _h\). Heuristically speaking, the result reads \(|d\gamma - d\gamma _h| \le O(h^2)\), which may be found in the literature (see e.g. [11]). Here and hereafter, we denote the surface elements of \(\varGamma \) and \(\varGamma _h\) by \(d\gamma \) and \(d\gamma _h\), respectively.

Theorem 8.1

Let \(S\in \mathcal S_h\) and f be an integrable function on S. Then we have

$$\begin{aligned} \left| \int _{\pi (S)}f\,d\gamma - \int _S f\circ \pi \, d\gamma _h \right| \le Ch^2 \int _S |f|\,d\gamma , \end{aligned}$$

where C is a constant depending only on N and \(\varGamma \).

Proof

Let \((U_r, y_r, \varphi _r)\) be a local coordinate that contains \(S\cup \pi (S)\). We omit the subscript r and use the abbreviation \(\partial _i = \frac{\partial }{\partial y_i}\,(i=1,\dots ,N)\). We represent the surface integral using the parametrization \(\varPhi \) as follows:

$$\begin{aligned} \int _{\pi (S)}f\,d\gamma = \int _{S'} f(\varPhi (y')) \sqrt{\mathrm {det}\,G}\,dy', \end{aligned}$$

where \(G=(G_{ij})_{1\le i,j\le N-1}\) denotes the Riemannian metric tensor given by \(G_{ij} = \partial _i\varPhi \cdot \partial _j\varPhi \) (dot means the inner product in \(\mathbb R^N\)). Similarly, noting that \(\pi \circ \varPhi _h = \varPhi \), one obtains

$$\begin{aligned} \int _{S}f\circ \pi \,d\gamma _h = \int _{S'} f(\varPhi (y')) \sqrt{\mathrm {det}\,G_h}\,dy', \end{aligned}$$

where \(G_h\) is given by \(G_{h,ij} = \partial _i\varPhi _h \cdot \partial _j\varPhi _h\). Then we assert that:

$$\begin{aligned} \Vert G_h - G\Vert _{L^\infty (S')} \le Ch^2. \end{aligned}$$

To prove this, noting that \(\varPhi _h = \varPhi + t^*n\circ \varPhi \), we compute each component of \(G_h-G\) as follows:

$$\begin{aligned} G_{h,ij} - G_{ij}&= \partial _i\varPhi \cdot \partial _j(\varPhi _h - \varPhi ) + \partial _j\varPhi \cdot \partial _i(\varPhi _h - \varPhi ) + \partial _i(\varPhi _h - \varPhi ) \cdot \partial _j(\varPhi _h - \varPhi ) \\&= \partial _i\varPhi \cdot \partial _j(t^*n\circ \varPhi ) + \partial _j\varPhi \cdot \partial _i(t^*n\circ \varPhi ) + \partial _i(t^*n\circ \varPhi ) \cdot \partial _j(t^*n\circ \varPhi ) \\&=: I_1 + I_2 + I_3. \end{aligned}$$

For \(I_1\), we notice that \(\partial _i\varPhi \) is a tangent vector so that \(\partial _i\varPhi \cdot n\circ \varPhi =0\). This yields

$$\begin{aligned} I_1 = \partial _i\varPhi \cdot t^*\partial _j(n\circ \varPhi ) = t^*\partial _i\varPhi \cdot (\partial _jn + \partial _j\varphi \,\partial _Nn)|_{\varPhi }, \end{aligned}$$

which is estimated by \(\tilde{C}_{0E}h^2(1 + C_1)(C_2 + C_1C_2)\) thanks to Proposition 8.2. \(I_2\) can be bounded in the same manner. To estimate \(I_3\), we observe that

$$\begin{aligned} \partial _i(t^*n\circ \varPhi ) = (\partial _i t^*) n\circ \varPhi + t^* \big ( \partial _in + \partial _i\varphi \,\partial _Nn \big )|_{\varPhi }, \end{aligned}$$

which is bounded by \(\tilde{C}_{1E}h + \tilde{C}_{0E}h^2(C_2 + C_1C_2) \le Ch\). Similarly one gets \(|\partial _j(t^*n\circ \varPhi )| \le Ch\), hence it follows that \(|I_3| \le Ch^2\). Therefore, \(|G_{h,ij} - G_{ij}| \le Ch^2\), which proves the assertion.

