Processing math: 46%
Research article Special Issues

Griffith energies as small strain limit of nonlinear models for nonsimple brittle materials

  • We consider a nonlinear, frame indifferent Griffith model for nonsimple brittle materials where the elastic energy also depends on the second gradient of the deformations. In the framework of free discontinuity and gradient discontinuity problems, we prove existence of minimizers for boundary value problems. We then pass to a small strain limit in terms of suitably rescaled displacement fields and show that the nonlinear energies can be identified with a linear Griffith model in the sense of Γ-convergence. This complements the study in [39] by providing a linearization result in arbitrary space dimensions.

    Citation: Manuel Friedrich. Griffith energies as small strain limit of nonlinear models for nonsimple brittle materials[J]. Mathematics in Engineering, 2020, 2(1): 75-100. doi: 10.3934/mine.2020005

    Related Papers:

    [1] Caterina Ida Zeppieri . Homogenisation of high-contrast brittle materials. Mathematics in Engineering, 2020, 2(1): 174-202. doi: 10.3934/mine.2020009
    [2] Stefano Almi, Giuliano Lazzaroni, Ilaria Lucardesi . Crack growth by vanishing viscosity in planar elasticity. Mathematics in Engineering, 2020, 2(1): 141-173. doi: 10.3934/mine.2020008
    [3] Francesco Maddalena, Danilo Percivale, Franco Tomarelli . Signorini problem as a variational limit of obstacle problems in nonlinear elasticity. Mathematics in Engineering, 2024, 6(2): 261-304. doi: 10.3934/mine.2024012
    [4] Luca Placidi, Emilio Barchiesi, Francesco dell'Isola, Valerii Maksimov, Anil Misra, Nasrin Rezaei, Angelo Scrofani, Dmitry Timofeev . On a hemi-variational formulation for a 2D elasto-plastic-damage strain gradient solid with granular microstructure. Mathematics in Engineering, 2023, 5(1): 1-24. doi: 10.3934/mine.2023021
    [5] Giacomo Canevari, Arghir Zarnescu . Polydispersity and surface energy strength in nematic colloids. Mathematics in Engineering, 2020, 2(2): 290-312. doi: 10.3934/mine.2020015
    [6] Federico Cluni, Vittorio Gusella, Dimitri Mugnai, Edoardo Proietti Lippi, Patrizia Pucci . A mixed operator approach to peridynamics. Mathematics in Engineering, 2023, 5(5): 1-22. doi: 10.3934/mine.2023082
    [7] Raimondo Penta, Ariel Ramírez-Torres, José Merodio, Reinaldo Rodríguez-Ramos . Effective governing equations for heterogenous porous media subject to inhomogeneous body forces. Mathematics in Engineering, 2021, 3(4): 1-17. doi: 10.3934/mine.2021033
    [8] Patrick Dondl, Martin Jesenko, Martin Kružík, Jan Valdman . Linearization and computation for large-strain visco-elasticity. Mathematics in Engineering, 2023, 5(2): 1-15. doi: 10.3934/mine.2023030
    [9] Fernando Farroni, Giovanni Scilla, Francesco Solombrino . On some non-local approximation of nonisotropic Griffith-type functionals. Mathematics in Engineering, 2022, 4(4): 1-22. doi: 10.3934/mine.2022031
    [10] Miyuki Koiso . Stable anisotropic capillary hypersurfaces in a wedge. Mathematics in Engineering, 2023, 5(2): 1-22. doi: 10.3934/mine.2023029
  • We consider a nonlinear, frame indifferent Griffith model for nonsimple brittle materials where the elastic energy also depends on the second gradient of the deformations. In the framework of free discontinuity and gradient discontinuity problems, we prove existence of minimizers for boundary value problems. We then pass to a small strain limit in terms of suitably rescaled displacement fields and show that the nonlinear energies can be identified with a linear Griffith model in the sense of Γ-convergence. This complements the study in [39] by providing a linearization result in arbitrary space dimensions.


    Mathematical models in solids mechanics typically do not predict the mechanical behavior correctly at every scale, but have a certain limited range of applicability. A central example in that direction are models for hyperelastic materials in nonlinear (finite) elasticity and their linear (infinitesimal) counterparts. The last decades have witnessed remarkable progress in providing a clear relationship between different models via Γ-convergence [30]. In their seminal work [33], Dal Maso, Negri and Percivale performed a nonlinear-to-linear analysis in terms of suitably rescaled displacement fields and proved the convergence of minimizers for corresponding boundary value problems. This study has been extended in various directions, including different growth assumptions on the stored energy densities [1], the passage from atomistic-to-continuum models [13,55], multiwell energies [2,54], plasticity [51], and viscoelasticity [43].

