Skip to main content
Log in

A projection method for Navier–Stokes equations with a boundary condition including the total pressure

  • Published:
Numerische Mathematik Aims and scope Submit manuscript

Abstract

We consider a projection method for time-dependent incompressible Navier–Stokes equations with a total pressure boundary condition. The projection method is one of the numerical calculation methods for incompressible viscous fluids often used in engineering. In general, the projection method needs additional boundary conditions to solve a pressure-Poisson equation, which does not appear in the original Navier–Stokes problem. On the other hand, many mechanisms generate flow by creating a pressure difference, such as water distribution systems and blood circulation. We propose a new additional boundary condition for the projection method with a Dirichlet-type pressure boundary condition and no tangent flow. We demonstrate stability for the scheme and establish error estimates for the velocity and pressure under suitable norms. A numerical experiment verifies the theoretical convergence results. Furthermore, the existence of a weak solution to the original Navier–Stokes problem is proved by using the stability.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Subscribe and save

Springer+ Basic
$34.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or eBook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3

Similar content being viewed by others

Notes

  1. If \(d=2\), then we define

    $$\begin{aligned} v\times w:=v_x w_y-v_y w_x\in {\mathbb R}\qquad \text{ for } \text{ all } v=(v_x,v_y),w=(w_x,w_y)\in {\mathbb R}^2. \end{aligned}$$
  2. If \(d=2\), then \(\nabla \times v\) and \((\nabla \times v)\times w\) denote the scalar and vector functions, respectively, defined as follows: for all \(v=(v_x,v_y),w=(w_x,w_y)\in {\mathbb R}^2\),

    $$\begin{aligned} \nabla \times v:=\partial _x v_y-\partial _y v_x,\quad (\nabla \times v)\times w:=(w_y(\partial _y v_x-\partial _x v_y), w_x(\partial _x v_y-\partial _y v_x)). \end{aligned}$$

    .

  3. If one replaces the bilinear form a with \(a_\gamma (u,v):=(\nabla \times u,\nabla \times v) + \gamma (\text{ div }\,u, \text{ div }\,v)\) for a fixed \(0 < \gamma \ne 1\), then the corresponding strong form is different from (1.3), but Theorems 2.13, 2.15, and 2.20 hold since we have for all \(\varphi \in H\), \(a_\gamma (\varphi ,\varphi ) \ge (\min \{1,\gamma \}/c_{a})\Vert \varphi \Vert _1^2\) by Lemma 2.2 (cf. [8, Lemma 2.11]). If we choose a smaller value as \(\gamma \), the coefficients of Theorems 2.132.15, and  2.20 are larger (see (3.4)).

  4. Here, it holds that for all \(i,j=1,\ldots ,d\) and \(k=1,2,\ldots ,N\),

    $$\begin{aligned} (g_k)_i:=\sum _{l=1}^d\frac{\partial (u^*_k)_l}{\partial x_i}(u^*_{k-1})_l -(u^*_{k-1})_i\text{ div }\,u^*_k,\qquad (h_k)_{ij}:= -(u^*_k)_i(u^*_{k-1})_j. \end{aligned}$$
  5. Since \(p_2=2+\varepsilon \) and \(p_3=3\), we have \(p_2/{\tilde{q}}_2 = \varepsilon /2\) and \(p_3/{\tilde{q}}_3=1/2\).

References

  1. Badia, S., Codina, R.: Convergence analysis of the FEM approximation of the first order projection method for incompressible flows with and without the inf-sup condition. Numer. Math. 107, 533–557 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  2. Begue, C., Conca, C., Murat, F., Pironneau, O.: A nouveau sur les equations de Stokes et de Navier–Stokes avec des conditions aux limites sur la pression. C. R. Acad. Sc Paris, Serie I 304, 23–28 (1987)

    MATH  Google Scholar 

  3. Begue, C., Conca, C., Murat, F., Pironneau, O.: Les équations de Stokes et de Navier–Stokes avec des conditions aux limites sur la pression. In: H. Brezis, J.L. Lions (eds.) Nonlinear Partial Differential Equations and Their Applications, College de France, Seminar, vol. IX, pp. 179–264. Longman Scientific and Technical (1988)

  4. Benedict, R.P.: Fundamentals of Temperature, Pressure, and Flow Measurements, 3rd edn. Wiley, New York (1984)

    Book  Google Scholar 

  5. Bernard, J.M.: Non-standard Stokes and Navier–Stokes problems: existence and regularity in stationary case. Math. Meth. Appl. Sci. 25, 627–661 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  6. Bernard, J.M.: Time-dependent Stokes and Navier–Stokes problems with boundary conditions involving pressure, existence and regularity. Nonlinear.: Anal Real World Appl. 4, 805–839 (2003)

    MathSciNet  MATH  Google Scholar 

  7. Bernardi, C., Canuto, C., Maday, Y.: Spectral approximations of the Stokes equations with boundary conditions on the pressure. SIAM J. Numer. Anal. 28(2), 333–362 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  8. Bernardi, C., Rebollo, T.C., Yakoubi, D.: Finite element discretization of the Stokes and Navier–Stokes equations with boundary conditions on the pressure. SIAM J. Numer. Anal. 53(3), 1256–1279 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  9. Bertoluzza, S., Chabannes, V., Prud’homme, C., Szopos, M.: Boundary conditions involving pressure for the Stokes problem and applications in computational hemodynamics. Comput. Methods Appl. Mech. Engrg. 322, 58–80 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  10. Boyer, F., Fabrie, P.: Mathematical Tools for the Study of the Incompressible Navier–Stokes Equations and Related Models. Springer-Verlag, Berlin (2013)

