Calculus of Variations and Partial Differential Equations (2021), doi: 10.1007/s00526-021-02000-x. by ) com- 3 . For × ν 3 P denotes R , 1 } (Ω ) r Ω X L ∂ k tr( P 1 3 1  ) − dev sym Curl 3 2 × ) = 0 on 3 X ν R , = × : (Ω r P L . dev sym Curl X k 3 2 k P = p which is determined by the condi- in terms of D dev sym( P , 3 ) -Korn inequality for incompatible tensor 1 if 3 and Patrizio Neff p S,dSC × L ). Here, dev 3 K 3 R dev sym Curl × k 2 3 R + (Ω; , and the boundary condition is understood in a r ) 0 such that the inequality , r > 3 L 3 1 × > X 3 1 ∈ ) R + , P , r 1 p (Ω p Abstract Ω L ≤ -estimate implies a classical Korn’s inequality with weak p, k p July 6, 2021 and a deviatoric-symmetric generalization of Poincar´e’sin- ( r 1 L c P 3 matrix u = × (dev sym Curl; Ω c sym ) = 0. In that case one can replace , 3 k = D Ω. If no boundary conditions are imposed then the estimate holds ) , Stefan M¨uller × dev sym Curl ∂ , p, r P ∗ 3 Primary: 35A23; Secondary: 35B45, 35Q74, 46E35. , ∞ | c 1 P 0 , R 1 ⊂ ) , (3). ≤ 3 W [1 ) × . The new so 3 3 ∈ ) ∈ 3 × ∈ R r 3 × P R 3 , A R , (Ω; (Ω p = p (Ω L L ,p P k 1 ∈ − P k P W (dev sym Curl; Ω { be an open and bounded set with Lipschitz boundary and outward unit normal k = 3 P Peter Lewintan = 0 and dev sym Curl R , p, r 1 0 we establish an improved version of the generalized P ⊂ W Korn’s inequality, Poincar´e’sinequality, Lions lemma, Neˇcasestimate, incompatibility, Curl- in the new Banach space ∞ P dimensions with conformally invariant dislocation energy The proof relies on a representation of theWe third also derivatives discuss D applications of the new inequality to the relaxed micromorphic model, to Cosserat Let Ω Specifically, there exists a constant < p < dev sym Curl the deviatoric (trace-free) part of a 3 suitable weak sense.tively This open, estimate non-empty also subsetafter holds Γ true taking if the the quotient boundary with condition the is finite-dimensional only space satisfied on a rela- models with theincompatible weakest linear form elasticity. of the curvature energy, to gradient plasticity with plastic spin and to where tions sym k equality by choosing bined with the Lions lemma and the Neˇcasestimate. boundary conditions by choosing 1 fields holds for all tensor fields Korn inequalities for incompatible tensor fields in three space Corresponding author, email: [email protected]. Chair for Nonlinear Analysis and Modeling, FacultyHausdorff of Center Mathematics, for University Mathematics of and Duisburg-Essen, Thea-Leymann-Str. Institute 9, for Applied Mathematics, Universit¨atBonn, Endenicher Allee 60, 53115

∗ 1 2

AMS 2020 subjectKeywords: classification: spaces, gradient plasticity,incompatibility dislocation tensor, density, trace-free Korn’s relaxed inequality, micromorphic conformal Killing model, vectors. Cosserat elasticity, Kr¨oner’s 45127 Essen, Bonn, Germany arXiv:2011.10573v3 [math.AP] 5 Jul 2021 Jul 5 [math.AP] arXiv:2011.10573v3 1 Introduction 1.1 Overview Korn’s second inequality provides an Lp-estimate of a gradient vector field (modulo a constant) in terms of the symmetric part of the derivative. This can be generalized to general fields P if one adds a term in Curl P on the right hand side [105, 104, 126]. For recent refined estimates which involve only the deviatoric part of sym P and Curl P , see [102, 11, 103]. Here, we show that P can be estimated in dimension n = 3 in terms of sym P and sym Curl P or even sym P and dev sym Curl P . The difference is that we need to subtract not only constants but also certain affine or quadratic skew-symmetric fields in the kernel of the operators sym Curl and dev sym Curl, respectively. To set the stage we recall the notation for the relevant Lie groups used in this paper and their Lie algebras and indicate how our new inequalities relate to (infinitesimal) conformal invariance. We denote the space n n of (n n)-matrices by R × and we denote the groups of proper orthogonal matrices, and matrices with determinant× 1 by n n T n n SO(n) = Q R × Q Q = 1 , SL(n) = B R × det B = 1 . (1.1a) { ∈ | } { ∈ | } The corresponding Lie algebras of skew-symmetric and trace-free matrices are denoted by n n T n n so(n) = A R × A = A , sl(n) = D R × tr D = 0 . (1.1b) { ∈ | − } { ∈ | } Let Ω R3 be open, bounded and simply connected. A C1-map ϕ :Ω R3 is conformal if its differential preserves⊂ the scalar product up to dilations, i.e., if for all x there exist →λ(x) 0 and Q(x) SO(3) such that Dϕ(x) = λ(x)Q(x). It is well-known that conformal maps are smooth and≥ the non-constant∈ conformal maps form a finite-dimensional manifold. The vector fields in the tangent space of the identity map are called conformal Killing fields (or infinitesimally conformal maps) and are characterized by the condition dev sym Du = 0. (1.2) In fact the solutions u of (1.2) are certain quadratic polynomials, see (1.10) below for an explicit formula. In the Cosserat theory, the curvature expression dev sym Curl A 2 for A :Ω so(3) (1.3a) k k → can be expressed equivalently as dev sym Daxl(A) 2. (1.3b) k k and the latter expression has been termed “conformal curvature”, consistent with (1.2) for u = axl(A). Therefore, we call the generalized curvature expression dev sym Curl P 2 (1.3c) k k conformal dislocation energy. Upon restricting P so(3) we recover (1.3b), see also (1.41). ∈ 1.2 The classical Korn’s inequalities First inequalities of this type were identified by more than hundred years ago, cf. [87, 88, 89], where they were derived for applications in linear elasticity. It is worth mentioning that after his graduation in 1890 Korn studied in under the supervision of Henri Poincar´e. For Korn’s biography including his pioneering work in telephotography we refer to [106, 90] but also [135, p. 182f]. We start by summarizing the inequalities which bear Korn’s name. In the following, let n 2 and Ω Rn be a bounded Lipschitz domain. Korn’s first inequality (in Lp) with vanishing boundary values≥ 1 reads ⊂ 1, p n Du Lp(Ω, n×n) c sym Du Lp(Ω, n×n) u W (Ω, R ). (1.4) k k R ≤ k k R ∀ ∈ 0 1In fact, the estimate is also true for functions with vanishing boundary values on a relatively open (non-empty) subset of the boundary.

2 It can be deduced from Korn’s second inequality (in Lp), which does not require boundary conditions:

 1, p n u W 1, p(Ω, n) c u Lp(Ω, n) + sym Du Lp(Ω, n×n) u W (Ω, R ). (1.5) k k R ≤ k k R k k R ∀ ∈ From the latter inequality also the following version follows

1, p n inf Du A Lp(Ω, n×n) c sym Du Lp(Ω, n×n) u W (Ω, R ). (1.6) A so(n)k − k R ≤ k k R ∀ ∈ ∈ For n 3 these inequalities can be improved to inequalities which only require the trace-free part of sym Du on the≥ right hand side. One has

1, p n Du Lp(Ω, n×n) c devn sym Du Lp(Ω, n×n) u W (Ω, R ), (1.7) k k R ≤ k k R ∀ ∈ 0 1 n n where devn X := X n tr(X) 1 denotes the deviatoric (trace-free) part of the square matrix X R × . Moreover, − · ∈

 1, p n u W 1, p(Ω, n) c u Lp(Ω, n) + devn sym Du Lp(Ω, n×n) u W (Ω, R ), (1.8) k k R ≤ k k R k k R ∀ ∈ as well as

1, p n u Πu W 1, p(Ω, n×n) c devn sym Du Lp(Ω, n×n) u W (Ω, R ) (1.9) k − k R ≤ k k R ∀ ∈ where Π is an arbitrary projection from W 1, p(Ω, Rn) onto the space of conformal Killing vectors (or in- finitesimal conformal mappings), i.e., the finite-dimensional kernel of devn sym D, which is given by quadratic polynomials of the form

1 2 n ϕC (x) = a, x x a x + A x + β x + b, with A so(n), a, b R and β R, (1.10) − 2 k k ∈ ∈ ∈ cf. [137, 124, 82, 37, 138, 146]. The situation is quite different in the planar case n = 2, since the condition dev2 sym Du 0 becomes the system of Cauchy-Riemann equations and the corresponding kernel is infinite- dimensional,≡ so that an adequate quantitative version of the trace-free classical Korn’s inequality does not hold true. However, in [62] it is proved that

1, p 2 Du Lp(Ω, 2×2) c dev2 sym Du Lp(Ω, 2×2) u W (Ω, R ), k k R ≤ k k R ∀ ∈ 0 but this result ceases to be valid if the homogeneous Dirichlet conditions are prescribed only on a part of the boundary, cf. the counterexample in [11, sec. 6.6]. For the limiting cases p = 1 and p = Korn-type inequalities fail, since from the counterexamples traced ∞R R back in [33, 98, 132, 112] it follows that Ω sym Du dx does not dominate each of the quantities Ω ∂iuj dx 1, 1 n k k | | for any vector field u W0 (Ω, R ). Nevertheless, Poincar´e-type inequalities estimating certain integral norms of the deformation∈ u in terms of the total variation of the symmetric strain tensor sym Du are still true. For Poincar´e-type inequalities for functions of bounded deformation involving only the deviatoric part of the symmetrized gradient we refer to [60]. These Korn inequalities are crucial for a priori estimates in linear elasticity and fluid mechanics, so that they are cornerstones for well-posedness results in linear elasticity (L2-setting) and the Stokes-problem (Lp- setting), cf. [81] and [29] for a modern elaboration, whereas the trace-free equivalents found applications in micropolar Cosserat-type models [83, 124, 82, 62] and general relativity [37]. The Korn inequalities generalize to many different settings, including the geometrically nonlinear coun- terpart [58, 101], mixed growth conditions [32], incompatible fields (also with dislocations) [115, 126, 11, 105, 104, 102, 63, 34, 69] and trace-free infinitesimal strain measures [37, 82, 137, 138, 62, 146, 11, 102, 103]. For trace-free Korn’s inequalities in pseudo-Euclidean space see [154] and for trace-free Korn inequalities on manifolds see [37, 80]. It is also possible to consider tangential boundary conditions, cf. [39, 143, 12, 13, 42].

3 Other generalizations are applicable to Orlicz-spaces [61, 19, 18, 26, 59] and SBD functions with small jump sets [55, 25, 56], thin domains [100, 71, 77, 113] as well as the case of non-constant coefficients [117, 92, 127, 136]. Moreover Korn-type inequalities are valid on H¨olderand John domains, see [84, 43, 40,3, 108, 41] and also the recent monograph [2] which relates those Korn inequalities to the existence of a right inverse of the divergence operator, to the Stokes equations and other inequalities. Piecewise Korn-type inequalities subordi- nate to a FEM-mesh and involving jumps across element boundaries have also been investigated, see e.g. [21, 99]. In the recent paper [147] the authors established a Korn inequality involving the BMO-seminorms which is valid on all bounded domains and with a constant depending only on the dimension. Here we focus on inequalities for general tensor fields.

1.3 Korn-type inequalities for incompatible tensor fields Classical Korn’s inequalities require compatibility, i.e., a gradient Du (the Jacobian matrix). Generalizations of such estimates to general fields P then need a control of the distance of P to a gradient by adding the incompatibility measure (the dislocation density tensor) Curl P . The matrix Curl operation is to be understood as row-wise application of the classical curl to vectors. Even though the usual Curl operation on R3 has a natural extension to all dimensions, the case n = 3 deserves our special attention, not only from the viewpoint of modeling but also since the matrix Curl then returns a square matrix in three dimensions. As 1, p 3 3 direct generalization of Korn’s first inequality (1.4) we have for all P W (Curl; Ω, R × ) ∈ 0  P p 3×3 c sym P p 3×3 + Curl P p 3×3 , (1.11) k kL (Ω,R ) ≤ k kL (Ω,R ) k kL (Ω,R ) cf. [126] for p = 2 and [105] for all p > 1. Furthermore, the version (1.6) generalizes to  3×3 3×3 3×3 inf P Ae Lp(Ω,R ) c sym P Lp(Ω,R ) + Curl P Lp(Ω,R ) (1.12) Ae so(3)k − k ≤ k k k k ∈ 1, p 3 3 for all P W (Curl; Ω, R × ), cf. [105]. These estimates also hold true in all dimensions n 2 with an adequate∈ understanding of the matrix Curl, [104]. However, in two dimensions even stronger≥ estimates 1 2 2 hold true, cf. [63] and its nonlinear counterpart in [115], so especially, for fields P L (Ω, R × ) with 1 2 2 2 2 ∈ Curl P L (Ω, R ) it follows that P L (Ω, R × ) and ∈ ∈  P 2 2×2 c sym P 2 2×2 + Curl P 1 2 (1.13) k kL (Ω,R ) ≤ k kL (Ω,R ) k kL (Ω,R ) R under the normalization condition Ω skew P dx = 0, cf. [63]. However, this is essentially a result for the divergence, since Div is a rotated Curl in two dimensions.2 Indeed, the authors of [63] make use of the fact 1 2 2 1 that a vector field u L (Ω, R ) satisfying div u H− (Ω) belongs to H− (Ω) with ∈ ∈

u −1 2 c ( u 1 2 + div u −2 ) (1.14) k kH (Ω,R ) ≤ k kL (Ω,R ) k kH (Ω) which follows from [22]. For the geometrically nonlinear counterpart of (1.13) in a mixed-growth setting in two dimensions we refer the reader to [67] and higher-dimensional analogues can be found in [93, 34]. Improvements of the Korn inequalities for incompatible tensor fields (1.11) and (1.12) towards the trace- 1, p 3 3 free cases are also valid. For all P W (Curl; Ω, R × ), where ∈ 0 1, p 3 3 p 3 3 p 3 3 W (Curl; Ω, R × ) := P L (Ω, R × ) Curl P L (Ω, R × ),P ν = 0 on ∂Ω (1.15) 0 { ∈ | ∈ × } one has  P p 3×3 c dev sym P p 3×3 + dev Curl P p 3×3 , (1.16) k kL (Ω,R ) ≤ k kL (Ω,R ) k kL (Ω,R ) 2 2 The Babuˇska-Aziz theorem implies that over a planar Lipschitz domain Ω ⊂ R it holds

kDuk 2 2×2 ≤ c kdiv uk 2 L (Ω,R ) L (Ω,R) 1 2 R for all u ∈ H0 (Ω, R ) such that Ω div u dx = 0, cf. [81, Section 6].

4 cf. [11] for p = 2 and [102] for all p > 1. Moreover, we have  3×3 3×3 3×3 inf P T Lp(Ω,R ) c dev sym P Lp(Ω,R ) + dev Curl P Lp(Ω,R ) (1.17) T KdS,dC k − k ≤ k k k k ∈ 1, p 3 3 for all P W (Curl; Ω, R × ), cf. [102], where the kernel on the right hand side is given by ∈ 3 3   KdS,dC = T :Ω R × T (x) = Anti Ae x + β x + b + axl A,e x + γ 1, { → | 3 Ae so(3), b R , β, γ R , (1.18) ∈ ∈ ∈ } where Anti : R3 so(3) is the canonical identification (consistent with the vector product) of R3 and the → vectorspace of skew-symmetric matrices so(3) and axl : so(3) R3 is its inverse. The appearance of the → 1, p 3 3 dev Curl operator on the right hand side would suggest to extend the Banach space W (Curl; Ω, R × ) to p-integrable tensor fields P with p-integrable dev Curl P , but this would not be a new space. Indeed, in [102] 3 3 the authors showed that for all P D 0(Ω, R × ) and all m Z one has ∈ ∈ m, p 3 3 m, p 3 3 Curl P W (Ω, R × ) dev Curl P W (Ω, R × ). (1.19) ∈ ⇔ ∈ Note, that the estimates (1.16) and (1.17) are strictly restricted to the case of three dimensions since the deviatoric operator acts on square matrices and only in the three-dimensional setting the matrix Curl op- erator returns again a square matrix. On the other hand, the corresponding weaker estimates in terms of devn sym P Lp + Curl P Lp hold true in all dimensions n 3, cf. [11] for p = 2 and [103] for all p > 1. k For compatiblek kP = Dku we get back from (1.11), (1.12≥), (1.16) and (1.17) the corresponding classical Korn inequalities. Recently, Gmeineder and Spector [69] extended inequality (1.11) to the case where sym P is generalized to any linear operator A(P ) such that A(Du) is a first order elliptic operator, thus including also one result of [102] with dev sym P . The objective of the present paper is to further improve on estimate (1.11) by showing that it already suffices to consider the symmetric or even the trace-free symmetric part of the Curl. More precisely, for all 1, p, r 3 3 P W (dev sym Curl; Ω, R × ) where ∈ 0 1, p, r 3 3 W0 (dev sym Curl; Ω, R × ) := p 3 3 r 3 3 P L (Ω; R × ) dev sym Curl P L (Ω; R × ), dev sym(P ν) = 0 on ∂Ω { ∈ | ∈ × } and 1 1 1 3 r [1, ), + , r > 1 if p = ∈ ∞ r ≤ p 3 2 there exists a constant c = c(p, Ω, r) > 0 such that one has  P p 3×3 c sym P p 3×3 + dev sym Curl P r 3×3 . (1.20) k kL (Ω,R ) ≤ k kL (Ω,R ) k kL (Ω,R ) If no boundary conditions are imposed then we show  3×3 3×3 3×3 inf P T Lp(Ω,R ) c sym P Lp(Ω,R ) + dev sym Curl P Lr (Ω,R ) , (1.21) T KS,dSC k − k ≤ k k k k ∈ where the kernel is given by

3 3 1 2 KS,dSC = T :Ω R × T (x) = Anti Ae x + β x + b + d, x x d x , { → | − 2 k k 3 Ae so(3), b, d R , β R . (1.22) ∈ ∈ ∈ } Remark 1.1. The right-hand side of (1.20) provides a norm on smooth, compactly supported functions 3 3 P C0∞(Ω, R × ). Indeed, sym P 0 implies P = A C0∞(Ω, so(3)), so that by Nye’s formula (1.26)1 the ∈ ≡ ∈ 3 condition dev sym Curl A 0 reads already dev sym Da 0 with a := axl A C0∞(Ω, R ), where 3 ≡ ≡ ∈ T axl : so(3) R associates to a skew-symmetric matrix A so(3) the vector axl A := ( A23,A13, A12) . The trace-free→ Korn’s inequality (1.7) then gives Da 0.∈ Hence, a = axl A is a constant− vector field,−P = A is a constant skew-symmetric matrix field, and with the≡ boundary condition we obtain P 0. ≡

5 Remark 1.2. On the other hand, there are no such estimates in terms of

dev P + sym Curl P , dev sym P + sym Curl P or dev sym P + dev sym Curl P k k k k k k k k k k k k due to the example P = ζ 1 for which Curl P = Anti( ζ), so that the corresponding right-hand sides would vanish, since here we· have dev P = dev sym P− = 0∇ and also sym Curl P = dev sym Curl P = 0.

