Skip to main content

Computability of the Solutions to Navier-Stokes Equations via Effective Approximation

  • Chapter
  • First Online:
Book cover Complexity and Approximation

Part of the book series: Lecture Notes in Computer Science ((LNTCS,volume 12000))

Abstract

As one of the seven open problems in the addendum to their 1989 book Computability in Analysis and Physics, Pour-El and Richards proposed “... the recursion theoretic study of particular nonlinear problems of classical importance. Examples are the Navier-Stokes equation, the KdV equation, and the complex of problems associated with Feigenbaum’s constant.” In this paper, we approach the question of whether the Navier-Stokes Equation admits recursive solutions in the sense of Weihrauch’s Type-2 Theory of Effectivity. A natural encoding (“representation”) is constructed for the space of divergence-free vector fields on 2-dimensional open square \(\varOmega = (-1, 1)^2\). This representation is shown to render first the mild solution to the Stokes Dirichlet problem and then a strong local solution to the nonlinear inhomogeneous incompressible Navier-Stokes initial value problem uniformly computable. Based on classical approaches, the proofs make use of many subtle and intricate estimates which are developed in the paper for establishing the computability results.

The paper is dedicated to the memory of Professor Ker-I Ko.

The third author is supported by grant NRF-2017R1E1A1A03071032.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 54.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 69.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    We use \(q\in [1,\infty ]\) to denote the norm index, \(P\) for the pressure field, \(p\) for polynomials, \(\mathcal {P}\) for a set of trimmed and mollified tuples of the latter, and \(\mathbb {P}\) for the Helmholtz Projection.

References

  1. Beggs, E., Costa, J.F., Tucker, J.V.: Axiomatising physical experiments as oracles to algorithms. Philos. Trans. R. Soc. A: Math. Phys. Eng. Sci. 370, 3359–3384 (2012)

    Article  Google Scholar 

  2. Boyer, F., Fabrie, P.: Mathematical Tools for the Study of the Incompressible Navier-Stokes Equations and Related Models. Spring Applied Mathematical Sciences. Springer, Heidelberg (2013). https://doi.org/10.1007/978-1-4614-5975-0

    Book  MATH  Google Scholar 

  3. Giga, Y.: Weak and strong solutions of the Navier-Stokes initial value problem. Publ. RIMS Kyoto Univ. 19, 887–910 (1983)

    Article  MathSciNet  Google Scholar 

  4. Giga, Y.: Time and spatial analyticity of solutions of the Navier-Stokes equations. Commun. Partial Differ. Equ. 8, 929–948 (1983)

    Article  MathSciNet  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  6. Girault, V., Raviart, P.A.: Finite Element Methods for Navier-Stokes Equations. Springer Series in Computational Mathematics, vol. 5. Springer, New York (1986). https://doi.org/10.1007/978-3-642-61623-5

    Book  MATH  Google Scholar 

  7. Kawamura, A., Steinberg, F., Ziegler, M.: Complexity of Laplace’s and Poisson’s equation. Bull. Symb. Logic 20(2), 231 (2014). Full version to appear in Mathem. Structures in Computer Science (2016)

    Google Scholar 

  8. Landriani, G.S., Vandeven, H.: Polynomial approximation of divergence-free functions. Math. Comput. 52, 103–130 (1989)

    Article  MathSciNet  Google Scholar 

  9. Mclean, W.: Strongly Elliptic Systems and Boundary Integral Equations. Cambridge University Press, London (2000)

    MATH  Google Scholar 

  10. Pazy, A.: Semigroups of Linear Operators and Applications to Partial Differential Equations. Springer, New York (1983). https://doi.org/10.1007/978-1-4612-5561-1

    Book  MATH  Google Scholar 

  11. Pour-El, M.B., Richards, J.I.: The wave equation with computable initial data such that its unique solution is not computable. Adv. Math. 39(4), 215–239 (1981)

    Article  MathSciNet  Google Scholar 

  12. Pour-El, M.B., Richards, J.I.: Computability in Analysis and Physics. Springer, New York (1989)

    Book  Google Scholar 

  13. Pour-El, M.B., Zhong, N.: The wave equation with computable initial data whose unique solution is nowhere computable. Math. Logic Q. 43(4), 499–509 (1997)

