Skip to main content

Advertisement

Log in

Weighted Korn and Poincaré-Korn Inequalities in the Euclidean Space and Associated Operators

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

Abstract

We prove functional inequalities on vector fields \(u: {{\mathbb {R}}}^d \rightarrow {{\mathbb {R}}}^d\) when \({{\mathbb {R}}}^d\) is equipped with a bounded measure \(\hbox {e}^{-\phi } \,\mathrm {d}x\) that satisfies a Poincaré inequality, and study associated self-adjoint operators. The weighted Korn inequality compares the differential matrix Du, once projected orthogonally to certain finite-dimensional spaces, with its symmetric part \(D^s u\) and, in an improved form of the inequality, an additional term \(\nabla \phi \cdot u\). We also consider Poincaré-Korn inequalities for estimating a projection of u by \(D^s u\) and zeroth-order versions of these inequalities obtained using the Witten-Laplace operator. The constants depend on geometric properties of the potential \(\phi \) and the estimates are quantitative and constructive. These inequalities are motivated by kinetic theory and related with the Korn inequality (1906) in mechanics, which compares Du and \(D^s u\) on a bounded domain.

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

Similar content being viewed by others

References

  1. Agmon, S.: Lectures on exponential decay of solutions of second-order elliptic equations: bounds on eigenfunctions of \(N\)-body Schrödinger operators, vol. 29 of Mathematical Notes. Princeton University Press, Princeton, NJ; University of Tokyo Press, Tokyo, 1982

  2. Ahmed, Z.M., Stroock, D.W.: A Hodge theory for some non-compact manifolds. J. Differ. Geom. 54(1), 177–225, 2000

    Article  MathSciNet  Google Scholar 

  3. Bakry, D., Émery, M.: Diffusions hypercontractives. In: Séminaire de probabilités, XIX, 1983/84, vol. 1123 of Lecture Notes in Math. Springer, Berlin, pp. 177–206, 1985

  4. Bardos, C., Golse, F., Nguyen, T.T., Sentis, R.: The Maxwell-Boltzmann approximation for ion kinetic modeling. Phys. D 376(377), 94–107, 2018

    Article  MathSciNet  Google Scholar 

  5. Bauer, S., Pauly, D.: On Korn’s first inequality for mixed tangential and normal boundary conditions on bounded Lipschitz domains in \(\mathbb{R}^N\). Ann. Univ. Ferrara Sez. VII Sci. Mat. 62, 2, pp. 173–188, 2016

  6. Bauer, S., Pauly, D.: On Korn’s first inequality for tangential or normal boundary conditions with explicit constants. Math. Methods Appl. Sci. 39(18), 5695–5704, 2016

  7. Carrapatoso, K., Dolbeault, J., Hérau, F., Mischler, S., Mouhot, C., Schmeiser, C.: Special macroscopic modes and hypocoercivity. arXiv:2105.04855, 2021

  8. Ciarlet, P. G.: Mathematical elasticity. Vol. I, vol. 20 of Studies in Mathematics and its Applications. North-Holland Publishing Co., Amsterdam, 1988

  9. Ciarlet, P.G.: Linear and Nonlinear Functional Analysis with Applications. Society for Industrial and Applied Mathematics, Philadelphia (2013)

    MATH  Google Scholar 

  10. Courtade, T.A., Fathi, M.: Stability of the Bakry-Émery theorem on \(\mathbb{R}^n\). J. Funct. Anal. 279(2), 108523, 2020

    Article  MathSciNet  Google Scholar 

  11. Cycon, H.L., Froese, R.G., Kirsch, W., Simon, B.: Schrödinger Operators with Application to Quantum Mechanics and Global Geometry. Texts and Monographs in Physics. Springer, Berlin (1987)

    Book  Google Scholar 

  12. Desvillettes, L., Villani, C.: On a variant of Korn’s inequality arising in statistical mechanics. ESAIM Control Optim. Calc. Var. 8,: 603–619 (electronic). A tribute to J. L, Lions, 2002

  13. Desvillettes, L., Villani, C.: On the trend to global equilibrium for spatially inhomogeneous kinetic systems: the Boltzmann equation. Invent. Math. 159(2), 245–316, 2005

    Article  ADS  MathSciNet  Google Scholar 

  14. Dolbeault, J., Volzone, B.: Improved poincaré inequalities. Nonlinear Anal. Theory Methods Appl. 75(16), 5985–6001, 2012

    Article  Google Scholar 

  15. Duan, R.: Hypocoercivity of linear degenerately dissipative kinetic equations. Nonlinearity 24(8), 2165–2189, 2011

    Article  ADS  MathSciNet  Google Scholar 

  16. Duvaut, G., Lions, J.-L.: Inequalities in mechanics and physics. Springer, Berlin-New York, Translated from the French by C, p. 219. W. John, Grundlehren der Mathematischen Wissenschaften, 1976

  17. Escobedo, M., Kavian, O.: Variational problems related to self-similar solutions of the heat equation. Nonlinear Anal. 11(10), 1103–1133, 1987

    Article  MathSciNet  Google Scholar 

  18. Figalli, A.: A geometric lower bound on Grad’s number. ESAIM Control Optim. Calc. Var. 15(3), 569–575, 2009

  19. Friedrichs, K.O.: On the boundary-value problems of the theory of elasticity and Korn’s inequality. Ann. Math. 2(48), 441–471, 1947

  20. Grad, H.: On Boltzmann’s \(H\)-theorem. J. Soc. Ind. Appl. Math. 13, 259–277, 1965

  21. Helffer, B., Nier, F.: Hypoelliptic Estimates and Spectral Theory for Fokker-Planck Operators and Witten Laplacians. Lecture Notes in Mathematics, vol. 1862. Springer, Berlin (2005)