    Book  MATH  Google Scholar 

  11. Cho, J.R.: Projection of the rotation form Navier–Stokes equation onto the half-staggered grid. J. Mech. Sci. Technol. 30, 3159–3164 (2016)

    Article  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  13. 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 

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

    Book  MATH  Google Scholar 

  15. Goldstein, R.J.: Fluid Mechanics Measurements. Hemisphere Pub Corp, London (1983)

    Google Scholar 

  16. Gresho, P.M.: Incompressible fluid dynamics: Some fundamental formulation issues. Annu. Rev. Fluid Mech. 23(1), 413–453 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  17. Guermond, J.L., Minev, P., Shen, J.: Error analysis of pressure-correction schemes for the time-dependent Stokes equations with open boundary conditions. SIAM J. Numer. Anal. 43(1), 239–258 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  18. Guermond, J.L., Minev, P., Shen, J.: An overview of projection methods for incompressible flows. Comput. Methods Appl. Mech. Engrg. 195, 6011–6045 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  19. Guermond, J.L., Quartapelle, L.: Calculation of incompressible viscous flows by an unconditionally stable projection FEM. J. Comput. Phys. 132, 12–33 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  20. Guermond, J.L., Quartapelle, L.: On the approximation of the unsteady Navier–Stokes equations by finite element projection methods. Numer. Math. 80, 207–238 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  21. Guermond, J.L., Shen, J.: On the error estimates for the rotational pressure-correction projection methods. Math. Comput. 73(248), 1719–1737 (2004)

    Article  MathSciNet  MATH  Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

  23. Heywood, J.G., Rannacher, R.: Finite-element approximation of the nonstationary Navier–Stokes problem part IV: Error analysis for second-order time discretization. SIAM J. Numer. Anal. 27(2), 353–384 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  24. Holman, J.P.: Experimental Methods for Engineers, 7th edn. McGraw-Hill, Bengaluru (2001)

    Google Scholar 

  25. Kangro, U., Nicolaides, R.: Divergence boundary conditions for vector Helmholtz equations with divergence constraints. Math. Modelling Numer. Anal. 33(3), 479–492 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  26. Kim, T.: Regularity of solutions to the Navier–Stokes equations with a nonstandard boundary condition. Acta Math. Appl. Sin-E. 31(3), 707–718 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  27. Kim, T., Cao, D.: Some properties on the surfaces of vector fields and its application to the Stokes and Navier–Stokes problems with mixed boundary conditions. Nonlinear Anal. 113, 94–114 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  28. Prohl, A.: Projection and Quasi-Compressibility Methods for Solving the Incompressible Navier–Stokes Equations. Teubner, B.G (1997)

  29. Rannacher, R.: On Chorin’s projection method for the incompressible Navier–Stokes equations. In: Heywood, J.G., Masuda, K., Rautmann, R., Solonnikov, V.A. (eds.) The Navier–Stokes Equations II — Theory and Numerical Methods, pp. 167–183. Springer, Berlin Heidelberg (1992)

    Chapter  MATH  Google Scholar 

  30. Rathakrishnan, E.: Instrumentation, Measurements, and Experiments in Fluids, 2nd edn. CRC Press, New York (2017)

    Google Scholar 

  31. Shen, J.: On error estimates of projection methods for Navier–Stokes equations: First-order schemes. SIAM J. Numer. Anal. 29(1), 57–77 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  32. Shen, J.: Remarks on the pressure error estimates for the projection methods. Numer. Math. 67, 513–520 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  33. Temam, R.: Sur l’approximation de la solution des équations de Navier–Stokes par la méthode des pas fractionnaires (II). Arch. Rational Mech. Anal. 33, 377–385 (1969)

    Article  MathSciNet  MATH  Google Scholar 

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

  35. Tropea, C., Yarin, A.L., Foss, J.F. (eds.): Springer Handbook of Experimental Fluid Mechanics. Springer Handbooks, Springer (2007)

  36. Wentong, S., Weilin, Y., Lucheng, J.: Effect of inlet total pressure non-uniform distribution on aerodynamic performance and flow field of turbine. Int. J. Mech. Sci. 148, 714–729 (2018)

    Article  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.

This work was supported by JSPS KAKENHI Grant Number 19J20514.

Proofs of Lemmas 2.4, 2.6, 3.2, 3.3, and Corollary 2.22

Proofs of Lemmas 2.4, 2.6, 3.2, 3.3, and Corollary 2.22

The purpose of this appendix is to provide the Proofs of Lemmas 2.4, 2.6, 3.2, 3.3, and Corollary 2.22. The continuity of the operator T follows from Lemmas 2.1 and 2.2 [14, Corollary 4.1]. By using Lemma 2.2 again, we prove the second inequality of Lemma 2.4.

Proof of Lemma 2.4

By Lemmas 2.1 and 2.2, there exists a unique solution \((w,r)\in H\times {L^2(\varOmega )}\) to (2.1) for all \(e\in {L^2(\varOmega )}^d\) and T is a continuous operator;

$$\begin{aligned} \Vert w\Vert _1+\Vert r\Vert _0\le c_1\Vert e\Vert _{H^*} \le c_1\Vert e\Vert _0 \end{aligned}$$

for a constant \(c_1>0\) independent of e [14, Corollary 4.1]. It is easy to check that T is a linear operator.