    Article  MathSciNet  Google Scholar 

  14. Patel, M.K., Markatos, N.C., Cross, M.: A critical evaluation of seven discretization schemes for convection-diffusion equations. Int. J. Numer. Meth. Fluids 5(3), 225–244 (1985)

    Article  Google Scholar 

  15. Brattka, V., Presser, G.: Computability on subsets of metric spaces. Theor. Comput. Sci. 305, 43–76 (2003)

    Article  MathSciNet  Google Scholar 

  16. Smith, W.D.: On the uncomputability of hydrodynamics. NEC preprint (2003)

    Google Scholar 

  17. Soare, R.I.: Computability and recursion. Bull. Symb. Logic 2, 284–321 (1996)

    Article  MathSciNet  Google Scholar 

  18. Sohr, H.: The Navier-Stokes Equations: An Elementary Functional Analytic Approach. Birkhäuser Advanced Texts. Birkhäuser, New York (2001)

    Book  Google Scholar 

  19. Sun, S.M., Zhong, N., Ziegler, M.: On computability of Navier-Stokes’ equation. In: Beckmann, A., Mitrana, V., Soskova, M. (eds.) CiE 2015. LNCS, vol. 9136, pp. 334–342. Springer, Cham (2015). https://doi.org/10.1007/978-3-319-20028-6_34

    Chapter  MATH  Google Scholar 

  20. Tao, T.: Finite time blowup for an averaged three-dimensional Navier-Stokes equation. J. Am. Math. Soc. 29, 601–674 (2016)

    Article  MathSciNet  Google Scholar 

  21. Temam, R.: Navier-Stokes Equations: Theory and Numerical Analysis. North-Holland Publishing Company, New York (1977)

    MATH  Google Scholar 

  22. Weihrauch, K.: Computable Analysis: An Introduction. Springer, New York (2000). https://doi.org/10.1007/978-3-642-56999-9

    Book  MATH  Google Scholar 

  23. Weihrauch, K., Zhong, N.: Is wave propagation computable or can wave computers beat the Turing machine? Proc. Lond. Math. Soc. 85(2), 312–332 (2002)

    Article  MathSciNet  Google Scholar 

  24. Weihrauch, K., Zhong, N.: Computing the solution of the Korteweg-de Vries equation with arbitrary precision on Turing machines. Theor. Comput. Sci. 332, 337–366 (2005)

    Article  MathSciNet  Google Scholar 

  25. Weihrauch, K., Zhong, N.: Computing Schrödinger propagators on Type-2 Turing machines. J. Complex. 22(6), 918–935 (2006)

    Article  Google Scholar 

  26. Weihrauch, K., Zhong, N.: Computable analysis of the abstract Cauchy problem in Banach spaces and its applications I. Math. Logic Q. 53, 511–531 (2007)

    Article  MathSciNet  Google Scholar 

  27. Wiegner, M.: The Navier-Stokes equations – a never-ending challenge? Jahresbericht der Deutschen Mathematiker Vereinigung (DMV) 101(1), 1–25 (1999)

    MATH  Google Scholar 

  28. Zhong, N.: Computability structure of the Sobolev spaces and its applications. Theor. Comput. Sci. 219, 487–510 (1999)

    Article  MathSciNet  Google Scholar 

  29. Ziegler, M., Brattka, V.: Computability in linear algebra. Theor. Comput. Sci. 326, 187–211 (2004)

    Article  MathSciNet  Google Scholar 

  30. Ziegler, M.: Physically-relativized Church-Turing hypotheses: physical foundations of computing and complexity theory of computational physics. Appl. Math. Comput. 215(4), 1431–1447 (2009)

    MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ning Zhong .