    Book  Google Scholar 

  22. Helffer, B., Sjöstrand, J.: On the correlation for Kac-like models in the convex case. J. Stat. Phys. 74(1–2), 349–409, 1994

    Article  ADS  MathSciNet  Google Scholar 

  23. Hérau, F., Nier, F.: Isotropic hypoellipticity and trend to equilibrium for the Fokker-Planck equation with a high-degree potential. Arch. Ration. Mech. Anal. 171(2), 151–218, 2004

    Article  MathSciNet  Google Scholar 

  24. Horgan, C.O.: Korn’s inequalities and their applications in continuum mechanics. SIAM Rev. 37(4), 491–511, 1995

  25. Johnsen, J.: On the spectral properties of Witten-Laplacians, their range projections and Brascamp-Lieb’s inequality. Integral Equ. Oper. Theory 36(3), 288–324, 2000

  26. Kato, T.: Schrödinger operators with singular potentials. Israel J. Math. 13(1972), 135–148, 1973

    MATH  Google Scholar 

  27. Kondratiev, V.A., Oleinik, O.A.: On Korn’s inequalities. C. R. Acad. Sci. Paris Sér. I Math. 308(16), 483–487, 1989

  28. Korn, A.: Die Eigenschwingungen eines elastischen Körpers mit ruhender Oberfläche. Akad. der Wissensch., Munich, Math. phys. KI. 36, 351, 1906

  29. Korn, A.: Solution générale du problème d’équilibre dans la théorie de l’élasticité, dans le cas ou les efforts sont donnés à la surface. Annales de la Faculté des sciences de Toulouse: Mathématiques 2ième Série, 10, 165–269, 1908

  30. Korn, A.: Über einige Ungleichungen, welche in der Theorie der elastischen und elektrischen Schwingungen eine Rolle spielen. Krak. Anz. 705–724, 1909

  31. Ledoux, M.: The Concentration of Measure Phenomenon. Mathematical Surveys and Monographs, vol. 89. American Mathematical Society, Providence (2001)

  32. Lewicka, M., Müller, S.: On the optimal constants in Korn’s and geometric rigidity estimates, in bounded and unbounded domains, under Neumann boundary conditions. Indiana Univ. Math. J. 65(2), 377–397, 2016

  33. Mouhot, C., Russ, E., Sire, Y.: Fractional Poincaré inequalities for general measures. J. Math. Pures Appl. (9) 95(1), 72–84, 2011

  34. Neff, P., Pauly, D., Witsch, K.-J.: Poincaré meets Korn via Maxwell: extending Korn’s first inequality to incompatible tensor fields. J. Differ. Equ. 258(4), 1267–1302, 2015

  35. Nitsche, J.A.: On Korn’s second inequality. RAIRO Anal. Numér. 15(3), 237–248, 1981

  36. Persson, A.: Bounds for the discrete part of the spectrum of a semi-bounded Schrödinger operator. Math. Scand. 8, 143–153, 1960

    Article  MathSciNet  Google Scholar 

  37. Reed, M., Simon, B.: Methods of modern mathematical physics. II. Fourier analysis, self-adjointness. Academic Press [Harcourt Brace Jovanovich, Publishers], New York-London, 1975

  38. Simader, C.G.: Essential self-adjointness of Schrödinger operators bounded from below. Math. Z. 159(1), 47–50, 1978

    Article  MathSciNet  Google Scholar 

  39. Sjöstrand, J.: Correlation asymptotics and Witten Laplacians. Algebra i Analiz 8(1), 160–191, 1996

    MathSciNet  MATH  Google Scholar 

  40. Witten, E.: Supersymmetry and Morse theory. J. Differ. Geom. 17(4), 661–692, 1982

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

This work has been partially supported by the Projects EFI (K.C., J.D., ANR-17-CE40-0030) and Kibord (K.C., J.D., S.M., ANR-13-BS01-0004) of the French National Research Agency (ANR). C.M. and S.M. acknowledge partial funding by the ERC grants MATKIT 2011-2016 and MAFRAN 2017-2022. Moreover C.M. is very grateful for the hospitality at Université Paris-Dauphine.

© 2020 by the authors. This paper may be reproduced, in its entirety, for non-commercial purposes.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jean Dolbeault.

Additional information

Communicated by D. Kinderlehrer.

Publisher's Note

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

Appendices

Appendix A: Extensions, Gometric Observations and Motivation from Kinetic Theory

1.1 On the assumptions and some generalizations

Remark 11

The fact that the Poincaré inequality (H3) implies that \(\hbox {e}^{-\phi } \,\mathrm {d}x\) has an average \(\left<x\right>\) and a variance \(\langle |x|^2 \rangle \) is classical (see, e.g., [31, Corollary 3.2]). Indeed (H3) yields a concentration property of \(\hbox {e}^{-\phi (x)} \,\mathrm {d}x\) via a concentration function. A direct application of Fatou’s Lemma allows to extend (H3) to the set \(W^{1,\infty }\) of uniformly Lipschitz functions, which includes \(x\mapsto x_j\) for all \(j\in \left\{ 1,\ldots , d\right\} \). This directly gives \(\int _{{{\mathbb {R}}}^d} |x|^2\,\hbox {e}^{-\phi (x)} \,\mathrm {d}x \le d \,C_{\mathrm {\scriptscriptstyle P}}\) and the integrability of x w.r.t. \(\hbox {e}^{-\phi (x)} \,\mathrm {d}x\) follows by the Cauchy-Schwarz inequality. By induction, we get under (H3) alone that the set of all polynomial functions \({{\mathbb {R}}}[x]\) is included in \(L^2\) and even \(H^1\).