    In the present contribution, we are interested in an analogous analysis for materials undergoing fracture. Based on the variational approach to quasistatic crack evolution by Francfort and Marigo [37], where the displacements and the (a priori unknown) crack paths are determined from an energy minimization principle, we consider an energy functional of Griffith-type. Such variational models of brittle fracture, which comprise an elastic energy stored in the uncracked region of the body and a surface contribution comparable to the size of the crack of codimension one, have been widely studied both at finite and infinitesimal strains, see [7,18,32,34,38,45,48] without claim of being exhaustive. We refer the reader to [11] for a general overview.

    In this context, first results addressing the question of a nonlinear-to-linear analysis have been obtained in [52,53] in a two-dimensional evolutionary setting for a fixed crack set or a restricted class of admissible cracks, respectively. Subsequently, the problem was studied in [44] from a different perspective. Here, a simultaneous discrete-to-continuum and nonlinear-to-linear analysis is performed for general crack geometries, but under the simplifying assumption that all deformations are close to the identity mapping.

    Eventually, a result in dimension two without a priori assumptions on the crack paths and the deformations, in the general framework of free discontinuity problems (see [35]), has been derived in [39]. This analysis relies fundamentally on delicate geometric rigidity results in the spirit of [22,46]. At this point, the geometry of crack paths in the plane is crucially exploited and higher dimensional analogs seem to be currently out of reach. In spite of the lack of rigidity estimates, the goal of this contribution is to perform a nonlinear-to-linear analysis for brittle materials in the spirit of [39] in higher space dimensions. This will be achieved by starting from a slightly different nonlinear model for so-called nonsimple materials.

    Whereas the elastic properties of simple materials depend only on the first gradient, the notion of a nonsimple material refers to the fact that the elastic energy depends additionally on the second gradient of the deformation. This idea goes back to Toupin [57,58] and has proved to be useful in modern mathematical elasticity, see e.g., [8,9,14,36,43,50], since it brings additional compactness and rigidity to the problem. In a similar fashion, we consider here a Griffith model with an additional second gradient in the elastic part of the energy. This leads to a model in the framework of free discontinuity and gradient discontinuity problems.

    The goal of this contribution is twofold. We first show that the regularization allows to prove existence of minimizers for boundary value problems without convexity properties for the stored elastic energy. In particular, we do not have to assume quasiconvexity [4]. Afterwards, we identify an effective linearized Griffith energy as the Γ-limit of the nonlinear and frame indifferent models for vanishing strains. In this context, it is important to mention that, in spite of the formulation of the nonlinear model in terms of nonsimple materials, the effective limit is a 'standard' Griffith functional in linearized elasticity depending only on the first gradient. A similar justification for the treatment of nonsimple materials has recently been discussed in [43] for a model in nonlinear viscoelasticity.

    The existence result for boundary value problems at finite strains is formulated in the space GSBV22(Ω;Rd), see (2.2) below, consisting of the mappings for which both the function itself and its derivative are in the class of generalized special functions of bounded variation [6]. The relevant compactness and lower semicontinuity results stated in Theorem 3.3 essentially follow from a study on second order variational problems with free discontinuity and gradient discontinuity [16]. Another key ingredient is the recent work [42] which extends the classical compactness result due to Ambrosio [3] to problems without a priori bounds on the functions.

    Concerning the passage to the linearized system, the essential step is to establish a compactness result in terms of suitably rescaled displacement fields which measure the distance of the deformations from the identity. Whereas in [39] this is achieved by means of delicate geometric rigidity estimates, the main idea in our approach is to partition the domain into different regions in which the gradient is 'almost constant'. This construction relies on the coarea formula in BV and is the fundamental point where the presence of a second order term in the energy is used to pass rigorously to a linear theory. The linear limiting model is formulated on the space of generalized special functions of bounded deformation GSBD2, which has been studied extensively over the last years, see e.g., [19,20,21,23,24,25,26,27,28,31,40,41,45,49].

    The paper is organized as follows. In Section 2 we first introduce our nonlinear model for nonsimple brittle materials and state our main results: We first address the existence of minimizers for boundary value problems at finite strains. Then, we present a compactness and Γ-convergence result in the passage from the nonlinear to the linearized theory. Here, we also discuss the convergence of minima and minimzers under given boundary data. Section 3 is devoted to some preliminary results about the function spaces GSBV and GSBD. In particular, we present a compactness result in GSBV22 involving the second gradient (see Theorem 3.3). Finally, Section 4 contains the proofs of our results.

