Skip to main content
Log in

Uniqueness of Plane Stationary Navier–Stokes Flow Past an Obstacle

  • Published:
Archive for Rational Mechanics and Analysis Aims and scope Submit manuscript

Abstract

We study the exterior problem for stationary Navier–Stokes equations in two dimensions describing a viscous incompressible fluid flowing past an obstacle. It is shown that, at small Reynolds numbers, the classical solutions constructed by Finn and Smith are unique in the class of D-solutions (that is, solutions with finite Dirichlet integral). No additional symmetry or decay assumptions are required. This result answers a long-standing open problem. In the proofs, we developed the ideas of the classical Ch. Amick paper (Acta Math. 1988).

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.

Similar content being viewed by others

Notes

  1. \(\partial {\mathcal {E}}\) being of class \(C^{2+\alpha }\), \(\alpha > 0\) would be sufficient. Such regularity is required in [7] for the construction of Finn-Smith solutions.

  2. Stokes himself gave the following explanation: the pressure of the cylinder on the fluid continuously tends to increase the quantity of fluid which it carries with it, while the friction of the fluid at a distance from the cylinder continually tends to diminish it. In the case of a sphere, these two causes eventually counteract each other, and the motion becomes uniform. But in the case of a cylinder, the increase in the quantity of the fluid carried continually gains on the decrease due to the friction of the surrounding fluid, and the quantity carried increases indefinitely as the cylinder moves on ([31], p. 65).

  3. This convergence is uniform on every bounded set.

  4. For example, in the recent paper [15] it was written: “The question of the uniqueness of weak solutions for small data is even more open in two-dimensional exterior domains... For two-dimensional exterior domains with nonempty boundary, we would a priori also expect the existence of infinitely many weak solutions parameterized by some parameter”.

  5. Physically, finiteness of the Dirichlet integral means that the total energy dissipation rate in the fluid is finite.

  6. This result is not trivial since in general the finiteness of the Dirichlet integral does not guarantee even the boundedness of the mapping, for example, the function \(f(z)=\bigl (\ln (|z|)\bigr )^{\frac{1}{3}}\) has a finite Dirichlet integral in \({{\mathcal {E}}}\).

  7. Throughout the rest of the paper, we always assume that such subtraction of a constant has been carried out.

  8. Similar facts concerning the integrability of \(E_{ij}\) are collected in [27, § 2–3].

References

  1. Amick, C.J.: On Leray’s problem of steady Navier-Stokes flow past a body in the plane. Acta Math. 161, 71–130, 1988

    Article  MathSciNet  Google Scholar 

  2. Bourgain, J., Korobkov, M., Kristensen, J.: On the Morse-Sard property and level sets of Sobolev and BV functions Rev. Mat. Iberoam. 29, 1–23, 2013

    Article  MathSciNet  Google Scholar 

  3. Coifman, R.R., Lions, J.L., Meier, Y., Semmes, S.: Compensated compactness and Hardy spaces. J. Math. Pures App. IX Sér. 72, 247–286, 1993

    MathSciNet  MATH  Google Scholar 

  4. Ferone, A., Russo, R., Tartaglione, A.: The Stokes Paradox in Inhomogeneous Elastostatics. J. Elasticity 142(1), 35–52, 2020

    Article  MathSciNet  Google Scholar 

  5. Finn, R.: On the exterior stationary problem for the Navier-Stokes equations, and associated perturbation problems. Arch. Rational. Mech. Anal. 19, 363–406, 1965

    Article  ADS  MathSciNet  Google Scholar 

  6. Finn, R., Smith, D.R.: On the linearized hydrodynamical equations in two dimensions. Arch. Ration. Mech. Anal. 25, 1–25, 1967

    Article  MathSciNet  Google Scholar 

  7. Finn, R., Smith, D.R.: On the stationary solutions of the Navier-Stokes equations in two dimensions. Arch. Ration. Mech. Anal. 25, 26–39, 1967

    Article  MathSciNet  Google Scholar 

  8. Fujita, H.: On the existence and regularity of the steady-state solutions of the Navier-Stokes squation. J. Fac. Sci. Univ. Tokyo 9(1A), 59–102, 1961

    MATH  Google Scholar 

  9. Galdi, G.P., Sohr, H.: On the asymptotic structure of plane steady flow of a viscous fluid in exterior domains. Arch. Ration. Mech. Anal. 131, 101–119, 1995

    Article  MathSciNet  Google Scholar 

  10. Galdi, G.P.: An Introduction to the Mathematical Theory of the Navier-Stokes Equations. Steady-State Problems. Springer, Berlin 2011

    MATH  Google Scholar 

  11. Gilbarg, D., Weinberger, H.F.: Asymptotic properties of Leray’s solution of the stationary two-dimensional Navier-Stokes equations. Russian Math. Surv. 29, 109–123, 1974

    Article  ADS  MathSciNet  Google Scholar 

  12. Gilbarg, D., Weinberger, H.F.: Asymptotic properties of steady plane solutions of the Navier–Stokes equation with bounded Dirichlet integral. Ann. Scuola Norm. Sup. Pisa Cl. Sci. 5(4), 381–404, 1978