1.4 Proof ideas for Korn inequalities There exist many different proofs of the classical Korn’s inequalities, cf. the discussions in [28, 126,5, 86, 129, 70, 78, 79, 57, 50, 51, 49, 14, 134, 85, 150, 23, 149, 151, 140] as well as [29, Sect. 6.15] and the references contained therein. A rather concise and elegant argument, see [64, 44, 39, 68] and also advocated by P. G. Ciarlet and his coworkers [28, 29, 30, 27, 31] uses the Lions lemma resp. Neˇcasestimate, the compact embedding W 1, p Lp and the well-known representation of the second distributional derivatives of the displacement u by⊂⊂ a linear combination of the first derivatives of the symmetrized gradient Du, namely

∂ ∂ u = ∂ (sym Du) + ∂ (sym Du) ∂ (sym Du) , (1.23a) i j k j ik i jk − k ij i.e. D2u = L(D sym Du) with a constant coefficient linear operator L. (1.23b)

Also the trace-free Korn’s inequalities can be deduced in such a way, relying on the “higher order” analogues of the differential relation (1.23): 2 D∆u = L(D devn sym Du), (1.24) cf. [37] for the case p = 2 and [146] for all p > 1. The first and the last author used a similar reasoning in their series of papers [105, 104, 102, 103] to obtain the Korn inequalities for incompatible tensor fields mentioned above. In particular, the gradient of a skew-symmetric matrix field A can be expressed as linear combination of the entries of the matrix Curl:

DA = L(Curl A), (1.25) which in three dimensions reads exactly as Nye’s formula [130, eq.(7)]: 1 Curl A = tr(Daxl A) 1 (Daxl A)T , resp. Daxl A = (tr[Curl A])1 (Curl A)T . (1.26) − 2 − Furthermore, the second derivatives of a skew-symmetric matrix field A are given by linear combinations of the entries of the derivative of the deviatoric matrix Curl:

D2A = L(D dev Curl A) (1.27) which was used in the proof of the trace-free case [102]. The expression (1.26)1 admits a counterpart on the group of orthogonal matrices O(3) and even in higher spatial dimensions, see e.g. [125]. Note in passing, that the representation of the kernel of sym Du 0 can either be deduced from (1.23) or (1.26) and yields the class RM of infinitesimal rigid motions ≡

3 RM := Ae x + b Ae so(3), b R . (1.28) { | ∈ ∈ } Indeed, assuming sym Du 0 ≡ • (1.23) implies that D2u 0, so that u has to be affine with u RM, equivalently, ≡ ∈ • since Du = A(x) with a skew-symmetric matrix field A, we obtain Curl A = Curl Du 0, so that by (1.26) we deduce Daxl A 0 and hence A const, i.e., again u RM. ≡ ≡ ≡ ∈

6 Summarizing, the following differential relations connecting higher order derivatives have been used in the distributional sense for • classical Korn: D2u = L(D sym Du) cf. [30, 64, 44] 2 • trace-free classical Korn: D∆u = L(D devn sym Du) cf. [37, 146] • incompatible Korn: DA = L(Curl A) cf. [105, 104] • trace-free incompatible Korn: D2A = L(D dev Curl A) cf. [102] D2(A + ζ 1) = L(DCurl(A + ζ 1)) cf. [102, 103] · · D3(A + ζ 1) = L(D2 dev Curl(A + ζ 1)) cf. [102] · · • symmetrized incompatible Korn: D2A = L(D sym Curl A) present paper • conformally invariant incompatible Korn: D3A = L(D2 dev sym Curl A) denoting by u a displacement vector field, by A a skew-symmetric tensor field, by ζ a scalar field and by L a corresponding linear operator with constant coefficients. Moreover, we have by [102] for a general field P

DCurl P = L(Ddev Curl P ) . (1.29)

1.5 Motivation for Korn type estimates for incompatible tensor fields Korn type inequalities for incompatible tensor fields originally motivated from infinitesimal gradient plasticity with plastic spin as well as in the linear relaxed micromorphic elasticity, see e.g. [45, 48, 119, 116, 121, 123, 122, 46, 47, 139, 118, 120, 66, 133, 110, 73, 74, 75, 52, 155, 11] and the references contained therein.

1.5.1 Application to the relaxed micromorphic model The relaxed micromorphic model is a novel micromorphic framework [121, 119] that allows e.g. the description of microstructure-related frequency band-gaps [36] through a homogenized linear model. The goal is to find 3 3 3 3 3 the displacement u:Ω R R and the non-symmetric micro-distortion field P :Ω R R × minimizing ⊆ → ⊆ → Z W (Du, P, Curl P ) + f, u dx , such that (u, P ) H1(Ω) H(Curl), Ω ∈ × where the energy W is defined as 1 1 W = Ce sym (Du P ) , sym (Du P ) 3×3 + Cmicro sym P, sym P 3×3 2 − − R 2 R 2 (1.30) 1 µLc + Cc skew (Du P ) , skew (Du P ) 3×3 + L Curl P, Curl P 3×3 . 2 − − R 2 R th th Here, Ce, Cmicro : Sym(3) Sym(3) are classical 4 order elasticity tensors, Cc : so(3) so(3) is a 4 order → → rotational coupling tensor, Lc 0 is a characteristic length scale, µ is a typical effective shear modulus and 3 3 3 3 ≥ L: R × R × . The associated Euler-Lagrange equations read → Div [Ce sym (Du P ) + Cc skew (Du P )] = f, − − 2 (1.31) Ce sym (Du P ) + Cc skew (Du P ) Cmicro sym P µ L Curl[L Curl P ] = 0. − − − − c The generalized moment balance (1.31)2 can be seen as a tensorial Maxwell problem due to the Curl[L Curl P ] operation, cf. [20]. The most general quadratic representation of the curvature energy is given by

L Curl P, Curl P (1.32)

3 3 3 3 where L : R × R × is a non-standard fourth order tensor with 45 independent entries acting on the → 3 3 non-symmetric second order tensor Curl P R × . Since Curl P transforms as a second order tensor under ∈

7 rotations of the coordinate system, cf. [116, 118], assuming a certain degree of anisotropy allows one to reduce the complexity of L. Notably, the most general isotropic quadratic expression of the curvature energy is given by α α dev sym Curl P 2 + α skew Curl P 2 + 3 tr2(Curl P ), (1.33) 1k k 2k k 3 + 3 3 with three free parameters α1, α2, α3 R . Here we have used the orthogonal decomposition of R × into orthogonal pieces, namely ∈

3 3 R × = [sl(3) Sym(3)] so(3) R 1 (1.34a) ∩ ⊕ ⊕ · 3 3 so that for any square matrix X R × we have ∈ 1 X = dev sym X + skew X + tr(X) 1, (1.34b) 3 · where dev sym, skew, tr are orthogonal projections on the vector space sl(3) Sym(3) of symmetric trace free ∩ matrices, the space so(3) of skew-symmetric matrices, and the space R 1 of spherical tensors, respectively. In order to reduce complexity in the model one might be tempted to· replace (1.32) with

Lb sym Curl P, sym Curl P (1.35) where Lb : Sym(3) Sym(3) is now a classical positive definite fourth order elasticity tensor, whose repre- sentation for all anisotropy→ classes is completely known. A weak formulation of the static problem

Div [Ce sym (Du P ) + Cc skew (Du P )] = f, − − (1.36) 2 Ce sym (Du P ) + Cc skew (Du P ) Cmicro sym P µ L Curl[Lb sym Curl P ] = 0 , − − − − c 3 3 2 3 3 2 3 3 is naturally formulated in the space H(sym Curl; Ω, R × ) := P L (Ω, R × ) sym Curl P L (Ω, R × ) and our new result shows that this problem is well-posed for{ a suitable∈ prescription| of tangential∈ boundary} data. Returning to (1.33), the problem may be even further “relaxed” by requiring only to control

dev sym Curl P 2. (1.37) k k 3 3 2 3 3 In this case, the natural space to consider is the Hilbert space H(dev sym Curl; Ω, R × ) := P L (Ω, R × ) 2 3 3 { ∈ | dev sym Curl P L (Ω, R × ) and our result (1.20) implies that the weak formulation is still well-posed. ∈ } Remark 1.3 (Nothing new in plane strain). Note, that due to the structure of the three-dimensional matrix Curl operator in plain strain, i.e., assuming that     Pb11(x, y) Pb12(x, y) 0 0 0 ∗ Pb(x, y, z) = P (x, y) P (x, y) 0 , Curl Pb = 0 0  (1.38) b21 b22 ∗ 0 0 0 0 0 0

3 3 the operation sym or dev sym is not leaving the classical H(Curl; Ω, R × ) space, since

2 2 1 2 dev3 sym Curl Pb = sym Curl Pb = Curl Pb . (1.39) k k k k 2k k Hence, new properties to be discovered are strictly three-dimensional in nature.

1.5.2 Cosserat model with weakest curvature energy – conformally invariant curvature The use of the dislocation density tensor Curl P in the relaxed micromorphic model allows a smooth transition in the modeling to the classical linear Cosserat model. Indeed, letting formally Cmicro in the relaxed → ∞

8 micromorphic model (1.30), i.e., assuming P = A so(3) is skew-symmetric, the (isotropic) elastic Cosserat free energy can be written as ∈ Z 2 2 λ 2 µ sym Du +µc skew(Du A) + tr (Du) Ω k k k − k 2 (1.40) α + α dev sym Curl A 2 + α skew Curl A 2 + 3 tr2(Curl A) dx min . 1k k 2k k 3 →

In [123, 122] it has been shown that choosing α1 > 0, α2 = α3 = 0 is mandatory for offering bounded stiffness in bending and torsion for arbitrary small specimen. This corresponds to the conformally invariant curvature case dev sym Curl A 2 = dev sym Daxl(A) 2 . (1.41) k k k k Well-posedness results are then based on the trace-free Korn’s inequality [122]. Finally, letting the Cosserat couple modulus µc in (1.40), one obtains the so-called modified inde- terminate couple stress model [124, 66] → ∞ Z 2 λ 2 2 µ sym Du + tr (Du) + α1 dev sym Dcurl u dx min . (1.42) Ω k k 2 |k {z k} → = sym Dcurl u 2 k k conformally invariant curvature3

In [124] this curvature energy has been obtained by a passage from a discrete model to a continuum model- ing, invoking a “micro-randomness” assumption, which introduces an additional invariance property beyond isotropy.

1.5.3 Application to gradient plasticity with plastic spin Experiments with differently sized specimens have revealed a pronounced size-effect in elasto-plastic transfor- mations [52, 73, 74, 75] which cannot be described with classical phenomenological elasto-plasticity models. For the sake of simplicity we assume in the following the additive decomposition of the displacement gradient Du into non-symmetric elastic (recoverable) and non-symmetric plastic (permanent) distortions e and P , respectively: Du = e + P, ε := sym Du = sym e + sym P = εe + εp , (1.43) under the side condition of plastic incompressibility tr(P ) = tr(εp) 0. A simplified framework for size- independent plasticity can be sketched, based on the introduction of≡ the total free energy, which consist of elastic contributions and local hardening Z

W (Du, P ) = e sym (Du P ) , sym (Du P ) 3×3 + hard sym P, sym P 3×3 + f, u dx C R C R Ω | − {z − } | {z } elastically stored energy local hardening Z

= Ce (ε εp) , (ε εp) 3×3 + Chard εp, εp + f, u dx, (1.44) Ω − − R where Ce, Chard are classical positive definite fourth order tensors acting on symmetric arguments. We are working here in a phenomenological modeling context. A variational approach to single crystals with dislocations, different from out presented phenomenological viewpoint, has been explored in [145, 144] based on [114]. Our term Chard sym P, sym P gives rise to the usual Prager-type backstress term (linear kinematic hardening) which appears ubiquitous in the literature. The appearance of sym(Du P ) and sym P instead of Du P and P alone is dictated by linearized frame indifference of the model. Equilibrium− of forces −

Div Ce sym(Du P ) = Div Ce(ε εp) = f (1.45) − − 3tr(sym Dcurl u) = tr(Dcurl u) = div curl u ≡ 0.

9 appears from variation of (1.44) with respect to the displacement u. It remains to postulate a “flow rule”, i.e., an evolution for the plastic variable P . This equation appears as gradient flow with respect to P in the form ˙ P = F ( DP W (Du, P )) = F (Ce sym(Du P ) Chard sym P ) (1.46) − − − 3 3 3 3 together with suitable initial conditions for P and boundary conditions for u, where F : R × R × is → monotone, i.e., F (X) F (Y ),X Y 3×3 0 and maps symmetric arguments to trace-free symmetric − − R ≥ arguments, the increment P˙ is determined to be trace-free symmetric and (1.46) can be therefore recast as

ε˙p = F (Ce (ε εp) Chard εp) , tr(εp) = 0 . (1.47) − − In order to extend the modeling framework to incorporate size-dependence, let us focus on the introduction of energetic length scales. In this case, one augments the total energy (1.44) by some terms involving space derivatives of the plastic distortion P or the plastic strain εp := sym P , for simplicity

DP 2 or Dε 2 . (1.48) k k k pk

Accordingly, based on (1.48)1, the evolution law (1.46) needs to be adapted to ˙ P = F ( DP W (Du, P, DP )) = F (Ce sym(Du P ) Chard sym P + ∆P ) (1.49) − − − and suitable boundary conditions for the plastic distortion P , here Dirichlet clamping P ∂Ω 0, cf. [76]. For initial condition P (0) Sym(3) (1.49) can again be recast into4 | ≡ ∈ ε˙p = F (Ce (ε εp) Chard εp + ∆εp) , εp = 0. (1.50) − − |∂Ω Such a model is already able to predict that smaller samples are relatively stiffer. However, the simple gradient terms in (1.48) lack a microscopical justification. However, ∆P can be seen as regularization term as in [54]. Since plasticity is mediated by dislocation movements it inspires that a physically more suitable description is given by considering the dislocation density tensor Curl P and, in first approximation, a simple quadratic function thereof to replace (1.48). Hence, the total stored energy can be written Z

Ce sym (Du P ) , sym (Du P ) + Chard sym P, sym P + Curl P, Curl P + f, u dx . (1.51) Ω − −

Since Curl is self-adjoint with suitable tangential boundary conditions P ν ∂Ω = 0, the evolution law turns into × | ˙ P = F (Ce sym(Du P ) Chard sym P Curl Curl P ) . (1.52) − − − Note that (1.52) is necessarily an evolution for a non-symmetric plastic distortion P since the contribution Curl Curl P does not have any symmetry properties. Such models are called gradient plasticity models with plastic spin or distortion gradient plasticity, cf. [46, 45, 47, 48, 110, 115, 118, 139, 155]. A closer look at (1.51) reveals that the energy provides a uniform control over

2 2 2 Chard sym P, sym P + Curl P c sym P + Curl P (1.53) k k ≥ k k k k where Chard is assumed to be positive definite. That the right-hand side of (1.53) provides a norm on 3 3 smooth, compactly supported functions P C∞(Ω, R × ) was first noted in [118]. Indeed, sym P 0 implies ∈ 0 ≡ P = A C0∞(Ω, so(3)), so that using Nye’s formula (1.26)2 we deduce from Curl A 0 that Daxl A 0. Hence, axl∈ A is a constant vector field, P = A is a constant skew-symmetric matrix≡ field, and with≡ the boundary condition we obtain P 0. Thus, controlling the plastic strain sym P in L2 and the dislocation density tensor Curl P L2, together≡ with suitable tangential boundary conditions5 P ν = 0 on ∂Ω one ∈ × 4 Laplace component-wise and observe that ∆εp ∈ Sym(3) for εp ∈ Sym(3). 5In the context of gradient plasticity, the boundary conditions P × ν = 0 postulates “no flux of the Burgers vector across the boundary surface” and is referred to as “micro-hard”, cf. [75].

10 controls the full plastic distortion P L2. The result led to a sequel of well-posedness results in gradient plasticity with plastic spin in H(Curl),∈ cf. [46, 45, 47, 48, 118]. However, choosing a simple quadratic energy in Curl P in (1.51) is, for many situations, not suitable. A major scientific question is therefore, how to replace Curl P 2 in order to capture important physical features. Let us write (Curl P ) for this extension. Experimentalk k evidence suggests to use with sub- quadratic growth and whereG the behavior at zero can be differentiable or not. Indeed, in [155] itG is argued to consider the one-homogeneous expression (Curl P ) = Curl P or (Curl P ) = Curl P log Curl P , see also [133, 35, 131]. G k k G k k · k k It is furthermore possible to extract some geometrical information from the dislocation density tensor (on the mesoscale). The indices i and j of (Curl P )ij determine the orientation of the Burger’s vector and the dislocation line, respectively. The diagonal components of Curl P describe screw dislocations and the off- diagonal components describe edge dislocations. For an overview on dislocations in the framework of different types of generalized continua we refer the reader to [95,1, 97] and the references therein. Lazar [94, 96], see also [120], has used the decomposition of the dislocation density tensor into SO(3)-irreducible pieces 1 Curl P = dev sym Curl P + skew Curl P + tr(Curl P ) 1 (1.54) | {z } | {z } 3 · “tentor” “trator” | {z } “axitor” i.e., the orthogonal decomposition (1.34). So, the axitor corresponds to the sum of all possible screw dislocations, the trator to “skew-symmetric” edge dislocations and the tentor describes a combination of “symmetric” edge-dislocations and single screw-dislocations, cf. [120]. In addition, for compatible X = Du the decomposition (1.34) reads 1 Du = dev sym Du + skew Du + tr(Du) 1 . (1.55) 3 · | {z } 1| {z } | {z } shear = 2 Anti(curl u) = div u anti-conformal part, rotation volumetric part shape-change | {z } conformal part, no shape change

The introduced nomenclature coming from the fact that dev sym Du = 0 implies that u = ϕC is an infinites- imal conformal mapping, see (1.10).6

(R 1 + so(3))⊥ = sl(3) Sym(3), “anti-conformal” · ∩

dev3 sym Du = 0 Du R 1 + so(3) u = φC ⇔ ∈ · ⇔ R 1 + so(3), “conformal” ·

Figure 1: Orthogonal decomposition and infinitesimal conformal mappings, see (1.10).

For improved transparency in the physical modeling, we may now assume an additively decomposed ansatz

6Note, that besides the divergence and the curl of a vector field also the term dev sym Du has a physical interpretation, namely as the shear, since for “a cube of moving fluid, the shear [of the velocity of that fluid] represents the rate at which each side is deviating from a square, and the nature of that deviation”, cf. [141] where the authors also make use of the “natural” decomposition (1.55), cf. [141, eq. (6)].

11 for : G 1 (Curl P ) = (dev sym Curl P + skew Curl P + tr(Curl P ) 1) G G 3 · = 1(dev sym Curl P ) + 2(skew Curl P ) + 3(tr(Curl P ) 1) (1.56) G G G α · = α dev sym Curl P q1 + α skew Curl P q2 + 3 tr(Curl P ) q3 1k k 2k k 3 | | where in the last step we considered a simple isotropic example. Our novel result (1.20) shows that under the conditions α > 0, α = α 0, q > 1 one can control the plastic distortion P in Lq1 . 1 2 3 ≥ 1 1.5.4 Incompatible linear elasticity Instead of the classical linear elasticity problem Z 1 3 7 Ce sym Du, sym Du + Div G, u dx min u H (Ω, R ), (1.57) Ω | {z } → ∈ f

that is 1 3 Div Ce sym Du = Div G, u Γ = u H (Ω, R ), (1.58) | e ∈ we may consider the corresponding incompatible nonlinear elasticity problem Z r r 1, 2, r 3 3 Ce sym e, sym e G, e + µ Lc dev sym Curl e dx min e W (dev sym Curl; Ω, R × ), Ω − k k → ∈ (1.59a) in other words the strong form of the second order Euler-Lagrange equations formally reads   r dev sym Curl e 8 Ce sym e + r µ L Curl = G, tr(Ce sym e) = tr(G), (1.59b) c dev sym Curl e r 2 k k − under the (consistent) symmetrized tangential boundary condition

dev sym(e ν) Γ = dev sym(Du ν) (1.59c) × | e × 1 3 1 3 3 where 1 < r 2, ue H (Ω, R ) and G H0 (Ω, R × ) are prescribed and ν is the outward unit vector field to ∂≤Ω. According∈ to our Theorem∈ 3.5, the solution to (1.59) is unique with respect to the 1, 2, r 3 3 6 non-symmetric elastic distortion e W (dev sym Curl; Ω, R × ). Note that for 5 r 2 it holds 1, 2, r 3 3 1∈, 2 3 3 ≤ ≤ W (dev sym Curl; Ω, R × ) % W (Ω, R × ). The formulation (1.59) might therefore be useful in prob- lems with fracture. Furthermore, the stored energy in (1.59) is always bounded above by the corresponding energy in (1.57) for the compatible case. Replacing e = Du and taking the divergence, recovers (1.58)1. In the same spirit, in [10] the authors have considered the non-variational second-order problem

2 Ce sym e + 2µLc inc (sym e) = G, inc (sym e) ν = 0 on ∂Ω, (1.60) which also looks for a “relaxation” of linear elasticity and determines a unique solution sym e L2(Ω), inc (sym e) L2(Ω), where inc P := Curl[(Curl P )T ] and further properties of the inc operator∈ will be ∈ discussed below. Replacing sym e = sym Du and taking the divergence would also recover (1.58)1.