    In this section we introduce our model and present the main results. We start with some basic notation. Throughout the paper, ΩRd is an open and bounded set. The notations Ld and Hd1 are used for the Lebesgue measure and the (d1)-dimensional Hausdorff measure in Rd, respectively. We set Sd1={xRd:|x|=1}. For an Ld-measurable set ERd, the symbol χE denotes its indicator function. For two sets A,BRd, we define AB=(AB)(BA). The identity mapping on Rd is indicated by id and its derivative, the identity matrix, by IdRd×d. The sets of symmetric and skew symmetric matrices are denoted by Rd×dsym and Rd×dskew, respectively. We set sym(F)=12(FT+F) for FRd×d and define SO(d)={RRd×d:RTR=Id,detR=1}.

    In this subsection we introduce our nonlinear model and discuss the existence of minimizers for boundary value problems.

    Function spaces: To introduce our Griffith-type model for nonsimple materials, we first need to introduce the relevant spaces. We use standard notation for GSBV functions, see [6,Section 4] and [32,Section 2]. In particular, we let

    GSBV2(Ω;Rd)={yGSBV(Ω;Rd): yL2(Ω;Rd×d), Hd1(Jy)<+}, (2.1)

    where y(x) denotes the approximate differential at Ld-a.e. xΩ and Jy the jump set. We define the space

    GSBV22(Ω;Rd):={yGSBV2(Ω;Rd): yGSBV2(Ω;Rd×d)}. (2.2)

    The approximate differential and the jump set of y will be denoted by 2y and Jy, respectively. (To avoid confusion, we point out that in the paper [32] the notation GSBV22(Ω;Rd) was used differently, namely for GSBV2(Ω;Rd)L2(Ω;Rd).)

    A similar space has been considered in [15,16] to treat second order free discontinuity functionals, e.g., a weak formulation of the Blake & Zissermann model [10] of image segmentation. We point out that the functions are allowed to exhibit discontinuities. Thus, the analysis is outside of the framework of the space of special functions with bounded Hessian SBH(Ω), considered in problems of second order energies for elastic-perfectly plastic plates, see e.g., [17].

    Nonlinear Griffith energy for nonsimple materials: We let W:Rd×d[0,+) be a single well, frame indifferent stored energy functional. More precisely, we suppose that there exists c>0 such that

    (i)  W continuous and C3 in a neighborhood of SO(d),(ii)  Frame indifference: W(RF)=W(F) for all FRd×d,RSO(d),(iii)  W(F)cdist2(F,SO(d))  for all FRd×d, W(F)=0 iff FSO(d). (2.3)

    We briefly note that we can also treat inhomogeneous materials where the energy density has the form W:Ω×Rd×d[0,+). Moreover, it suffices to assume WC2,α, where C2,α is the Hölder space with exponent α(0,1], see Remark 4.2 for details.

    Let κ>0 and β(23,1). For ε>0, define the energy Eε(,Ω):GSBV22(Ω;Rd)[0,+] by

    Eε(y,Ω)={ε2ΩW(y(x))dx+ε2βΩ|2y(x)|2dx+κHd1(Jy) if JyJy+ else. (2.4)

    Here and in the following, the inclusion JyJy has to be understood up to an Hd1-negligible set. Since W grows quadratically around SO(d), the parameter ε corresponds to the typical scaling of strains for configurations with finite energy.

    Due to the presence of the second term, we deal with a Griffith-type model for nonsimple materials. As explained in the introduction, elastic energies which depend additionally on the second gradient of the deformation were introduced by Toupin [57,58] to enhance compactness and rigidity properties. In the present context, we add a second gradient term for a material undergoing fracture. This regularization effect acts on the entire intact region ΩJy of the material. This is modeled by the condition JyJy.

    The goal of this contribution is twofold. We first show that the regularization allows to prove existence of minimizers for boundary value problems without convexity properties of W. The main result of the present work is then to identify a linearized Griffith energy in the small strain limit ε0 which is related to the nonlinear energies Eε through Γ-convergence. We point out that the effective limit is a 'standard' Griffith model in linearized elasticity depending only on the first gradient, see (2.14) below, although we start with a nonlinear model for nonsimple materials.

    We observe that the condition JyJy is not closed under convergence in measure on Ω. In fact, consider, e.g., Ω=(1,1)2,Ω1=(1,0)×(1,1),Ω2=(0,1)×(1,1), and for δ0 the configurations

    yδ(x1,x2)=(x1,x2)χΩ1+(2x1+δ,x2)χΩ2      for (x1,x2)Ω.

    Then Jyδ=Jyδ={0}×(1,1) for δ>0 and yδy0 in measure on Ω as δ0. However, there holds =Jy0Jy0={0}×(1,1). Therefore, we need to pass to a relaxed formulation.