Next, we show the second inequality of Lemma 2.4. By the first equation of (2.1), it holds that for all \(\varphi \in V\),

$$\begin{aligned} a(w,\varphi )=(r,\text{ div }\,\varphi )+(e,\varphi )=(e,\varphi ). \end{aligned}$$

By Lemma 2.2, we have

$$\begin{aligned} \frac{1}{c_a}\Vert w\Vert _1^2 \le \Vert w\Vert _a^2=a(w,w)=(e,w)\le \Vert e\Vert _{V^*}\Vert w\Vert _1, \end{aligned}$$

which implies that \(\Vert w\Vert _1\le c_a\Vert e\Vert _{V^*}\). On the other hand, by Lemma 2.2, it holds that for all \(e\in {L^2(\varOmega )}^d\),

$$\begin{aligned} \Vert e\Vert _{V^*} =\sup _{0\ne \varphi \in V}\frac{(e,\varphi )}{\Vert \varphi \Vert _1} =\sup _{0\ne \varphi \in V}\frac{a(w,\varphi )}{\Vert \varphi \Vert _1} \le \sup _{0\ne \varphi \in V}\frac{c_a\Vert w\Vert _1\Vert \varphi \Vert _1}{\Vert \varphi \Vert _1} =c_a\Vert w\Vert _1. \end{aligned}$$

\(\square \)

In order to prove Lemma 2.6 and Corollary 2.22, we define \({\tilde{p}}_d, {\tilde{q}}_d\) as

$$\begin{aligned} {\tilde{p}}_d:= \frac{2p_d}{p_d+2} = \frac{1}{\frac{1}{2} + \frac{1}{p_d}},\qquad {\tilde{q}}_d:= \frac{{\tilde{p}}_d}{1 - {\tilde{p}}_d} = \frac{1}{1 - \frac{1}{2} - \frac{1}{p_d}}. \end{aligned}$$

Here, since \(p_2=2+\varepsilon \) and \(p_3=3\), we have \(({\tilde{p}}_2, {\tilde{q}}_2) = (\frac{4+2\varepsilon }{4+\varepsilon }, 2+\frac{4}{\varepsilon })\) and \(({\tilde{p}}_2, {\tilde{q}}_2) = (\frac{6}{5}, 6)\). By the Sobolev embeddings [10, Theorem III.2.33], it holds that \({H^1(\varOmega )}\subset L^{{\tilde{q}}_d}(\varOmega )\), \({H^1(\varOmega )}\subset L^{p_d}(\varOmega )\), \(H^2(\varOmega ) \subset L^\infty (\varOmega )\) and the embeddings are continuous.

Proof of Lemma 2.6

(i) For all \(u \in L^{p_d}(\varOmega )^d, v, w \in H\), we have

$$\begin{aligned} |d(u,v,w)|&\le \int _\varOmega \left| u\cdot ((w\cdot \nabla )v-(v\cdot \nabla )w +v\,\text{ div }\,w-w\,\text{ div }\,v)\right| dx\\&\le c_1\Vert u\Vert _{L^{p_d}} (\Vert w\Vert _{L^{{\tilde{q}}_d}}\Vert \nabla v\Vert _0 +\Vert v\Vert _{L^{{\tilde{q}}_d}}\Vert \nabla w\Vert _0 +\Vert v\Vert _{L^{{\tilde{q}}_d}}\Vert \text{ div }\,w\Vert _0 +\Vert w\Vert _{L^{{\tilde{q}}_d}}\Vert \text{ div }\,v\Vert _0)\\&\le {\tilde{c}}_1\Vert u\Vert _{L^{p_d}}\Vert v\Vert _1\Vert w\Vert _1 \end{aligned}$$

for two constants \(c_1,{\tilde{c}}_1>0\), which implies the third inequality of Lemma 2.6.

(ii) For all \(u \in L^{p_d}(\varOmega )^d, v \in H, w \in H \cap H^2(\varOmega )^d\), we have

$$\begin{aligned}&|d(u,v,w)|\\&\quad \le c_2\Vert u\Vert _0 (\Vert w\Vert _{L^\infty }\Vert \nabla v\Vert _0 +\Vert v\Vert _{L^{p_d}}\Vert \nabla w\Vert _{L^{{\tilde{q}}_d}} +\Vert v\Vert _{L^{p_d}}\Vert \text{ div }\,w\Vert _{L^{{\tilde{q}}_d}} +\Vert w\Vert _{L^\infty }\Vert \text{ div }\,v\Vert _0)\\&\quad \le {\tilde{c}}_2\Vert u\Vert _0\Vert v\Vert _1\Vert w\Vert _2 \end{aligned}$$

for two constants \(c_2,{\tilde{c}}_2>0\).

(iii) For all \(u \in {H^1(\varOmega )}^d, v \in H \cap H^2(\varOmega )^d, w \in H\), we have

$$\begin{aligned} |d(u,v,w)|\le & {} \int _\varOmega |((\nabla \times u)\times v)\cdot w|dx \le c_3\Vert \nabla \times u\Vert _0\Vert v\Vert _{L^\infty }\Vert w\Vert _0\\\le & {} {\tilde{c}}_3\Vert u\Vert _1\Vert v\Vert _2\Vert w\Vert _0 \end{aligned}$$

for two constants \(c_3,{\tilde{c}}_3>0\).