Now we use the following crude estimate for perturbation of determinants (cf. [17, equation (3.13)]): if A and B are \(N\times N\) matrices such that \(|A_{ij}| \le a\) and \(|B_{ij}|\le b\) for all ij, then

$$\begin{aligned} |\mathrm {det}\,(A+B) - \mathrm {det}\,A| \le N!N(a+b)^{N-1}b. \end{aligned}$$

Combining this with the assertion above and also with \(\sqrt{\alpha }- \sqrt{\beta }= (\alpha - \beta )/(\sqrt{\alpha }+ \sqrt{\beta })\), we obtain

$$\begin{aligned} \Vert \sqrt{\mathrm {det}\,G} - \sqrt{\mathrm {det}\,G_h}\Vert _{L^\infty (S')} \le Ch^2. \end{aligned}$$

In addition, note that \(\sqrt{\mathrm {det}\,G} = \sqrt{1 + |\nabla '\varphi |^2} \ge 1\). Consequently,

$$\begin{aligned} \left| \int _{\pi (S)}f\,d\gamma - \int _S f\circ \pi \, d\gamma _h \right| \le Ch^2\,\int _{S'} |f(\varPhi (y'))| \sqrt{\mathrm {det}\,G} \,dy' = Ch^2 \int _{\pi (S)}|f|\,d\gamma , \end{aligned}$$

which proves the theorem. \(\square \)

Remark 8.3

Adding up the results of the theorem for all \(S\in \mathcal S_h\) yields

$$\begin{aligned} \left| \int _\varGamma f\,d\gamma - \int _{\varGamma _h} f\circ \pi \, d\gamma _h \right| \le Ch^2 \int _\varGamma |f|\,d\gamma . \end{aligned}$$

It also follows that \(|\int _{\varGamma _h} f\circ \pi \, d\gamma _h| \le C\int _\varGamma |f|\,d\gamma \). Choosing in particular \(|f|^p\) as the integrand gives \(\Vert f\circ \pi \Vert _{L^p(\varGamma _h)} \le C^{1/p}\Vert f\Vert _{L^p(\varGamma )}\) for \(p\in [1,\infty ]\).

Let f be a smooth function given on \(\varGamma \). Then its transformation to \(\varGamma _h\) is defined by \(f\circ \pi \). However, if f is extended to a neighborhood of \(\varGamma \), e.g. to \(\varGamma (\delta )\), then we may also consider f’s natural trace on \(\varGamma _h\). The next theorem provides error estimation of these two quantities.

Theorem 8.2

Let \(f\in W^{1,p}(\varGamma (\delta _1))\), where \(p\in [1,\infty ]\) and \(\delta _1 \in [\tilde{C}_{0E}h^2, 2\tilde{C}_{0E}h^2]\). Then,

$$\begin{aligned} \Vert f - f\circ \pi \Vert _{L^p(\varGamma _h)} \le C\delta _1^{1-1/p}\Vert f\Vert _{W^{1,p}(\varGamma (\delta _1))}, \end{aligned}$$

where C is a constant depending only on \(p,\,N,\,\Omega \).