    Proposition 2.1 (Relaxation). Let ΩRd be open and bounded. Suppose that W satisfies (2.3). Then the relaxed functional ¯Eε(,Ω):GSBV22(Ω;Rd)[0,+] defined by

    ¯Eε(y,Ω)=inf{lim infnEε(yn,Ω):yny in measure on Ω}

    is given by

    ¯Eε(y,Ω)=ε2ΩW(y(x))dx+ε2βΩ|2y(x)|2dx+κHd1(JyJy). (2.5)

    The result is proved in Subsection 4.1. Clearly, ¯Eε(,Ω) is lower semicontinuous with respect to the convergence in measure. We point out that this latter property has essentially been shown in [16], cf. Theorem 3.2.

    In the following, our goal is to study boundary value problems. To this end, we suppose that there exist two bounded Lipschitz domains ΩΩ. We will impose Dirichlet boundary data on DΩ:=ΩΩ. As usual for the weak formulation in the framework of free discontinuity problems, this will be done by requiring that configurations y satisfy y=g on Ω¯Ω for some gW2,(Ω;Rd). From now on, we write Eε()=Eε(,Ω) and ¯Eε()=¯Eε(,Ω) for notational convenience. The following result about existence of minimizers will be proved in Subsection 4.1.

    Theorem 2.2 (Existence of minimizers). Let ΩΩRd be bounded Lipschitz domains. Suppose that W satisfies (2.3), and let gW2,(Ω;Rd). Then the minimization problem

    infyGSBV22(Ω;Rd){¯Eε(y): y=g on Ω¯Ω} (2.6)

    admits solutions.

    The main goal of the present work is the identification of an effective linearized Griffith energy in the small strain limit. In this subsection, we formulate the relevant compactness result. Let ΩΩ be bounded Lipschitz domains. The limiting energy is defined on the space of generalized special functions of bounded deformation GSBD2(Ω). For basic properties of GSBD2(Ω) we refer to [31] and Section 3.3 below. In particular, for uGSBD2(Ω), we denote by e(u)=12(uT+u) the approximate symmetric differential and by Ju the jump set.

    The general idea in linearization results in many different settings (see, e.g., [2,13,33,43,44,52,54,55]) is the following: Given a sequence (yε)ε with supεEε(yε)<+, define displacement fields which measure the distance of the deformations from the identity, rescaled by the small parameter ε, i.e.,

    uε=1ε(yεid). (2.7)

    It turns out, however, that in general no compactness can be expected if the body may undergo fracture. Consider, e.g., the functions yε=idχΩB+RxχB, for a small ball BΩ and a rotation RSO(d), RId. Then |uε|,|uε| on B as ε0. The main idea in our approach is the observation that this phenomenon can be avoided if the deformation is rotated back to the identity on the set B. This will be made precise in Theorem 2.3(a) below where we pass to piecewise rotated functions. For such functions, we can control at least the symmetric part of uε for the rescaled displacement fields defined in (2.7). This will allow us to derive a compactness result in the space GSBD2(Ω), see Theorem 2.3(b).

    Recall the definition of GSBV22(Ω;Rd) in (2.2). To account for boundary data hW2,(Ω;Rd), we introduce the spaces

    Sε,h={yGSBV22(Ω;Rd): y=id+hε on Ω¯Ω},GSBD2h={uGSBD2(Ω):u=h on Ω¯Ω}. (2.8)

    Recall β(23,1) and the definition of ¯Eε=¯Eε(,Ω) in (2.5). For definition and basic properties of Caccioppoli partitions we refer to Section 3.1. In particular, for a set of finite perimeter EΩ, we denote by E its essential boundary and by (E)1 the points where E has density one, see [6,Definition 3.60].

    Theorem 2.3 (Compactness). Let γ(23,β). Assume that W satisfies (2.3), and let hW2,(Ω;Rd). Let (yε)ε be a sequence satisfying yεSε,h and supε¯Eε(yε)<+.

    (a) (Piecewise rotated functions) There exist Caccioppoli partitions (Pεj)j of Ω and corresponding rotations (Rjε)jSO(d) such that the piecewise rotated functions yrotεGSBV22(Ω;Rd) given by

    yrotε:=j=1RjεyεχPjε (2.9)

    satisfy

    (i)  yrotε=id+hε on Ω¯Ω(ii)  Hd1((JyrotεJyrotε)(JyεJyε))Hd1((Ωj=1Pεj)Jyε)Cεβγ,(iii)  sym(yrotε)IdL2(Ω)Cε,(iv)  yrotεIdL2(Ω)Cεγ (2.10)

    for a constant C>0 independent of ε.