7 T div(G u) = Div G, u 3 + G, Du 3×3 , to fix notation. R R 8Since tr(Curl S) = 0 for S ∈ Sym(3).

12 2 Notations and technical preliminaries

n Pn Let n 2. For vectors a, b R we consider the scalar product a, b := i=1 ai bi R, the (squared) ≥ 2 ∈ n n ∈ norm a := a, a and the dyadic product a b := (ai bj)i,j=1,...,n R × . Similarly, the scalar product for k k n n P⊗n ∈ 2 matrices P,Q R × is given by P,Q := i,j=1 Pij Qij R and the (squared) Frobenius-norm by P := ∈ T ∈ k k P,P . Moreover, P := (Pji)i,j=1,...,n stands for the transposition of the matrix P = (Pij)i,j=1,...,n. We 1 T  make use of the orthogonal decomposition of the latter into the symmetric part sym P := 2 P + P and the 1 T  n n skew-symmetric part skew P := 2 P P . We denote by sl(n) := X R × tr(X) = 0 the Lie-algebra − n n T { ∈ | n n} T of trace-free matrices, with Sym(n) := X R × X = X and by so(n) := A R × A = A the { ∈ | } { ∈ | − } Lie-Algebra of skew-symmetric matrices. For the identity matrix we write 1, so that the trace of a squared 1 matrix P is tr P := P, 1 . The deviatoric (trace-free) part of P is given by devn P := P tr(P ) 1 and in − n three dimensions we will suppress its index, i.e., we write dev instead of dev3. n k, p By D 0(Ω) we denote the space of distributions on a bounded Lipschitz domain Ω R and by W − (Ω) k, p0 p ⊂ the dual space of W0 (Ω), where p0 = p 1 is the dual H¨olderexponent to p. − 2.1 The three-dimensional case

In R3 we further make use of the vector product : R3 R3 R3. For a fixed vector a R3 the cross product a . is linear in the second component, so× that there× exists→ a unique matrix Anti(a)∈ such that × 3 a b =: Anti(a) b b R . (2.1) × ∀ ∈ T Direct computations show that for a = (a1, a2, a3) the matrix Anti(a) is of the form

 0 a a  − 3 2 Anti(a) =  a3 0 a1 , (2.2) a a −0 − 2 1 so that with Anti : R3 so(3) we have a canonical identification of R3 with the vector space of skew- symmetric matrices so(3).→ This algebraic approach to the cross product facilitates some of the traditional proofs of vector algebra, cf. [153, 142, 107, 72]. Indeed, also the notations Ta, W (a) or even [a] are used for Anti(a), but the latter emphasizes that we deal with a skew-symmetric matrix. Furthermore,× the vector product can be written as

T T 3 a b = Anti(a) b = b a = Anti(b) a = (a Anti(b)) a, b R . (2.3) × − × − ∀ ∈ The inverse of Anti will be called axl : so(3) R3 and it associates to a skew-symmetric matrix A so(3) the (axial) vector axl A := ( A ,A , A →)T , so that ∈ − 23 13 − 12 3 A b = axl(A) b b R . (2.4) × ∀ ∈ The identification of the vector product with a suitable matrix product allows us to generalize the vector 3 3 3 3 product in R to a vector product of a vector b R and a matrix P R × from the left and from the right: ∈ ∈ b P := Anti(b) P and P b := P Anti(b), (2.5) × × Thus, b P is given by a column-wise vector multiplication ×   b P = Anti(b) P e1 P e2 P e3 = Anti(b) P e1 Anti(b) P e2 Anti(b) P e3 × | | | | (2.6) = b (P e ) b (P e ) b (P e ), × 1 | × 2 | × 3

13 whereas in P b we have a row-wise vector multiplication ×  T T   T T   T T  (P e1) (P e1) Anti(b) ((P e1) b) T T T T (2.3) T × T P b = (P e2)  Anti(b) = (P e2) Anti(b) = ((P e3) b)  . (2.7) × (P T e )T (P T e )T Anti(b) ((P T e ) × b)T 3 3 3 × For the identity matrix we obtain

3 1 b = 1 Anti(b) = Anti(b) b R . (2.8) × ∀ ∈ Formally, Nye’s formula (1.26) is a consequence of the following algebraic identity9:

3 Anti(a) Anti(b) = (Anti(a)) b = b a b, a 1 = b a tr(b a) 1 a, b R (2.9) × ⊗ − ⊗ − ⊗ ∀ ∈ and the second identity (1.26)1 comes from the converse expression

1 b a = (Anti(a)) b + b, a 1 = (Anti(a)) b tr((Anti(a)) b) 1 (2.10) ⊗ × × − 2 × where we have used (2.2) tr((Anti(a)) b)) = tr(Anti(a) Anti(b)) = Anti(a) Anti(b), 1 = Anti(a), Anti(b) = 2 a, b . (2.11) × − − In addition, for all b R3 we obtain ∈ (2.5) (2.9) Anti(b) Anti(b) = Anti(b) b = b b b 21 , (2.12) × ⊗ − k k so that (2.12) 2  2 Anti(b) Anti(b) Anti(b) = b b b 1 b = b Anti(b) . (2.13) (2.5) ⊗ − k k × −k k

Consequently, for a symmetric matrix S we have tr(S b) = 0 for any b R3, since × ∈ (2.5) T S Sym(3) tr(S b) = S b, 1 = S Anti(b), 1 = Anti(b),S ∈ = 0 , (2.14) × × and similarly

T T (2.5) tr((S b) b) b) = (S b) b) b, 1 = Anti(b) Anti(b) S Anti(b), 1 × × × × × × (2.13) S Sym(3) = S, Anti(b) Anti(b) Anti(b) = b 2 S, Anti(b) ∈ = 0 . (2.15) − k k Furthermore, we can consider the vector multiplication on both sides:

b P b = Anti(b) P Anti(b) . (2.16) × × However, from the viewpoint of application it is more convenient to look at

(P b)T b = (P Anti(b))T Anti(b) = Anti(b) P T Anti(b) = b P T b. (2.17) × × − − × × In particular, for a skew-symmetric matrix A so(3) and a symmetric matrix S Sym(3) we have ∈ ∈ (A b)T b = b A b and (S b)T b = b S b . (2.18) × × × × × × − × × Observation 2.1. For a, b R3 we have ∈ 1 2 a 2 b 2 dev sym(Anti(a) b) 2 a 2 b 2. (2.19) 2k k k k ≤ k × k ≤ 3k k k k 9This algebraic relation is already contained in [142, p. 691 (ii)].

14 Proof. Considering the dev sym parts on both sides of (2.9) we obtain

(2.9) 1 1 dev sym(Anti(a) b) = dev sym(b a) = sym(a b) tr(b a) 1 = sym(a b) a, b 1. (2.20) × ⊗ ⊗ − 3 ⊗ · ⊗ − 3 · Since,

1 1 1 1 1 2 sym(a b) 2 = a b + b a 2 = a b 2 + a b, b a = a 2 b 2 + a, b , (2.21) k ⊗ k 4k ⊗ ⊗ k 2k ⊗ k 2 ⊗ ⊗ 2k k k k 2 taking the squared norm on both sides of (2.20) we obtain

2 (2.20) 2 1 2 2 2 dev sym(Anti(a) b) = sym(a b) + a, b 1 a, b sym(a b), 1 (2.22) k × k k ⊗ k 9 k k − 3 ⊗ (2.21) 1 1 2 1 2 2 2 1 1 2 = a 2 b 2 + a, b + a, b a, b = a 2 b 2 + a, b . 2k k k k 2 3 − 3 2k k k k 6 The right hand side is bounded from above by 2 a 2 b 2 and from below by 1 a 2 b 2. These bounds are 3 k k k k 2 k k k k sharp if a is parallel to b and if a is perpendicular to b, respectively.  Remark 2.2. Due to the identification of skew-symmetric matrices with vectors in R3 the relation (2.19) reads also 1 1 A 2 Ae 2 dev sym(A Ae) 2 A 2 Ae 2 A, Ae so(3). (2.23) 8k k k k ≤ k k ≤ 6k k k k ∀ ∈ Indeed, setting a := axl A and a := axl Ae, we have A Ae = Anti(a) a, so that the estimate follows from (2.19) e ×e in combination with the identities A 2 = 2 a 2 and Ae 2 = 2 a 2. The bounds in (2.23) are sharp. The k k k k k k kek upper bound is achieved for A = Ae and the lower bound is achieved, e.g., for A = Anti(e1) = e3 e2 e2 e3 ⊗ − ⊗ and Ae = Anti(e2) = e1 e3 e3 e1. ⊗ − ⊗ 3 3 3 In [102] we used moreover for P R × and b R the relation ∈ ∈ dev(P b) = 0 P b = 0. (2.24) × ⇔ × Here, we use a similar equivalence.

3 3 3 Observation 2.3. For P R × and b R we have ∈ ∈ dev sym(P b) = 0 sym(P b) = 0 . (2.25) × ⇔ × Remark 2.4. Surely, (2.25) is not equivalent to the condition P b = 0, cf. the example in (2.8). × Proof of Observation 2.3. We decompose P into its symmetric and skew-symmetric part, i.e.,

3 P = S + Anti(a), for some S Sym(3), a R . ∈ ∈ and obtain

1 (2.14) 1 dev sym(P b) = sym(P b) tr(P b) 1 = sym(P b) tr(Anti(a) b) 1 × × − 3 × × − 3 × (2.11) 2 = sym(P b) + a, b 1. (2.26) × 3

3 3 Moreover, for any matrix P R × it holds ∈ (2.5) (2.3) (P b) b = P Anti(b) b = P (b b) = 0 . (2.27) × ×

15 Thus, we obtain

(2.26) 2  (2.27) 2 2 b, dev sym(P b) b = b, sym(P b) + a, b 1 b = a, b b , (2.28) × × 3 3 k k and the statement follows from the identity

2 (2.26) 2 2 2 b sym(P b) = b dev sym(P b) b a, b 1 k k × k k × − 3k k (2.28) 2 = b dev sym(P b) b, dev sym(P b) b 1 . (2.29) k k × − × Applying the Cauchy-Bunyakovsky-Schwarz inequality on the right hand side of (2.29) we obtain

(2.29)   dev sym(P b) sym(P b) 1 + √3 dev sym(P b) . (2.30) k × k ≤ k × k ≤ k × k 

2.2 Considerations from vector calculus The vector differential operator behaves algebraically like a vector, so that, formally, the derivative, the ∇ 3 divergence and the curl of a vector field a D 0(Ω, R ) can be expressed as ∈ T Da = a = ( a) , div a = a, = Da, 1 = tr(Da) and curl a = a ( ) = a. (2.31) ⊗ ∇ ∇ ⊗ ∇ × −∇ ∇ × Formally, the Laplace operator behaves like a scalar with ∆ = 2. More generally, we can use multilinear expressions to definek∇k differential operators as follows. Let V and W be finite-dimensional vectorspaces and let Lin(V,W ) denote the space of linear maps from V to W . Let d d M : R ... R Lin(V,W ) be a multilinear map and denote by Mi1...ir := M(ei1 , . . . , eir ) the coeffiencts × × → d of M with respect to the standard basis e1, . . . , ed of R . We define a differential operator DM by X DM = Mi1...ir ∂i1 . . . ∂ir

i1,...,ir

d where each index ij runs from 1 to d. Let Ω R be open. Then DM maps a distribution f D 0(Ω,V ) to ∈ ∈ a distribution DM f D 0(Ω,W ). The following simple observation allows us to transfer algebraic identities into identities of vector∈ calculus. We have

d b R M(b, . . . , b) = 0 f D 0(Ω,V ) DM f = 0. (2.32) ∀ ∈ ⇐⇒ ∀ ∈

Indeed, since ∂i∂j = ∂j∂i in the sense of distributions, both assertions are equivalent to the statement that the symmetrized coefficients of M vanish. For example, the algebraic identity tr((a b) b) = 0 for all 3 3 × ⊗ a, b R translates into the identity div curl f = 0 for all f D 0(Ω, R ). Since M is multilinear we often use the∈ notation ∈ M( ,..., ) := D (2.33) ∇ ∇ M With this notation, (2.32) asserts that we can formally compute as if was a vector in Rd. Of special interest is the operator curl and its row-wise extension to∇ a matrix-valued operator Curl. Thus, formally, Curl P := P ( ) = P Anti( ) (2.34) × −∇ − ∇ 3 3 for P D 0(Ω, R × ) where the vector product acts row-wise, cf. (2.7). Surely, P or P would also be interesting∈ candidates to consider, but, among them, only the matrix Curl from∇ × (2.34)−∇ kills × the derivative 3 of a general vector field a D 0(Ω, R ), i.e., Curl Da 0. For symmetric tensor fields S D 0(Ω, Sym(3)) we obtain by (2.14) ∈ ≡ ∈ tr(Curl S) 0 . (2.35) ≡

16 3 Moreover, it holds for ζ D 0(Ω, R) and a D 0(Ω, R ) ∈ ∈ Curl(ζ 1) = Anti( ζ) and Curl Anti(a) = div a 1 (Da)T . (2.36) · − ∇ · − Note in passing that in three dimensions the matrix Curl returns again a square matrix. Furthermore, we make use of the incompatibility operator

(2.17) inc P := Curl([Curl P ]T ) = (P )T = P T = Anti( ) P T Anti( ). (2.37) × ∇ × ∇ −∇ × × ∇ − ∇ ∇ The last expression shows, in particular, that the incompatibility operator preserves symmetry:

(inc P )T = inc (P T ), inc sym P = sym inc P and inc skew P = skew inc P. (2.38)

Moreover, the incompatibility operator annihilates the symmetric displacement gradient since 1 inc (sym Da) = sym Da = ( a + a ) −∇ × × ∇ −2∇ × ∇ ⊗ ⊗ ∇ × ∇ 1 = [( ) a + a ( )] 0 . (2.39) −2 |∇{z × ∇} ⊗ × ∇ ∇ × ⊗ |∇{z × ∇} ≡ =0 =0 Note that this formal calculation was already carried out in Lagally’s monograph on vector calculus from 1928 [91, Ziff. 191]. The action of the incompatibility operator on spherical tensors and antisymmetric tensors is given by

inc (ζ 1) = ∆ζ 1 D2ζ Sym(3) and inc (Anti(a)) = Anti( div a) so(3) , (2.40) · · − ∈ − ∇ ∈ respectively. For symmetric tensor fields S we obtain, formally by (2.15), again

tr(inc Curl S) 0. (2.41) ≡ Remark 2.5. The incompatibility operator inc occurs in infinitesimal strain dislocation models, e.g., in the modeling of dislocated crystals or in the modeling of elastic materials with dislocations, since the strain cannot be a symmetrized gradient of a vector field as soon as dislocations are present and the notion of incompatibility is at the basis of a new paradigm to describe the inelastic effects, cf. [47,8, 95,9,7, 109]. Furthermore, the equation inc sym e 0 is equivalent to the Saint-Venant compatibility condition(s)10 defining the relation between the displacement≡ vector field u and the symmetric strain sym e, more precisely:

inc sym e 0 sym e = sym Du (2.42) ≡ ⇔ in simply connected domains, cf. [4, 109]. For investigations over multiply connected domains see e.g. [152, 65].

2.3 Linear combinations of higher derivatives Our analysis relies on a number of apparently hitherto unnoticed identities which arise from the interaction of the matrix Curl operator with the algebraic splitting (1.34): 1 X = dev sym X + skew X + tr(X) 1. 3 · In particular, we have the following identities.

10Those compatibility conditions can be found in the third appendix §32 p. 597 et seq. of the third edition of the lecture notes R´esistance des corps solides given by Navier and extended with several notes and appendices by Barr´ede Saint-Venant and published as R´esum´edes Le¸consdonn´ees`al’Ecole´ des Ponts et Chauss´ees sur l’Application de la M´ecanique, vol. I, Paris, 1864. Their coordinate-free version are contained in Lagally’s monograph on vector calculus from 1928 [91, Ziff. 191].

17 Lemma 2.6. Let A D 0(Ω, so(3)). Then ∈ (a) the entries of D2A are linear combinations of the entries of D sym Curl A. (b) the entries of D3A are linear combinations of the entries of D2 dev sym Curl A.

Proof. By Nye’s formula (1.26)1 we have sym Curl A = tr(Daxl A) 1 sym(Daxl A). (2.43) − Taking the trace on both sides we obtain tr(sym Curl A) = 2 tr(Daxl A) and inserting this identity into (2.43) we get 1 sym(Daxl A) = tr(sym Curl A)1 sym Curl A. (2.44) 2 − Moreover, by the relation (1.23) used for the proof of the classical Korn’s inequality, we obtain

2 (1.23) (2.44) D axl A = L(D sym Daxl A) = L1(D sym Curl A). (2.45) In other words, the entries of D2A are linear combinations of the entries of Dsym Curl A which establishes part (a). To prove (b) we make use of the incompatibility operator inc , since it kills the symmetric displacement gradient, cf. (2.39). Consider now the deviatoric part on both sides of (2.43): 1 dev sym Curl A = tr(Daxl A) 1 sym(Daxl A) . (2.46) 3 − Applying inc on both sides, we obtain in view of (2.39) and (2.40):

3 inc dev sym Curl A = ∆ tr(Daxl A) 1 D2 tr(Daxl A) (2.47) · − or, equivalently,

2 3 2 D tr(Daxl A) = tr(inc dev sym Curl A) 1 3 inc dev sym Curl A = L2(D dev sym Curl A), (2.48) 2 · − where we have used that the entries of inc B are, of course, linear combinations of the entries of D2B, so that by (2.46) we have 2 2 D sym(Daxl A) = L3(D dev sym Curl A). (2.49) The conclusion of part (b) then follows using the relation (1.23):

3 (1.23) 2 (2.49) 2 D axl A = L(D sym Daxl A) = L4(D dev sym Curl A).  The algebraic considerations above provide information on higher derivatives of P in negative Sobolev spaces. To obtain Lp-estimates for P we use the following deep result. Theorem 2.7 (Lions lemma and Neˇcasestimate). Let Ω Rn be a bounded Lipschitz domain. Let m Z d m 1, p ⊂ d n m, p d ∈ and p (1, ). Then f D 0(Ω, R ) and Df W − (Ω, R × ) imply f W (Ω, R ). Moreover, ∈ ∞ ∈ ∈ ∈  f m, p d c f m−1, p d + Df m−1, p d×n , (2.50) k kW (Ω,R ) ≤ k kW (Ω,R ) k kW (Ω,R ) with a constant c = c(m, p, n, d, Ω) > 0. For a proof we refer to [6, Proposition 2.10 and Theorem 2.3] and [15]. However, for our discussions the heart of the matter is the estimate (2.50), see Neˇcas[128, Th´eor`eme1]. The case m = 0 is already contained in [24]; for an alternative proof, see [111, Lemma 11.4.1] and [16, Chapter IV] as well as [17] and [39]. For further historical remarks, see the discussions in [28,5] and the references contained therein. Since we only have information on higher order derivatives of P we will use the following consequence of Theorem 2.7.