Editor information

Editors and Affiliations

Appendices

A Proof of Proposition 1

(a) For a divergence-free and boundary-free polynomial, its coefficients must satisfy a system of linear equations. In the following, we derive explicitly this system of linear equations in the 2-dimensional case, i.e. \(\varOmega = (-1, 1)^2\). Let \(\varvec{p}=(p_1, p_2)=\big (\sum _{i, j=0}^{N}a^1_{i, j}x^iy^j, \sum _{i, j=0}^{N}a^2_{i, j}x^iy^j\big )\) be a divergence-free and boundary-free polynomial of real coefficients. (If the degree of \(p_1\) or \(p_2\) is less than N, then zeros are placed for the coefficients of missing terms). Then, by definition,

$$\begin{aligned} \nabla \cdot \varvec{p}= & {} \frac{\partial p_1}{\partial x}+\frac{\partial p_2}{\partial y} \\= & {} \sum _{1\le i\le N, 0\le \le N}ia^1_{i, j}x^{i-1}y^j + \sum _{0\le i\le N, 1\le \le N}ja^2_{i, j}x^{i}y^{j-1} \\= & {} \sum _{0\le i, j\le N-1}[(i+1)a^1_{i+1, j}+(j+1)a^2_{i, j+1}]x^{i}y^j \\&+ \sum _{0\le i\le N-1}(i+1)a^1_{i+1, N}x^iy^N + \sum _{0\le j\le N-1}(j+1)a^2_{N, j+1}x^Ny^j \\\equiv & {} 0 \quad \hbox {on} \, \varOmega \end{aligned}$$

which implies that all coefficients in \(\nabla \cdot \varvec{p}\) must be zero; or equivalently, Eq. (7) holds true. Turning to the boundary conditions, along the line \(x=1\), since

$$\begin{aligned} \varvec{p} (1, y) = \big (\sum \nolimits _{j=0}^N(\sum \nolimits _{i=0}^N a^1_{i, j})y^j, \sum \nolimits _{j=0}^N(\sum \nolimits _{i=0}^N a^2_{i, j})y^j \big ) \end{aligned}$$

is identically zero, it follows that \(\sum _{i=0}^N a^1_{i, j}=\sum _{i=0}^N a^2_{i, j}=0\) for \(0\le j\le N\). There are similar types of restrictions on the coefficients of \(\varvec{p}\) along the lines \(x=-1\), \(y=1\), and \(y=-1\). In summary, \(\varvec{p}\) vanishes on \(\partial \varOmega \) if and only if for all \(0\le j,i\le N\), both (8) and (9) hold true.

In the 3-dimensional case, a similar calculation shows that a polynomial triple \(\varvec{p}(x, y, z)=\big (p_1(x, y, z), p_2(x, y, z), p_3(x, y, z)\big )\) is divergence-free and boundary-free if and only if its coefficients satisfies a system of linear equations with integer coefficients.

(b) In [8] it is shown that for any real number \(s\ge 3\) and for any function \(\varvec{ w}\in \mathcal {N}_{div}^s\cap H_{2,0}^{1,\sigma }(\varOmega )^d\), the following holds:

$$\begin{aligned} \inf _{\varvec{p}\in \mathcal {N}^1_{\text {div}}\bigcap \mathcal {P}_N^0(\varOmega )^d}\Vert \varvec{ w}-\varvec{p}\Vert _{H_{2}^{s}(\varOmega )^d}\le CN^{-2}\Vert \varvec{ w}\Vert _{H_{2}^{s}(\varOmega )^d} \end{aligned}$$

where \(\varOmega =(-1, 1)^d\),

$$\begin{aligned} \mathcal {N}_{\text {div}}^s =\{ \varvec{ w}\in H_{2}^{s}(\varOmega )^d \, | \, \nabla \cdot \varvec{w}=0\},\quad \mathcal {P}_N^0(\varOmega )=\mathcal {P}_N(\varOmega )\bigcap H_{2,0}^{1,\sigma }(\varOmega ), \end{aligned}$$