Remark 12

As a consequence of Remark 11 and of the strong Poincaré inequality (21), we directly get that for all \(i,j \in \left\{ 1,\ldots d\right\} \), \(\int _{{{\mathbb {R}}}^d} |x_i\,\partial _j \phi (x)|\,\hbox {e}^{-\phi (x)} \,\mathrm {d}x <\infty \). It is indeed sufficient to apply the strong Poincaré inequality to \(x \mapsto x_j\), which is in \(H^1\). Note that this gives sense to all quantities of Theorem 1, e.g., \(\left\| \nabla \phi \cdot R\right\| \) for any infinitesimal rotation R.

Remark 13

There are many sufficient conditions for the Poincaré inequality of Assumption (H3). When \(\phi \) is uniformly convex, it is shown in [3] that \(C_{\mathrm P}\) is greater or equal than the convexity constant, hence leading to fully explicit estimates in the two main theorems. If we only assume that \(\lim _{|x| \mapsto \infty } |\nabla \phi (x)|=+\infty \), then \(\Lambda \) is in fact an operator with compact resolvent, hence with discrete spectrum, and (H3) follows. Another, less stringent, sufficient condition on \(\phi \) is \(\liminf _{|x| \rightarrow \infty } \big (\frac{1}{2} |\nabla \phi (x)|^2-\Delta \phi (x)\big ) >c\) for some \(c>0\) (it implies the Poincaré inequality from the Persson-Agmon formula of [36] or [1, Theorem 3.2]). We note that this last assumption is satisfied by any regular function which coincides with \(x \mapsto \alpha \,|x|+\beta \) outside of a large centred ball, where \(\alpha \) and \(\beta \) are normalization constants.

Remark 14

Assumptions (H1)–(H2)–(H3) may be satisfied in other geometries than the one of the whole Euclidean space \({{\mathbb {R}}}^d\). In particular, given an open, smooth, bounded and connected subset \(\Omega \) of \({{\mathbb {R}}}^d\), we observe that these hypotheses are satisfied by the potential \(\phi (x)=\exp \big (1/d^2(x,\partial \Omega )\big )\), where \(d(x,\partial \Omega )\) denotes the usual Euclidean distance from x to \(\partial \Omega \). Here \({{\mathbb {R}}}^d\) is replaced by \(\Omega \) equipped with the measure \(\hbox {e}^{-\phi (x)} \,\mathrm {d}x\). It is an open question to understand how our results could be extended to usual boundary problems with potentials mimicking walls at the boundary of \(\Omega \).

1.2 Rigidity constants and defects of axisymmetry

In the proofs of Theorems 1 and 4 (see also “Appendix B.5”), we used the two rigidity constants \(C_{\mathrm {\scriptscriptstyle RV}}\) and \(C_{\mathrm {\scriptscriptstyle RV0}}\) defined by (11) and (17), depending on the level of regularity in each case, to measure the defects of axisymmetry (note that \(C_{\mathrm {\scriptscriptstyle RV}}\le C_{\mathrm {\scriptscriptstyle RV0}}\) since \(\Lambda \ge \mathrm {Id}\)). We used also \(C_{\mathrm {\scriptscriptstyle RD}}\) defined in (12) but remark that \(C_{\mathrm {\scriptscriptstyle RV}}\) and \(C_{\mathrm {\scriptscriptstyle RD}}\) are not directly comparable either, because \({\mathfrak {M}}_\phi ^c \ne D {{\mathcal {R}}}_\phi ^c\). Other ways of measuring the default of axisymmetry of the potential \(\phi \) can be considered.

\(\textcircled {1}\) One can consider, again, a rigidity of vector fields constant, but this time defined alternatively by

$$\begin{aligned} C_{\mathrm {\scriptscriptstyle RVL}}^{-1}=\min _{ A\,x \in {{\mathcal {R}}}_\phi ^c \setminus \left\{ 0\right\} } \frac{\left\| \nabla \phi \cdot A\,x\right\| ^2}{\left\| A\,x\right\| ^2}\quad \text{ when }\quad {{\mathcal {R}}}_\phi ^c \ne \left\{ 0\right\} \quad \text{ and } \quad C_{\mathrm {\scriptscriptstyle RVL}}=0\quad \text{ otherwise }\,. \end{aligned}$$

This leads to the modified Poincaré-Korn inequality

$$\begin{aligned}&\inf _{A\,x\in {{\mathcal {R}}}_\phi }\Vert u- \left<u\right>-A\,x\Vert ^2=\Vert u- \left<u\right>- {{\mathbb {P}}}_\phi (u)\Vert ^2 \nonumber \\&\quad \le C_{\mathrm {\scriptscriptstyle PKL}}'\,\Vert D^s u\Vert ^2 +2\,C_{\mathrm {\scriptscriptstyle RVL}}\,\Vert \nabla \phi \cdot (u-\left<u\right>)\Vert ^2 \end{aligned}$$
(48)

with an explicit bound for the constant \(C_{\mathrm {\scriptscriptstyle PKL}}'\) using (9) and the method of proof of (13). Once more, the existence of \(C_{\mathrm {\scriptscriptstyle RVL}}\) follows from the injectivity of \(A\,x \mapsto \nabla \phi \cdot A\,x\) on \({{\mathcal {R}}}_\phi ^c\) and the fact that \({{\mathcal {R}}}_\phi ^c\) is of finite dimension. The main advantage of this approach is to preserve a continuity property with respect to axisymmetry, which can be stated as follows: a small perturbation of a radial potential \(\phi \) gives rise to a small constant \(C_{\mathrm {\scriptscriptstyle RVL}}\), the limiting case being \({{\mathcal {R}}}_\phi ^c=\left\{ 0\right\} \) and \(C_{\mathrm {\scriptscriptstyle RVL}}=0\). The main drawback is that the symmetric operator associated to (48) is neither local nor differential because of the term \(\left<u\right>\) which appears in the right-hand side of (48).