18 Corollary 2.8. Let Ω Rn be a bounded Lipschitz domain, m Z and p (1, ). Denote by Dkf the ⊂ ∈ d ∈ k∞ m k, p d nk collection of all distributional derivatives of order k. Then f D 0(Ω, R ) and D f W − (Ω, R × ) ∈ ∈ imply f W m, p(Ω, Rd). Moreover, ∈  k  f m, p d c f m−1, p d + D f − × k , (2.51) k kW (Ω,R ) ≤ k kW (Ω,R ) k kW m k, p(Ω,Rd n ) with a constant c = c(m, p, n, d, Ω) > 0.

Proof. The assertion f W m, p(Ω, Rd) and the estimate (2.51) follow by inductive application of Theorem 2.7 to Dlf with l = k ∈1, k 2,..., 0. − −  Lemma 2.9. Let A Lp(Ω, so(3)). Then ∈ (a) sym Curl A 0 in the distributional sense if and only if A = Anti(Ae x + b) almost everywhere in Ω, ≡ 1 2 (b) dev sym Curl A 0 in the distributional sense if and only if A = Anti Ae x+b+β x+ d, x x 2 d x almost everywhere≡ in Ω − k k with constant Ae so(3), b, d R3 and β, γ R. ∈ ∈ ∈ Remark 2.10. It is seen already from the calculations of the kernels that there can not be corresponding Korn type inequalities in terms of dev sym P + sym Curl P or dev sym P + dev sym Curl P . The kernels would be infinite-dimensional,k since allk restrictingk informationk k on ζ wouldk getk lost. Indeed, wek have

(1.26) sym Curl(A + ζ 1) = tr(Daxl A)1 sym(Daxl A) · − so that sym Curl(A + ζ 1) 0 or dev sym Curl(A + ζ 1) 0 allow ζ to be arbitrary. · ≡ · ≡ Remark 2.11. Solutions of Lemma 2.9 have already been partially indicated in the literature, cf. [137, 11]. We include their full deduction here for the convenience of the reader.

Proof of Lemma 2.9. The “if”-parts follow from a direct calculation using Nye’s formula (1.26):

(a) Curl(Anti(Ae x + b)) = Ae,

1 2 (b) D(Anti Ae x+b+β x+ d, x x 2 d x ) = Ae+β 1+ d, x 1+x d d x = (β+ d, x )1+Ae+Anti(d x), − k k 1  ⊗ − ⊗ × hence, Curl(Anti Ae x + b + β x + d, x x d x 2 ) = 2(β + d, x )1 + Ae + Anti(d x), − 2 k k × Now, we will focus on the “only if”-directions. By (2.44) the condition sym Curl A 0 implies sym(Daxl A) 0, so that the usual calculation for Korn’s inequality, cf. (1.23), gives that Daxl≡ A must be a constant skew-symmetric≡ matrix. Thus,

A = Anti(Ae x + b) for some Ae so(3) and b R3, which establishes (a). ∈ ∈ Considering now dev sym Curl A 0 we obtain by (2.48) that D2 tr(Daxl A) 0. Hence, ≡ ≡ 1 tr(Daxl A) = β + d, x (2.52) 3 for some d R3 and β R. Definea ¯ by ∈ ∈ 1 a¯(x) = β x + d, x x d x 2. (2.53) − 2 k k

19 Then

D¯a = β 1 + d, x 1 + x d d x = (β + d, x ) 1 + Anti(d x) ⊗ − ⊗ × and

sym D¯a = (β + d, x ) 1. (2.54) Thus, by (2.46) we have

(2.54) (2.46) 1 (2.52) sym(D(axl A a¯)) = sym(Daxl A) (β + d, x ) 1 = tr(Daxl A)1 (β + d, x ) 1 = 0 . − − 3 − Again, (1.23) gives that D(axl A a¯) must be a constant skew-symmetric matrix and we have − axl A = Axe + b +a, ¯ for some Ae so(3) and b R3, and statement (b) follows from the representation (2.53). ∈ ∈  Remark 2.12. The conclusion of (b) also follows directly from Nye’s formula and is connected to infinitesimal conformal maps. Indeed, we have

(1.26) dev sym Curl Anti(a) = 1 dev sym Da, (2.55) − so that dev sym Curl Anti(a) 0 dev sym Da 0 a = ϕ , (2.56) ≡ ⇔ ≡ ⇔ C denoting by ϕC an infinitesimal conformal map, so that the expression in (b) of Lemma 2.9 follows from the expression for infinitesimal conformal maps (1.10).

3 New incompatible Korn type inequalities

3 3 3 Lemma 3.1. Let Ω R be a bounded Lipschitz domain, 1 < p < and P D 0(Ω, R × ). Then either of the conditions ⊂ ∞ ∈

p 3 3 1, p 3 3 (a) sym P L (Ω, R × ) and sym Curl P W − (Ω, R × ), ∈ ∈ p 3 3 1, p 3 3 (b) sym P L (Ω, R × ) and dev sym Curl P W − (Ω, R × ), ∈ ∈ p 3 3 implies P L (Ω, R × ). Moreover, we have the estimates ∈  P p 3×3 c skew P −1, p 3×3 k kL (Ω,R ) ≤ k kW (Ω,R )  + sym P p 3×3 + sym Curl P −1, p 3×3 , (3.1a) k kL (Ω,R ) k kW (Ω,R )  P p 3×3 c skew P −1, p 3×3 k kL (Ω,R ) ≤ k kW (Ω,R )  + sym P p 3×3 + dev sym Curl P −1, p 3×3 , (3.1b) k kL (Ω,R ) k kW (Ω,R ) always with a constant c = c(p, Ω) > 0. Remark 3.2. Clearly, condition (b) is weaker than condition (a) and (3.1b) implies (3.1a). Furthermore, (3.1b) implies the estimate  P p 3×3 c skew P −1, p 3×3 k kL (Ω,R ) ≤ k kW (Ω,R )  (3.2) + sym P p 3×3 + Curl P −1, p 3×3 k kL (Ω,R ) k kW (Ω,R )

20 in [105, Lemma 3.1] as well as the estimate  P p 3×3 c skew P −1, p 3×3 k kL (Ω,R ) ≤ k kW (Ω,R )  (3.3) + sym P p 3×3 + dev Curl P −1, p 3×3 k kL (Ω,R ) k kW (Ω,R ) in [102, Lemma 3.6], but not the estimate

 1 P p 3×3 c skew P + tr P 1 −1, p 3×3 k kL (Ω,R ) ≤ k 3 · kW (Ω,R )  (3.4) + dev sym P p 3×3 + dev Curl P −1, p 3×3 k kL (Ω,R ) k kW (Ω,R ) in [102, Lemma 3.6] which uses dev sym P rather than sym P on the right hand side. The point is, that we cannot improve (3.1) to an estimate which involves dev sym P instead of sym P on the right hand side, cf. Remark 1.2.

p 3 3 Proof of Lemma 3.1. By the previous remark it suffices to establish the assertion P L (Ω; R × ) under condition (b) and to prove the estimate (3.1b). ∈ We will follow the same line of reasoning as in the proof of [105, Lemma 3.1] and start by considering the orthogonal decomposition

P = sym P + skew P.

p 3 3 To deduce skew P L (Ω, R × ) under assumption (b) we consider ∈ 2 D dev sym Curl skew P −3, p 3×33 c dev sym Curl(P sym P ) −1, p 3×3 k kW (Ω,R ) ≤ k − kW (Ω,R ) c ( dev sym Curl P −1, p 3×3 + Curl sym P −1, p 3×3 ) ≤ k kW (Ω,R ) k kW (Ω,R ) c ( dev sym Curl P −1, p 3×3 + sym P p 3×3 ). (3.5) ≤ k kW (Ω,R ) k kL (Ω,R ) 2 3, p 3 33 Hence, D dev sym Curl skew P W − (Ω, R × ) and it follows from Lemma 2.6 (b) that ∈ 3 3, p 3 34 D skew P W − (Ω, R × ). (3.6) ∈ p 3 3 Now, we apply Corollary 2.8 to skew P and we deduce that skew P L (Ω, R × ) and ∈ 3 skew P p 3×3 c ( skew P −1, p 3×3 + D skew P −3 3×34 ) k kL (Ω,R ) ≤ k kW (Ω,R ) k kW , p(Ω,R ) Lem. 2.6 (b) 2 c ( skew P −1, p 3×3 + D dev sym Curl skew P −3, p 3×33 ≤ k kW (Ω,R ) k kW (Ω,R ) (3.5) c ( skew P −1, p 3×3 (3.7) ≤ k kW (Ω,R ) + sym P p 3×3 + dev sym Curl P −1, p 3×3 ). k kL (Ω,R ) k kW (Ω,R )  The rigidity results follow by eliminating the corresponding first term on the right-hand side of (3.1).

Theorem 3.3. Let Ω R3 be a bounded Lipschitz domain and 1 < p < . There exists a constant ⊂ p 3 3 ∞ c = c(p, Ω) > 0 such that for all P L (Ω, R × ) ∈ (a)  3×3 3×3 −1 3×3 inf P T Lp(Ω,R ) c sym P Lp(Ω,R ) + sym Curl P W , p(Ω,R ) (3.8a) T KS,SC k − k ≤ k k k k ∈ (b)  3×3 3×3 −1 3×3 inf P T Lp(Ω,R ) c sym P Lp(Ω,R ) + dev sym Curl P W , p(Ω,R ) (3.8b) T KS,dSC k − k ≤ k k k k ∈

21 where the kernels are given by

3 3 3 KS,SC = T :Ω R × T (x) = Anti(Ae x + b), Ae so(3), b R , (3.9a) { → | ∈ ∈ } 3 3 1 2 KS,dSC = T :Ω R × T (x) = Anti Ae x + β x + b + d, x x d x , { → | − 2 k k 3 Ae so(3), b, d R , β R , (3.9b) ∈ ∈ ∈ }

Remark 3.4. Setting ea = axl(Ae) we have for the linear functions in the kernels Anti(Ae x) = Anti(Anti(a) x) = x a a x = 2 skew(x a) (3.10) e ⊗ e − e ⊗ ⊗ e so that KS,SC can be alternatively written as

3 3 3 KS,SC = T :Ω R × T (x) = skew(x a) + Anti(b), a, b R . (3.9a’) { → | ⊗ e e ∈ }

Furthermore, the elements of KS,dSC are connected to infinitesimal conformal mappings ϕC via

3 3 KS,SC = T :Ω R × T (x) = Anti(ϕC (x)), with dev sym DϕC 0 (3.9b’) { → | ≡ } cf. Remark 2.12.

Proof of Theorem 3.3. We first prove the formulae for the kernels KS,Sc and KS,dSC . If

p 3 3 P KS,SC := P L (Ω, R × ) sym P = 0 a.e. and sym Curl P = 0 in the dist. sense , (3.11) ∈ { ∈ | } then P = skew P and sym Curl skew P = 0. Thus (3.9a) follows by virtue of Lemma 2.9 (a). Similarly, the formula (3.9b) follows from Lemma 2.9 (b). The estimates (3.8a) and (3.8b) now follow from Lemma 3.1, the fact that the kernels are finite-dimensional p 1,p and the compactness of the embedding L (Ω) , W − (Ω), see, for example, the proofs [105, 104, 102] or [29, Theorem 6.15-3] for similar reasoning. For→ the convenience of the reader we provide the details for the argument for the estimate (3.8a). The proof of (3.8b) is analogous. By e1, . . . , eM we denote a basis of KS,SC , and by `1, . . . , `M we denote the corresponding dual basis of linear functionals on KS,SC which is characterized by the conditions `α(ej) := δαj. (3.12) Then, the Hahn-Banach theorem in a normed vector space (see e.g. [29, Theorem 5.9-1]), allows us to extend p 3 3 `α to continuous linear forms - again denoted by `α - on the Banach space L (Ω, R × ), 1 α M. Note that ≤ ≤ T K T = 0 ` (T ) = 0 α 1,...,M . (3.13) ∀ ∈ S,SC ⇔ α ∀ ∈ { } We claim that

M ! X P p 3×3 c sym P p 3×3 + sym Curl P −1, p 3×3 + ` (P ) . (3.14) k kL (Ω,R ) ≤ k kL (Ω,R ) k kW (Ω,R ) | α | α=1

p 3 3 Indeed, if this inequality is false, there exists a sequence Pk L (Ω, R × ) with the properties ∈ M ! X 1 P p 3×3 = 1 and sym P p 3×3 + sym Curl P −1,p 3×3 + ` (P ) < . k kkL (Ω,R ) k kkL (Ω,R ) k kkW (Ω,R ) | α k | k α=1

p 3 3 Hence, (for a subsequence) Pk *P ∗ in L (Ω, R × ) and we have sym P ∗ 0 and sym Curl P ∗ 0 in ≡ ≡ the distributional sense but also `α(P ∗) = 0 for all α = 1,...,M, so that P ∗ 0. Since the embedding p 3 3 1, p 3 3 ≡ 1, p 3 3 L (Ω, R × ) , W − (Ω, R × ) is compact we get skew Pk skew P ∗ 0 in W − (Ω, R × ). Thus, →1, p 3 3 → ≡ Pk 0 in W − (Ω, R × ) and this yields to a contradiction with (3.1a). Hence (3.14) holds. →

22 p 3 3 Now consider the projection πa : L (Ω, R × ) KS,SC given by → M X πa(P ) := `j(P ) ej. (3.15) j=1

(3.12) We obtain `α(P πa(P )) = 0 for all 1 α M, so that (3.8a) follows after inserting P πa(P ) into (3.14): − ≤ ≤ −

3×3 3×3 inf P T Lp(Ω,R ) P πa(P ) Lp(Ω,R ) T KS,SC k − k ≤ k − k ∈  c sym P p 3×3 + sym Curl P −1, p 3×3 . ≤ k kL (Ω,R ) k kW (Ω,R )  Finally, we show that the estimates in Theorem 3.3 can be improved to estimates for P itself, and not just for P T , if we impose a natural boundary condition which annihilates the relevant kernels. We focus− on the improvement of (3.8b) because this already implies the improved estimate for (3.8a). For a weak definition of boundary values of certain linear combinations of P it is not sufficient to assume only 1, p 3 3 p 3 3 dev sym Curl P W − (Ω, R × ). Indeed, this condition is satisfied for every P L (Ω, R × ). We thus consider, for p ∈(1, ) and r [1, ) the spaces ∈ ∈ ∞ ∈ ∞ 1, p, r 3 3 p 3 3 r 3 3 W (dev sym Curl; Ω, R × ) := P L (Ω, R × ) dev sym Curl P L (Ω, R × ) . (3.16) { ∈ | ∈ } Equipped with the norm

P := P p 3×3 + dev sym Curl P r 3×3 (3.17) k kp,r,dSC k kL (Ω,R ) k kL (Ω,R ) this space becomes a Banach space. In terms of scaling the natural relation between p and r is p = r∗ where 3 r∗ is the Sobolev exponent of r. To properly treat the borderline case p = 1∗ = 2 we make the following assumptions 1 1 1 3 r [1, ), + , r > 1 if p = . (3.18) ∈ ∞ r ≤ p 3 2 We assume that Ω R3 is a bounded domain with Lipschitz boundary. To define boundary conditions for certain linear combinations⊂ of the components of P in the distributional sense, we first recall that for q (1, ) the space C1(Ω) is dense in W 1, q(Ω) and there exists a linear bounded and surjective trace ∈ ∞ 1, q 1 1/q, q operator Tr : W (Ω) W − (∂Ω) which is uniquely characterized by the condition Trf = f ∂Ω for all 1 → 1 1/q, q 1,| q f C (Ω). Moreover there exists a linear, bounded extension operator E : W − (∂Ω) W (Ω) with Tr∈ E = id. If follows from the divergence theorem and the density of C1(Ω) that for all i =→ 1, 2, 3 ◦ Z Z 2 1, q ∂if dx = Trf νi d f W (Ω) (3.19) Ω ∂Ω H ∀ ∈ where ν denotes the outer normal of ∂Ω (which exist 2 a.e. on ∂Ω) and 2 the two-dimensional Hausdorff H 1 H 1 measure. For p (1, ) we denote by p0 the dual exponent given by p + p0 = 1. The dual of the space 0 0 ∈ ∞ 0 1 1/p , p 1/p, p 1/p, p W − (∂Ω) = W (∂Ω) is denoted by W − (∂Ω). In order to introduce a weak definition of the boundary values of dev sym[P ν] we assume that r satisfies (3.18). Then we can define a bounded map 1, p, r 3 ×3 1/p, p 3 3 S : W (dev sym Curl; Ω; R × ) W − (∂Ω; R × ) by →

Z 0 : 1/p, p 3 3 SP,Q ∂Ω = dev sym Curl P, EQ P, Curl dev sym EQ dx Q W (∂Ω, R × ). (3.20) Ω − ∀ ∈

Here the extension operator is applied componentwise. If a C1(Ω; R3) and b W 1, q(Ω, R3) then it follows from (3.19) that ∈ ∈ Z Z curl a, b a, curl b dx = a ( ν), Tr b d 2 . (3.21) Ω − ∂Ω × − H

23 1 3 3 Using this identity, the fact that Curl acts row-wise and (2.7), one easily deduces that for P C (Ω; R × ) 1/p, p0 3 3 ∈ and Q W (∂Ω, R × ) ∈ Z 2 SP,Q ∂Ω = dev sym[P ( ν)],Q d . (3.22) ∂Ω × − H

1 3 3 Thus, for P C (Ω; R × ) we have SP = dev sym[P ( ν)]. Let Γ be a relatively open subset of ∂Ω. We say that ∈ × −

1/p, p0 0 3 3 SP = 0 in Γ if SP,Q = 0 Q (W C )(∂Ω; R × ) with Q = 0 on ∂Ω Γ. ∀ ∈ ∩ \ 0 0 Note that (W 1/p, p C0)(∂Ω) is dense in W 1/p, p (∂Ω) since it contains Tr(C1(Ω)). We define ∩ 1, p, r 3 3 1, p, r 3 3 W (dev sym Curl; Ω, R × ) := P W (dev sym Curl; Ω, R × ) SP = 0 in Γ . (3.23) 0,Γ { ∈ | } In particular

1 3 3 1, p, r 3 3 T C (Ω; R × ) W (dev sym Curl; Ω, R × ) = dev sym[T ν] = 0 on Γ. (3.24) ∈ ∩ 0,Γ ⇒ × 1, p, r 3 3 Since S is continuous, the space W0,Γ (dev sym Curl; Ω, R × ) is a closed subspace of 1, p, r 3 3 W (dev sym Curl; Ω, R × ). Theorem 3.5. Let Ω R3 be a bounded Lipschitz domain, let 1 < p < and assume that r satisfies (3.18). Let Γ ∂Ω be relatively⊂ open and non-empty. Then there exists a constant∞ c = c(p, r, Ω, Γ) such that for all ⊂1, p, r 3 3 P W (dev sym Curl; Ω, R × ) we have ∈ 0,Γ  P p 3×3 c sym P p 3×3 + dev sym Curl P r 3×3 . (3.25) k kL (Ω,R ) ≤ k kL (Ω,R ) k kL (Ω,R ) Remark 3.6. Conti and Garroni [34] and Gmeineder and Spector [69] have shown that the estimate  P p 3×3 c sym P p 3×3 + Curl P r 3×3 k kL (Ω,R ) ≤ k kL (Ω,R ) k kL (Ω,R ) 3 R holds also in the borderline case r = 1 and p = 2 under the normalization condition Ω skew P dx = 0 similar to [63]. We do not know if Theorem 3.5 holds in this borderline case. To show Theorem 3.5 we use the following simple fact which will be proved after the proof of Theorem 3.5:

Lemma 3.7. Assume that Γ R3 has the following properties ⊆ 1. Γ is not discrete; 2. Γ is not contained in a line; 3. Γ is not contained in a circle.