  13. Guillod, J.: Steady solutions of the Navier–Stokes equations in the plane. arXiv:1511.03938

  14. Guillod, J., Wittwer, P.: Optimal asymptotic behavior of the vorticity of a viscous flow past a two-dimensional body. J. Math. Pures Appl. 118, 481–499, 2017

    Article  MathSciNet  Google Scholar 

  15. Guillod, J., Wittwer, P.: Existence and uniqueness of steady weak solutions to the Navier-Stokes equations in \( {\mathbb{R}}^{2}\). Proc. Am. Math. Soc. 146, 4429–4445, 2018

    Article  Google Scholar 

  16. Hillaret, M., Wittwer, P.: On the existence of solutions to the planar exterior Navier Stokes system. J. Differ. Equ. 255, 2996–3019, 2013

    Article  ADS  MathSciNet  Google Scholar 

  17. Hopf, E.: Ein Allgemeiner Endlichkeitsatz der Hydrodynamik. Math. Ann. 117, 764–775, 1941

    MathSciNet  MATH  Google Scholar 

  18. Korobkov, M.V., Pileckas, K., Russo, R.: On convergence of arbitrary \(D\)-solution of steady Navier-Stokes system in \(2D\) exterior domains. Arch. Ration. Mech. Anal. 233(1), 385–407, 2019

    Article  MathSciNet  Google Scholar 

  19. Korobkov, M.V., Pileckas, K., Russo, R.: On the steady Navier-Stokes equations in \(2D\) exterior domains. J. Differ. Equ. 269(3), 1796–1828, 2020

    Article  ADS  MathSciNet  Google Scholar 

  20. Korobkov, M.V., Pileckas, K., Russo, R.: Leray’s plane steady state solutions are nontrivial. Adv. Math. 269(3), 1796–1828, 2020

    MATH  Google Scholar 

  21. Ladyzhenskaya, O.A.: The Mathematical Theory of Viscous Incompressible Flow. Gordon & Breach, New York 1969

    MATH  Google Scholar 

  22. Lieb, E.H., Loss, M.: Analysis, Graduate Studies in mathematics, vol. 14, 2nd edn. AMS, Providence 2001

  23. Leray, J.: Étude de diverses équations intégrales non linéaires et de quelques problèmes que pose l’hydrodynamique. J. Math. Pures Appl. 12, 1–82, 1933

    MATH  Google Scholar 

  24. Nakatsuka, T.: On uniqueness of symmetric Navier-Stokes flows around a body in the plane. Adv. Differ. Equ. 20(3/4), 193–212, 2013

    MathSciNet  MATH  Google Scholar 

  25. Oseen, C.W.: Neuere Methoden und Ergebnisse in der Hydrodynamik. Akademische Verlagsgesellschaft m.b.H, Leipzig 1927

    MATH  Google Scholar 

  26. Russo, R.: On Stokes’ problem. In: Rannacher, R., Sequeira, A. (eds.) Advances in Mathematical Fluid Mechanics, pp. 473–511. Springer, Berlin 2010

    Chapter  Google Scholar 

  27. Sazonov, L.I.: On the asymptotic behavior of the solution to the two-dimensional stationary problem of the flow past a body far from it. Math. Notes 65(2), 202–207, 1999

    Article  MathSciNet  Google Scholar 

  28. Sazonov, L.I.: The justification of an asymptotic expansion for the solution of the two-dimensional flow problem at small Reynolds numbers. Izvestiya Math. 67(5), 975–1006, 1999

    Article  ADS  Google Scholar 

  29. Smith, D.R.: Estimates at infinity for stationary solutions of the Navier-Stokes equations in two dimensions. Arch. Rational Mech. Anal. 20, 341–372, 1965

    Article  ADS  MathSciNet  Google Scholar 

  30. Stein, E.: Harmonic Analysis: Real-Variables Methods. Orthogonality and Oscillatory Integrals. Princeton University Press, Princeton 1993

    Google Scholar 

  31. Stokes, G.G.: On the effect of the internal friction of fluids on the motion of pendulums. Trans. Cambridge Phil. Soc. 9, 8–106, 1851

    ADS  Google Scholar 

  32. Yudovich, V.I.: Eleven great problems of mathematical hydrodynamics. Mosc. Math. J. 3, 711–737, 2003

    Article  MathSciNet  Google Scholar 

  33. Yamazaki, M.: Unique existence of stationary solutions to the two-dimensional Navier-Stokes equations on exterior domains. In: Adachi, T., et al. (eds.) Mathematical Aspects on the Navier-Stokes Equations and Related Topics, Past and Future - In memory of Professor Tetsuro Miyakawa, Gakuto International Series, Mathematical Sciences and Applications, vol. 35, pp. 220–241. Gakkotosho, Tokyo 2011

    Google Scholar 

  34. Yamazaki, M.: Existence and uniqueness of weak solutions to the two-dimensional stationary Navier-Stokes exterior problem. J. Math. Fluid Mech. 20(4), 2019–2051, 2018

    Article  ADS  MathSciNet  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Mikhail Korobkov.