Proof

Since \(\varGamma (\delta _1) = \bigcup _{S\in \mathcal S_h} \pi (S,\delta _1)\), where \(\pi (S, \delta _1) = \{ x\in \varGamma (\delta _1) : \pi (x)\in \pi (S) \}\) denotes a tubular neighborhood of S, it suffices to prove that

$$\begin{aligned} \int _S |f - f\circ \pi |^p\,d\gamma \le C\delta _1^{p - 1} \int _{\pi (S, \delta _1)} |\nabla f|^p\,dx \qquad \forall S\in \mathcal S_h. \end{aligned}$$
(8.5)

To this end, using the notation in Theorem 8.1, we estimate the left-hand side of (8.5) by

$$\begin{aligned} \int _S |f - f\circ \pi |^p\,d\gamma&= \int _{S'} |f\circ \varPhi _h - f\circ \varPhi |^p \sqrt{\mathrm {det}\,G_h}\,dy' \\&\le C \int _{S'} |f\circ \varPhi _h - f\circ \varPhi |^p\,dy'. \end{aligned}$$

Here, for fixed \(y'\in S'\) we have

$$\begin{aligned} f(\varPhi _h(y')) - f(\varPhi (y'))&= \int _0^1 \frac{d}{ds} f\big ( \varPhi (y') + s(\varPhi _h(y') - \varPhi (y')) \big )\,ds \\&= \int _0^1 \frac{d}{ds} f\big ( \varPhi (y') + s\, t^*(y') n(\varPhi (y')) \big )\,ds \\&= \int _0^1 t^*(y') n(\varPhi (y')) \cdot \nabla f\big ( \varPhi (y') + s\, t^*(y') n(\varPhi (y')) \big )\,ds \\&= \int _0^{t^*(y')} n(\varPhi (y')) \cdot \nabla f\big ( \varPhi (y') + t n(\varPhi (y')) \big )\,dt. \end{aligned}$$

Because \(|t^*(y')| \le \tilde{C}_{0E}h^2 \le \delta _1\), it follows that

$$\begin{aligned} |f(\varPhi _h(y')) - f(\varPhi (y'))|&\le \int _{-\delta _1}^{\delta _1} \big | \nabla f\big ( \varPhi (y') + tn(\varPhi (y')) \big ) \big |\,dt \\&\le (2\delta _1)^{1-1/p} \left( \int _{-\delta _1}^{\delta _1} \big | \nabla f\big ( \varPhi (y') + tn(\varPhi (y')) \big ) \big |^p\,dt \right) ^{1/p}, \end{aligned}$$

where we have used Hölder’s inequality. Consequently,

$$\begin{aligned} \int _S |f - f\circ \pi |^p\,d\gamma \le C\delta _1^{p-1} \int _{S'\times [-\delta _1,\delta _1]} \big | \nabla f\big ( \varPhi (y') + tn(\varPhi (y')) \big ) \big |^p\,dy'dt. \end{aligned}$$
(8.6)

On the other hand, we observe that the N-dimensional transformation

$$\begin{aligned} \Psi : S'\times [-\delta _1, \delta _1] \rightarrow \pi (S, \delta _1); \quad (y', t) \mapsto \varPhi (y') + tn(\varPhi (y')) \end{aligned}$$
(8.7)

is bijective and smooth. Application of this transformation to the right-hand side of (8.5) leads to

$$\begin{aligned} \int _{\pi (S, \delta _1)} |\nabla f|^p\,dx = \int _{S'\times [-\delta _1, \delta _1]} \big | \nabla f\big ( \varPhi (y') + tn(\varPhi (y')) \big ) \big |^p\, |\mathrm {det}\, J| \,dy'dt, \end{aligned}$$

where \(J = (\partial _1\varPhi + t\partial _1(n\circ \varPhi ), \cdots , \partial _{N-1}\varPhi + t\partial _{N-1}(n\circ \varPhi ), n\circ \varPhi )\) denotes the Jacobi matrix of \(\Psi \). Letting \(\tilde{J} := (\partial _1\varPhi , \cdots , \partial _{N-1}\varPhi , n\circ \varPhi )\), we find that

$$\begin{aligned} \Vert J - \tilde{J}\Vert _{L^\infty (S'\times [-\delta _1, \delta _1])} \le C\delta _1, \end{aligned}$$