(ix) For all \(u \in H^2(\varOmega )^d, v, w\in H\), we have

$$\begin{aligned} |d(u,v,w)| \le c_4\Vert \nabla \times u\Vert _{L^{p_d}}\Vert v\Vert _{L^{{\tilde{q}}_d}}\Vert w\Vert _0 \le {\tilde{c}}_4\Vert u\Vert _2\Vert v\Vert _1\Vert w\Vert _0 \end{aligned}$$

for two constants \(c_4,{\tilde{c}}_4>0\). \(\square \)

Next, we prove Lemmas 3.2 and 3.3.

Proof of Lemma 3.2

It holds that for all \(\varphi \in H\) and \(k=1,2,\ldots ,N\),

$$\begin{aligned} \begin{aligned} \quad \langle R_k,\varphi \rangle _H = \tau \int ^1_0\left\langle s\frac{\partial ^2 u}{\partial t^2}(t_{k-1}+s\tau ),\varphi \right\rangle _H ds \le \sqrt{\frac{\tau }{3}}\Vert \varphi \Vert _1 \left\| \frac{\partial ^2 u}{\partial t^2}\right\| _{L^2(t_{k-1},t_k;H^*)}, \end{aligned} \end{aligned}$$

which implies that

$$\begin{aligned} \Vert {\bar{R}}_\tau \Vert _{L^2(H^*)}^2 =\sum ^N_{k=1}\tau \left( \sup _{0\ne \varphi \in H} \frac{\langle R_k,\varphi \rangle _H}{\Vert \varphi \Vert _1}\right) ^2 \le \frac{1}{3}\tau ^2 \left\| \frac{\partial ^2 u}{\partial t^2}\right\| _{L^2(H^*)}^2. \end{aligned}$$

Next, we show the second inequality of the conclusion. For all \(\varphi \in H\) and \(k=2,3,\ldots ,N\), we have

$$\begin{aligned}\begin{aligned} \langle D_\tau R_k,\varphi \rangle _H&= \tau \int ^1_0\int ^1_0\left\langle s_1\frac{\partial ^3 u}{\partial t^3}(t_{k-2}+s_1\tau +s_2\tau ), \varphi \right\rangle _H ds_1 ds_2\\&\le \sqrt{\frac{\tau }{3}}\Vert \varphi \Vert _1 \left\| \frac{\partial ^3 u}{\partial t^3}\right\| _{L^2(t_{k-2},t_k;H^*)}, \end{aligned} \end{aligned}$$

where we have used the coordinate transformation \((s_1,s_2)\mapsto ({\tilde{s}}_1,{\tilde{s}}_2):=(s_1+s_2,-s_1+s_2)\). Therefore, we obtain

$$\begin{aligned} \sum ^N_{k=2}\tau \Vert D_\tau R_k\Vert _{H^*}^2 \le \sum ^N_{k=2}\tau \frac{\tau }{3} \left\| \frac{\partial ^3 u}{\partial t^3}\right\| _{L^2(t_{k-2},t_k;H^*)}^2 \le \frac{2}{3}\tau ^2 \left\| \frac{\partial ^3 u}{\partial t^3}\right\| _{L^2(H^*)}^2. \end{aligned}$$

\(\square \)

Proof of Lemma 3.3

It holds that for all \(k = 1,2,\ldots ,N\) and \(t \in [t_{k-1}, t_k]\),

$$\begin{aligned} \Vert x(t_k) - x(t)\Vert _E \le \int ^{t_k}_{t} \left\| \frac{\partial x}{\partial t}(s)\right\| _E ds \le \sqrt{t_k - t} \left\| \frac{\partial x}{\partial t}\right\| _{L^2(t_{k-1},t_k;E)}, \end{aligned}$$

which implies that \(\Vert x - x_\tau \Vert _{L^\infty (E)} \le \sqrt{\tau } \left\| \frac{\partial x}{\partial t}\right\| _{L^2(E)}\) and

$$\begin{aligned} \Vert D_\tau x(t_k)\Vert _E =\frac{1}{\tau } \Vert x(t_k) - x(t_{k-1})\Vert _E \le \frac{1}{\sqrt{\tau }} \left\| \frac{\partial x}{\partial t}\right\| _{L^2(t_{k-1},t_k;E)}. \end{aligned}$$

On the other hand, we have

$$\begin{aligned} \begin{aligned} \Vert x-x_\tau \Vert _{L^2(E)}^2 \le \sum ^N_{k=1}\int ^{t_k}_{t_{k-1}}(t_k-t)dt \left\| \frac{\partial x}{\partial t}\right\| _{L^2(t_{k-1},t_k;E)}^2 \le \frac{1}{2}\tau ^2\left\| \frac{\partial x}{\partial t}\right\| _{L^2(E)}^2. \end{aligned} \end{aligned}$$

\(\square \)

We prove Corollary 2.22 by using the boundedness from Theorem 2.13 and the Aubin–Lions compactness lemma.