Ethics declarations

Conflicts of interest

The authors declare that they have no conflict of interest.

Additional information

Communicated by P. Constantin.

Publisher's Note

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

Appendices

Appendix I

Proof of Step 2

In this section we discuss the proof of the uniform pointwise estimate (5.8), that is,

$$\begin{aligned} |\mathbf{u }-(1,0)|=|{\mathbf{v }}|{\,\leqq \,}C{\varepsilon }\end{aligned}$$

for \(r{\,\geqq \,}1.\)

Step 2a. First of all, consider the “good” cone

$$\begin{aligned} {\mathcal {K}}^{\pm y}=\biggl \{r{\,\geqq \,}1,|\theta |\in \bigl (\frac{1}{5}\pi ,\frac{4}{5}\pi \bigr )\biggr \}, \end{aligned}$$

which is separated from the x-axis. Here one can simply follow the proof of [1, Theorem 19]. The key observation is that \(|\psi |{\,\geqq \,}cr>0\) in such a cone, and this fact, by virtue of smallness, gives

$$\begin{aligned} |\Phi -\frac{1}{2}|{\,\leqq \,}C{\varepsilon },\qquad |\gamma -\frac{1}{2}|=|\Phi -\frac{1}{2}-\omega \psi |{\,\leqq \,}C{\varepsilon }\end{aligned}$$

on good circles and by maximum principle for \(\omega \), implies \(|\omega | {\,\leqq \,}C{\varepsilon }r^{-1}\) in \({\mathcal {K}}^{\pm y}\), which easily gives the required smallness \(|\mathbf{u }-(1,0)|{\,\leqq \,}C{\varepsilon }\) here.

Now consider the more complicated region

$$\begin{aligned} \bigl \{r{\,\geqq \,}1,|\theta |{\,\leqq \,}\frac{1}{5}\pi \bigr \}. \end{aligned}$$

Here the previous estimate \(|\omega | {\,\leqq \,}C{\varepsilon }r^{-1}\) does not hold in general, so the arguments should be more delicate and subtle. The main ideas here are due to Amick [1] and Korobkov et al. [18, Remark 4.1].

Step 2b. We are going to use the uniform smallness of \(\gamma \)-function

$$\begin{aligned} \bigl |\gamma -\frac{1}{2}\bigr |=\bigl |\Phi -\frac{1}{2}-\omega \psi \bigr |{\,\leqq \,}C{\varepsilon }\end{aligned}$$

and the level sets of the stream function \(\psi \). On the first step here we show that the set \({\mathcal {C}} = \{r{\,\geqq \,}\frac{1}{2}, \psi = 0\}\) consists of exactly two smooth curves \({\mathcal {C}}_{\pm }\). Indeed, from Step 1, we know that

$$\begin{aligned} |\psi - y| {\,\leqq \,}C{\varepsilon }\end{aligned}$$
(A.1)

in \(\Omega _{\frac{1}{2},2}\). Using (5.1), there exists an angle \(\theta _1 \in [\frac{\pi }{10}, \frac{\pi }{5})\) such that

$$\begin{aligned} \int \limits _{1/4}^{+\infty } |\partial _r \mathbf{u }(r, \theta )|^2 rdr {\,\leqq \,}C{\varepsilon }^2 \end{aligned}$$

for \(\theta = \theta _1, -\theta _1, \pi - \theta _1, -\pi + \theta _1\). Hence, for such \(\theta \),

$$\begin{aligned} \int _{r_n}^{r_{n+1}} |\partial _r \mathbf{u }(r, \theta )| dr&{\,\leqq \,}C {\varepsilon }\left( \int _{r_n}^{r_{n+1}} \frac{dr}{r}\right) ^\frac{1}{2}, \\&{\,\leqq \,}C{\varepsilon }\end{aligned}$$

with \(r_n, n=-2,-1,\cdots \) given in Step 1. In view of (5.4), we obtain

$$\begin{aligned} |\mathbf{v }(r, \theta )| {\,\leqq \,}C{\varepsilon }\end{aligned}$$

for any \(r{\,\geqq \,}\frac{1}{2}\) and \(\theta = \theta _1, -\theta _1, \pi - \theta _1, -\pi + \theta _1\). By the definition of \(\psi \) and (A.1), we have

$$\begin{aligned} |\psi - y| {\,\leqq \,}C{\varepsilon }r \end{aligned}$$
(A.2)

pointwise on the set

$$\begin{aligned} {\mathcal {N}} := \bigl (\cup _{n {\,\geqq \,}-2} S_{r_n}\bigr )\cup \bigl \{r {\,\geqq \,}\frac{1}{2}, \theta = \theta _1, -\theta _1, \pi - \theta _1, -\pi + \theta _1\bigr \}. \end{aligned}$$
(A.3)