\(\mathcal {P}_N\) is the set of all d-tuples of real polynomials with d variables and degree less than or equal to N with respect to each variable, \(H_{2,0}^{1,\sigma }(\varOmega )\) is the closure in \(H_{2}^{1}(\varOmega )\) of \(C^\infty _0(\varOmega )\), and C is a constant independent of N. This estimate implies that every function \(\varvec{ w}\in L^{\sigma }_{2,0}\) can be approximated with arbitrary precision by divergence-free and boundary-free real polynomials as follows: for any \(n\in \mathbb {N}\), since \(\{ \varvec{ u}\in C^\infty _0(\varOmega )^d \, : \, \nabla \cdot \varvec{ u}=0\}\) is dense in \(L^{\sigma }_{2,0}\), there is a divergence-free \(C^\infty \) function \(\varvec{ u}\) with compact support in \(\varOmega \) such that \(\Vert \varvec{ w}-\varvec{u}\Vert _{L_{2}}\le 2^{-(n+1)}\). Then it follows from the above inequality that there exists a positive integer N and a divergence-free and boundary-free polynomial \(\varvec{p}\) of degree N with real coefficients such that \(\Vert \varvec{ u} - \varvec{ p}\Vert _{L_{2}}\le \Vert \varvec{ u}-\varvec{p}\Vert _{H^3(\varOmega )^d}\le 2^{-(n+1)}\). Consequently, \(\Vert \varvec{ w}-\varvec{ p}\Vert _{L_{2}}\le \Vert \varvec{ w}-\varvec{ u}\Vert _{L_{2}}+\Vert \varvec{ u}-\varvec{ p}\Vert _{L_{2}}\le 2^{-n}\).

It remains to show that \(\mathbb {Q}^{\sigma }_{0}[\mathbb {R}^2]\), the divergence-free and boundary-free polynomial tuples with rational coefficients, is dense (in \(L_{2}\)-norm) in the set of all polynomial tuples with real coefficients which are divergence-free on \(\varOmega \) and boundary-free on \(\partial \varOmega \). To this end we note that, according to part (a), the divergence-free and boundary-free polynomials can be characterized, independent of their coefficient field, in terms of a homogeneous system of linear equations with integer coefficients. Then it follows from the lemma below that the set of the rational solutions of this system is dense in the set of its real solutions. And since \(\varOmega \) is bounded (=relatively compact), the approximations to its coefficients of a polynomial yields (actually uniform) the approximations to the polynomial itself:

$$\begin{aligned} \sup _{\varvec{ x}\in \varOmega } |p_k(\varvec{ x})| \;\le \; \sum _{i, j=0}^{N} |a^k_{i,j}|\cdot M^{i+j} \quad \text { for } \varOmega \subseteq [-M,+M]^2 \quad \text {and } k=1,2 \end{aligned}$$

Lemma 7

Let \(A\in \mathbb {Q}^{m\times n}\) be a rational matrix. Then the set \(\mathrm {kernel}_{IQ}(A):=\{\varvec{ x}\in \mathbb {Q}^n:A\cdot \varvec{ x}=\varvec{ 0}\}\) of rational solutions to the homogeneous system of linear equations given by A is dense in the set \(\mathrm {kernel}_{\mathbb {R}}(A)\) of real solutions.

Proof

For \(d:=\mathrm {dim}\big (\mathrm {kernel}_{\mathbb {R}}(A)\big )\), Gaussian Elimination yields a basis \(B=(\varvec{ b}^1,\ldots ,\varvec{ b}^d)\) of \(\mathrm {kernel}(A)\); in fact it holds \(B\in \mathbb {Q}^{n\times d}\) and

$$\begin{aligned} \mathrm {kernel}_{\mathbb {F}}(A) \;=\; \mathrm {image}_{\mathbb {F}}(B) \;:=\; \big \{\lambda _1\varvec{ b}^1+\cdots +\lambda _d\varvec{ b}^d:\lambda _1,\ldots ,\lambda _d\in \mathbb {F}\big \} \end{aligned}$$

for every field \(\mathbb {F}\supseteq \mathbb {Q}\): Observe that the elementary row operations Gaussian Elimination employs to transform A into echelon form containing said basis B consist only of arithmetic (=field) operations! (We deliberately do not require B to be orthonormal; cf. [29, §3]). Now \(\mathrm {image}_{\mathbb {Q}}(B)\) is obviously dense in \(\mathrm {image}_{\mathbb {R}}(B)\).