because \(|t\partial _i(n\circ \varPhi )| = |t (\partial _in + \partial _i\varphi \partial _Nn)_{|\varPhi }| \le \delta _1(C_2 + C_1C_2)\). This implies

$$\begin{aligned} \Vert \mathrm {det}\, J - \mathrm {det}\,\tilde{J}\Vert _{L^\infty (S'\times [-\delta _1, \delta _1])} \le C\delta _1, \end{aligned}$$

which combined with \(\mathrm {det}\,\tilde{J} = K(y') \ge 1\) yields \(\mathrm {det}\,J \ge 1/2\) if h is sufficiently small. Therefore,

$$\begin{aligned} \int _{\pi (S, \delta _1)} |\nabla f|^p\,dx \ge \frac{1}{2} \int _{S'\times [-\delta _1, \delta _1]} \big | \nabla f\big ( \varPhi (y') + tn(\varPhi (y')) \big ) \big |^p \,dy'dt. \end{aligned}$$
(8.8)

The desired estimate (8.5) is now a consequence of (8.6) and (8.8). This completes the proof. \(\square \)

Finally, we show that the \(L^p\)-norm in a tubular neighborhood can be bounded in terms of its width. Such estimate is stated e.g. in [35, Lemma 2.1] or in [28, equation (3.6)]. However, since we could not find a full proof of this fact (especially for \(N=3\)) in the literature, we present it here.

Theorem 8.3

Under the same assumptions as in Theorem 8.2, we have

$$\begin{aligned} \Vert f\Vert _{L^p(\varGamma (\delta _1))} \le C (\delta _1 \Vert \nabla f\Vert _{L^p(\varGamma (\delta _1))} + \delta _1^{1/p} \Vert f\Vert _{L^p(\varGamma )}), \end{aligned}$$

where C is a constant depending only on \(p,\,N,\,\Omega \).

Proof

We adopt the same notation as in the proofs of Theorems 8.1 and 8.2. Then it suffices to prove that

$$\begin{aligned} \int _{\pi (S,\delta _1)} |f|^p\,dy \le C \left( \delta _1^p \int _{\pi (S,\delta _1)} |\nabla f|^p\,dy + \delta _1 \int _{\pi (S)} |f|^p\,d\gamma \right) \qquad \forall S\in \mathcal S_h. \end{aligned}$$

To this end, using the transformation \(\Psi \) given in (8.7) we express the left-hand side as

$$\begin{aligned} \int _{\pi (S,\delta _1)} |f|^p\,dy&= \int _{S'\times [-\delta _1,\delta _1]} |f(\Psi (y', t))|^p|\mathrm {det}\,J(y',t)|\,dy'dt \\&\le C \int _{S'\times [-\delta _1,\delta _1]} \big ( |f(\Psi (y', t)) - f(\varPhi (y'))|^p + |f(\varPhi (y'))|^p \big ) \,dy'dt \\&=: I_1 + I_2. \end{aligned}$$

For \(I_1\), we see from the same argument as before that

$$\begin{aligned} |f(\Psi (y', t)) - f(\varPhi (y'))|^p \le (2\delta _1)^{p-1} \int _{-\delta _1}^{\delta _1} \big | \nabla f\big ( \varPhi (y') + sn(\varPhi (y')) \big ) \big |^p\,ds, \end{aligned}$$

which yields

$$\begin{aligned} |I_1| \le C\delta _1^p \int _{S'\times [-\delta _1,\delta _1]} \big | \nabla f(\Psi (y',s)) \big |^p |\mathrm {det}\,J(y',s)|\,dy'ds = C\delta _1^p \int _{\pi (S,\delta _1)} |\nabla f|^p\,dy. \end{aligned}$$

For \(I_2\), it follows that

$$\begin{aligned} |I_2|= & {} 2C\delta _1 \int _{S'} |f(\varPhi (y'))|^p\,dy' \le 2C\delta _1 \int _{S'} |f(\varPhi (y'))|^p \sqrt{\mathrm {det}\,G}\,dy'\\= & {} 2C\delta _1 \int _{\pi (S)} |f|^p\,d\gamma . \end{aligned}$$

We have thus obtained the desired estimate, which completes the proof. \(\square \)

Appendix 2: Error of n and \(n_h\)

Let us prove that \(|n\circ \pi - n_h| \le O(h)\) on \(\varGamma _h\) and also that, when \(N=2\), it is improved to \(O(h^2)\) if the consideration is restricted to the midpoint of edges.