When \(\lambda \) is small, this implies that \({\mathcal {C}}\) must intersect \({\mathcal {N}}\) within the cone \({\mathcal {K}} := \{r{\,\geqq \,}\frac{1}{2}, |\theta |< \frac{\pi }{10} \ \text {or}\ |\pi - \theta | < \frac{\pi }{10}\}\). Moreover, by (5.4), \({\mathcal {C}}\) intersects each \(S_{r_n}, n{\,\geqq \,}-2\) at exactly two points, one with \(x>0\) and another with \(x<0\). On the level set \({\mathcal {C}}\) we have \(\Big |\frac{|\mathbf{u }|^2}{2} - \frac{1}{2}\Big | = |\gamma - \frac{1}{2} - q| {\,\leqq \,}C{\varepsilon }\) (see (5.3), (5.6) ), hence

$$\begin{aligned} \Big ||\mathbf{u }| - 1\Big | {\,\leqq \,}C{\varepsilon }\quad \text{ on } {\mathcal {C}}. \end{aligned}$$
(A.4)

As a consequence, \(|\nabla \psi |=|{\mathbf{u }}|\ne 0\) on the set \({\mathcal {C}}\), that is, \({\mathcal {C}}\) is a regular curve in the case when \(\lambda \) is small.

Further, when \(\lambda \) is small, \({\mathcal {C}}\) cannot contain any closed curve \({\mathcal {L}}\) by an elegent idea of Amick’s. To explain this, suppose \({\mathcal {C}}\) contains a closed curve \({\mathcal {L}}\). Let \({\mathcal {U}}\) denote the domain bounded by \({\mathcal {L}}\), then

$$\begin{aligned} \left( \int _{\mathcal {U}} \omega ^2\, \text {d}x\text {d}y\right) ^\frac{1}{2} |{\mathcal {U}}|^\frac{1}{2} {\,\geqq \,}\left| \int _{\mathcal {U}} \omega \, \text {d}x\text {d}y\right| = \left| \int _{\mathcal {U}} \Delta \psi \, \text {d}x\text {d}y\right| = \left| \int _{\mathcal {L}} \partial _n \psi \,ds\right| = \\ =\int _{\mathcal {L}} |\nabla \psi |\,ds=\int _{\mathcal {L}} |{\mathbf{u }}|\,ds\overset{(A.4)}{{\,\geqq \,}}\frac{1}{2}|{\mathcal {L}}|. \end{aligned}$$

(Note, that on \({\mathcal {L}}\) one of the identities \(\partial _n \psi \equiv |\nabla \psi |\) or \(\partial _n \psi \equiv -|\nabla \psi |\) holds, because \({\mathcal {L}}\) is regular closed level set of \(\psi \).) This implies

$$\begin{aligned} |{\mathcal {L}}| {\,\leqq \,}C {\varepsilon }|{\mathcal {U}}|^\frac{1}{2}, \end{aligned}$$

which contradicts with the isoperimetric inequality when \({\varepsilon }\) is small. Hence \({\mathcal {C}}\) must consist of two smooth curves starting from \(r=\frac{1}{2}\) and extending to infinity, each contained in \({\mathcal {K}}\cap \{x>0\}\) and \({\mathcal {K}}\cap \{x<0\}\) respectively. We denote these two curves by \({\mathcal {C}}_{\pm }\).

Step 2c. Here we show that

$$\begin{aligned} |\mathbf{v }| {\,\leqq \,}C{\varepsilon }\quad \text{ along } {\mathcal {C}}_\pm . \end{aligned}$$
(A.5)

Take any point \(z \in {\mathcal {C}}_+\) (\({\mathcal {C}}_-\) would be similar) with \(r=|z| {\,\geqq \,}1\). By Lemma 6, there exist good circles \(S_{{\tilde{r}}_n}\) centered at z with \({\tilde{r}}_n \in [2^{-n-1}r, 2^{-n}r), n=1,2, \cdots \), such that

$$\begin{aligned} |\mathbf{u } - \bar{\mathbf{u }}^{(n)}| {\,\leqq \,}C{\varepsilon }\quad \text{ on } S_{{\tilde{r}}_n}. \end{aligned}$$

Here \(\bar{\mathbf{u }}^{(n)}\) is the average of \(\mathbf{u }\) on \(S_{{\tilde{r}}_n}\). Since \(S_{{\tilde{r}}_n}\) must intersect \({\mathcal {C}}_+\) and on \({\mathcal {C}}_+\) the inequality (A.4) holds, we obtain

$$\begin{aligned} \left| |\bar{\mathbf{u }}^{(n)}| - 1\right| {\,\leqq \,}C{\varepsilon }\end{aligned}$$

on each \(S_{{\tilde{r}}_n}\). From this, (2.1) of Lemma 6 implies

$$\begin{aligned} \left| |\bar{\mathbf{u }}^{(\rho )}| - 1\right| {\,\leqq \,}C{\varepsilon }, \end{aligned}$$

where \(\bar{\mathbf{u }}^{(\rho )}\) is the mean value of \(\mathbf{u }\) on circles \(S_\rho (z)\) centered at z with radius \(\rho {\,\leqq \,}\frac{r}{2}\). Let \(\varphi ^{(n)}\) be the angle of \(\bar{\mathbf{u }}^{(n)}\). By Lemma 7, we have, for any \(n,m{\,\geqq \,}1\),