Let A so(3), b, d R3, β R and consider the function f : R3 R3 given by ∈ ∈ ∈ → 1 f(x) = Ax + βx + b + d, x x d x 2. h i − 2 k k Then f = 0 on Γ = A = 0, b = d = 0, β = 0. (3.26) ⇒ Proof of Theorem 3.5. We first show that

1, p, r 3 3 KS,dSC W (dev sym Curl; Ω; R × ) = 0 . (3.27) ∩ 0,Γ { }

Then the assertion will follow by a standard argument from (3.8b) and the fact that KS,dSC is finite- 1, p, r 3 3 dimensional while W0,Γ (dev sym Curl; Ω; R × ) is closed.

24 To show (3.27), let T K W 1, p, r. Then T is smooth and thus (3.24) implies that ∈ S,dSC ∩ 0,Γ dev sym[T ν] = 0 on Γ. Since the elements of KS,dSC are skew-symmetric it follows from Observation 2.1 × 3 and the formula for KS,dSC that there exist Ae so(3), b, d R and β R such that ∈ ∈ ∈ 1 2 axl T (x) = Ae x + β x + b + d, x x d x = 0 for all x Γ . − 2 k k ∈ Now Lemma 3.7 implies that all coefficients vanish and hence T 0. This concludes the proof of (3.27). ≡ 1, p, r 3 3 Assume now that (3.25) does not hold. Then there exists a sequence Pk W0,Γ (dev sym Curl; Ω; R × ) such that ∈ sym P p 3×3 + dev sym Curl P r 3×3 0 k kkL (Ω,R ) k kkL (Ω,R ) → and P p 3×3 = 1. k kkL (Ω,R ) 0 1, p r0 r The assumption (3.18) on r implies that W0 (Ω) embeds continuously into L (Ω). Hence L (Ω) embeds 1, p continuously into W − (Ω). Thus it follows from (3.8b) that there exist T K such that k ∈ S,dSC

P T p 3×3 0. k k − kkL (Ω,R ) → p In particular, the sequence Tk is bounded in L and since KS,dSC is finite-dimensional, there exists a p 3 3 T KS,dSC and a subsequence such that Tk T in L (Ω, R × ). Thus (for the same subsequence) ∈ p 3 3 → r Pk T in L (Ω, R × ). Moreover dev sym Curl Pk converges to zero in L and dev sym Curl T = 0. → 1, p, r 3 3 1, p, r 3 3 Since W0,Γ (dev sym Curl; Ω; R × ) is a closed subspace of W (dev sym Curl; Ω; R × ) it follows that 1, p, r 3 3 p 3 3 T W (dev sym Curl; Ω; R × ). Hence (3.27) implies that T = 0 and thus Pk 0 in L (Ω, R × ). This ∈ 0,Γ → contradicts the hypothesis P p = 1. k kkL  Remark 3.8. Estimate (3.25) does not hold true in other dimensions, since only in three dimensions the matrix Curl returns a square matrix.

3 Proof of Lemma 3.7. Since Γ is not discrete there existsx ¯ R and xk Γ x¯ such that limk xk =x ¯. The map g(x) := f(¯x + x) has the same form as f (with different∈ values∈ of\{ the parameters} A, b,→∞ d, β). Thus we may assume without loss of generality thatx ¯ = 0. Since f is continuous we get f(0) = 0 and hence b = 0. Since 1 0 = f(x), x = β x 2 + d, x x 2 x Γ 0 h i k k 2h ik k ∀ ∈ \{ } 1 we deduce that β + 2 d, x = 0 for all x Γ 0 . Considering points xk Γ 0 with xk 0 we see that β = 0 and h i ∈ \{ } ∈ \{ } → d, x = 0 x Γ. h i ∀ ∈ If d = 0 then f(x) = Ax. If A = 0 then the kernel of A is a line since A so(3). Thus Γ would be contained in a line which contradicts6 our assumption. Hence for d = 0 we get A =∈ 0 and we are done. If d = 0 then Γ is contained in the hyperplane perpendicular to d. Since d, f(x) = 0 for all x Γ and 6 ∈ A is skew-symmetric we get 2 A d, x + d 2 x 2 = 0 for all x Γ. This implies that k k k k ∈ 2 2 1 A d x + A d = k k x Γ. d 2 d 4 ∀ ∈ k k k k Since A is skew-symmetric, the vector A d is contained in the plane perpendicular to d. It follows that Γ is 1 either a point (and hence discrete) or a circle (x Γ and A d lie in the same hyperplane) with center 2 A d ∈ − d A d k k and radius k d 2k which contradicts our assumption.  k k It is well-known, that Korn’s inequality and Poincar´e’sinequality are not equivalent, however, due to the presence of the Curl we get back both inequalities from our general result (3.25). Indeed, in the compatible case P = Du we recover a tangential Korn inequality.

25 Corollary 3.9. Let Ω R3 be a bounded Lipschitz domain, 1 < p < and Γ a relatively open non- ⊂ ∞ empty subset in ∂Ω. There exists a constant c = c(p, Ω, Γ) > 0 such that for all u W 1, p(Ω, R3) with dev sym(Du ν) = 0 on Γ we have ∈ ×

Du p 3×3 c sym Du p 3×3 . (3.28) k kL (Ω,R ) ≤ k kL (Ω,R ) Proof. This follows from Theorem 3.5 by setting P = Du. 

Remark 3.10. This boundary condition is rather weak. If Γ is flat, then the condition dev sym(Du ν) Γ = 0 × | implies that u = α x + b along Γ with α R and b R3, see Appendix A.1.4. ∈ ∈ For skew-symmetric P = Anti(a) we recover from (3.25) a Poincar´e’sinequality involving only the devia- toric (trace-free) part of the symmetrized gradient. Such a Poincar´e-type inequality can also be generalized to functions of bounded deformation, cf. [60].

3 1,p 3 Corollary 3.11. Let Ω R be a bounded Lipschitz domain and 1 < p < . Set WΓ,0 (Ω, R ) := a 1,p 3 ⊂ ∞ 1, p { 3∈ W (Ω; R ) Tr a = 0 on Γ . There exists a constant c = c(p, Ω, Γ) > 0 such that for all a WΓ,0 (Ω, R ), we have | } ∈ a p 3 c dev sym Da p 3×3 . (3.29) k kL (Ω,R ) ≤ k kL (Ω,R ) Proof. This follows from Theorem 3.5 by setting P = Anti(a) and the following observations: dev sym(Anti(a) ν) = 0 a = 0 on Γ, Curl(Anti(a)) = L(Da) and the form of Anti(a), cf. (2.2). × ⇔  The results of Theorem 3.5 , Corollary 3.9 and Corollary 3.11 can be graphically summarized as follows.

conformally invariant incompatible Korn

1, p, r 3 3 P p c ( sym P p + dev sym Curl P r ) P W (dev sym Curl; Ω, R × ) ∥ ∥L ≤ ∥ ∥L ∥ ∥L ∀ ∈ 0,Γ

P = D u P = Anti( a)

Korn with weak boundary conditions trace-free symmetrized Poincaré 1, p R3 Du Lp c sym Du Lp u W (Ω, ) 1, p 3 ∥ ∥ ≤ ∥ ∥ ∀ ∈ a p c dev sym Da p a W (Ω, R ) with dev sym(Du ν) = 0 on Γ ∂Ω ∥ ∥L ≤ ∥ ∥L ∀ ∈ Γ,0 × ⊆

4 Comparison of the spaces W 1, p(sym Curl) and W 1, p(dev sym Curl)

Using the linear expression of the entries of DCurl P in terms of the entries of Ddev Curl P the authors of 3 3 [102] showed that for all P D 0(Ω, R × ) and all m Z one has ∈ ∈ m, p 3 3 m, p 3 3 Curl P W (Ω, R × ) dev Curl P W (Ω, R × ). (4.1) ∈ ⇔ ∈ One might, therefore, wonder whether the spaces W 1, p(dev sym Curl; Ω; R3) := W 1, p, p(dev sym Curl; Ω; R3) and W 1, p(sym Curl; Ω; R3) are actually identical, where 1, p 3 3 p 3 3 p 3 3 W (sym Curl; Ω, R × ) := P L (Ω, R × ) sym Curl P L (Ω, R × ) . { ∈ | ∈ } 1, p 3 3 1, p 3 3 We first note that clearly W (sym Curl; Ω, R × ) W (dev sym Curl; Ω, R × ) and that the natural 1, p ⊂ 3 3 norm P Lp + dev sym Curl P Lp on W (dev sym Curl; Ω, R × ) is weaker than the natural norm on k k k k

26 1, p 3 3 W (sym Curl; Ω, R × ). Thus, in view of the open mapping theorem, the two spaces are identical if and only if the two natural norms are equivalent. In view of the second estimate in (2.30) (which follows directly form (2.29) by dividing by b 2) one might expect that this is really the case. Indeed, using the reasoning in Section 2.2, we see that thek algebraick identity (2.29) shows that

∆ sym Curl P = L(D2 dev sym Curl P ) (4.2) in the sense of distributions. The identity (4.2) yields interior estimates for all compactly contained subsets Ω0 of Ω of the form

sym Curl P p 0 C(Ω0)( dev sym Curl P p + P p ) (4.3) k kL (Ω ) ≤ k kL (Ω) k kL (Ω) but we will see in the proof of Theorem 4.3 assertion4 below that this is not enough to obtain equivalence of the norm on the full set Ω because we do not impose boundary conditions on P . To illustrate the obstruction to a global estimate, let use consider the following example. Let D be the 2 2,2 2,2 1,2 unit ball in R and consider the spaces W (D) and W∆ (D) := u W (D) ∆u = 0 with norms 2 { ∈ | } u 2,2 = u 2 + Du 2 + D u 2 and u = u 2 + Du 2 + ∆u 2 , respectively. k kW k kL (Ω) k kL (Ω) k kL (Ω) k k∆ k kL (Ω) k kL (Ω) k kL (Ω) Since ∆ is an elliptic operator, we have interior estimates u W 2,2(Ω0) C(Ω0) u ∆, but the norms are not k(xk1+ik x2) ≤ k k equivalent since for the harmonic functions fk(x) := (e ) we get limk u W 2,2 / u ∆ = . The reason: while the symbol σ(ξ) = (ξ2 +ξ2) of the operator< ∆ has no non-trivial→∞ realk zeroesk (thisk k is ellipticity),∞ − 1 2 it does have the non-trivial complex zeroes ξ1 = k, ξ2 = i k. This allows us to construct the ‘bad’ functions fk. A similar analysis of the action of the matrix-valued symbols of the operators sym Curl and dev sym Curl 3 on C will allow us below to construct maps Pk which show that the norms dev sym Curl P Lp(Ω) + P Lp(Ω) k k k k and sym Curl P Lp(Ω) + P Lp(Ω) are not equivalent if Ω is a bounded domain. By contrast, one can use k k k k p 3 3 3 Fourier transform to show that the norms are equivalent for periodic P or P L (R , R × ), which we show for the convenience of the reader in Appendix A.6. For the latter purposes we∈ start with the following proposition. Proposition 4.1. Let V be a finite-dimensional vectorspace and denote by Lin(V,V ) the space of linear maps from V to V . Let A and Ae be linear maps from Rn to Lin(V,V ). Assume that

n Ae(ξ)a = 0 ξ R 0 a ker A(ξ) (4.4) ∀ ∈ \{ } ∀ ∈ and n dim ker A is constant on R 0 . (4.5) \{ } Define differential operators by

n n X X = A( ) := A(ej)∂j and e = Ae( ) := Ae(ej)∂j. A ∇ A ∇ j=1 j=1

Then for each p (1, ) there exists a constant c = c(p) such that ∈ ∞ p n ef Lp( n,V ) c f Lp( n,V ) f L (R ,V ) (4.6) kA k R ≤ kA k R ∀ ∈ and

p n ef Lp( n,V ) c f Lp( n,V ) f L (T ,V ) (4.7) kA k T ≤ kA k T ∀ ∈ in the distributional sense.

Proof. This is well-known, cf. e.g. [53, pp. 1362–1365] or [148, Section IV.3]. We recall the argument for the convenience of the reader. We focus on (4.6), the proof of (4.7) is analogous. If suffices to show (4.6)

27 n for f C∞(R ,V ). Then the general case follows by approximation. We will construct a linear bounded ∈ c operator : Lp(Rn,V ) Lp(Rn,V ) such that M → n ef = f f C∞(R ,V ). (4.8) A MA ∀ ∈ c For ξ Rn 0 we define ∈ \{ } P(ξ): V V as the orthogonal projection onto ker A(ξ) (4.9) → and we define Q(ξ): V V by → Q(ξ)A(ξ) = Id P(ξ), Q 0 on (range A(ξ))⊥. (4.10) − ≡ It follows from (4.5) that ξ P(ξ) is smooth and homogeneous of degree zero on Rn 0 , while ξ Q(ξ) 7→ \{ } 7→ is smooth and homogeneous of degree 1. For ξ Rn 0 define − ∈ \{ } M(ξ) := Ae(ξ)Q(ξ). (4.11) n 1 n n Then M is homogeneous of degree zero and smooth on the unit sphere S − of R . For f C∞(R ,V ) define ∈ c 1 f = ( )− M f (4.12) M F F where denotes the Fourier transform. By the Mikhlin-H¨ormandermultiplier theorem has a unique F M extension to a bounded operator on Lp(Rn,V ). Moreover we have M(ξ)A(ξ) = Ae(ξ)Q(ξ)A(ξ) = Ae(ξ) Ae(ξ)P(ξ) = Ae(ξ). (4.13) − Here we used the assumption (4.4) in the last identity. Now (4.8) follows from the definition of . M  On bounded sets we will make use of the following behavior. 3 k Proposition 4.2. Let Ω R be bounded, open and non-empty. Let z = x1 + i x2 and let qk(x) = z . Then ⊂ kqk 1 Lp(Ω, ) lim k − k C = . k q p ∞ →∞ k kkL (Ω,C) 3 2 2 2 1 Proof. For δ > 0 define Uδ := x R x1 + x2 < δ . Since qk 1/qk δ− on Ω Uδ we have { ∈ | } | − | ≤ \ kqk 1 Lp(Ω U , ) lim k − k \ δ C = . (4.14) k qk Lp(Ω U , ) ∞ →∞ k k \ δ C Now the assertion follows from the fact that

qk Lp(Ω U , ) lim k k \ δ C = 1 k q p →∞ k kkL (Ω,C) whenever δ > 0 is so small that Ω U has positive measure. \ 2δ  With these preparations in hand we arrive at our final result. Theorem 4.3. The following assertions hold for p (1, ). ∈ ∞ 3 p 3 3 3 1. (whole space R ) There exists a constant c = c(p) such that for P L (R , R × ) ∈ sym Curl P p 3 3×3 c dev sym Curl P p 3 3×3 ; (4.15) k kL (R ,R ) ≤ k kL (R ,R ) 3 3 3 p 3 3 3 2. (periodic functions) if T = R /Z then for all P L (T , R × ) ∈ sym Curl P p 3 3×3 c dev sym Curl P p 3 3×3 ; (4.16) k kL (R ,R ) ≤ k kL (R ,R ) p 3 3 3. (half-spaces) if Ω is a half-space then for P L (Ω, R × ) the seminorms sym Curl P Lp(Ω, 3×3) and ∈ k k R dev sym Curl P p 3×3 are not equivalent; k kL (Ω,R ) 4. (bounded sets) If Ω R3 is a bounded, open, non-empty set then ⊂ 1, p 3 1, p 3 W (sym Curl; Ω, R ) = W (dev sym Curl; Ω, R ). 6

28 Notation In this subsection we use the notation

3 X 3 a, b := ajbj for a, b C . ∈ j=1 P Note that this is different from the usual sesquilinear form j ajbj where z denotes the complex conjugate of a complex number z. In particular a, a is not nonnegative on C3. Proof. The first and second assertion for the whole space and periodic functions follow from the estimate (2.30) 3 3 3 ξ R Pb R × sym(Pb ξ) (1 + √3) dev sym(Pb ξ) , (4.17) ∀ ∈ ∀ ∈ k × k ≤ k × k the fact that dim Pb dev sym(Pb ξ) = 0 = 4 is independent of ξ for ξ R3 0 and Proposition 4.1, { | × } ∈ \{ } applied to the operators A(ξ)P = dev sym(P ξ) and Ae(ξ)P = sym(P ξ). To prove the third and fourth assertion we× first show that ×

3 3 3 ξ C Pb C × : dev sym(Pb ξ) = 0 and sym(Pb ξ) = 0. (4.18) ∃ ∈ ∃ ∈ × × 6 Then the assertion will follow by standard arguments. One such example is given by

0 0 1 1 i 1 0 − − Pb = 0 0 i  and ξ = i Pb ξ = 1 i 0 (4.19a) 0 i 0 0 ⇒ × 0 0 i − so that

sym(Pb ξ) = i 1 but dev sym(Pb ξ) = 0. (4.19b) × · × Further examples which fulfill (4.18) can be found splitting Pb into the symmetric and skew-symmetric part: Pb = Sb + Anti(ba). By (2.26) 2 dev sym(Pb ξ) = sym(Pb ξ) + a, ξ 1. (4.20) × × 3 b

3 3 3 Thus it suffices to find Pb C × and ξ C such that dev sym(Pb ξ) = 0 and ba, ξ = 0. Indeed, the example in (4.19a) satisfies∈ these conditions.∈ × 6

3×3 3×3 Now we show that in a half-space the seminorms sym Curl Lp(Ω,R ) and dev sym Curl Lp(Ω,R ) are not equivalent. Since the operators dev sym Curl andk sym Curl·k interact naturallyk with rotations·k it suffices to consider the half-space 3 Ω = x R x1 < 0 . { ∈ | } Note that the norms are equivalent for real-valued fields P if and only if they are equivalent for complex- valued fields P . Let ξ and Pb be as in (4.19a). For a constant vector b R3 and a scalar function ζ ∈ we have curl(b ζ) = b ( ζ). Since Curl acts row-wise we have for a constant matrix Pb the identity × −∇ Curl(Pb ζ) = Pb ζ. Thus for all k N − × ∇ ∈   Curl Pbek ξ,x = Pb ek ξ,x = kek ξ,x (Pb ξ) (4.21) − × ∇ − × so that with (4.19b) we have     sym Curl Pbek ξ,x = i k ek ξ,x 1 and dev sym Curl Pbek ξ,x = 0. (4.22) − ·