\(\textcircled {2}\) In a bounded domain \(\Omega \subset {{\mathbb {R}}}^d\), with flat metric (i.e., for a constant potential \(\phi \)) considered in [12], the authors use Grad’s number. Let us explain how to adapt this method in our context under, e.g., the additional condition

$$\begin{aligned} \lim _{|x| \rightarrow \infty }\frac{D^2\phi (x)}{\lfloor \nabla \phi (x)\rceil ^2}=0\,. \end{aligned}$$

This property implies that the multiplication operator by \(D^2 \phi \) is relatively compact with respect to \(-\Delta _\phi \) acting on vector fields, with essential spectrum in \([C_{\mathrm {\scriptscriptstyle P}}, \infty )\). The spectrum in \([0,C_{\mathrm {\scriptscriptstyle P}})\) is then a pure point spectrum and the kernel is finite dimensional. For any antisymmetric matrix \(A\), there exists an affine space \(\mathfrak {V}_A\) of functions \(v\in H^1\) solving the Witten-Hodge problem

$$\begin{aligned} {{\,\mathrm{div_{\phi }}\,}}v=0\,,\quad D^av=A\,. \end{aligned}$$

The Witten-Hodge inequality asserts that

$$\begin{aligned} \inf _{v \in \mathfrak {V}_A}\Vert D^s v\Vert ^2 \le c_{\mathrm H}\,|A|^2 \end{aligned}$$
(49)

for some constant \(c_{\mathrm H}\in (0,\infty )\). The reverse inequality amounts to the existence of Grad’s number such that

$$\begin{aligned} C_{\mathrm G}^{-1} :=\inf _{A\in {\mathfrak {M}}_\phi ^c,\,|A|=1,\,v \in \mathfrak {V}_A}\;\Vert D^sv\Vert ^2. \end{aligned}$$

The existence of \(C_{\mathrm G}\) as well as a quantitative positive lower bound could be establish using mass transport theory exactly as in [12]. Of course, \(C_{\mathrm G}\) is well defined only when \({{\mathcal {R}}}_\phi ^c\ne \{0\}\), i.e., under the condition that \(\phi \) is not radially symmetric. Inequality (49) is natural in differential geometry and more specifically in De Rham cohomology theory: we refer to [2, 21] for further developments on this topic. In bounded domains, how to measure the symmetry defect by Grad’s number in view of Korn type inequalities is at the core of [12] but has also been studied in [18]. This approach differs from ours. Using \(C_{\mathrm G}\) and (8), the inequality (13) can be proved along a similar strategy as in [12], although with different constants.

1.3 An elementary application in kinetic theory

The main motivation for this paper comes from kinetic equations involving a confining potential studied in [7]. Also see [12, Section 2] and [4, 15] for applications of Korn inequalities to kinetic equations. As an example, let us consider the linear relaxation model of BGK-type

$$\begin{aligned} \partial _tf+v\cdot \nabla _x f -\nabla _x \phi \cdot \nabla _v f =\mathcal Lf :=G_f-f, \end{aligned}$$
(50)

where f(txv) is an unknown distribution function for a system of particles depending on time \(t \ge 0\), position \(x\in {{\mathbb {R}}}^d\) and velocity \(v\in {{\mathbb {R}}}^d\), and where \(G_f\) is defined by

$$\begin{aligned} G_f:=(\rho +\mathbf {u}\cdot v)\,\mu \quad&\text { where }\quad \rho (t,x) :=\int _{{{\mathbb {R}}}^d} f(t,x,v)\,\,\mathrm {d}v\,, \\&\text { and }\quad \mathbf {u}(t,x) :=\int _{{{\mathbb {R}}}^d}v\,f(t,x,v)\,\,\mathrm {d}v\,. \end{aligned}$$

Here \(\mu (v):=(2\,\pi )^{-d/2}\,\hbox {e}^{-|v|^2/2}\) while \(\rho \) and \(\mathbf{u}\) are respectively the macroscopic density and the average velocity associated with f. The collision kernel admits \(d+1\) conserved moments, in the sense that \(\int _{{{\mathbb {R}}}^d}\mathcal Lf(t,x,v)\,\,\mathrm {d}v=0=\int _{{{\mathbb {R}}}^d}v_i\,\mathcal Lf(t,x,v)\,\,\mathrm {d}v\) for any \(i=1,\ldots ,d\) and \(f \in L^1((1+|v|)\mathrm{d} v)\).

A natural question is to look for equilibria of (50). A quick glance at the equation shows that \({{\mathcal {M}}}(x,v) :=\hbox {e}^{-\phi (x)}\,\mu (v)\) is one of them. Korn inequalities provide us with a complete answer.

Proposition 15

Under Assumptions (H1)–(H2)–(H3), all equilibria of (50) in \(L^2({\mathcal {M}}^{-1}\,\mathrm {d}x\,\mathrm {d}v)\) take the form \(f(x,v)=\big (( R(x)\cdot v)+c\big )\,{{\mathcal {M}}}\) for some \(R \in {{\mathcal {R}}}_\phi \) and \(c \in {{\mathbb {R}}}\).