B Proof of Lemma 1

Note that \(\gamma _n*\mathbb {T}_k\varvec{p} = (\gamma _n*\mathbb {T}_kp_1, \gamma _n*\mathbb {T}_kp_2)\). For each \(\varvec{p}\in \mathbb {Q}^{\sigma }_{0}[\mathbb {R}^2]\) and \(n\ge k\), since

$$\begin{aligned} \frac{\partial (\gamma _n*\mathbb {T}_kp_1)}{\partial x}&(x, y) = \frac{\partial }{\partial x}\int ^{1}_{-1}\int ^{1}_{-1}\gamma _n(x-s, y-t)\cdot \mathbb {T}_kp_1(s, t)\,ds\,dt \nonumber \\&= \int ^{1-2^{-k}}_{-1+2^{-k}}\left[ \int ^{1-2^{-k}}_{-1+2^{-k}}\frac{\partial \gamma _n}{\partial x}(x-s, y-t)\cdot \mathbb {T}_kp_1(s, t)\,ds\right] \,dt \nonumber \\&= \int ^{1-2^{-k}}_{-1+2^{-k}}\left[ \int ^{1-2^{-k}}_{-1+2^{-k}}-\frac{\partial \gamma _n}{\partial s}(x-s, y-t)\cdot \mathbb {T}_kp_1(s, t)\,ds\right] \,dt \end{aligned}$$
(37)

for \(\mathbb {T}_kp_1=0\) in the exterior region of \(\varOmega _k\) including its boundary \(\partial \varOmega _k\). Note that \(\frac{\partial \gamma _n}{\partial s}\) is continuous on \(\mathbb {R}^2\); \(\frac{\partial \gamma _n}{\partial s}(x-s, y-t)\cdot \mathbb {T}_kp_1(s, t)\) is continuous on \([-1, 1]^2\) for any given \(x, y\in \mathbb {R}\); \(\frac{\partial \mathbb {T}_kp_1}{\partial s}(s, t)\) is continuous in \((-1+2^{-n}, 1-2^{-n})\) and \(\mathbb {T}_kp_1\) is continuous on \([-1+2^{-n}, 1-2^{-n}]\) for any given \(t\in [-1; 1]\). Thus, we can apply the integration by parts formula to the integral

$$\int ^{1-2^{-k}}_{-1+2^{-k}}-\frac{\partial \gamma _n}{\partial s}(x-s, y-t)\cdot \mathbb {T}_kp_1(s, t)\,ds$$

as follows:

$$\begin{aligned}&\quad \int ^{1-2^{-k}}_{-1+2^{-k}} -\frac{\partial \gamma _n}{\partial s}(x-s, y-t)\cdot \mathbb {T}_kp_1(s, t)\,ds \nonumber \\&\quad = -\gamma _n(x-s, y-t)\cdot \mathbb {T}_kp_1(s, t)\big | ^{1-2^{-k}}_{-1+2^{-k}} \nonumber \\&\qquad \qquad \qquad +\int ^{1-2^{-k}}_{-1+2^{-k}}\gamma _n(x-s, y-t) \cdot \frac{\partial \mathbb {T}_kp_1}{\partial s}(s, t)\,ds \nonumber \\&\quad = \int ^{1-2^{-k}}_{-1+2^{-k}}\gamma _n(x-s, y-t)\cdot \frac{\partial \mathbb {T}_kp_1}{\partial s}(s, t)\,ds \end{aligned}$$
(38)

Then it follows from (37) and (38) that for any \((x, y)\in \varOmega \),

$$\begin{aligned} \frac{\partial \gamma _n*\mathbb {T}_kp_1}{\partial x}(x, y)= \int ^{1-2^{-k}}_{-1+2^{-k}}\int ^{1-2^{-k}}_{-1+2^{-k}}\gamma _n(x-s, y-t)\cdot \frac{\partial \mathbb {T}_kp_1}{\partial s}(s, t)\,ds\,dt . \end{aligned}$$

A similar calculation yields that for any \((x, y)\in \varOmega \),

$$\begin{aligned} \frac{\partial \gamma _n*\mathbb {T}_kp_2}{\partial y}(x, y)= \int ^{1-2^{-k}}_{-1+2^{-k}}\int ^{1-2^{-k}}_{-1+2^{-k}}\gamma _n(x-s, y-t)\cdot \frac{\partial \mathbb {T}_kp_2}{\partial t}(s, t)\,ds\,dt . \end{aligned}$$