29 Let η C∞(B(0, 2)) be a cut-off function such that η = 1 in B(0, 1) and consider the functions ∈ c

1 k ξ,x Pk(x) = Pbe η(x). (4.23) k Then     η(x) k ξ,x 1 k ξ,x (4.21) k ξ,x 1 Curl Pk(x) = Curl Pbe e Pb Anti( η) = e Pb ξ + Pb Anti( η) k − k ∇ − × k ∇ and with (4.19b) we obtain 1 dev sym Curl P (x) C ekx1 sup η (4.24a) k k k ≤ k k∇ k and

k x1+i k x2 sym Curl Pk = i e 1 in B(0, 1). (4.24b) − · From this we easily conclude that sym Curl Pk p/ dev sym Curl Pk p which shows claim3. Finally, we prove the last assertionk 4. Let Ωk bek a bounded, open,k → non-empty ∞ set. It suffices to show 1, p 3 3 3×3 3×3 3×3 that in W (sym Curl; Ω, C × ) the norms Lp(Ω,C ) + sym Curl Lp(Ω,C ) and Lp(Ω,C ) + k·k k ·k 1, p k·k 3 3 dev sym Curl Lp(Ω, 3×3) are not equivalent. Indeed, this implies that also in W (sym Curl; Ω, R × ) the k ·k C 3×3 3×3 3×3 3×3 norms Lp(Ω,R ) + sym Curl Lp(Ω,R ) and Lp(Ω,R ) + dev sym Curl Lp(Ω,R ) are not equivalent. Thus, sincek·k the identityk map ·k k·k k ·k

1, p 3 3 1, p 3 3 i : W (sym Curl; Ω, R × ) W (dev sym Curl; Ω, R × ) → is continuous it then follows from the open mapping theorem that

1, p 3 3 1, p 3 3 W (dev sym Curl; Ω, R × ) = W (sym Curl; Ω, R × ). 6 Let ξ and Pˆ be again as in (4.19a). Set z = x1 + ix2 and

t ξ,x tz Pt(x) := Pbe = Pbe . (4.25) Then as in (4.22) dev sym Curl Pt = 0 (4.26) and tz sym Curl Pt = i t e 1. (4.27) − · k Let Qk(x) = Pb z . Taking the k-th derivative of (4.26) and (4.27) and evaluating at t = 0 we get, for all k N, ∈ k 1 dev sym Curl Qk = 0, sym Curl Qk = i k z − 1. (4.28) − · It follows from Proposition 4.2 that

sym Curl Qk p 3×3 lim k kL (Ω,C ) = . k Q p 3×3 + dev sym Curl Q p 3×3 ∞ →∞ k kkL (Ω,C ) k kkL (Ω,C ) This concludes the proof of the theorem.  Remark 4.4. Assertion4 of Theorem 4.3 is complemented by the following two strict inclusions:

1, p 3 3 1, p 3 3 1, p 3 3 W (Ω, R × ) $ W (Curl; Ω, R × ) $ W (sym Curl; Ω, R × ). (4.29)

To see that the first inclusion is strict, we may use functions of the form Pk = Duk where uk = w(kx) and w : R3 R3 is periodic, to see that the corresponding norms are not equivalent. To see that the second → 3 inclusion is strict we can use functions of the form Pk(x) = ζ(kx) 1 where ζ : R R is periodic, and observe that sym Curl P = k sym(Anti( ζ)(kx)) = 0. · → k ∇

30 Acknowledgment The authors thank Ionel-Dumitrel Ghiba, University of Iasi, Romania, for helpful discussions and also the anonymous referee for his valuable comments and suggestions. This work was initiated in the framework of the Priority Programme SPP 2256 ’Variational Methods for Predicting Complex Phenomena in Engineering Structures and Materials’ funded by the Deutsche Forschungsgemeinschaft (DFG, German research foun- dation), Project-ID 422730790, by a collaboration of projects ’Mathematical analysis of microstructure in supercompatible alloys’ (Project-ID 441211072) and ’A variational scale-dependent transition scheme - from Cauchy elasticity to the relaxed micromorphic continuum’ (Project-ID 440935806). Peter Lewintan and Pa- trizio Neff were supported by the Deutsche Forschungsgemeinschaft (Project-ID 415894848). Stefan M¨uller has also been supported by the Deutsche Forschungsgemeinschaft through the Hausdorff Center for Mathe- matics (GZ EXC 2047/1, Projekt-ID 390685813) and the collaborative research centre ’The mathematics of emerging effects’ (CRC 1060, Projekt-ID 211504053).

References

[1] A. Acharya. “Driving forces and boundary conditions in continuum dislocation mechanics”. Proceedings of the Royal Society of . Series A. Mathematical, Physical and Engineering Sciences 459.2034 (2003). Pp. 1343–1363. issn: 1364-5021; 1471-2946/e. [2] G. Acosta and R. G. Dur´an. Divergence Operator and Related Inequalities. New York, NY: Springer, 2017. [3] G. Acosta, R. G. Dur´an,and M. A. Muschietti. “Solutions of the divergence operator on John domains”. Advances in Mathematics 206.2 (2006). Pp. 373–401. [4] C. Amrouche, P. G. Ciarlet, L. Gratie, and S. Kesavan. “On Saint Venant’s compatibility conditions and Poincar´e’s lemma”. Comptes Rendus. Math´ematique.Acad´emiedes Sciences, Paris 342.11 (2006). Pp. 887–891. [5] C. Amrouche, P. G. Ciarlet, and C. Mardare. “On a lemma of Jacques-Louis Lions and its relation to other fundamental results”. Journal de Math´ematiques Pures et Appliqu´ees(9) 104.2 (2015). Pp. 207–226. [6] C. Amrouche and V. Girault. “Decomposition of vector spaces and application to the Stokes problem in arbitrary dimension”. Czechoslovak Mathematical Journal 44(119).1 (1994). Pp. 109–140. [7] S. Amstutz and N. Van Goethem. “Analysis of the incompatibility operator and application in intrinsic elasticity with dislocations”. SIAM Journal on Mathematical Analysis 48.1 (2016). Pp. 320–348. [8] S. Amstutz and N. Van Goethem. “Incompatibility-governed elasto-plasticity for continua with dislocations”. Proceedings of the Royal Society of London. A. Mathematical, Physical and Engineering Sciences 473.2199 (2017). P. 21, Id/No 20160734. [9] S. Amstutz and N. Van Goethem. “The incompatibility operator: From Riemann’s intrinsic view of geometry to a new model of elasto-plasticity”. CIM Series in Mathematical Sciences (2019). Ed. by J. F. Rodrigues and M. Hinterm¨uller. Pp. 33–70. [10] S. Amstutz and N. Van Goethem. “Existence and asymptotic results for an intrinsic model of small-strain incompatible elasticity”. Discrete and Continuous Dynamical Systems. Series B 10 (2020). Pp. 3769–3805. [11] S. Bauer, P. Neff, D. Pauly, and G. Starke. “Dev-Div- and DevSym-DevCurl-inequalities for incompatible square tensor fields with mixed boundary conditions”. ESAIM: Control, Optimisation and Calculus of Variations 22.1 (2016). Pp. 112– 133. [12] S. Bauer and D. Pauly. “On Korn’s first inequality for mixed tangential and normal boundary conditions on bounded N Lipschitz domains in R ”. Annali dell’Universit`adi Ferrara. Sezione VII. Scienze Matematiche 62.2 (2016). Pp. 173– 188. [13] S. Bauer and D. Pauly. “On Korn’s first inequality for tangential or normal boundary conditions with explicit constants.” Mathematical Methods in the Applied Sciences 39.18 (2016). Pp. 5695–5704. [14] B. Bernstein and R. A. Toupin. “Korn inequalities for the sphere and circle”. Archive for Rational Mechanics and Analysis 6 (1960). Pp. 51–64. [15] W. Borchers and H. Sohr. “On the equations rot v = g and div u = f with zero boundary conditions”. Hokkaido Mathematical Journal 19.1 (1990). Pp. 67–87. [16] F. Boyer and P. Fabrie. Mathematical Tools for the Study of the Incompressible Navier-Stokes Equations and Related Models. Vol. 183. Applied Mathematical Sciences. Springer, New York, 2013. [17] J. H. Bramble. “A proof of the inf-sup condition for the Stokes equations on Lipschitz domains”. Mathematical Models and Methods in Applied Sciences 13.3 (2003). Pp. 361–371.

31 [18] D. Breit, A. Cianchi, and L. Diening. “Trace-free Korn inequalities in Orlicz spaces”. SIAM Journal on Mathematical Analysis 49.4 (2017). Pp. 2496–2526. [19] D. Breit and L. Diening. “Sharp conditions for Korn inequalities in Orlicz spaces”. Journal of Mathematical Fluid Mechanics 14.3 (2012). Pp. 565–573. [20] S. C. Brenner, F. Li, and L.-Y. Sung. “A locally divergence-free interior penalty method for two-dimensional curl-curl problems”. SIAM Journal of Numerical Analysis 46.3 (2008). Pp. 1190–1211. [21] S. C. Brenner. “Korn’s inequalities for piecewise H1 vector fields”. Mathematics of Computation 73.247 (2004). Pp. 1067– 1087. [22] H. Brezis and J. Van Schaftingen. “Boundary estimates for elliptic systems with L1-data”. Calculus of Variations and Partial Differential Equations 30.3 (2007). Pp. 369–388. [23] S. Campanato. “La disuguaglianza di Korn in elasticit`a”. Conference on Esistenza e Stabilit`ain Elasticit`a. CISM, Udine, 18–23 June 1971. Unpublished, proof contained in [38]. [24] L. Cattabriga. “Su un problema al contorno relativo al sistema di equazioni di Stokes”. Rendiconti del Seminario Matematico della Universit`adi Padova. The Mathematical Journal of the University of Padova 31 (1961). Pp. 308–340. [25] A. Chambolle, S. Conti, and G. Francfort. “Korn-Poincar´einequalities for functions with a small jump set”. Indiana University Mathematics Journal 65.4 (2016). Pp. 1373–1399. [26] A. Cianchi. “Korn type inequalities in Orlicz spaces”. Journal of Functional Analysis 267.7 (2014). Pp. 2313–2352. [27] P. G. Ciarlet. “An introduction to differential geometry with applications to elasticity”. Journal of Elasticity 78-79.1-3 (2005). Pp. 3–201. [28] P. G. Ciarlet. “On Korn’s inequality”. Chinese Annals of Mathematics. Series B 31.5 (2010). Pp. 607–618. [29] P. G. Ciarlet. Linear and Nonlinear Functional Analysis with Applications. Society for Industrial and Applied Mathe- matics, Philadelphia, PA, 2013. [30] P. G. Ciarlet and P. Ciarlet Jr. “Another approach to linearized elasticity and a new proof of Korn’s inequality”. Mathematical Models and Methods in Applied Sciences 15.2 (2005). Pp. 259–271. [31] P. G. Ciarlet, M. Malin, and C. Mardare. “On a vector version of a fundamental lemma of J. L. Lions”. Chinese Annals of Mathematics. Series B 39.1 (2018). Pp. 33–46. [32] S. Conti, G. Dolzmann, and S. M¨uller.“Korn’s second inequality and geometric rigidity with mixed growth conditions”. Calculus of Variations and Partial Differential Equations 50.1-2 (2014). Pp. 437–454. [33] S. Conti, D. Faraco, and F. Maggi. “A new approach to counterexamples to L1 estimates: Korn’s inequality, geometric rigidity, and regularity for gradients of separately convex functions”. Archive for Rational Mechanics and Analysis 175.2 (2005). Pp. 287–300. [34] S. Conti and A. Garroni. “Sharp rigidity estimates for incompatible fields as consequence of the Bourgain Brezis div-curl result”. Comptes Rendus. Math´ematique.Acad´emiedes Sciences, Paris 359.2 (2021). Pp. 155–160. [35] S. Conti and M. Ortiz. “Dislocation microstructures and the effective behavior of single crystals”. Archive for Rational Mechanics and Analysis 176.1 (2005). Pp. 103–147. [36] M. V. d’Agostino, G. Barbagallo, I.-D. Ghiba, B. Eidel, P. Neff, and A. Madeo. “Effective description of anisotropic wave dispersion in mechanical band-gap metamaterials via the relaxed micromorphic model”. Journal of Elasticity 139 (2020). Pp. 299–329. [37] S. Dain. “Generalized Korn’s inequality and conformal Killing vectors”. Calculus of Variations and Partial Differential Equations 25.4 (2006). Pp. 535–540. [38] G. Del Piero and R. Rizzoni. “Weak local minimizers in finite elasticity”. Journal of Elasticity 93.3 (2008). Pp. 203–244. [39] L. Desvillettes and C. Villani. “On a variant of Korn’s inequality arising in statistical mechanics”. ESAIM: Control, Optimization and Calculus of Variations 8 (2002). Pp. 603–619. [40] L. Diening, M. R˚uˇziˇcka, and K. Schumacher. “A decomposition technique for John domains”. Annales AcademiæScien- tiarum Fennicæ. Mathematica 35.1 (2010). Pp. 87–114. [41] Z. Ding and B. Li. “A conformal Korn inequality on H¨olderdomains”. Journal of Mathematical Analysis and Applications 481.1 (2020). P. 14, Id/No 123440. [42] S. Dom´ınguezand N. Nigam. Korn’s inequality and Jones eigenpairs. 2019. arXiv: 1912.00133 [math.AP]. [43] R. G. Dur´anand M. A. Muschietti. “The Korn inequality for Jones domains”. Electronic Journal of Differential Equations 127 (2004). Pp. 1–10. [44] G. Duvaut and J.-L. Lions. Les In´equations en M´ecanique et en Physique. Dunod, Paris, 1972. [45] F. Ebobisse, K. Hackl, and P. Neff. “A canonical rate-independent model of geometrically linear isotropic gradient plasticity with isotropic hardening and plastic spin accounting for the Burgers vector”. Continuum Mechanics and Thermodynamics 31.5 (2019). Pp. 1477–1502.

32 [46] F. Ebobisse and P. Neff. “Existence and uniqueness for rate-independent infinitesimal gradient plasticity with isotropic hardening and plastic spin”. Mathematics and Mechanics of Solids 15.6 (2010). Pp. 691–703. [47] F. Ebobisse and P. Neff. “A fourth order gauge-invariant gradient plasticity model for polycrystals based on Kr¨oner’s incompatibility tensor”. Mathematics and Mechanics of Solids 25.2 (2020). Pp. 129–159. [48] F. Ebobisse, P. Neff, and S. Forest. “Well-posedness for the microcurl model in both single and polycrystal gradient plasticity”. International Journal of Plasticity 107 (2018). Pp. 1–26. [49] D. M. Eidus. “On a mixed problem of the theory of elasticity”. Russian. Doklady Akademii Nauk SSSR 76 (1951). Pp. 181–184. [50] G. Fichera. “Sull’esistenza e sul calcolo delle soluzioni dei problemi al contorno, relativi all’equilibrio di un corpo elastico”. Italian. Annali della Scuola Normale Superiore di Pisa. Classe di Scienze. Serie III 4.248 (1950). Pp. 35–99. [51] G. Fichera. “Existence Theorems in Elasticity”. Linear Theories of Elasticity and Thermoelasticity. Ed. by C. Truesdell. Springer-Verlag Heidelberg, 1972, pp. 347–389. [52] N. A. Fleck, G. M. M¨uller,M. F. Ashby, and J. W. Hutchinson. “Strain gradient plasticity: Theory and experiment”. Acta Metallurgica et Materialia 42.2 (1994). Pp. 475–487. [53] I. Fonseca and S. M¨uller.“A-quasiconvexity, lower semicontinuity, and Young measures”. SIAM Journal on Mathematical Analysis 30.6 (1999). Pp. 1355–1390. [54] G. Francfort and S. M¨uller.“Combined effects of homogenization and singular perturbations in elasticity”. Journal f¨ur die reine und angewandte Mathematik 454 (1994). Pp. 1–35. [55] M. Friedrich. “A Korn-type inequality in SBD for functions with small jump sets”. Mathematical Models and Methods in Applied Sciences 27.13 (2017). Pp. 2461–2484. [56] M. Friedrich. “A piecewise Korn inequality in SBD and applications to embedding and density results”. SIAM Journal on Mathematical Analysis 50.4 (2018). Pp. 3842–3918. [57] K. O. Friedrichs. “On the boundary-value problems of the theory of elasticity and Korn’s inequality”. Annals of Math- ematics. Second Series 48 (1947). Pp. 441–471. [58] G. Friesecke, R. D. James, and S. M¨uller.“A theorem on geometric rigidity and the derivation of nonlinear plate theory from three-dimensional elasticity”. Communications on Pure and Applied Mathematics 55.11 (2002). Pp. 1461–1506. [59] M. Fuchs. “Generalizations of Korn’s inequality based on gradient estimates in Orlicz spaces and applications to varia- tional problems in 2D involving the trace free part of the symmetric gradient”. Journal of Mathematical Sciences (New York) 167.3 (2010). Pp. 418–434. [60] M. Fuchs and S. Repin. “Some Poincar´e-type inequalities for functions of bounded deformation involving the deviatoric part of the symmetric gradient.” Journal of Mathematical Sciences (New York) 178.3 (2011). Pp. 367–372. [61] M. Fuchs. “Korn inequalities in Orlicz spaces”. Irish Mathematical Society Bulletin 65 (2010). Pp. 5–9. [62] M. Fuchs and O. Schirra. “An application of a new coercive inequality to variational problems studied in general relativity and in Cosserat elasticity giving the smoothness of minimizers”. Archiv der Mathematik 93.6 (2009). Pp. 587–596. [63] A. Garroni, G. Leoni, and M. Ponsiglione. “Gradient theory for plasticity via homogenization of discrete dislocations”. Journal of the European Mathematical Society 12.5 (2010). Pp. 1231–1266. [64] G. Geymonat and P. Suquet. “Functional spaces for Norton-Hoff materials”. Mathematical Methods in the Applied Sciences 8.2 (1986). Pp. 206–222. [65] G. Geymonat and F. Krasucki. “Some remarks on the compatibility conditions in elasticity”. Rendiconti Accademia Nazionale delle Scienze detta dei XL 29 (2005). Pp. 175–181. [66] I.-D. Ghiba, P. Neff, A. Madeo, and I. M¨unch. “A variant of the linear isotropic indeterminate couple-stress model with symmetric local force-stress, symmetric nonlocal force-stress, symmetric couple-stresses and orthogonal boundary conditions”. Mathematics and Mechanics of Solids 22.6 (2017). Pp. 1221–1266. [67] J. Ginster. “Strain-gradient plasticity as the Γ-limit of a nonlinear dislocation energy with mixed growth”. SIAM Journal on Mathematical Analysis 51.4 (2019). Pp. 3424–3464. [68] V. Girault and P.-A. Raviart. Finite Element Methods for Navier-Stokes Equations. Vol. 5. Springer Series in Compu- tational Mathematics. Theory and algorithms. Springer-Verlag, Berlin, 1986. [69] F. Gmeineder and D. Spector. “On Korn-Maxwell-Sobolev Inequalities”. Journal of Mathematical Analysis and Appli- cations 502.125226 (2021). [70] J. Gobert. “Une in´egalit´efondamentale de la th´eoriede l’´elasticit´e”.French. Bulletin de la Soci´et´eRoyale des Sciences de Li`ege 31 (1962). Pp. 182–191. [71] Y. Grabovsky and D. Harutyunyan. “Korn inequalities for shells with zero Gaussian curvature”. Annales de l’Institut Henri Poincar´e.Analyse Non Lin´eaire 35.1 (2018). Pp. 267–282.