Proof

Write \(f=h\,{\mathcal {M}}\) with \(h \in L^2({\mathcal {M}}\,\mathrm {d}x\,\mathrm {d}v)\), \(\rho =r\,\hbox {e}^{-\phi }\), \(\mathbf{u}=u\,\hbox {e}^{-\phi }\), so that equation (50) reads

$$\begin{aligned} \partial _th+v\cdot \nabla _x h -\nabla _x \phi \cdot \nabla _v h = L(h) :=h-r-u\cdot v. \end{aligned}$$
(51)

The restriction of L to \(L^2(\mu \,\mathrm {d}v)\) is \(L=-\,\Pi ^\bot \) where \(\Pi \) is the orthogonal projection onto \(\mathrm {Span} \{1, v_1, \ldots , v_d\}\). We compute

$$\begin{aligned}&\frac{d}{dt}\iint _{{{\mathbb {R}}}^d\times {{\mathbb {R}}}^d}|h|^2\,{{\mathcal {M}}}\,\mathrm {d}x\,\mathrm {d}v = 2\iint _{{{\mathbb {R}}}^d\times {{\mathbb {R}}}^d}(L h)\,h\,{{\mathcal {M}}}\,\mathrm {d}x\,\mathrm {d}v \\&\quad = -\,2 \iint _{{{\mathbb {R}}}^d \times {{\mathbb {R}}}^d} \left| \Pi ^\bot h \right| ^2\, {{\mathcal {M}}}\,\mathrm {d}x \,\mathrm {d}v \end{aligned}$$

and deduce that any stationary solution of (51) takes the form \(h(x,v)=r(x)+u(x)\cdot v \). Equation (51) then reads

$$\begin{aligned} v\cdot \nabla _x(r+u\cdot v)=\nabla _x\phi \cdot u. \end{aligned}$$

Integrating the latter equation against respectively 1, \(v_i\) and \(v_i\,v_j\) with \(i \not =j\) in \(L^2(\mu \,\mathrm {d}v)\) yields

$$\begin{aligned} \mathrm{i)}\quad \nabla _x\cdot u-\nabla _x \phi \cdot u=0,\quad \mathrm{ii)}\quad D^s u=0,\quad \mathrm{iii)}\quad \nabla r=0. \end{aligned}$$

From iii) we get that there exists \(c \in {{\mathbb {R}}}\) such that \(r=c\). As for i), an integration by parts gives

$$\begin{aligned}&0=\int _{{{\mathbb {R}}}^d} (\nabla _x\cdot u-\nabla _x \phi \cdot u) \left<u\right>\cdot x\,\hbox {e}^{-\phi (x)} \,\mathrm {d}x= -\,\int _{{{\mathbb {R}}}^d} u\cdot \nabla \left( \left<u\right>\cdot x\right) \,\hbox {e}^{-\phi (x)} \,\mathrm {d}x\\&\quad = -\,\int _{{{\mathbb {R}}}^d} u\cdot \left<u\right>\,\hbox {e}^{-\phi (x)} \,\mathrm {d}x=-\,\left<u\right>^2, \end{aligned}$$

so that \(\left<u\right>=0\). Note also that taking the trace in ii) yields \(\nabla _x\cdot u=0\) so that i) reads \(\nabla _x \phi \cdot u=0\). Using this and (19) in Theorem 4 shows that \(u=R\) with \(R={{\mathbb {P}}}_\phi (u)\in {{\mathcal {R}}}_\phi \). Hence \(h=R(x)\cdot v+c\) and \(f(x,v)=\big (( R(x)\cdot v)+c\big )\,{{\mathcal {M}}}\). The reciprocal is straightforward, which completes the proof. \(\square \)

Appendix B: Additional Details on Computations

1.1 Functions, derivatives and projections

We denote by f a generic scalar function on \({{\mathbb {R}}}^d\) and by \(u:{{\mathbb {R}}}^d\rightarrow {{\mathbb {R}}}^d\) a generic vector field, so that \(\nabla f=(\partial _if)_{i=1}^d\) is a vector field and \(Du=(\partial _ju_i)_{i,j=1}^d\) takes values in \({\mathfrak {M}}\). The symmetric and the antisymmetric differentials of u, respectively \(D^su=\big ((D^su)_{ij}\big )_{i,j=1}^d\) and \(D^au=\big ((D^au)_{ij}\big )_{i,j=1}^d\) are defined by

$$\begin{aligned} (D^su)_{ij}:=\tfrac{1}{2}\left( \partial _ju_i+\partial _iu_j\right) \quad \text{ and }\quad (D^au)_{ij}:=\tfrac{1}{2}\left( \partial _ju_i-\partial _iu_j\right) \end{aligned}$$

so that \(D^su+D^au=Du\).

The orthogonal projection \({{\mathbb {P}}}\) of vector-valued functions is defined as follows. Let \((A_{ij})_{1\le i<j\le d}\) be a basis of \(\mathfrak M^a\) whose elements are

$$\begin{aligned} A_{ij}=\big ((\delta _{ij}-\delta _{ji})\,\delta _{ki}\,\delta _{j\ell } \big )_{k,\ell =1}^d \end{aligned}$$

and \((R_{ij})_{1\le i<j\le d}\) the orthonormal (in the \(L^2\) sense) basis of \({\mathcal {R}}\) given by

$$\begin{aligned} R_{ij}(x)=Z_{ij}^{-1}\,A_{ij}\,x \end{aligned}$$

whose coordinates are all 0 except the ith and the jth ones, with respective values \(-x_j/Z_{ij}\) and \(x_i/Z_{ij}\), i.e.,

$$\begin{aligned} R_{ij}(x)^\perp =Z_{ij}^{-1}\,\big (0,\ldots ,0,-x_j,0, \ldots ,0,x_i,0,\ldots ,0\big )\,, \end{aligned}$$

and where the normalization constant is \(Z_{ij}=\left( \int _{{{\mathbb {R}}}^d}(x_i^2+x_j^2)\,\hbox {e}^{-\phi (x)}\,\mathrm {d}x\right) ^{1/2}\). With these notations, \({{\mathbb {P}}}u\) is the vector field

$$\begin{aligned} x\mapsto {{\mathbb {P}}}u(x):=\sum _{1\le i<j\le d}{\mathsf {c}}_{ij}\,R_{ij}(x), \end{aligned}$$

where the coefficients are computed, for all integers i, j such that \(1\le i<j\le d\), as

$$\begin{aligned} \textstyle {\mathsf {c}}_{ij}= \int _{{{\mathbb {R}}}^d}u(x)\cdot R_{ij}(x)\,\hbox {e}^{-\phi (x)}\,\mathrm {d}x= \frac{1}{Z_{ij}}\int _{{{\mathbb {R}}}^d}\left( x_i\,u_j(x)-x_j\,u_i(x)\right) \, \hbox {e}^{-\phi (x)}\,\mathrm {d}x\,. \end{aligned}$$