    (b) (Compactness of rescaled displacement fields) There exists a subsequence (not relabeled) and a function uGSBD2h such that the rescaled displacement fields uεGSBV22(Ω;Rd) defined by

    uε:=1ε(yrotεid) (2.11)

    satisfy

    (i)  uεu   a.e. in ΩEu,(ii)  e(uε)e(u)  weakly in L2(ΩEu;Rd×dsym),(iii)  Hd1(Ju)lim infε0Hd1(Juε)lim infε0Hd1(JyεJyε),(iv)  e(u)=0  on Eu,   Hd1((EuΩ)Ju)=Hd1(Ju(Eu)1)=0, (2.12)

    where Eu:={xΩ:|uε(x)|} is a set of finite perimeter.

    Here and in the sequel, we follow the usual convention that convergence of the continuous parameter ε0 stands for convergence of arbitrary sequences {iε}i with iε0 as i, see [12,Definition 1.45]. The compactness result will be proved in Subsection 4.2.

    Note that (2.10)(ⅰ) implies yrotεSε,h. In view of (2.10)(ⅱ), the frame indifference of the elastic energy, and γ<β, one can show that the Griffith-type energy (2.5) of yrotε is asymptotically not larger than the one of yε. The control on the symmetric part of the derivative (2.10)(ⅲ) is essential to obtain compactness in GSBD2(Ω) for the sequence (uε)ε. Property (2.10)(ⅳ) will be needed to control higher order terms in the passage to linearized elastic energies, see Theorem 2.7 below.

    The presence of the set Eu is due to the compactness result in GSBD2(Ω), see [26] and Theorem 3.4. In principle, the phenomenon that the sequence is unbounded on a set of positive measure can be avoided by generalizing the definition of (2.11): In [45,Theorem 6.1] and [39,Theorem 2.2] it has been shown that, by subtracting in (2.11) suitable translations on a Caccioppoli partition of Ω related to yε, one can achieve Eu=. This construction, however, is limited so far to dimension two. As discussed in [26], the presence of Eu is not an issue for minimization problems of Griffith energies since a minimizer can be recovered by choosing u affine on Eu with e(u)=0, cf. (2.12)(ⅳ). We also note that EuΩ, i.e., Eu(Ω¯Ω)=.

    Definition 2.4 (Asymptotic representation). We say that a sequence (yε)ε with yεSε,h is asymptotically represented by a limiting displacement uGSBD2h, and write yεu, if there exist sequences of Caccioppoli partitions (Pεj)j of Ω and corresponding rotations (Rjε)jSO(d) such that (2.10) and (2.12) hold for some fixed γ(23,β), where yrotε and uε are defined in (2.9) and (2.11), respectively.

    Theorem 2.3 shows that for each (yε)ε with supε¯Eε(yε)<+ there exists a subsequence (ykε)k and uGSBD2h such that ykεu as k. We speak of asymptotic representation instead of convergence, and we use the symbol , in order to emphasize that Definition 2.4 cannot be understood as a convergence with respect to a certain topology. In particular, the limiting function u for a given (sub-)sequence (yε)ε is not determined uniquely, but depends fundamentally on the choice of the sequences (Pεj)j and (Rjε)j. To illustrate this phenomenon, we consider an example similar to [39,Example 2.4].

    Example 2.5 (Nonuniqueness of limits). Consider Ω=(0,3)×(0,1), Ω=(1,3)×(0,1), Ω1=(0,2)×(0,1), Ω2=(2,3)×(0,1), h0, and

    yε(x)=xχΩ1(x)+ˉRεxχΩ2(x)    for xΩ,

    where ˉRεSO(2) with ˉRε=Id+Aε+O(ε2) for some AR2×2skew. Then two possible alternatives are

    (1)  Pε1=Ω1,  Pε2=Ω2,  Rε1=Id,  Rε2=ˉR1ε,(2)  ˜Pε1=Ω,  ˜Rε1=Id.

    Letting uε=ε1(2j=1RεjyεχPεjid) and ˜uε=ε1(yεid), we find the limits u0 and ˜u(x)=AxχΩ2(x), respectively.

    We refer to [39,Section 2.3] for a further discussion about different choices of the involved partitions and rigid motions. Here, we show that it is possible to identify uniquely the relevant notions e(u) and Ju of the limit. This is content of the following lemma.

    Lemma 2.6 (Characterization of limiting displacements). Suppose that a sequence (yε)ε satisfies yεu1 and yεu2, where u1,u2GSBD2h, u1u2. Let Eu1,Eu2Ω be the sets given in (2.12). Then

    (a) e(u1)=e(u2) Ld-a.e. on Ω(Eu1Eu2).