Lemma 9.1

Let n and \(n_h\) be the outer unit normals to \(\varGamma \) and \(\varGamma _h\) respectively. Then there holds

$$\begin{aligned} \Vert n\circ \pi - n_h\Vert _{L^\infty (\varGamma _h)} \le Ch. \end{aligned}$$
(9.1)

If in addition \(N=2\), \(\varGamma \in C^{2,1}\), and \(m_S\) denotes the midpoint of \(S\in \mathcal S_h\), then

$$\begin{aligned} \sup _{S\in \mathcal S_h}|n\circ \pi (m_S) - n_h(m_S)| \le Ch^2. \end{aligned}$$
(9.2)

Here, C is a constant depending only on N and \(\varGamma \).

Proof

Let \(S\in \mathcal S_h\) be arbitrary and let \((U_r, y_r, \phi _r)\) be a local coordinate that contains \(S\cup \pi (S)\). We omit the subscript r in the following. One sees that n and \(n_h\) are represented as

$$\begin{aligned} n(y', \varphi (y'))= & {} \frac{1}{\sqrt{1 + |\nabla '\varphi |^2}} \begin{pmatrix}\nabla '\varphi \\ -1\end{pmatrix}, \\ n_h(y', \varphi _h(y'))= & {} \frac{1}{\sqrt{1 + |\nabla '\varphi _h|^2}} \begin{pmatrix}\nabla '\varphi _h \\ -1\end{pmatrix}, \quad y'\in \varDelta . \end{aligned}$$

A direct computation gives

$$\begin{aligned} |n(y', \varphi (y')) - n_h(y', \varphi _h(y'))| \le 2|\nabla '(\varphi (y') - \varphi _h(y'))| \le 2C_{1E}h. \end{aligned}$$
(9.3)

This combined with the observation that

$$\begin{aligned} |n\circ \pi (y', \varphi _h(y')) - n(y', \varphi (y'))|&= |n\circ \pi (y', \varphi _h(y')) - n\circ \pi (y', \varphi (y'))| \\&\le C_1|\pi (y', \varphi _h(y')) - \pi (y', \varphi (y'))| \\&\le C_1 \Vert \nabla \pi \Vert _{L^\infty (\varGamma (\delta ))} |\varphi _h(y') - \varphi (y')| \\&\le C_1 \Vert \nabla \pi \Vert _{L^\infty (\varGamma (\delta ))} C_{0E}h^2 \end{aligned}$$

proves (9.1). When \(N=2\) and \(\varGamma \in C^{2,1}\), by using Taylor expansion, we find that \(m_S\) is a point of super-convergence such that \(|\frac{d\varphi _h}{dy_1} (b(m_S)) - \frac{d\varphi }{dy_1} (b(m_S))| \le Ch^2\), which improves (9.3) to \(O(h^2)\). Thus (9.2) is proved. \(\square \)

Remark 9.1

In the case \(N=3\), it is known that the barycenter of a triangle is not a point of super-convergence for the derivative of linear interpolations; see [34, p. 1930]. For this reason, (9.2) holds only for \(N=2\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kashiwabara, T., Oikawa, I. & Zhou, G. Penalty method with P1/P1 finite element approximation for the Stokes equations under the slip boundary condition. Numer. Math. 134, 705–740 (2016). https://doi.org/10.1007/s00211-016-0790-5

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00211-016-0790-5

Mathematics Subject Classification

Navigation