The orthogonal projection \(\mathfrak {P}\) of a matrix-valued function \({\mathfrak {F}}\) is defined as

$$\begin{aligned} \mathfrak {P}\,{\mathfrak {F}}:=\sum _{1\le i<j\le d}{\mathsf {d}}_{ij}\,A_{ij} \end{aligned}$$

where the coefficients are computed, for all integers i, j such that \(1\le i<j\le d\), as

$$\begin{aligned} \textstyle {\mathsf {d}}_{ij}= \frac{1}{2}\int _{{{\mathbb {R}}}^d}\mathfrak F(x):A_{ij}\,\hbox {e}^{-\phi (x)}\,\mathrm {d}x= \frac{1}{2}\int _{{{\mathbb {R}}}^d} \big (\mathfrak F_{ij}(x)-{\mathfrak {F}}_{ji}(x)\big )\, \hbox {e}^{-\phi (x)}\,\mathrm {d}x\,. \end{aligned}$$

As a a consequence, we deduce that \(\mathfrak {P}{\mathfrak {F}}=\left<\mathfrak F^a\right>\) and \(\mathfrak {P}(Du) = \left<D^a u\right>\) for any \(u \in H^1\).

A matrix \(A\in D{{\mathcal {R}}}_\phi ^c\) is such that \(A\in {\mathfrak {M}}^a\) and for any \(B\in D{{\mathcal {R}}}_\phi \subset {\mathfrak {M}}^a\),

$$\begin{aligned} \textstyle 0=\int _{{{\mathbb {R}}}^d}A\,x\cdot B\,x\,\hbox {e}^{-\phi (x)}\,\mathrm {d}x= \sum _{i,j,k=1}^dA_{ij}\,B_{ik} \int _{{{\mathbb {R}}}^d}x_j\,x_k\,\hbox {e}^{-\phi (x)}\,\mathrm {d}x\,. \end{aligned}$$

A matrix \(A\in {\mathfrak {M}}_\phi ^c\) is such that \(A\in {\mathfrak {M}}^a\) and for any \(B\in D{{\mathcal {R}}}_\phi ={\mathfrak {M}}_\phi \subset {\mathfrak {M}}^a\),

$$\begin{aligned} \textstyle 0=\int _{{{\mathbb {R}}}^d}A:B\,\hbox {e}^{-\phi (x)}\,\mathrm {d}x= A:B=\sum _{i,j}^dA_{ij}\,B_{ij}. \end{aligned}$$

Based on these two definitions, it is clear that \(D{{\mathcal {R}}}_\phi ^c\) and \({\mathfrak {M}}_\phi ^c\) generically differ.

1.2 Operators

Let us give some details on the differential operators \(-\Delta _\phi \) and \(-\Delta _S\) associated respectively with the quadratic forms \(f\mapsto \Vert \nabla f\Vert ^2\) and \(u\mapsto \Vert D^s u\Vert ^2\).

\(\rhd \) Using \(\nabla _{\!\phi }u := \nabla \cdot u- \nabla \phi \cdot u\), the Witten-Laplace operator \(\Delta _\phi \) on functions is such that \(\Vert \nabla f\Vert ^2=(f,-\Delta _\phi f)\) and takes the form

$$\begin{aligned} \Delta _\phi f= \hbox {e}^\phi \,\nabla \cdot \left( \nabla f\,\hbox {e}^{-\phi }\right) = \nabla _{\!\phi }\cdot \nabla f = \Delta f - \nabla \phi \cdot \nabla f\,. \end{aligned}$$

\(\rhd \) By definition of \(D^su\) and using integration by parts, we have

$$\begin{aligned} \begin{array}{rl} (-\Delta _S u,u)&{}{}=2\int _{{{\mathbb {R}}}^d}|D^su|^2\,\text{ e}^{-\phi }\,\mathrm {d}x=\frac{1}{2}\sum _{i,j=1}^d \int _{{{\mathbb {R}}}^d}\left( \partial _iu_j+\partial _ju_i\right) ^2\text{ e}^{-\phi }\,\mathrm {d}x\\ &{}{}=-\frac{1}{2}\sum _{i,j=1}^d\int _{{{\mathbb {R}}}^d}u_j\, \partial _i\left( (\partial _iu_j+\partial _ju_i)\,\text{ e}^{-\phi }\right) \,\mathrm {d}x \\ &{}{}\quad -\frac{1}{2}\sum _{i,j=1}^d\int _{{{\mathbb {R}}}^d}u_i\, \partial _j\left( (\partial _iu_j+\partial _ju_i)\, \text{ e}^{-\phi }\right) \,\mathrm {d}x\\ &{}{}=-\sum _{i,j=1}^d\int _{{{\mathbb {R}}}^d}u_j\, \partial _i\left( (\partial _iu_j+\partial _ju_i)\, \text{ e}^{-\phi }\right) \,\mathrm {d}x\\ &{}{}=-\sum _{i=1}^d\int _{{{\mathbb {R}}}^d}\big (u_i\,\Delta u_i+u_i\,\partial _{ij}u_j\big )\,\text{ e}^{-\phi }\,\mathrm {d}x\\ &{}{}\quad +\sum _{i=1}^d\int _{{{\mathbb {R}}}^d}u_i\, \big ((\nabla \phi \cdot \nabla )\,u_i +2\,(D^su\,\nabla \phi )_i\big )\, \text{ e}^{-\phi }\,\mathrm {d}x\\ &{}{}=-\int _{{{\mathbb {R}}}^d}u\cdot \big (\Delta u+\nabla (\nabla \cdot u)-(\nabla \phi \cdot \nabla )\,u -2\,D^su\,\nabla \phi \big )\,\mathrm {d}x\,. \end{array} \end{aligned}$$