$$\begin{aligned} |\varphi ^{(n)} - \varphi ^{(m)}|&{\,\leqq \,}C \int _{|\zeta -z| {\,\leqq \,}\frac{r}{2}} \frac{|\nabla \omega |}{|\zeta -z|} + |\nabla \mathbf{u }|^2 \, \text {d}\zeta _1 \text {d}\zeta _2 \nonumber \\&{\,\leqq \,}C \left( \int _{1{\,\leqq \,}|\zeta -z| {\,\leqq \,}\frac{r}{2}} \frac{|\nabla \omega |}{|\zeta -z|} \text {d}\zeta _1 \text {d}\zeta _2\right) \end{aligned}$$
(A.6)
$$\begin{aligned}&\quad +\, C \left( \int _{|\zeta -z| {\,\leqq \,}1} \frac{|\nabla \omega |}{|\zeta -z|} \text {d}\zeta _1 \text {d}\zeta _2\right) + C{\varepsilon }^2 \nonumber \\&{\,\leqq \,}C {\varepsilon }. \end{aligned}$$
(A.7)

In this last line we have used (5.2) and (5.7) for the first and second terms in the penultimate inequality respectively. By construction, the circle \(S_{{\tilde{r}}_1}\) must intersect the set \({\mathcal {N}}\) (see (A.3) ) on which \(|\mathbf{v }| {\,\leqq \,}C{\varepsilon }\). Hence, \(|\bar{\mathbf{u }}^{(1)} - \mathbf{e }_1| {\,\leqq \,}C{\varepsilon }\). Letting \(n=1, m\rightarrow \infty \) in (A.6), and using (A.4), we get \(|\mathbf{v }(z)| {\,\leqq \,}C{\varepsilon }\). Note that this implies that the slope of \({\mathcal {C}}_\pm \) is small.

Step 2d. Now take arbitrary point \(z_1 = (x_1,y_1) \in \{r{\,\geqq \,}1, |\theta |<\frac{\pi }{5}, x>0\}\) (the case \(x<0, |\pi - \theta |< \frac{\pi }{5}\) is similar) and show that

$$\begin{aligned} |\mathbf{v }(z_1)| {\,\leqq \,}C{\varepsilon }. \end{aligned}$$

Let \(z_2 = (x_1,y_2) \in {\mathcal {C}}_+\). To simplify notations, let’s change the coordinate system, namely, let’s move the coordinate origin to the point \(z_2\), so now

$$\begin{aligned} z_2=(x_1,y_2)=(0,0)\in {\mathcal {C}}_+, \quad z_1=(x_1,y_1)=(0,y_1). \end{aligned}$$

Consider the case \(y_1 >0\), that is, when the point \(z_1\) is above the \({\mathcal {C}}_+\) curve, so that \(\psi (z_1) > 0\) (the opposite case \(y_1<0\) case is quite similar). Let

$$\begin{aligned} R = y_1. \end{aligned}$$

Using Lemma 6, we find two good circles \(S_{R_1}(z_1)\) and \(S_{R_2}(z_2)\) centered at \(z_1\) and \(z_2\) respectively, with radii

$$\begin{aligned} R_1, R_2 \in \bigl (\frac{2R}{3},\frac{3R}{4}\bigr ). \end{aligned}$$

Clearly,

$$\begin{aligned} S_{R_1}(z_1)\cap S_{R_2}(z_2)\ne \emptyset \ne {\mathcal {C}}_+\cap S_{R_2}(z_2). \end{aligned}$$

As a result, on both \(S_{R_1}(z_1)\) and \(S_{R_2}(z_2)\) we have

$$\begin{aligned} |\mathbf{v }| {\,\leqq \,}C{\varepsilon }. \end{aligned}$$
(A.8)

Note that \(\psi = 0\) on \({\mathcal {C}}_+\). We can integrate \(\nabla (\psi -y) = \mathbf{v }^\perp \) starting from \(z_2\in {\mathcal {C}}_+\) along arcs of \({\mathcal {C}}_+\), \(S_{R_2}(z_2)\) and \(S_{R_1}(z_1)\). This process gives \(|\psi - y| {\,\leqq \,}C{\varepsilon }R\) on \(S_{R_1}(z_1)\). As a consequence, by virtue of the evident inequalities

$$\begin{aligned} \frac{1}{4}R{\,\leqq \,}\min \limits _{(x,y)\in S_{R_1}(z_1)}y<\max \limits _{(x,y)\in S_{R_1}(z_1)}y<\frac{7}{4} R, \end{aligned}$$

we have

$$\begin{aligned} \left| \frac{\psi }{y} - 1\right| {\,\leqq \,}C{\varepsilon }\end{aligned}$$
(A.9)

on \(S_{R_1}(z_1)=\partial B_{R_1}(z_1)\).