33 3 [72] J. Gross, G. Trenkler, and S.-O. Troschke. “The vector cross product in C ”. International Journal of Mathematical Education in Science and Technology 30.4 (1999). Pp. 549–555. [73] M. E. Gurtin. “On the plasticity of single crystals: free energy, microforces, plastic-strain gradients”. Journal of the Mechanics and Physics of Solids 48.5 (2000). Pp. 989–1036. [74] M. E. Gurtin and L. Anand. “A theory of strain-gradient plasticity for isotropic, plastically irrotational materials. I. Small deformations”. Journal of the Mechanics and Physics of Solids 53.7 (2005). Pp. 1624–1649. [75] M. E. Gurtin and A. Needleman. “Boundary conditions in small-deformation, single-crystal plasticity that account for the Burgers vector”. Journal of the Mechanics and Physics of Solids 53.1 (2005). Pp. 1–31. [76] W. Han and B. D. Reddy. Plasticity. Mathematical Theory and Numerical Analysis. Second edition. Vol. 9. Interdisci- plinary Applied Mathematics. Springer, New York, 2013. [77] D. Harutyunyan. “Sharp weighted Korn and Korn-like inequalities and an application to washers”. Journal of Elasticity 127.1 (2017). Pp. 59–77. [78] I. Hlav´aˇcekand J. Neˇcas.“On inequalities of Korn’s type. I: Boundary-value problems for elliptic systems of partial differential equations”. Archive for Rational Mechanics and Analysis 36 (1970). Pp. 305–311. [79] I. Hlav´aˇcekand J. Neˇcas.“On inequalities of Korn’s type. II: Applications to linear elasticity”. Archive for Rational Mechanics and Analysis 36 (1970). Pp. 312–334. [80] M. Holst, J. Kommemi, and G. Nagy. Rough solutions of the Einstein constraint equations with nonconstant mean curvature. 2007. arXiv: 0708.3410 [gr-qc]. [81] C. O. Horgan. “Korn’s inequalities and their applications in continuum mechanics”. SIAM Review. A Publication of the Society for Industrial and Applied Mathematics 37.4 (1995). Pp. 491–511. [82] J. Jeong and P. Neff. “Existence, uniqueness and stability in linear Cosserat elasticity for weakest curvature conditions”. Mathematics and Mechanics of Solids 15.1 (2010). Pp. 78–95. [83] J. Jeong, H. Ram´ezani,I. M¨unch, and P. Neff. “A numerical study for linear isotropic Cosserat elasticity with conformally invariant curvature”. Zeitschrift f¨urAngewandte Mathematik und Mechanik 89.7 (2009). Pp. 552–569. [84] R. Jiang and A. Kauranen. “Korn’s inequality and John domains”. Calculus of Variations and Partial Differential Equations 56.109 (2017). Pp. 1–18. [85] Y. Kato. “The coerciveness for integro-differential quadratic forms and Korn’s inequality”. Nagoya Mathematical Journal 73 (1979). Pp. 7–28. [86] V. A. Kondrat’ev and O. A. Oleinik. “Boundary-value problems for a system in elasticity theory in unbounded domains. Korn’s inequalities”. Russian Mathematical Surveys 43.5 (1988). Pp. 65–119. [87] A. Korn. “Abhandlungen zur Elastizit¨atstheorieII: Die Eigenschwingungen eines elastischen K¨orpers mit ruhender Oberfl¨ache”. German. Sitzungsberichte der Mathematisch-physikalischen Klasse der K¨oniglichbayerischen Akademie der Wissenschaften zu M¨unchen 36 (1906). Pp. 351–402. [88] A. Korn. “Solution g´en´eraledu probl`emed’´equilibredans la th´eoriede l’´elasticit´edans le cas o`ules efforts donn´es`ala surface.” French. Annales de la Facult´edes Sciences de l’Universit´ede Toulouse pour les Sciences Math´ematiqueset les Sciences Physiques. 2i`emeS´erie 10 (1908). Pp. 165–269. [89] A. Korn. “Uber¨ einige Ungleichungen, welche in der Theorie der elastischen und elektrischen Schwingungen eine Rolle spielen”. Bulletin International de l’Acad´emiedes Sciences de Cracovie (1909). Pp. 705–724. [90] T. Korn and E. P. Korn. Trailbrazer to , The Story of Arthur Korn. Charles Scribner’s Sons, 1950. [91] M. Lagally. Vorlesungen ¨uber Vektor-Rechnung. German. Ed. by E. Hilb. Mathematik und ihre Anwendungen in Mono- graphien und Lehrb¨uchern. Akademische Verlagsgesellschaft m.B.H., Leipzig, 1928.

[92] J. Lankeit, P. Neff, and D. Pauly. “Uniqueness of integrable solutions to ∇ζ = Gζ, ζ|Γ = 0 for integrable tensor coefficients G and applications to elasticity”. Zeitschrift f¨urangewandte Mathematik und Physik 64.6 (2013). Pp. 1679– 1688. [93] G. Lauteri and S. Luckhaus. “Geometric rigidity estimates for incompatible fields in dimension ≥ 3” (2017). arXiv: 1703.03288 [math.AP]. [94] M. Lazar. “An elastoplastic theory of dislocations as a physical field theory with torsion.” Journal of Physics A: Mathematical and General 35.8 (2002). Pp. 1983–2004. [95] M. Lazar. “Dislocations in generalized continuum mechanics”. Mechanics of generalized continua. One hundred years after the Cosserats. Papers based on the presentations at the EUROMECH colloquium 510, Paris, France, May 13–16, 2009. Ed. by G. A. Maugin and A. V. Metrikine. New York, NY: Springer, 2010, pp. 235–244. [96] M. Lazar and C. Anastassiadis. “The gauge theory of dislocations: static solutions of screw and edge dislocations”. Philosophical Magazine 89.03 (2009). Pp. 199–231.

34 [97] M. Lazar and G. A. Maugin. “Dislocations in gradient elasticity revisited”. Proceedings of the Royal Society of London. Series A. Mathematical, Physical and Engineering Sciences 462.2075 (2006). Pp. 3465–3480.

[98] K. de Leeuw and H. Mirkil. “A priori estimates for differential operators in L∞ norm”. Illinois Journal of Mathematics 8 (1964). Pp. 112–124. [99] A. Lew, P. Neff, D. Sulsky, and M. Ortiz. “Optimal BV estimates for a discontinuous Galerkin method for linear elasticity”. Applied Mathematics Research Express 2004.3 (2004). Pp. 73–106. [100] M. Lewicka and S. M¨uller.“The uniform Korn-Poincar´einequality in thin domains”. Annales de l’Institut Henri Poincar´e. Analyse Non Lin´eaire 28.3 (2011). Pp. 443–469. [101] M. Lewicka and S. M¨uller.“On the optimal constants in Korn’s and geometric rigidity estimates, in bounded and unbounded domains, under Neumann boundary conditions”. Indiana University Mathematics Journal 65.2 (2016). Pp. 377–397. [102] P. Lewintan and P. Neff. “Lp-trace-free generalized Korn inequalities for incompatible tensor fields in three space dimensions”. submitted (2020). arXiv: 2004.05981 [math.AP]. [103] P. Lewintan and P. Neff. “Lp-trace-free version of the generalized Korn inequality for incompatible tensor fields in arbitrary dimensions”. Zeitschrift f¨urangewandte Mathematik und Physik 72.127 (2021). [104] P. Lewintan and P. Neff. “Lp-versions of generalized Korn inequalities for incompatible tensor fields in arbitrary dimen- sions with p-integrable exterior derivative”. to appear in ComptesRendus (2021). arXiv: 1912.11551 [math.AP]. [105] P. Lewintan and P. Neff. “Neˇcas-Lionslemma revisited: An Lp-version of the generalized Korn inequality for incompatible tensor fields”. Mathematical Methods in the Applied Sciences (2021). Pp. 1–12. [106] F. Litten. Arthur Korn – Kurzbiographie. German. http://litten.de/fulltext/korn.htm. Accessed: 2020-04-27. [107] S. Liu and G. Trenkler. “Hadamard, Khatri-Rao, Kronecker and other matrix products”. International Journal of Information and Systems Sciences 4.1 (2008). Pp. 160–177. [108] F. L´opez-Garc´ıa.“Weighted generalized Korn inequalities on John domains”. Mathematical Methods in the Applied Sciences 41.17 (2018). Pp. 8003–8018. issn: 0170-4214; 1099-1476/e. [109] G. B. Maggiani, R. Scala, and N. Van Goethem. “A compatible-incompatible decomposition of symmetric tensors in Lp with application to elasticity”. Mathematical Methods in the Applied Sciences 38.18 (2015). Pp. 5217–5230. [110] A. Menzel and P. Steinmann. “On the continuum formulation of higher gradient plasticity for single and polycrystals”. Journal of the Mechanics and Physics of Solids 48.8 (2000). Pp. 1777–1796. [111] M. Mitrea and M. Wright. Boundary Value Problems for the Stokes System in Arbitrary Lipschitz Domains. Vol. 344. Paris: Soci´et´eMath´ematique de France (SMF), 2012. [112] B. S. Mityagin. “On second mixed derivative”. Russian. Doklady Akademii Nauk SSSR 123 (1958). Pp. 606–609. [113] T.-H. Miura. Navier-Stokes equations in a curved thin domain, Part I: uniform estimates for the Stokes operator. 2020. arXiv: 2002.06343 [math.AP]. [114] S. M¨ullerand M. Palombaro. “Existence of minimizers for a polyconvex energy in a crystal with dislocations”. Calculus of Variations and Partial Differential Equations 31.4 (2008). Pp. 473–482. [115] S. M¨uller,L. Scardia, and C. I. Zeppieri. “Geometric rigidity for incompatible fields, and an application to strain-gradient plasticity”. Indiana University Mathematics Journal 63.5 (2014). Pp. 1365–1396. [116] I. M¨unch and P. Neff. “Rotational invariance conditions in elasticity, gradient elasticity and its connection to isotropy”. Mathematics and Mechanics of Solids 23.1 (2018). Pp. 3–42. [117] P. Neff. “On Korn’s first inequality with non-constant coefficients”. Proceedings of the Royal Society of Edinburgh Section A: Mathematics 132.01 (2002). Pp. 221–243. [118] P. Neff, K. Chelmi´nski,and H.-D. Alber. “Notes on strain gradient plasticity: finite strain covariant modelling and global existence in the infinitesimal rate-independent case”. Mathematical Models and Methods in Applied Sciences 19.02 (2009). Pp. 307–346. [119] P. Neff, B. Eidel, M. V. d’Agostino, and A. Madeo. “Identification of scale-independent material parameters in the relaxed micromorphic model through model-adapted first order homogenization”. Journal of Elasticity 139 (2020). Pp. 269–298. [120] P. Neff, I.-D. Ghiba, M. Lazar, and A. Madeo. “The relaxed linear micromorphic continuum: well-posedness of the static problem and relations to the gauge theory of dislocations”. The Quarterly Journal of Mechanics and Applied Mathematics 68.1 (2015). Pp. 53–84. [121] P. Neff, I.-D. Ghiba, A. Madeo, L. Placidi, and G. Rosi. “A unifying perspective: the relaxed linear micromorphic continuum”. Continuum Mechanics and Thermodynamics 26.5 (2014). Pp. 639–681. [122] P. Neff and J. Jeong. “A new paradigm: the linear isotropic Cosserat model with conformally invariant curvature energy”. Zeitschrift f¨urAngewandte Mathematik und Mechanik 89.2 (2009). Pp. 107–122.

35 [123] P. Neff, J. Jeong, and A. Fischle. “Stable identification of linear isotropic Cosserat parameters: bounded stiffness in bending and torsion implies conformal invariance of curvature”. Acta Mechanica 211.3-4 (2010). Pp. 237–249. [124] P. Neff, J. Jeong, and H. Ram´ezani.“Subgrid interaction and micro-randomness–Novel invariance requirements in infinitesimal gradient elasticity”. International Journal of Solids and Structures 46.25 (2009). Pp. 4261–4276. [125] P. Neff and I. M¨unch. “Curl bounds Grad on SO(3)”. ESAIM: Control, Optimisation and Calculus of Variations 14.1 (2008). Pp. 148–159. [126] P. Neff, D. Pauly, and K.-J. Witsch. “Poincar´emeets Korn via Maxwell: Extending Korn’s first inequality to incompatible tensor fields”. Journal of Differential Equations 258.4 (2015). Pp. 1267–1302. [127] P. Neff and W. Pompe. “Counterexamples in the theory of coerciveness for linear elliptic systems related to generalizations of Korn’s second inequality”. Zeitschrift f¨urAngewandte Mathematik und Mechanik 94.9 (2014). Pp. 784–790. (k) [128] J. Neˇcas.“Sur les normes ´equivalentes dans Wp (Ω) et sur la coercivit´edes formes formellement positives”. Equations´ aux d´eriv´eespartielles. Les Presses de l’Universit´ede Montreal, 1966, pp. 102–128. [129] J. A. Nitsche. “On Korn’s second inequality”. RAIRO Analyse Num´erique 15.3 (1981). Pp. 237–248. [130] J. F. Nye. “Some geometrical relations in dislocated crystals”. Acta Metallurgica 1 (1953). Pp. 153–162. [131] N. Ohno and D. Okumura. “Higher-order stress and grain size effects due to self-energy of geometrically necessary dislocations”. Journal of the Mechanics and Physics of Solids 55.9 (2007). Pp. 1879–1898.

[132] D. Ornstein. “A non-equality for differential operators in the L1 norm”. Archive for Rational Mechanics and Analysis 11 (1962). Pp. 40–49. [133] A. Panteghini, L. Bardella, and C. F. Niordson. “A potential for higher-order phenomenological strain gradient plasticity to predict reliable response under non-proportional loading”. Proceedings A 475.2229 (2019). Pp. 20190258, 21. [134] L. E. Payne and H. F. Weinberger. “On Korn’s inequality”. Archive for Rational Mechanics and Analysis 8 (1961). Pp. 89–98. [135] M. Pinl. “Kollegen in einer dunklen Zeit”. Jahresbericht der DMV 71 (1969). Pp. 167–228. [136] W. Pompe. “Korn’s first inequality with variable coefficients and its generalization”. Commentationes Mathematicae Universitatis Carolinae 44.1 (2003). Pp. 57–70. [137] Y. G. Reshetnyak. “Estimates for certain differential operators with finite-dimensional kernel”. Siberian Mathematical Journal 11.2 (1970). Pp. 315–326. [138] Y. G. Reshetnyak. Stability Theorems in Geometry and Analysis. Springer Science+Business Media Dordrecht, 1994. [139] M. R¨ogerand B. Schweizer. “Strain gradient visco-plasticity with dislocation densities contributing to the energy”. Mathematical Models and Methods in Applied Sciences 27.14 (2017). Pp. 2595–2629. [140] A. Romano. “On the necessity of Korn’s inequality”. STAMM 2000 : Symposium on Trends in the Application of Mathematics to Mechanics. Ed. by P.-E. O’donoghue and J.-N. Flavin. Elsevier, Paris, 2000, pp. 166–173. [141] J. D. Romano and R. H. Price. “Why no shear in “Div, grad, curl and all that”?” American Journal of Physics 80.6 (2012). Pp. 519–524. [142] T. G. Room. “The composition of rotations in Euclidean three-space”. American Mathematical Monthly 59 (1952). Pp. 688–692. [143] E. I. Ryzhak. “Korn’s constant for a parallelepiped with a free face or pair of faces”. Mathematics and Mechanics of Solids 4.1 (1999). Pp. 35–55. [144] R. Scala and N. van Goethem. “A variational approach to single crystals with dislocations”. SIAM Journal on Mathe- matical Analysis 51.1 (2019). Pp. 489–531. [145] R. Scala and N. Van Goethem. “Analytic and geometric properties of dislocation singularities”. Proceedings of the Royal Society of Edinburgh. Section A. Mathematics 150.4 (2020). Pp. 1609–1651. [146] O. D. Schirra. “New Korn-type inequalities and regularity of solutions to linear elliptic systems and anisotropic variational problems involving the trace-free part of the symmetric gradient”. Calculus of Variations and Partial Differential Equations 43.1 (2012). Pp. 147–172. [147] D. E. Spector and S. J. Spector. BMO and elasticity: Korn’s inequality; local uniqueness in tension. 2020. arXiv: 2004. 02368 [math.AP]. [148] E. M. Stein. Singular Integrals and Differentiability Properties of Functions. Princeton, N.J.: Princeton University Press. XIV, 1970. [149] A. Tiero. “On inequalities of Korn, Friedrichs, Magenes-Stampacchia-Neˇcasand Babuˇska-Aziz”. Zeitschrift f¨urAnalysis und ihre Anwendungen 20.1 (2001). Pp. 215–222. [150] A. Tiero. “On Korn’s inequality in the second case”. Journal of Elasticity 54.3 (1999). Pp. 187–191.

36 [151] T. W. Ting. “Generalized Korn’s inequalities”. The Tensor Society. Tensor. New Series 25 (1972). Pp. 295–302, Com- memoration volumes for Prof. Dr. Akitsugu Kawaguchi’s seventieth birthday, Vol. II. [152] T. W. Ting. “St.Venant’s compatibility conditions and basic problems in elasticity”. Rocky Mountain Journal of Math- ematics 7 (1977). Pp. 47–52. [153] G. Trenkler. “The vector cross product from an algebraic point of view”. Discussiones Mathematicae. General Algebra and Applications 21.1 (2001). Pp. 67–82. [154] W. Wang. “Korn’s inequality and Donati’s theorem for the conformal Killing operator on pseudo-Euclidean space”. Journal of Mathematical Analysis and Applications 345.2 (2008). Pp. 777–782. [155] S. Wulfinghoff, S. Forest, and T. B¨ohlke. “Strain gradient plasticity modeling of the cyclic behavior of laminate mi- crostructures”. Journal of the Mechanics and Physics of Solids 79 (2015). Pp. 1–20.

A Appendix A.1 Geometrical interpretation of tangential boundary conditions A.1.1 The case P ν = 0 following [75] × In this appendix we provide the reader with the development of Gurtin and Needleman [75] adapted to our notation. Let 3 ν ∈ R be a unit vector. For the projection onto the plane perpendicular to ν we can consider one of the following matrix representations: (2.9) (2.12) T Pν = 1 − ν ⊗ ν = − Anti(ν) × ν = − Anti(ν) Anti(ν) = Anti(ν) Anti(ν) . (A.1) The last expression shows directly that the vector product of Pν and ν commutes: (2.13) ν × Pν = − Anti(ν) Anti(ν) Anti(ν) = Pν Anti(ν) = Pν × ν = Anti(ν) ∈ so(3) (A.2) kνk=1 and we have moreover for any (3 × 3)-matrices P and H:

(A.2) (A.3) P × ν, H = P Anti(ν),H = − P,H Anti(ν) = − P,H × ν = P Pν Anti(ν),H = − P Pν ,H × ν . and also

2 (A.1) kP Pν k = P Pν ,P Pν = P Anti(ν) Anti(ν),P Anti(ν) Anti(ν) = − P Anti(ν),P Anti(ν) Anti(ν) Anti(ν) (A.4) (2.13) = P Anti(ν),P Anti(ν) = kP × νk2 (A.5) kνk=1 (A.1) = − P,P Anti(ν) Anti(ν) = P,P (1 − ν ⊗ ν) = P,P − P, P ν ⊗ ν = kP k2 − P ν, P ν (A.6) = kP k2 − kP νk2 . (A.7)

Thus, P Pν = 0 if and only if P × ν = 0. The latter condition can be tested by applying the scalar product with deviatoric (trace-free) matrices, it holds: P × ν, D = 0 ∀D with tr D = 0 ⇔ P × ν = 0. (A.8) 3 This implies dev(P × ν) = 0 if and only if P × ν = 0 and extends, of course, to the case of arbitrary non-zero vector ν ∈ R , cf. also our Observation 2.2 in [102] and shows kdev(P × ν)k ≤ kP × νk ≤ C · kdev(P × ν)k. (A.9) To establish (A.8), let H be an arbitrary matrix and consider the trace-free matrix D := H − tr(H)ν ⊗ ν. By the assumption we have 0 = P × ν, D = P × ν, H − tr(H)ν ⊗ ν = P × ν, H + tr(H) P, (ν ⊗ ν) × ν = P × ν, H . (A.10) Since H is arbitrary, it follows P × ν = 0.