so that \(-\Delta _S\) is given, for an arbitrary vector field \(u\in {{\mathcal {C}}}_c^\infty ({{\mathbb {R}}}^d;{{\mathbb {R}}}^d)\), by

$$\begin{aligned} -\Delta _S\,u=-\,D^s_\phi \cdot D^su= -\big (\Delta u+\nabla (\nabla \cdot u) -(\nabla \phi \cdot \nabla )\,u-2\,D^su\,\nabla \phi \big )\,. \end{aligned}$$

1.3 Gaussian measure

In the normalized centred Gaussian case \(\phi (x)=\frac{1}{2}\,|x|^2+\frac{d}{2}\,\ln (2\pi )\) corresponding to (5), the basic constants are \(C_{\mathrm {\scriptscriptstyle P}}=1\) (which is the optimal constant in the Gaussian Poincaré inequality), either \(C_\phi =1+4\,d\) and \(C_\phi '=4\,\sqrt{d\,(1+4\,d)}\) if \(d\ge 2\), or \(C_\phi =8\) and \(C_\phi '=8\,\sqrt{2} \) if \(d=1\), as a limit case.

Let \(u(x)=(1-x_2^2,x_1\,x_2,0,\ldots 0)^\perp \). By elementary computations, we find that

$$\begin{aligned}&Du= \begin{pmatrix} \begin{array}{cc} 0&{}-\,2\,x_2\\ x_2&{}x_1\end{array}&{}{\mathbf{0}}\\ {\mathbf{0}}&{}{\mathbf{0}} \end{pmatrix},\quad D^su= \begin{pmatrix} \begin{array}{cc}0&{}-\,\frac{1}{2}\,x_2\\ -\,\frac{1}{2}\,x_2&{}x_1\end{array}&{}{\mathbf{0}}\\ {\mathbf{0}}&{}{\mathbf{0}} \end{pmatrix},\\&D^au= \begin{pmatrix} \begin{array}{cc}0&{}-\,\frac{3}{2}\,x_2\\ \frac{3}{2}\,x_2&{}0\end{array}&{}{\mathbf{0}}\\ {\mathbf{0}}&{}{\mathbf{0}} \end{pmatrix} \end{aligned}$$

where \(\mathbf{0}\) denotes \(2\times (d-2)\), \((d-2)\times 2\), and \((d-2)\times (d-2)\) null matrices. After integration against the normalized centred Gaussian measure, we have

$$\begin{aligned}&\left<u\right>=0,\quad \Vert u\Vert ^2=3\,,\quad {{\mathbb {P}}}(u)=0\,,\quad \mathfrak {P}(Du)=0=\left<D^au\right>\,,\\&\Vert D^su\Vert ^2=\frac{3}{2}\,,\quad \Vert D^au\Vert ^2=\frac{9}{2}\,, \quad \Vert Du\Vert ^2=6\,. \end{aligned}$$

This proves that \(C_{\mathrm {\scriptscriptstyle K}}=4\) in \(\Vert Du-\mathfrak {P}(Du)\Vert ^2 \le C_{\mathrm {\scriptscriptstyle K}}\,\Vert D^s u\Vert ^2\) and \(C_{\mathrm {\scriptscriptstyle PK}}=2\) in \(\Vert u-\left<u\right> - {{\mathbb {P}}}(u)\Vert ^2 \le C_{\mathrm {\scriptscriptstyle PK}}\,\Vert D^s u\Vert ^2\) are both optimal.

1.4 Estimates for the\(D(\Lambda )\)-toolbox and consequences

Here we give some details on the computation of \(C_{\mathrm {\scriptscriptstyle B}}\) in the proof of Proposition 8 in Sect. 3.1. Let \(A=\big \Vert \lfloor \nabla \phi \rceil \,\nabla f\big \Vert \), \(B=\big \Vert \lfloor \nabla \phi \rceil ^2\,f\big \Vert \), \(Z=\Vert \xi \Vert \) and let \({c'}^2=C_\phi '\ge C_\phi =c^2\). Inequalities (36) and (37) amount to

$$\begin{aligned} B\le \tfrac{4}{3}\,c\,A+\tfrac{5}{3}\,{c'}^2\,Z\quad \text{ and } \quad A^2\le B\,Z+\tfrac{1}{2\,c}\,A\,B+\tfrac{{c'}^2}{2\,c}\,A\,Z\,. \end{aligned}$$

Taking the equality case in the first inequality, we find that

$$\begin{aligned} A^2-4\,A\,\big (c+{c'}^2/c\big )\,Z-5\,{c'}^2\,Z^2\le 0 \end{aligned}$$

which means that

$$\begin{aligned} \textstyle A\le \left( 2\,\sigma +\sqrt{4\,\sigma ^2+5}\,\right) c'\,Z \quad \text{ with }\quad \sigma =\frac{c}{c'}+\frac{c'}{c}\ge 2\,. \end{aligned}$$