Proof of Corollary 2.22

By the first and third equations of (PM), it holds that for all \(v\in V\) and \(k=1,2,\ldots ,N\),

$$\begin{aligned} \begin{aligned} (D_\tau u_k,v)+a(u^*_k,v)+(g_k,v)+(h_k,\nabla v)&=\langle f_k,v\rangle _H-\int _{\varGamma _2}p^b_kv\cdot nds, \end{aligned} \end{aligned}$$

where \(g_k\) and \(h_k\) are definedFootnote 4 by

$$\begin{aligned} g_k:= (\nabla u^*_k)^T u^*_{k-1} - u^*_{k-1}\text{ div }\,u^*_k,\qquad h_k:= -u^*_k(u^*_{k-1})^T, \end{aligned}$$

which implies that for all \(v\in V\) and \(\theta \in C^\infty _0(0,T)\),

$$\begin{aligned} \begin{aligned}&\int ^T_0\left( \left( \frac{\partial {\hat{u}}_\tau }{\partial t},v\right) +a({\bar{u}}^*_\tau ,v) +({\bar{g}}_\tau ,v)+({\bar{h}}_\tau ,\nabla v)\right) \theta dt \!\\&\quad =\!\int ^T_0\left( \langle {\bar{f}}_\tau ,v\rangle _H - \int _{\varGamma _2}{\bar{p}}^b_\tau v\cdot nds\right) \theta dt. \end{aligned} \end{aligned}$$
(A.1)

Here, \({\bar{f}}_\tau \rightarrow f\) strongly in \(L^2(H^*)\) and \({\bar{p}}^b_\tau \rightarrow p^b\) strongly in \(L^2({H^1(\varOmega )})\) as \(\tau \rightarrow 0\). By Theorem 2.13 and Lemma 3.1, there exists a constant \(c_1>0\) such that

$$\begin{aligned} \begin{aligned} \Vert {\bar{u}}_\tau \Vert _{L^\infty (L^2)}^2 \!+\! \Vert {\bar{u}}^*_\tau \Vert _{L^\infty (L^2)}^2 \!+\! \Vert {\bar{u}}^*_k\Vert _{L^2(H^1)}^2 \!+\! \tau \left\| \frac{\partial {\hat{u}}_\tau }{\partial t}\right\| _{L^2(L^2)}^2 \!\!+\! \frac{1}{\tau }\Vert {\bar{u}}_\tau -{\bar{u}}^*_\tau \Vert _{L^2(L^2)}^2 \!\le \! c_1. \end{aligned} \end{aligned}$$
(A.2)

In particular, it holds that

$$\begin{aligned} \Vert u^*_1\Vert _0^2 + \tau \Vert u^*_1\Vert _1^2 + \Vert u_1-u_0\Vert _0^2 + \Vert u_1-u^*_1\Vert _0^2 \le c_1, \end{aligned}$$
(A.3)

which implies that \(\Vert u^*_1-u_0\Vert _0 \le \Vert u^*_1-u_1\Vert _0 + \Vert u_1-u_0\Vert _0 \le 2\sqrt{c_1}\). Furthermore, by the first equation of (PM) and Lemmas 2.2, 2.6, we have

$$\begin{aligned} \begin{aligned} \Vert u^*_1-u_0\Vert _{V^*}&= \sup _{0\ne v\in H}\frac{\tau }{\Vert v\Vert _1} \left| -a(u^*_1,v) - d(u_0,u^*_1,v) + \langle f_1, v\rangle _H\right| \\&\le c_a\tau \Vert u^*_1\Vert _1 + c_d\tau \Vert u_0\Vert _{L^{p_d}}\Vert u^*\Vert _1 + \tau \Vert f_1\Vert _{H^*} \le c_2\sqrt{\tau }. \end{aligned} \end{aligned}$$
(A.4)

where \(c_2:=\sqrt{c_1}(c_a + c_d\Vert u_0\Vert _{L^{p_d}}) + \Vert f\Vert _{L^2(H^*)}\). Let \(u^\circ _0:=u^*_1\), \(u^\circ _k:=u^*_k\) for all \(k=1,2,\ldots ,N\) and let \({\hat{u}}^\circ _\tau \) be the piecewise linear interpolant of \((u^\circ _k)_{k=0}^N \subset H\).

From the uniform estimates (A.2), one can show that there exist a sequence \((\tau _k)_{k\in {\mathbb N}}\) and three functions \(u\in L^2(H)\cap L^\infty ({L^2(\varOmega )}^d)\cap W^{1,4/p_d}(V^*)\) (in particular, \(u\in C([0,T];V^*)\)), \(g\in L^{4/p_d}(L^{{\tilde{p}}_d}(\varOmega )^d)\) and \(h\in L^{4/p_d}({L^2(\varOmega )}^{d\times d})\) such that \(\tau _k\rightarrow 0\) and

$$\begin{aligned} {\bar{u}}^*_{\tau _k}\rightarrow u&\quad \text {weakly in }L^2(H), \end{aligned}$$
(A.5)
$$\begin{aligned}&\quad \text {strongly in }L^2({L^2(\varOmega )}^d), \end{aligned}$$
(A.6)
$$\begin{aligned} {\hat{u}}^\circ _{\tau _k}\rightarrow u&\quad \text {strongly in }L^2({L^2(\varOmega )}^d), \end{aligned}$$
(A.7)
$$\begin{aligned}&\quad \text {strongly in }C([0,T];V^*), \end{aligned}$$
(A.8)
$$\begin{aligned} {\hat{u}}_{\tau _k} \rightarrow u&\quad \text {strongly in }L^2({L^2(\varOmega )}^d), \end{aligned}$$
(A.9)
$$\begin{aligned}&\quad \text {weakly in }W^{1,4/p_d}(V^*), \end{aligned}$$
(A.10)
$$\begin{aligned} {\bar{g}}_{\tau _k} \rightharpoonup g&\quad \text {weakly in }L^{4/p_d}(L^{{\tilde{p}}_d}(\varOmega )^d), \end{aligned}$$
(A.11)
$$\begin{aligned} {\bar{h}}_{\tau _k} \rightharpoonup h&\quad \text {weakly in }L^{4/p_d}({L^2(\varOmega )}^{d\times d}), \end{aligned}$$
(A.12)