    (b) If additionally (yε)ε is a minimizing sequence, i.e.,

    ¯Eε(yε)infˉySε,h¯Eε(ˉy)+ρε     with ρε0as ε0, (2.13)

    then e(u1)=e(u2) Ld-a.e. on Ω, and Ju1=Ju2 up to an Hd1-negligible set.

    Note that property (a) is consistent with Example 2.5. Example 2.5 also shows that the property Ju1=Ju2 is not satisfied in general but some extra condition, e.g., the one in (2.13), is necessary. We refer to Example 4.3 below for an illustration that in case (a) the strains are not necessarily the same inside Eu1Eu2. The result will be proved in Subsection 4.4.

    We now show that the nonlinear energies of Griffith-type can be related to a linearized Griffith model in the small strain limit by Γ-convergence. We also discuss the convergence of minimizers for boundary value problems. Given bounded Lipschitz domains ΩΩ, we define the energy E:GSBD2(Ω)[0,+) by

    E(u)=Ω12Q(e(u))+κHd1(Ju), (2.14)

    where κ>0, and Q:Rd×d[0,+) is the quadratic form Q(F)=D2W(Id)F:F for all FRd×d. In view of (2.3), Q is positive definite on Rd×dsym and vanishes on Rd×dskew.

    For the Γ-limsup inequality, more precisely for the application of the density result stated in Theorem 3.6, we make the following geometrical assumption on the Dirichlet boundary DΩ=ΩΩ: there exists a decomposition Ω=DΩNΩN with

    DΩ,NΩ relatively open,   Hd1(N)=0,   DΩNΩ=,   (DΩ)=(NΩ), (2.15)

    and there exist ˉδ>0 small and x0Rd such that for all δ(0,ˉδ) there holds

    Oδ,x0(DΩ)Ω, (2.16)

    where Oδ,x0(x):=x0+(1δ)(xx0).

    We now present our main Γ-convergence result. Recall Definition 2.4, as well as the definition of the nonlinear energies in (2.4) and (2.5). Moreover, recall the spaces Sε,h and GSBD2h in (2.8) for hW2,(Ω;Rd).

    Theorem 2.7 (Passage to linearized model). Let ΩΩRd be bounded Lipschitz domains. Suppose that W satisfies (2.3) and that (2.15)–(2.16) hold. Let hW2,(Ω;Rd).

    (a) (Compactness) For each sequence (yε)ε with yεSε,h and supεEε(yε)<+, there exists a subsequence (not relabeled) and uGSBD2h such that yεu.

    (b) (Γ-liminf inequality) For each sequence (yε)ε, yεSε,h, with yεu for some uGSBD2h we have

    lim infε0Eε(yε)E(u).

    (c) (Γ-limsup inequality) For each uGSBD2h there exists a sequence (yε)ε, yεSε,h, such that yεu and

    limε0Eε(yε)=E(u).

    The same statements hold with ¯Eε in place of Eε.

    We point out that we identify a 'standard' Griffith energy in linearized elasticity although we departed from a nonlinear model for nonsimple materials. As a corollary, we obtain the convergence of minimizers for boundary value problems.

    Corollary 2.8 (Minimization problems). Consider the setting of Theorem 2.7. Then

    infˉySε,hEε(ˉy)  minuGSBD2hE(u) (2.17)

    as ε0. Moreover, for each sequence (yε)ε with yεSε,h satisfying

    Eε(yε)infˉySε,hEε(ˉy)+ρε     with ρε0as ε0, (2.18)

    there exist a subsequence (not relabeled) and uGSBD2h with E(u)=minvGSBD2hE(v) such that yεu.

    The results announced in this subsection will be proved in Subsection 4.3.

    In this section we collect some fundamental properties about (generalized) special functions of bounded variation and deformation. In particular, we recall and prove some results for GSBV22 and GSBD2 that will be needed for the proofs in Section 4.

    We say that a partition (Pj)j of an open set ΩRd is a Caccioppoli partition of Ω if jHd1(Pj)<+, where Pj denotes the essential boundary of Pj (see [6,Definition 3.60]). The local structure of Caccioppoli partitions can be characterized as follows (see [6,Theorem 4.17]).

    Theorem 3.1. Let (Pj)j be a Caccioppoli partition of Ω. Then

    j(Pj)1ij(PiPj)

    contains Hd1-almost all of Ω.

    Here, (P)1 denote the points where P has density one (see again [6,Definition 3.60]). Essentially, the theorem states that Hd1-a.e. point of Ω either belongs to exactly one element of the partition or to the intersection of exactly two sets Pi, Pj.