Now we are going to prove that the last estimate is valid not only for boundary circle \(S_{R_1}(z_1)\), but for all points of the disk \(B_{R_1}(z_1)\). Recall that

$$\begin{aligned} |\gamma - q -\frac{1}{2}| = \left| \frac{|\nabla \psi |^2}{2} - \psi \Delta \psi - \frac{1}{2} \right| {\,\leqq \,}C{\varepsilon }\end{aligned}$$
(A.10)

pointwisely holds in the region \(\{r{\,\geqq \,}1\}\) (see (5.3), (5.6) ). Denote \(\psi = y(1+S)\) in \(B_{R_1}(z_1)\). Since \({\mathcal {C}}_+\cap B_{R_1}(z_1)=\emptyset \) and \(\psi \) is positive on \( B_{R_1}(z_1)\), by construction we have that the values \((1+S)\) are positive on \(B_{R_1}(z_1)\) as well. Assume at the moment, that S has positive maximum at the interior point of the disk \(B_{R_1}(z_1)\). Then at this maximum point \(\nabla S=0\) and \(\Delta S{\,\leqq \,}0\), therefore,

$$\begin{aligned} |\nabla \psi |^2 - 2\psi \Delta \psi = (1+S)^2-2\psi y\Delta S{\,\geqq \,}(1+S)^2, \end{aligned}$$

which, by virtue of (A.10), implies \((1+S)^2 {\,\leqq \,}1 + C{\varepsilon },\) and consequently,

$$\begin{aligned} S {\,\leqq \,}C{\varepsilon }\end{aligned}$$

at any maximum point of S inside the disk \(B_{R_1}(z_1)\). Similarly, a consideration for the negative minimal points of S gives \(S {\,\geqq \,}-C{\varepsilon }\). Hence, taking into account (A.9), we have proved

$$\begin{aligned} \left| \frac{\psi }{y} - 1\right| {\,\leqq \,}C{\varepsilon }\end{aligned}$$
(A.11)

in \(B_{R_1}(z_1)\). In particular,

$$\begin{aligned} |\psi -y|{\,\leqq \,}C{\varepsilon }R\qquad \text{ in } \ B_{R_1}(z_1). \end{aligned}$$

Next, one observes that

$$\begin{aligned} \Delta (\sqrt{\psi } - \sqrt{y}) = \frac{2\psi \Delta \psi - |\nabla \psi |^2}{4 \psi ^\frac{3}{2}} + \frac{1}{4y^\frac{3}{2}}. \end{aligned}$$

Hence, using (A.10) and (A.11), we get

$$\begin{aligned} |\Delta (\sqrt{\psi } - \sqrt{y})| {\,\leqq \,}C{\varepsilon }y^{-\frac{3}{2}} {\,\leqq \,}C{\varepsilon }R^{-\frac{3}{2}} \end{aligned}$$
(A.12)

in the disc \(B_{R_1}(z_1)\). On the other hand, (A.11) implies

$$\begin{aligned} |\sqrt{\psi } - \sqrt{y}| {\,\leqq \,}C{\varepsilon }R^{\frac{1}{2}} \end{aligned}$$
(A.13)

in \(B_{R_1}(z_1)\). Now, applying the standard estimates for Laplac operator in the unit disk and scaling to the function \(\sqrt{\psi } - \sqrt{y}\) with (A.12)–(A.13), we obtain

$$\begin{aligned} |\nabla (\sqrt{\psi } - \sqrt{y})| {\,\leqq \,}C{\varepsilon }R^{-\frac{1}{2}} \end{aligned}$$

inside \(\frac{1}{2}B_{R_1}(z_1)\), that implies the required estimate \(|\nabla \psi (z_1) - \mathbf{e }_1^\perp |=|\mathbf{u }-\mathbf{e }_1| {\,\leqq \,}C{\varepsilon }\), see [1, Proof of Theorem 27, page 118].

\(\square \)

Appendix II

Proof of Lemma 16

Let \(w_r = \mathbf{w } \cdot \mathbf{e }_r\) be the radial component of \(\mathbf{w }\). We need the following classical inequality:

$$\begin{aligned} \frac{d}{dr} \int _0^{2\pi } |\mathbf{w }(r, \theta ) - \bar{\mathbf{w }}(r)|^2 \text {d}\theta&= \int _0^{2\pi } 2w_r \cdot (\mathbf{w } - \bar{\mathbf{w }}(r)) \text {d}\theta \nonumber \\&{\,\leqq \,}\int _0^{2\pi } \left[ r|w_r|^2 + \frac{|\mathbf{w } - \bar{\mathbf{w }}(r)|^2}{r} \right] \text {d}\theta \nonumber \\&{\,\leqq \,}\int _0^{2\pi } |\nabla \mathbf{w }|^2 r\text {d}\theta . \end{aligned}$$
(B.1)