as \(k\rightarrow \infty \). Here, we note that \({\bar{u}}^*_{\tau _k}\), \({\hat{u}}^\circ _{\tau _k}\) and \({\hat{u}}_{\tau _k}\) possess a common limit function. Indeed, the weak convergence (A.5) of \({\bar{u}}^*_\tau \) immediately follows from the uniform estimates (A.2). Since we have \(1/{\tilde{p}}_d = 1/2 + 1/p_d\), \(p_d/4 = 1/2 + p_d/(2{\tilde{q}}_d)\), and

$$\begin{aligned} \Vert {\bar{u}}^*_\tau \Vert _{L^{2{\tilde{q}}_d/p_d}(L^{p_d})} \le \Vert {\bar{u}}^*_\tau \Vert _{L^2(L^{{\tilde{q}}_d})}^{p_d/{\tilde{q}}_d} \Vert {\bar{u}}^*_\tau \Vert _{L^\infty (L^2)}^{1-p_d/{\tilde{q}}_d} \le c_3\Vert {\bar{u}}^*_\tau \Vert _{L^2(H^1)}^{p_d/{\tilde{q}}_d} \Vert {\bar{u}}^*_\tau \Vert _{L^\infty (L^2)}^{1-p_d/{\tilde{q}}_d} \end{aligned}$$

for a constant \(c_3>0\) (cf. [10, Theorem II.5.5]),Footnote 5 it holds that

$$\begin{aligned}&\Vert {\bar{g}}_\tau \Vert _{L^{4/p_d}(L^{{\tilde{p}}_d})} \le \! c_4\left( \!\tau \sum ^N_{k=1} \Vert u^*_k\Vert _1^{4/{p_d}}\Vert u^*_{k-1}\Vert _{L^{p_d}}^{4/{p_d}}\!\right) ^{{p_d}/4}\\&\quad \le c_4 \left( \Vert {\bar{u}}^*_\tau \Vert _{L^2(H^1)}^2 + \Vert {\bar{u}}^*_\tau \Vert _{L^{2{\tilde{q}}_d/p_d}(L^{p_d})}^2 + \tau ^{p_d/{\tilde{q}}_d}\Vert u_0\Vert _{L^{p_d}}^2\right) \\&\quad \le c_4\left( \Vert {\bar{u}}^*_\tau \Vert _{L^2(H^1)}^2 + c_3\Vert {\bar{u}}^*_\tau \Vert _{L^2(H^1)}^{2p_d/{\tilde{q}}_d} \Vert {\bar{u}}^*_\tau \Vert _{L^\infty (L^2)}^{2-2p_d/{\tilde{q}}_d} + \tau ^{p_d/{\tilde{q}}_d}\Vert u_0\Vert _{L^{p_d}}^2\right) ,\\&\qquad \Vert {\bar{h}}_\tau \Vert _{L^{4/{p_d}}(L^2)} \le c_5\!\!\left( \tau \sum ^N_{k=1} \Vert u^*_k\Vert _1^{4/{p_d}}\Vert u^*_{k-1}\Vert _{L^{p_d}}^{4/{p_d}}\right) ^{{p_d}/4}\\&\quad \le c_5\left( \Vert {\bar{u}}^*_\tau \Vert _{L^2(H^1)}^2 + c_3\Vert {\bar{u}}^*_\tau \Vert _{L^2(H^1)}^{2p_d/{\tilde{q}}_d} \Vert {\bar{u}}^*_\tau \Vert _{L^\infty (L^2)}^{2-2p_d/{\tilde{q}}_d} + \tau ^{p_d/{\tilde{q}}_d}\Vert u_0\Vert _{L^{p_d}}^2\right) \end{aligned}$$

for constants \(c_4\) and \(c_5\). Hence, by (A.2), the weak convergences (A.11) and (A.12) hold. Moreover, since there exists a constant \(c_6>0\) such that \(|(g_k,v)| \le \Vert g_k\Vert _{L^{{\tilde{p}}_d}}\Vert v\Vert _{L^{{\tilde{q}}_d}} \le c_6\Vert g_k\Vert _{L^{{\tilde{p}}_d}}\Vert v\Vert _1\) for all \(k=1,2,\ldots ,N\) and \(v\in {H^1(\varOmega )}^d\), we have