A.1.2 The compatible case Du ν = 0 × Let Γ be a relatively open (non-empty) connected subset of the boundary ∂Ω and assume that P = Du is compatible. The condition Du × ν|Γ ≡ 0 is equivalent to Du Pν |Γ ≡ 0 which can also be written as Du τ |Γ ≡ 0 for all tangential directions on Γ, meaning that all tangential derivatives of u along Γ are vanishing. Thus, u has to be constant along Γ, since for any curve d 0 γ : [0, 1] → Γ on Γ we have ds u(γ(s)) = Du(γ(s)) γ (s) = 0, cf. [68, p. 35].

37 A.1.3 The case sym(P ν) = 0 × 3 If sym(P × ν) = 0, then there exists a vector a ∈ R so that P × ν = Anti(a). Hence, a × ν = Anti(a) ν = (P × ν) ν = P Anti(ν) ν = P (ν × ν) = 0 (A.11) and a has to be of the form a = α · ν with α ∈ R. Thus we have P × ν = α · Anti(ν) = α · 1 × ν and we conclude

sym(P × ν) = 0 ⇔ P × ν = α · Anti(ν) ⇔ (P − α · 1) × ν = 0 ⇔ (P − α · 1) Pν = 0 for an α ∈ R. (A.12) In a similar way to (A.8), it follows that sym(P × ν),D = 0 ∀D with tr D = 0 ⇔ sym(P × ν) = 0 (A.13) so that again we deduce that dev sym(P × ν) = 0 if and only if sym(P × ν) = 0, cf. our Observation 2.3.

A.1.4 The compatible case sym(Du ν) = 0 × Let Γ be a relatively open (non-empty) connected subset of the boundary ∂Ω and assume that P = Du is compatible. By the previous observation the condition sym(Du × ν)|Γ = 0 is fulfilled if and only if there exists a function ζ :Γ → R such that (Du−ζ ·1) Pν |Γ = 0. If u = α x+b, then it is clear that this boundary condition is satisfied. On a flat portion of the boundary we 2 will establish also a converse statement. Indeed, let (after possible rotation) Γ ⊆ R ×{0} be simply connected and γ : [0, 1] → Γ. d 0 0 R 1 0 By the previous observation we have ds u(γ(s)) = Du(γ(s)) γ (s) = ζ(γ(s))γ (s), so that u(γ(1)) = u(γ(0)) + 0 ζ(γ(s))γ (s)ds and for a closed curve γ we deduce

 0   0 0  1 1 ζ(γ(s)) e1, γ (s) 1 (ζb(γ1(s), γ2(s)), 0), (γ (s), γ (s)) 2 Z Z Z 1 2 R 0 = ζ(γ(s))γ0(s)ds = ζ(γ(s)) e , γ0(s) ds = 0 0 ds (A.14)  2   (0, ζb(γ1(s), γ2(s))), (γ1(s), γ2(s)) 2  0 0 0 0 R ζ(γ(s)) e3, γ (s) 0

2 where in the last step we have used, that γ ⊂ R × {0} has vanishing third component and we have set ζb(x, y) = ζ(x, y, 0). Since 2 T T (A.14) is valid for all connected curves γ ⊂ Γ ⊆ R × {0} the vector fields (ζ,b 0) and (0, ζb) have to be conservative. Thus,     ζb 0 0 = curl2 D = −ζ,b and 0 = curl2 D = ζ,b (A.15) 0 y ζb x where curl2 D v = v2, x − v1, y, so that we conclude ζ = ζb ≡ const and set ζ(x, y, 0) = α. The previous observation imply d ds [u(γ(s)) − αγ(s)] ≡ 0 for all admissible curves γ meaning that along Γ the function u has to be of a form u(x, y, 0) = T T α · (x, y, 0) + (b1, b2, 0) .

A.2 Some basic identities We outline some basic identities which played useful roles in our considerations:

38 1. from linear algebra: 2. and their formal equivalents from calculus: (a) a ⊗ b dyadic product, (a) Da = a ⊗ ∇, a, b = tr(a ⊗ b) scalar product, div a = a, ∇ = tr(Da), a × b = axl(b ⊗ a − a ⊗ b) vector product, curl a = a × (−∇) = 2 axl skew(Da), b × b = 0, curl ∇ζ ≡ 0, a × b, b = 0, div curl a ≡ 0, 2 skew(a ⊗ b) = − Anti(a × b), 2 skew(Da) = Anti(curl a), (b) P b, (b) Div P = P ∇, 1 b = b, Div(ζ · 1) = ∇ζ, Anti(a) b = a × b = − Anti(b) a, Div(Anti(a)) = − curl a = Anti(∇) a, A b = (axl A) × b, Div A = − curl axl A, (a ⊗ b)b = kbk2a, Div(Da) = ∆a, (b ⊗ a)b = a, b b = kbk2a + (a × b) × b, Div((Da)T ) = ∇ div a = ∆a + curl curl a, (c) P × b = P Anti(b) = −(Anti(b) P T )T , (c) Curl P = P × (−∇) = −P Anti(∇), (d) 1 × b = Anti(b) ∈ so(3), (d) Curl(ζ · 1) = − Anti(∇ζ) ∈ so(3), (e) (a ⊗ b) × b = 0, (e) Curl(Da) ≡ 0, 1 1 Curl((Da)T ) = Curl(sym Da) = − Curl(skew Da) 2 (b ⊗ a) × b = sym(a ⊗ b) × b = − skew(a ⊗ b) × b 2 = −b ⊗ axl skew(a ⊗ b), T 1 T = (Daxl skew Da) = 2 (D curl a) , (f) Room’s formulas: (f) Nye’s formulas: (Anti(a)) × b = b ⊗ a − b, a · 1, Curl(Anti(a)) = div a · 1 − (Da)T ,

A × b = b ⊗ axl A − b, axl A · 1, Curl A = tr(Daxl A) · 1 − (Daxl A)T , T 1 1 T (axl A) ⊗ b = (A × b) − 2 tr(A × b) · 1, Daxl(A) = 2 tr(Curl A) · 1 − (Curl A) , tr(A × b) = −2 axl A, b , tr(Curl A) = 2 div axl A, (g) tr(S × b) = 0, (g) tr(Curl S) ≡ 0, (h) (P × b)T × b = − Anti(b) P T Anti(b) = −b × P T × b, (h) inc (P ) = Curl[(Curl P )T ] = −∇ × P T × ∇, (i) 1 × bT × b = kbk2 · 1 − b ⊗ b ∈ Sym(3), (i) inc (ζ · 1) = ∆ζ · 1 − D2ζ ∈ Sym(3), T (j) (b ⊗ a) × b × b = 0, (j) inc ((Da)T ) ≡ 0, sym(a ⊗ b) × bT × b = 0, inc (sym Da) ≡ 0, skew(a ⊗ b) × bT × b = 0, inc (skew Da) ≡ 0, T (k) (Anti(a)) × b × b = − b, a Anti(b) ∈ so(3), (k) inc (Anti(a)) = − Anti(∇ div a) ∈ so(3), T (l) S × b × b ∈ Sym(3), (l) inc S ∈ Sym(3), T 2 tr(inc S) = ∆ tr(S) − div Div S, tr( S × b × b) = kbk tr(S) − S, b ⊗ b 3×3 , R 2 2 (m) dev(P × b) = P × b + 3 axl skew P, b · 1, (m) dev Curl P = Curl P − 3 div axl skew P · 1, (n) [(P × b)T × b]T = (P T × b)T × b, (n) [inc (P )]T = inc (P T ), sym[(P × b)T × b] = ((sym P ) × b)T × b, sym inc P = inc sym P , skew[(P × b)T × b] = ((skew P ) × b)T × b, skew inc P = inc skew P , (o) tr[((S × b) × b)T × b] = 0, (o) tr(inc Curl S) ≡ 0, (p) a × b, c = − a, c × b , 0 0 3 0 (q) a ⊗ b = 0 ⇔ dev sym(a ⊗ b) = 0, for ζ ∈ D (Ω, R), a ∈ D (Ω, R ), A ∈ D (Ω, so(3)) 0 0 3×3 ⇔ dev sym(Anti(a) × b) = 0, S ∈ D (Ω, Sym(3)) and P ∈ D (Ω, R ).

(r) dev(P × b) = 0 ⇔ P × b = 0, (s) dev sym(P × b) = 0 ⇔ sym(P × b) = 0,

3 3×3 for a, b ∈ R , S ∈ Sym(3), A ∈ so(3) and P ∈ R ,

The expression in (l) reads in more details (S × b)T × b = −b × S × b = − Anti(b) S Anti(b) 2 2 = S(b ⊗ b) + (b ⊗ b)S − kbk S − tr(S)b ⊗ b + (kbk tr(S) − S, b ⊗ b 3×3 ) · 1 , (A.16) R so that the formal equivalent for inc S has the form inc S = DDiv S + (DDiv S)T − ∆S − D2 tr(S) + (∆ tr(S) − div Div S) · 1 . (A.17)

39 A.3 The kernel of Curl, dev Curl and sym Curl 3 0 3×3 Lemma A.1. Let Ω ⊂ R be a simply connected open set and P ∈ D (Ω, R ). Then we have (a) Curl P ≡ 0 if and only if P = Du, (b) dev Curl P ≡ 0 if and only if P = α · Anti(x) + Du, (c) sym Curl P ≡ 0 if and only if P = ζ · 1 + Du. 0 0 3 where α ∈ R, ζ ∈ D (Ω, R) and u ∈ D (Ω, R ). Proof. Part (a) follows by the definition of the matrix Curl which acts row-wise on matrices and the fact that Ω is simply connected from the classical Poincar´elemma, cf. e.g. [29, Theorem 6.17-2] but also the historical remarks therein. Now, let dev Curl P ≡ 0 then Curl P = 2 α·1 with a scalar field α. Since Div Curl P ≡ 0 (in the sense of distributions) we get, taking the matrix Div on both sides, that 2 ∇α = 2 Div(α · 1) = Div Curl P ≡ 0 ⇒ α ≡ const . (A.18) Hence, Curl(P − α · Anti(x)) = Curl P − α · Curl(Anti(x)) = Curl P − 2 α · 1 ≡ 0, so that there exists a vector field u such that P − α · Anti(x) = Du. (A.19) Conversely, we have dev Curl(α · Anti(x) + Du) = α · dev Curl(Anti(x)) = α · dev(2 · 1) ≡ 0, so that part (b) follows. The conclusion of part (c) is obtained in a similar way. Indeed, if sym Curl P ≡ 0 then Curl P = Anti(a) for a vector field a. Taking the matrix Div on both sides we obtain − curl a = Div Anti(a) = Div Curl P ≡ 0. (A.20) Hence, (Ω is a simply connected) a = −∇ζ (A.21) for a scalar field ζ. Moreover, we have Curl(ζ · 1) = − Anti(∇ζ) = Anti(a) = Curl P. (A.22) Thus, Curl(P − ζ · 1) ≡ 0 and therefore there exists a vector field u such that P − ζ · 1 = Du. (A.23) Conversely, we have sym Curl(ζ · 1 + Du) = sym Curl(ζ · 1) = − sym(Anti(∇ζ)) ≡ 0, (A.24) which establishes part (c). 

Remark A.2 (Concerning the kernel of dev sym Curl). It is clear that, Anti(ϕC )+ζ·1+Du belongs to the kernel of dev sym Curl, denotig by ϕC the infinitesimal conformal maps, cf. (1.10). However, it is not clear wether these functions already represent the whole class.

A.4 The kernel of inc sym, inc skew and inc 3 0 3×3 Let Ω ⊂ R be a bounded domain and P ∈ D (Ω, R ). Then the Saint-Venant compatibility conditions give inc sym P ≡ 0 ⇔ sym P = sym Du, (A.25) 0 3 where u ∈ D (Ω, R ). Moreover, we have inc skew P ≡ 0 ⇔ skew P = α · Anti(x) + skew Dv (A.26) 0 3 where α ∈ R and v ∈ D (Ω, R ). However, it would be desirable to obtain in (A.26) an expression as in (A.25) without an additional term in Anti(x). This can be achieved, e.g., assuming additional (boundary) conditions. In a first observation, for 1 skew P = skew Dv we obtain axl skew P = axl skew Dv = 2 curl v, where curl v = 2 axl skew P always has a solution provided that div axl skew P ≡ 0. However, we will see, that inc (skew P ) ≡ 0 implies only div axl skew P ≡ const. Indeed, to establish (A.26) we make use of the expression inc Anti(a) = − Anti(∇ div a) 0 3 valid for all a ∈ D (Ω, R ). Thus, inc Anti(a) ≡ 0 ⇔ ∇ div a ≡ 0 ⇔ div a ≡ const . Therefore, there exists an α ∈ R such that the vector field a(x) − α · x is solenoidal, i.e., div(a(x) − α · x) ≡ 0. Hence, there 0 3 v(x) exists a vector potential v ∈ D (Ω, R ) such that a(x) − α · x = curl 2 . Since Anti(curl v) = 2 skew(Dv) we obtained the expression from (A.26) where we have used a = axl skew P . Conversely, we have inc (α · Anti(x) + skew Dv) ≡ 0, which establishes the relation from (A.26). Furthermore, inc P ≡ 0 ⇔ sym inc P ≡ 0 ∧ skew inc P ≡ 0 ⇔ inc sym P ≡ 0 ∧ inc skew P ≡ 0 (A.25) ⇔ sym P = sym Du ∧ skew P = α · Anti(x) + skew Dv (A.26) ⇔ P = α · Anti(x) + sym Du + skew Dv (A.27) 0 3 where u, v ∈ D (Ω, R ) and α ∈ R.

40 A.5 Rotations and the cross product By definition of Anti it holds Anti(a) b = a × b. Thus, for any rotation R ∈ SO(3) we have 3 Anti(R a) R b = (R a) × (R b) = R(a × b) = R Anti(a) b ∀a, b ∈ R (A.28) and therefore Anti(R a) R = R Anti(a) ⇒ Anti(R a) = R Anti(a)RT . (A.29) It follows that Anti(R a) Anti(R b) = R Anti(a) Anti(b)RT (A.30a) but also dev(Anti(R a) Anti(R b)) = R[dev(Anti(a) Anti(b))]RT , (A.30b) sym(Anti(R a) Anti(R b)) = R[sym(Anti(a) Anti(b))]RT , (A.30c) dev sym(Anti(R a) Anti(R b)) = R[dev sym(Anti(a) Anti(b))]RT , (A.30d) tr(Anti(R a) Anti(R b)) = tr(Anti(a) Anti(b)) . (A.30e) Consequently, we have kAnti(R a) Anti(R b)k = kR Anti(a) Anti(b)RT k = kAnti(a) Anti(b)k (A.31a) and, due to the isotropy of dev, sym and dev sym, also kdev(Anti(R a) Anti(R b))k = kdev(Anti(a) Anti(b))k, (A.31b) ksym(Anti(R a) Anti(R b))k = ksym(Anti(a) Anti(b))k, (A.31c) kdev sym(Anti(R a) Anti(R b))k = kdev sym(Anti(a) Anti(b))k . (A.31d)

Since Anti(α a) = α Anti(a) for all α ∈ R and in regard with the invariance relations above and the quadratic homogeneity, the considerations in section 2.1 can also be obtained with a fixed skew-symmetric matrix, say for a = e3 or, equivalently A = Anti(e3) = e2 ⊗ e1 − e1 ⊗ e2. We demonstrate it in re-proving Observation 2.1, i.e.

1 2 2 2 2 2 2 3 kak kbk ≤ kdev sym(Anti(a) × b)k ≤ kak kbk ∀a, b ∈ R . (A.32) 2 3

By the previous discussion it suffices to establish this estimates already for fixed a = e3. Indeed, we have       0 −1 0 0 −b3 b2 −b3 0 b1 Anti(e3) × b = 1 0 0  b3 0 −b1 =  0 −b3 b2 (A.33) 0 0 0 −b2 b1 0 0 0 0 Hence, 2 2 2 1 2 1 2 kdev sym(Anti(e3) × b)k = b + b + b (A.34) 3 3 2 1 2 2 and therefore 1 2 2 2 2 kbk ≤ kdev sym(Anti(e3) × b)k ≤ kbk . (A.35) 2 3 From the estimate (A.32) it follows that for fixed b 6= 0 the linear map a 7→ dev sym(Anti(a) × b) is invertible. A specific inverse map can be obtained as follows. Set

(2.20) 1 1 M := dev sym(Anti(a) × b) = (a ⊗ b + b ⊗ a) − a, b · 1. 2 3 Thus, 1 1 2 tr M = 0, M b = kbk2a + a, b b, and Mb, b = a, b kbk2 2 6 3 and we have ! 2 1 Mb, b a = L M, where L M := M b − b . (A.36) b b kbk2 4 kbk2

A.6 Fourier transformation and equivalence of spaces 3 The Fourier transform of f on R is given by Z − 3 −ihξ,xi 3 Ff(ξ) = fb(ξ) = (2π) 2 e f(x) dx, ξ ∈ R . (A.37) 3 R 3 3 If f is sufficiently regular, then ∂dj f(ξ) = i ξj fb(ξ). Thus, for a sufficiently regular vector field v : R → R we have

curl\v(ξ) = i ξ × vb(ξ) = i vb(ξ) × (−ξ) (A.38)

41 3 3×3 and for a sufficiently regular matrix field P : R → R it follows

Curl\P (ξ) = −i Pb(ξ) × ξ = −i Pb(ξ) Anti(ξ). (A.39) Consequently,

F sym Curl P (ξ) = −i sym(Pb(ξ) × ξ) and F dev sym Curl P (ξ) = −i dev sym(Pb(ξ) × ξ) . (A.40) Recall, that by (2.30) we have the estimate √ 3 ∀ξ ∈ R : kdev sym(P × ξ)k ≤ ksym(P × ξ)k ≤ (1 + 3)kdev sym(P × ξ)k (A.41) which in regard with (A.40) gives

c kF dev sym Curl P k 2 3 ≤ kF sym Curl P k 2 3 ≤ C kF dev sym Curl P k 2 3 . (A.42) L (R ) L (R ) L (R ) 2 Since by Plancherel’s theorem the Fourier transformation is an isometry of spaces, i.e. the L -norm satisfies kfk 2 3 = L (R ) kfk 2 3 we conclude b L (R )

c kdev sym Curl P k 2 3 3×3 ≤ ksym Curl P k 2 3 3×3 ≤ C kdev sym Curl P k 2 3 3×3 . (A.43) L (R ,R ) L (R ,R ) L (R ,R ) 2 3 3×3 2 3 3×3 In other words dev sym Curl P ∈ L (R , R ) if and only if sym Curl P ∈ L (R , R ), thus, establishing the equivalence of spaces without boundary conditions 1,2 3 3×3 1,2 3 3×3 W (dev sym Curl; R , R ) = W (sym Curl; R , R ), as well as the norm equivalence

kP k 2 3 3×3 + ksym Curl P k 2 3 3×3 ≤ C(kP k 2 3 3×3 + kdev sym Curl P k 2 3 3×3 ) . L (R ,R ) L (R ,R ) L (R ,R ) L (R ,R ) Hence, we conclude

(A.36)    ξ     ξ  skew\P (ξ) = Anti L dev sym (skew\P (ξ)) × = Anti L dev sym (skew Pb(ξ)) × ξ/kξk kξk ξ/kξk kξk    ξ  = Anti L dev sym ((Pb − sym Pb)(ξ)) × ξ/kξk kξk (A.40)   −i    ξ  = Anti L F dev sym Curl P (ξ) − L dev sym (sym\P (ξ)) × . ξ/kξk kξk ξ/kξk kξk

p 3 Thus, standard multiplier estimates will give L estimates (for 1 < p < ∞) in the periodic setting and in the whole space R .

42