    For the general notions on SBV and GSBV functions and their properties we refer to [6,Section 4]. For ΩRd open and mN, we define GSBV2(Ω;Rm) as in (2.1), for general m. We denote by y the approximate differential and by Jy the set of approximate jump points of y, which is an Hd1-rectifiable set. We recall that GSBV2(Ω;Rm) is a vector space, see [32,Proposition 2.3]. In a similar fashion, we say ySBV2(Ω;Rm) if ySBV(Ω;Rm), yL2(Ω;Rm×d), and Hd1(Jy)<+.

    We define GSBV22(Ω;Rm) as in (2.2), for general m. For m=1 we write GSBV22(Ω). By definition, yGSBV2(Ω;Rm×d), and we use the notation 2y and Jy for the approximate differential and the jump set of y, respectively. Applying [32,Proposition 2.3] on y and y, we find that GSBV22(Ω;Rm) is a vector space. The following result is the key ingredient for the proof of Proposition 2.1.

    Theorem 3.2 (Compactness in GSBV22). Let ΩRd be open and bounded, and let mN. Let (yn)n be a sequence in GSBV22(Ω;Rm). Suppose that there exists a continuous, increasing function ψ:[0,)[0,) with limtψ(t)=+ such that

    supnN(Ωψ(|yn|)dx+Ω|2yn|2dx+Hd1(JynJyn))<+.

    Then there exist a subsequence, still denoted by (yn)n, and a function y[GSBV(Ω)]m with yGSBV2(Ω;Rm×d) such that for all 0<γ2γ12γ2 there holds

    (i)  ynya.e. in Ω,(ii)  ynya.e. Ω,(iii)  2yn2yweakly in L2(Ω;Rm×d×d),(iv)  γ1Hd1(Jy)+γ2Hd1(JyJy)lim infn(γ1Hd1(Jyn)+γ2Hd1(JynJyn)). (3.1)

    If in addition supnNynL2(Ω)<+, then yGSBV22(Ω;Rm).

    Proof. First, we observe that it suffices to treat the case m=1 since otherwise one may argue componentwise, see particularly [38,Lemma 3.1] how to deal with property (ⅳ). The result has been proved in [16,Theorem 4.4,Theorem 5.13,Remark 5.14] with the only difference that we just assume supnNΩψ(|yn|)dx<+ here instead of supnNynL2(Ω)<+. We briefly indicate the necessary adaptions in the proof of [16,Theorem 4.4] for m=1. To ease comparison with [16], we point out that in that paper the notation GSBV2(Ω) is used for functions u with uGSBV(Ω) and u[GSBV(Ω)]d.

    For kN, we define some φkC2(R) by φk(t)=t for t[k+1,k1], |φk(t)|=k for |t|>k+1, and 0φk1. By φkynL1(Ω)kLd(Ω) and by using an interpolation inequality one can check that (φkyn)n is bounded in BVloc(Ω), see [16,(4.8)]. Therefore, by a diagonal argument there exist a subsequence of (yn)n and functions wkBVloc(Ω) for all kN such that

    φkynwk   a.e. in Ω for all kN. (3.2)

    Since ψ is continuous and increasing, and |φk(t)||t| for all tR, we also get by Fatou's lemma

    ψ(|wk|)L1(Ω)lim infnψ(|φkyn|)L1(Ω)supnNΩψ(|yn|)dx<+. (3.3)

    Let Ek={|wk|<k1}. The properties of φk along with (3.2) imply

    ynwk   a.e. in Ek for all kN,     wk=wl    on Ek for all kl (3.4)

    By using (3.3) we observe that Ld(ΩEk)0 as k since limtψ(t)=+. This together with (3.4) shows that the measurable function y:ΩR defined by y:=limkwk satisfies y=wk on Ek for all kN and therefore

    yny   a.e. in Ω.

    The rest of the proof starting with [16,(4.10)] remains unchanged. In [16], it has been shown that yGSBV(Ω) and y[GSBV(Ω)]d. Since 2yL2(Ω;Rd×d) and Hd1(Jy)<+, we actually get yGSBV2(Ω;Rd). Finally, given an additional control on (yn)n in L2, we also find yL2(Ω;Rd) and Hd1(Jy)<+. This implies yGSBV22(Ω), see (2.2).

    We now proceed with a version of Theorem 3.2 without a priori bounds on the functions. We also take boundary data into account. The result relies on Theorem 3.2 and [42].

    Theorem 3.3 (Compactness in GSBV22 without a priori bounds). Let ΩΩRd be bounded Lipschitz domains, and let mN. Let gW2,(Ω;Rm). Consider (yn)nGSBV22(Ω;Rm) with yn=g on Ω¯Ω and

    supnN(Ω(|yn|2+|2yn|2)dx+Hd1(JynJyn))<+.