$$\begin{aligned}&\left\| \frac{\partial {\hat{u}}_\tau }{\partial t}\right\| _{L^{4/{p_d}}(V^*)}\\&\quad = \biggl \{\int ^T_0 \biggl (\sup _{0\ne v\in V}\frac{1}{\Vert v\Vert _1} \biggl |-a({\bar{u}}^*_\tau (t),v) -({\bar{g}}_\tau (t),v)-({\bar{h}}_\tau (t),\nabla v)\\&\qquad +\langle f_\tau (t),v\rangle _H -\int _{\varGamma _2}p^b_\tau (t) v\cdot nds\biggr |\biggr )^{4/{p_d}}dt\biggr \}^{{p_d}/4}\\&\quad \le \! T^{p_d/(2{\tilde{q}}_d)}(c_a\sqrt{c_1} \!+\! \Vert f\Vert _{L^2(H^*)} \!+\! \Vert p^b\Vert _{L^2(H^1)}) \!+\! c_6\Vert {\bar{g}}_\tau \Vert _{L^{4/{p_d}}(L^{{\tilde{p}}_d})} \!+\! \Vert {\bar{h}}_\tau \Vert _{L^{4/{p_d}}(L^2)},\\&\left\| \frac{\partial {\hat{u}}^\circ _\tau }{\partial t} -\frac{\partial {\hat{u}}_\tau }{\partial t}\right\| _{L^{4/{p_d}}(V^*)} \\&\quad = \left\{ \tau \sum ^N_{k=1}\left( \sup _{0\ne v\in V}\frac{ |(u^\circ _k-u^\circ _{k-1}-u_k+u_{k-1},v)| }{\tau \Vert v\Vert _1} \right) ^{4/{p_d}}\right\} ^{{p_d}/4}\\&\quad \le 2\left( \tau \sum ^N_{k=1}\sup _{0\ne v\in V} \frac{|(\nabla P_k,v)|^{4/{p_d}}}{\Vert v\Vert _1^{4/{p_d}}} + \tau \Vert u^*_1-u_0\Vert _{V^*}^{4/{p_d}} \right) ^{{p_d}/4}\\&\quad \le 2\left( \tau \sum ^N_{k=1}\sup _{0\ne v\in V} \frac{1}{\Vert v\Vert _1^{4/{p_d}}}\left| \int _{\varGamma _2} p^b(t_k) v\cdot nds\right| ^{4/{p_d}} \right) ^{{p_d}/4} + 2\tau ^{p_d/4}\Vert u^*_1-u_0\Vert _{V^*} \\&\quad \le 2\left( \tau \sum ^N_{k=1}\Vert p^b(t_k)\Vert _1^{4/{p_d}}\right) ^{{p_d}/4} + 2c_2\tau ^{p_d/4+1/2} \le 2T^{p_d/(2{\tilde{q}}_d)}(\Vert p^b\Vert _{L^2(H^1)} + c_2 T), \end{aligned}$$

and \(\Vert \frac{\partial {\hat{u}}_\tau }{\partial t}\Vert _{L^{4/{p_d}}(V^*)}\) and \(\Vert \frac{\partial {\hat{u}}^\circ _\tau }{\partial t}\Vert _{L^{4/{p_d}}(V^*)}\) are also bounded. Hence, (A.10) holds. Furthermore, \(\Vert {\hat{u}}^\circ _\tau \Vert _{L^2(H^1)}\) is bounded: by (A.2) and (A.3),

$$\begin{aligned} \begin{aligned} \Vert {\hat{u}}^\circ _\tau \Vert _{L^2(H^1)}^2&\!\le \! \sum ^N_{k=1}\tau (\Vert u^\circ _{k-1}\Vert _1^2 + \Vert u^\circ _k\Vert _1^2)\int ^1_0\{(1-s)^2+s^2\} ds\\&\!\le \! \frac{4}{3}\Vert {\bar{u}}^*_\tau \Vert _{L^2(H^1)}^2 + \frac{2c_1}{3} \!\le \! 2c_1, \end{aligned} \end{aligned}$$

which implies the strong convergence (A.7) of \({\hat{u}}^\circ _\tau \) in \(L^2({L^2(\varOmega )}^d)\) from the Aubin–Lions lemma [10, Theorem II.5.16 (i)]. Since we have for all \(t\in (t_{k-1},t_k), k=1,2,\ldots ,N\),

$$\begin{aligned} \begin{aligned} \Vert {\bar{u}}^*_\tau (t)-{\hat{u}}^\circ _\tau (t)\Vert _0&= \left| \frac{t_k-t}{\tau }\right| \Vert u^\circ _k-u^\circ _{k-1}\Vert _0\\&\le \Vert u^\circ _k-u_k\Vert _0+\tau \Vert D_\tau u_k\Vert _0+\Vert u_{k-1}-u^\circ _{k-1}\Vert _0,\\ \Vert {\bar{u}}^*_\tau (t)-{\hat{u}}_\tau (t)\Vert _0&\le \Vert {\bar{u}}^*_\tau (t)-{\bar{u}}_\tau (t)\Vert _0 +\Vert {\bar{u}}_\tau (t)-{\hat{u}}_\tau (t)\Vert _0\\&\le \Vert u^*_k-u_k\Vert _0+\tau \Vert D_\tau u_k\Vert _0, \end{aligned} \end{aligned}$$

the functions \({\bar{u}}^*_{\tau _k}\), \({\hat{u}}^\circ _{\tau _k}\) and \({\hat{u}}_{\tau _k}\) possess a common limit function u, and the strong convergences (A.6) and (A.9) hold: by (A.2) and (A.3),

$$\begin{aligned} \Vert {\bar{u}}^*_\tau -{\hat{u}}^\circ _\tau \Vert _{L^2(L^2)}\le & {} 2\sqrt{3}\Vert {\bar{u}}^*_\tau -{\bar{u}}_\tau \Vert _{L^2(L^2)} \!+\! \sqrt{3}\tau \left\| \frac{\partial {\hat{u}}_\tau }{\partial t}\right\| _{L^2(L^2)}\\&+\, \sqrt{3\tau } \Vert u_0-u^*_1\Vert _0 \le 5\sqrt{3c_1 \tau }, \Vert {\bar{u}}^*_\tau -{\hat{u}}_\tau \Vert _{L^2(L^2)}\\\le & {} \sqrt{2}\Vert {\bar{u}}^*_\tau -{\bar{u}}_\tau \Vert _{L^2(L^2)} +\sqrt{2}\tau \left\| \frac{\partial {\hat{u}}_\tau }{\partial t}\right\| _{L^2(L^2)} \le 2\sqrt{2c_1\tau }. \end{aligned}$$