Thus, for any \((x, y)\in \varOmega \) and \(n\ge k\),

$$\begin{aligned} \nabla \cdot&(\gamma _n*\mathbb {T}_k\varvec{p})(x, y) = \frac{\partial \gamma _n*\mathbb {T}_kp_1}{\partial x}(x, y)+ \frac{\partial \gamma _n*\mathbb {T}_kp_2}{\partial y}(x, y) \\&= \int ^{1-2^{-k}}_{-1+2^{-k}}\int ^{1-2^{-k}}_{-1+2^{-k}}\gamma _n(x-s, y-t)\cdot \left[ \frac{\partial \mathbb {T}_kp_1}{\partial s}+\frac{\partial \mathbb {T}_kp_2}{\partial t}\right] (s, t) \,ds\,dt \\&=0 \end{aligned}$$

for \(\mathbb {T}_k\varvec{p}=(\mathbb {T}_kp_1, \mathbb {T}_kp_2)\) is divergence-free on \(\varOmega _k\). This proves that for any \(\varvec{p}\in \mathbb {Q}^{\sigma }_{0}[\mathbb {R}^2]\) and \(n\ge k\), \(\gamma _n*\mathbb {T}_k\varvec{p}\) is divergence-free on \(\varOmega \).

C Proof of Lemma 2

Since for each \(\varvec{p}\in \mathbb {Q}^{\sigma }_{0}[\mathbb {R}^2]\) and \(k\in \mathbb {N}\), \(\gamma _n*\mathbb {T}_k\varvec{p}\rightarrow \mathbb {T}_k\varvec{p}\) effectively and uniformly on \(\varOmega _k\) as \(n\rightarrow \infty \), it suffices to show that \(\{\mathbb {T}_k\varvec{p}:k\in \mathbb {N}, \varvec{p}\in \mathbb {Q}^{\sigma }_{0}[\mathbb {R}^2]\}\) is dense in \(L^{\sigma }_{2,0}(\varOmega )\). On the other hand, since \(\mathbb {Q}^{\sigma }_{0}[\mathbb {R}^2]\) is dense in \(L^{\sigma }_{2,0}(\varOmega )\), we only need to show that for each \(\varvec{p}\in \mathbb {Q}^{\sigma }_{0}[\mathbb {R}^2]\) and \(m\in \mathbb {N}\), there is a \(k\in \mathbb {N}\) such that \(2^{-m}\ge \Vert \varvec{p}-\mathbb {T}_k\varvec{p}\Vert _\infty = \max \{|p_1(\varvec{ x})-\mathbb {T}_kp_1(\varvec{ x})|,|p_2(\varvec{ x})-\mathbb {T}_kp_2(\varvec{ x})|:\varvec{ x}\in \bar{\varOmega }\}\).

Since \(p_i\) is uniformly continuous on \(\bar{\varOmega }\), there exists a \(k\in \mathbb {N}\) such that \(|p_i(x,y)-p_i(x',y')|\le 2^{-m}\) whenever \(|x-x'|,|y-y'|\le 2^{-k+1}\), and, in particular, for \(x'=\frac{x}{1-2^{-k}}\) and \(y'=\frac{y}{1-2^{-k}}\). Also, since \(p_i(x,y)=0\) for \((x,y)\in \partial \varOmega \), \(|p_i(x,y)|\le 2^{-m-1}\) for all \((x,y)\in \varOmega \setminus \varOmega _k\). This establishes \(|p_i(x,y)-\mathbb {T}_kp_i(x,y)|\le 2^{-m}\) on \(\bar{\varOmega }\).

Rights and permissions

Reprints and permissions

Copyright information

© 2020 Springer Nature Switzerland AG

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Sun, SM., Zhong, N., Ziegler, M. (2020). Computability of the Solutions to Navier-Stokes Equations via Effective Approximation. In: Du, DZ., Wang, J. (eds) Complexity and Approximation. Lecture Notes in Computer Science(), vol 12000. Springer, Cham. https://doi.org/10.1007/978-3-030-41672-0_7

Download citation

  • DOI: https://doi.org/10.1007/978-3-030-41672-0_7

  • Published:

  • Publisher Name: Springer, Cham

  • Print ISBN: 978-3-030-41671-3

  • Online ISBN: 978-3-030-41672-0

  • eBook Packages: Computer ScienceComputer Science (R0)

Publish with us

Policies and ethics