By our assumption on the domain \({\mathcal {E}}\), we have \(\{r{\,\geqq \,}1\}\subset {\mathcal {E}}\). By integrating (B.1) on the interval \([r, \infty )\), and using the fact that \(\mathbf{w } \rightarrow \lambda \mathbf{e }_1\) uniformly at infinity, we obtain

$$\begin{aligned} \int _0^{2\pi } |\mathbf{w }(r, \theta ) - \bar{\mathbf{w }}(r)|^2 \text {d}\theta {\,\leqq \,}D_\lambda \end{aligned}$$
(B.2)

for any \(r{\,\geqq \,}1\). By (2.1) and (3.31), for \(1 {\,\leqq \,}r {\,\leqq \,}\frac{3}{2}\lambda ^{-1}\), we have

$$\begin{aligned} |\bar{\mathbf{w }}(r) - \lambda \mathbf{e }_1|&{\,\leqq \,}|\bar{\mathbf{w }}(r) - \bar{\mathbf{w }}(R_1)| + |\bar{\mathbf{w }}(R_1) - \lambda \mathbf{e }_1| \nonumber \\&{\,\leqq \,}C D_\lambda ^{\frac{1}{2}} \left( \log \frac{2}{\lambda r}\right) ^{\frac{1}{2}} + C D_\lambda ^{\frac{1}{2}} {\,\leqq \,}2C D_\lambda ^{\frac{1}{2}} \left( \log \frac{2}{\lambda r}\right) ^{\frac{1}{2}}. \end{aligned}$$
(B.3)

Here we have used that \(\log \frac{R_1}{r} {\,\leqq \,}\log \frac{4}{\lambda r} {\,\leqq \,}C \log \frac{2}{\lambda r}\) and \(\log \frac{2}{\lambda r} {\,\geqq \,}c\) for any \(r {\,\leqq \,}\frac{3}{2}\lambda ^{-1}\), for some absolute positive constants cC. Combining (B.2) and (B.3), we obtain

$$\begin{aligned} \int _0^{2\pi } |\mathbf{w } - \lambda \mathbf{e }_1|^2 \text {d}\theta&{\,\leqq \,}2\int _0^{2\pi } |\mathbf{w } - \bar{\mathbf{w }}|^2 \text {d}\theta + 2\int _0^{2\pi } |\bar{\mathbf{w }} - \lambda \mathbf{e }_1|^2 \text {d}\theta \\&{\,\leqq \,}2D_\lambda + C D_\lambda \log \frac{2}{\lambda r}\\&{\,\leqq \,}C D_\lambda \log \frac{2}{\lambda r} \end{aligned}$$

for \(1{\,\leqq \,}r {\,\leqq \,}\frac{3}{2}\lambda ^{-1}\). Integrating the above in r with respect to the measure rdr gives

$$\begin{aligned} \int _{\Omega _{\frac{1}{2}r, \frac{3}{2}r}} |\mathbf{w } - \lambda \mathbf{e }_1|^2 \, \text {d}x\text {d}y {\,\leqq \,}C r^2 D_\lambda \log \frac{2}{\lambda r} \end{aligned}$$

for any \(2{\,\leqq \,}r {\,\leqq \,}\lambda ^{-1}\). (Recall our notation \(\Omega _{r_1, r_2} := \{z: r_1<|z|<r_2\}\).) Now we use Ladyzhenskaya’s inequality to obtain

$$\begin{aligned} \int _{\Omega _{\frac{2}{3}r, \frac{4}{3}r}} |\mathbf{w }-\lambda \mathbf{e }_1|^4 \, \text {d}x\text {d}y&{\,\leqq \,}C\, \int _{\Omega _{\frac{1}{2}r, \frac{3}{2}r}} |\mathbf{w }-\lambda \mathbf{e }_1|^2 \, \text {d}x\text {d}y \ \Big ( \int _{\Omega _{\frac{1}{2}r, \frac{3}{2}r}} |\nabla \mathbf{w }|^2 \, \text {d}x\text {d}y \nonumber \\&\quad \quad \quad \quad \quad +\, \frac{1}{r^2} \int _{\Omega _{\frac{1}{2}r, \frac{3}{2}r}} |\mathbf{w }-\lambda \mathbf{e }_1|^2 \, \text {d}x\text {d}y \Big ) \nonumber \\&{\,\leqq \,}C r^2 D_\lambda ^2 \left( \log \frac{2}{\lambda r}\right) ^2 \end{aligned}$$
(B.4)

for any \(2{\,\leqq \,}r {\,\leqq \,}\lambda ^{-1}\). By Hölder’s inequality and (B.4), we have