It also holds that

$$\begin{aligned} \Vert {\bar{u}}^*_\tau -{\hat{u}}^\circ _\tau \Vert _{L^\infty (L^2)} \le \max _{k=1,2,\ldots ,N}(\Vert u^\circ _k\Vert _0+\Vert u^\circ _{k-1}\Vert _0) \le 2\sqrt{c_1}. \end{aligned}$$

Since \(\Vert {\hat{u}}^\circ _\tau \Vert _{L^\infty (L^2)}\) and \(\Vert \frac{\partial {\hat{u}}^\circ _\tau }{\partial t}\Vert _{L^{4/p_d}(V^*)}\) are bounded, we obtain the strong convergence (A.8) of \({\hat{u}}^\circ _\tau \) in \(C([0,T];V^*)\) [10, Theorem II.5.16 (ii)]. In particular, \({\hat{u}}^\circ _\tau (0)\) converges to u(0) in \(V^*\). On the other hand, by (A.4), \({\hat{u}}^\circ _\tau (0)=u^*_1\) converges to \(u_0\) in \(V^*\). Through the uniqueness of the limit in \(V^*\), we have indeed obtained that \(u(0)=u_0\).

From (A.1) with \(\varepsilon :=\varepsilon _k\), taking \(k\rightarrow \infty \), it holds that for all \(v\in V\) and \(\theta \in C^\infty _0(0,T)\),

$$\begin{aligned} \begin{aligned}&\, \int ^T_0\left( \left\langle \frac{\partial u}{\partial t}, \theta v\right\rangle _V + a(u,\theta v) + (g,\theta v) + (h,\nabla (\theta v))\right) dt\\&\quad =\int ^T_0\left( \langle f,\theta v\rangle _H - \int _{\varGamma _2}p^b\theta v\cdot nds\right) dt. \end{aligned} \end{aligned}$$

Next, we show that

$$\begin{aligned} g = (\nabla u)^T u - u\text{ div }\,u,\qquad h = -u(u)^T. \end{aligned}$$
(A.13)

We set \({\bar{v}}_\tau (t):=u^*_{k-1}\) for \(t\in (t_{k-1},t_k],k=1,2,\ldots ,N\). Then it holds that

$$\begin{aligned} \begin{aligned} \Vert {\bar{v}}_\tau -{\bar{u}}^*_\tau \Vert _{L^2(L^2)}&\!\le \! \left( \tau \sum _{k=1}^N(\Vert u^*_k-u_k\Vert _0 + \tau \Vert D_\tau u_k\Vert _0 + \Vert u_{k-1}-u^*_{k-1}\Vert _0)^2\right) ^{1/2}\\&\!\le \! 3\sqrt{3c_1 \tau }, \end{aligned} \end{aligned}$$

and hence it follows from (A.6) that \({\bar{v}}_{\tau _k}\rightarrow u\) strongly in \(L^2({L^2(\varOmega )}^d)\) as \(k\rightarrow \infty \). Since \(\nabla {\bar{u}}^*_\tau \rightharpoonup \nabla u\) weakly in \(L^2({L^2(\varOmega )}^{d\times d})\) and \(\text{ div }\,{\bar{u}}^*_\tau \rightharpoonup \text{ div }\,u\) weakly in \(L^2({L^2(\varOmega )})\) as \(k\rightarrow \infty \), we have

$$\begin{aligned} \begin{array}{rcll} {\bar{g}}_\tau = (\nabla {\bar{u}}^*_\tau )^T {\bar{v}}_\tau - {\bar{v}}_\tau \text{ div }\,{\bar{u}}^*_\tau &{}\rightharpoonup &{} (\nabla u)^T u - u\text{ div }\,u &{}\text {weakly in }L^1(L^1(\varOmega )^d),\\[4pt] {\bar{h}}_\tau = -{\bar{u}}^*_\tau ({\bar{v}}_\tau )^T &{}\rightarrow &{} -u(u)^T &{}\text {strongly in }L^1(L^1(\varOmega )^{d\times d}) \end{array} \end{aligned}$$

as \(k\rightarrow \infty \) (cf. [10, Proposition II.2.12]). On the other hand, we also know (A.11) and (A.12). The convergence in these spaces imply the convergence in the distributions sense, therefore (A.13) holds by the uniqueness of the limit in \({\mathcal {D}}'((0,T)\times \varOmega )\). Hence, it holds that for all \(v\in V\) and \(\theta \in C^\infty _0(0,T)\),

$$\begin{aligned} \begin{aligned}&\int ^T_0\left( \left\langle \frac{\partial u}{\partial t},v\right\rangle _V +a(u,v)+((\nabla u)u-u\text{ div }\,u,v) - (u(u)^T,\nabla v)\right) \theta dt\\&\quad =\int ^T_0\left( \langle f,v\rangle _H - \int _{\varGamma _2}p^bv\cdot nds\right) \theta dt, \end{aligned} \end{aligned}$$

which is equivalent to the following

$$\begin{aligned} \int ^T_0\left( \left\langle \frac{\partial u}{\partial t},v\right\rangle _V + a(u,v) + d(u,u,v)\right) \theta dt =\int ^T_0\left( \langle f,v\rangle _H - \int _{\varGamma _2}p^bv\cdot nds\right) \theta dt. \end{aligned}$$

\(\square \)

Rights and permissions

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Matsui, K. A projection method for Navier–Stokes equations with a boundary condition including the total pressure. Numer. Math. 152, 663–699 (2022). https://doi.org/10.1007/s00211-022-01323-x

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00211-022-01323-x

Mathematics Subject Classification

Navigation