On \([2,+\infty )\), the function \(\sigma \mapsto \sqrt{4\,\sigma ^2+5}/\sigma \) is monotone non-increasing, so that \(\sqrt{4\,\sigma ^2+5}\le \frac{1}{2}\,\sqrt{21}\,\sigma \). Using the monotonicity of \(c\mapsto c\,\sigma \) and \(c\le c'\), we also have \(c\,\sigma \le 2\,c'\). As a consequence, we have

$$\begin{aligned}&\textstyle A\le \left( 2+\tfrac{1}{2}\,\sqrt{21}\,\right) c\,\sigma \, \frac{c'}{c}\,Z\le \left( 4+\sqrt{21}\,\right) \frac{{c'}^2}{c}\,Z \le 9\,\frac{{c'}^2}{c}\,Z\,,\\&\textstyle B \le \frac{1}{3}\left( 4\,c\left( 2\,\sigma +\sqrt{4\,\sigma ^2+5}\,\right) +5\,c'\right) c'\,Z\le \left( 7+4\,\sqrt{7/3}\,\right) \le 14\,{c'}^2\,Z\,, \end{aligned}$$

that is, the bounds (38). Moreover, from \(\big \Vert D^2f\big \Vert ^2\le \tfrac{C}{d}+\tfrac{1}{2}\left( d+\sqrt{d^2+4\,C}\,\right) \Vert \xi \Vert ^2\) with \(C=\tfrac{277}{8}\left( C_\phi -1\right) {C_\phi '}^2\), we deduce that

$$\begin{aligned} C_{\mathrm {\scriptscriptstyle B}}=81\,\tfrac{{C_\phi '}^2}{C_\phi } +196\,{C_\phi '}^2+\tfrac{C}{d}+\tfrac{1}{2}\left( d+\sqrt{d^2+4\,C}\,\right) . \end{aligned}$$
(52)

1.5 Estimates on various constants

The constants \(C_\phi \) and \(C_\phi '\) appear in (6) as a consequence of (H2) while the Poincaré constant \(C_{\mathrm {\scriptscriptstyle P}}\) follows from Assumption (H3). According to (41), the constant in the Poincaré-Lions inequality (22) is given with \(C_{\mathrm {\scriptscriptstyle B}}\) as in (52) by \(C_{\mathrm {\scriptscriptstyle PL}}=(1+C_{\mathrm {\scriptscriptstyle P}})^2\,\big (1+C_\phi '\,\sqrt{C_{\mathrm {\scriptscriptstyle B}}}/4\big )^2\). From Proposition 5, we know that the strong Poincaré inequality (21) holds for some \(C_{\mathrm {\scriptscriptstyle SP}}\le C_\phi \,(1+C_{\mathrm {\scriptscriptstyle P}})\). As for the other constants in Theorems 1 and 2, we learn from the proofs in Sects. 4.1 and 4.2 that

$$\begin{aligned}&C_{\mathrm {\scriptscriptstyle K}}\le 1+4\,C_{\mathrm {\scriptscriptstyle PL}}\,,\quad C_{\mathrm {\scriptscriptstyle PK}}\le C_{\mathrm {\scriptscriptstyle P}}\,C_{\mathrm {\scriptscriptstyle K}}\,, \quad C_{\mathrm {\scriptscriptstyle SPK}}\le C_{\mathrm {\scriptscriptstyle SP}}(C_{\mathrm {\scriptscriptstyle K}}+3\,C_\phi \,C_{\mathrm {\scriptscriptstyle PK}})\,,\\&C_{\mathrm {\scriptscriptstyle PK}}'\le C_{\mathrm {\scriptscriptstyle PK}}+2\,C_{\mathrm {\scriptscriptstyle RV}}\,C_{\mathrm {\scriptscriptstyle SPK}}\quad \text{ and }\quad C_{\mathrm {\scriptscriptstyle K}}'\le C_{\mathrm {\scriptscriptstyle K}}(1+2\,C_{\mathrm {\scriptscriptstyle RD}}\,C_{\mathrm {\scriptscriptstyle SP}})\,. \end{aligned}$$

In Sect. 6.1, Lemma 10, using \(C_\phi '':=C_{\varepsilon }\) as in (H2) with \({\varepsilon }=1/(2\,\sqrt{C_{\mathrm {\scriptscriptstyle B}}})\), the constants in (46) are

$$\begin{aligned} C_{\mathrm {\scriptscriptstyle LPL}}\le 4\,(1+C_{\mathrm {\scriptscriptstyle P}})^2\,(1+C_\phi '')^2\quad \text{ and }\quad C_{\mathrm {\scriptscriptstyle RPL}}\le \left( 1+\tfrac{1}{4}\,C_\phi '\,{\textstyle \sqrt{C_{\mathrm {\scriptscriptstyle B}}/C_\phi }}\,\right) ^2\,. \end{aligned}$$

Finally, the constants in Theorem 4 are given by

$$\begin{aligned} C_{\mathrm {\scriptscriptstyle K0}}=1+4\,C_{\mathrm {\scriptscriptstyle LPL}}\,,\quad C_{\mathrm {\scriptscriptstyle PK0}}\le C_{\mathrm {\scriptscriptstyle PL}}\,(1+4\,C_{\mathrm {\scriptscriptstyle LPL}}) \quad \text{ and }\quad C_{\mathrm {\scriptscriptstyle PK0}}' \le C_{\mathrm {\scriptscriptstyle PK0}}(1+2\,C_{\mathrm {\scriptscriptstyle RV0}}\,C_\phi ). \end{aligned}$$

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Carrapatoso, K., Dolbeault, J., Hérau, F. et al. Weighted Korn and Poincaré-Korn Inequalities in the Euclidean Space and Associated Operators. Arch Rational Mech Anal 243, 1565–1596 (2022). https://doi.org/10.1007/s00205-021-01741-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00205-021-01741-5

Navigation