$$\begin{aligned} \int _{\Omega _{\frac{2}{3}r, \frac{4}{3}r}} |\mathbf{w } \cdot \nabla \mathbf{w }|^{\frac{4}{3}} \, \text {d}x\text {d}y&{\,\leqq \,}\left( \int _{\Omega _{\frac{2}{3}r, \frac{4}{3}r}} |\mathbf{w }|^4 \, \text {d}x\text {d}y \right) ^{\frac{1}{3}} \left( \int _{\Omega _{\frac{2}{3}r, \frac{4}{3}r}} |\nabla \mathbf{w }|^2 \, \text {d}x\text {d}y \right) ^{\frac{2}{3}} \nonumber \\&\lesssim \left( \int _{\Omega _{\frac{2}{3}r, \frac{4}{3}r}} |\mathbf{w }-\lambda \mathbf{e }_1|^4 \, \text {d}x\text {d}y \right) ^{\frac{1}{3}} \left( \int _{\Omega _{\frac{2}{3}r, \frac{4}{3}r}} |\nabla \mathbf{w }|^2 \, \text {d}x\text {d}y \right) ^{\frac{2}{3}} \nonumber \\&\quad +\, \left( \int _{\Omega _{\frac{2}{3}r, \frac{4}{3}r}} \lambda ^4 \, \text {d}x\text {d}y \right) ^{\frac{1}{3}} \left( \int _{\Omega _{\frac{2}{3}r, \frac{4}{3}r}} |\nabla \mathbf{w }|^2 \, \text {d}x\text {d}y \right) ^{\frac{2}{3}} \nonumber \\&{\,\leqq \,}C r^{\frac{2}{3}} D_\lambda ^{\frac{4}{3}} \left( \log \frac{2}{\lambda r}\right) ^{\frac{2}{3}} + C r^{\frac{2}{3}} \lambda ^{\frac{4}{3}} D_\lambda ^{\frac{2}{3}} \nonumber \\&{\,\leqq \,}C r^{\frac{2}{3}} \lambda ^{\frac{4}{3}} D_\lambda ^{\frac{2}{3}} \max \left\{ |\log \lambda |^{-\frac{2}{3}} \left( \log \frac{2}{\lambda r} \right) ^{\frac{2}{3}}, 1 \right\} \nonumber \\&{\,\leqq \,}C r^{\frac{2}{3}} \lambda ^{\frac{4}{3}} D_\lambda ^{\frac{2}{3}} \end{aligned}$$
(B.5)

for any \(2{\,\leqq \,}r {\,\leqq \,}\lambda ^{-1}\). Local regularity theory for Stokes system as shown in Section 2.3 yields the estimate

$$\begin{aligned} \Vert \nabla ^2 \mathbf{w }\Vert _{L^\frac{4}{3}(\Omega _{\frac{3}{4}r, \frac{5}{4}r})}&{\,\leqq \,}C \Big ( \frac{1}{r^2}\Vert \mathbf{w } - \lambda \mathbf{e }_1\Vert _{L^\frac{4}{3}(\Omega _{\frac{2}{3}r, \frac{4}{3}r})} + \frac{1}{r} \Vert \nabla \mathbf{w }\Vert _{L^\frac{4}{3}(\Omega _{\frac{2}{3}r, \frac{4}{3}r})} \\&\quad +\, \Vert \mathbf{w }\cdot \nabla \mathbf{w }\Vert _{L^\frac{4}{3}(\Omega _{\frac{2}{3}r, \frac{4}{3}r})} \Big ) \end{aligned}$$

where C is independent of r. Applying (B.4), (3.1) and (B.5) to the above inequality, we obtain

$$\begin{aligned} \Vert \nabla ^2 \mathbf{w }\Vert _{L^\frac{4}{3}(\Omega _{\frac{3}{4}r, \frac{5}{4}r})} {\,\leqq \,}C r^{-\frac{1}{2}} D_\lambda ^{\frac{1}{2}} \left( \log \frac{2}{\lambda r}\right) ^{\frac{1}{2}} \end{aligned}$$

for any \(2 {\,\leqq \,}r{\,\leqq \,}\lambda ^{-1}\), which clearly implies

$$\begin{aligned} \Vert \nabla ^2 \mathbf{w }\Vert _{L^1(\Omega _{\frac{3}{4}r, \frac{5}{4}r})}{\,\leqq \,}C D_\lambda ^{\frac{1}{2}} \left( \log \frac{2}{\lambda r}\right) ^{\frac{1}{2}}. \end{aligned}$$

Now, Sobolev space theory (see, e.g., [2, Lemma 4.3] ) gives the following bound for the variation of \(\mathbf{w }\):

$$\begin{aligned} \text {diam}\, \mathbf{w }(\Omega _{\frac{3}{4}r, \frac{5}{4}r})&{\,\leqq \,}C \Big ( \Vert \nabla \mathbf{w }\Vert _{L^2(\Omega _{\frac{3}{4}r, \frac{5}{4}r})} + \Vert \nabla ^2 \mathbf{w }\Vert _{L^1(\Omega _{\frac{3}{4}r, \frac{5}{4}r})} \Big ) \nonumber \\&{\,\leqq \,}C D_\lambda ^{\frac{1}{2}} \left( \log \frac{2}{\lambda r}\right) ^{\frac{1}{2}}, \end{aligned}$$
(B.6)

for any \(2 {\,\leqq \,}r{\,\leqq \,}\lambda ^{-1}\). Together with (B.3), (B.6) gives the desired bound

$$\begin{aligned} |\mathbf{w }(z) - \lambda \mathbf{e }_1| {\,\leqq \,}C D_\lambda ^{\frac{1}{2}} \left( \log \frac{2}{\lambda r}\right) ^{\frac{1}{2}} \end{aligned}$$
(B.7)

in the region \(2 {\,\leqq \,}r{\,\leqq \,}\lambda ^{-1}\). To finish the proof, we point out that for the region \(\{r {\,\leqq \,}2\} \cap {\mathcal {E}}\), due to Lemma 11, (B.7) also holds true. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Korobkov, M., Ren, X. Uniqueness of Plane Stationary Navier–Stokes Flow Past an Obstacle. Arch Rational Mech Anal 240, 1487–1519 (2021). https://doi.org/10.1007/s00205-021-01640-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00205-021-01640-9

Navigation