    Then we find a subsequence (not relabeled), modifications (zn)nGSBV22(Ω;Rm) satisfying zn=g on Ω¯Ω and

    (i)  zn=gonSn:={znyn}{2zn2yn},    where Ld(Sn)0 as n,(ii)  limnHd1((JznJzn)(JynJyn))=0, (3.5)

    as well as a limiting function yGSBV22(Ω;Rm) with y=g on Ω¯Ω such that

    (i)  zny in measure on Ω(ii)  znya.e. Ω and zny weakly in L2(Ω;Rm×d)(iii)  2zn2y weakly in L2(Ω;Rm×d×d)(iv)  Hd1(JyJy)lim infnHd1(JznJzn). (3.6)

    In general, it is indispensable to pass to modifications. Consider, e.g., the sequence yn=nχU for some set UΩ of finite perimeter. The idea in [42,Theorem 3.1], where this result is proved in the space GSBV2(Ω;Rm), relies on constructing modifications (zn)n by (cf. [42,(37)–(38)])

    zn=gχRn+j1(yntnj)χPnj (3.7)

    for Caccioppoli partitions Ω=j1PnjRn, and suitable translations (tnj)j1Rm, where

    (i)  limnLd(Rn)=0,(ii)  limnHd1(JznJyn)=limnHd1((RnΩ)Jyn)=0. (3.8)

    Proof of Theorem 3.3. We briefly indicate the necessary adaptions with respect to [42,Theorem 3.1] to obtain the result in the frame of GSBV22(Ω;Rm) involving second derivatives. First, by [42,Theorem 3.1] we find modifications (zn)n as in (3.7) satisfying zn=g on Ω¯Ω and yGSBV2(Ω;Rm) such that zny in measure on Ω, up to passing to a subsequence. By (3.8) we get (3.5).

    As zny in measure on Ω, [45,Remark 2.2] implies that there exists a continuous, increasing function ψ:[0,)[0,) with limtψ(t)=+ such that up to subsequence (not relabeled) supnNΩψ(|zn|)dx<+. Moreover, by the assumptions on yn, (3.5), and the fact that gW2,(Ω;Rm) we get that zn and 2zn are uniformly controlled in L2, as well as supnNHd1(JznJzn)<+. Then Theorem 3.2 yields yGSBV22(Ω;Rm). Along with (3.1) for γ1=γ2 we also get (3.6), apart from the weak convergence of (zn)n. The weak convergence readily follows from supnNznL2(Ω)supnNynL2(Ω)+gL2(Ω)<+.

    We refer the reader to [5] and [31] for the definition, notations, and basic properties of SBD and GSBD functions, respectively. Here, we only recall briefly some relevant notions which can be defined for generalized functions of bounded deformation: let ΩRd open and bounded. In [31,Theorem 6.2 and Theorem 9.1] it is shown that for uGSBD(Ω) the jump set Ju is Hd1-rectifiable and that an approximate symmetric differential e(u)(x) exists at Ld-a.e. xΩ. We define the space GSBD2(Ω) by

    GSBD2(Ω):={uGSBD(Ω):e(u)L2(Ω;Rd×dsym),Hd1(Ju)<+}.

    The space GSBD2(Ω) is a vector subspace of the vector space of Ld-measurable function, see [31,Remark 4.6]. Moreover, there holds GSBV2(Ω;Rd)GSBD2(Ω). The following compactness result in GSBD2 has been proved in [26].

    Theorem 3.4 (GSBD2 compactness). Let ΩRd be open, bounded. Let (un)nGSBD2(Ω) be a sequence satisfying

    supnN(e(un)L2(Ω)+Hd1(Jun))<+.

    Then there exists a subsequence (not relabeled) such that the set A:={xΩ:|un(x)|} has finite perimeter, and there exists uGSBD2(Ω) such that

    (i)  unu    in measure on ΩA,(ii)  e(un)e(u)   weakly in L2(ΩA;Rd×dsym),(iii)  lim infnHd1(Jun)Hd1(Ju(AΩ)). (3.9)

    We briefly remark that (3.9)(ⅰ) is slightly weaker with respect to (3.6)(ⅰ) in Theorem 3.3 (or the corresponding version in GSBV, see [42]) in the sense that there might be a set A where the sequence (un)n is unbounded, cf. the example below Theorem 3.3. This phenomenon is avoided in Theorem 3.3 by passing to suitable modifications which consists in subtracting piecewise constant functions, see (3.7). We point out that an analogous result in GSBD2 is so far only available in dimension two, see [45,Theorem 6.1]. We now state two density results.

    Theorem 3.5 (Density). Let ΩRd be a bounded Lipschitz domain. Let uGSBD2(Ω). Then there exists a sequence (un)nSBV2(Ω;Rd)L(Ω;Rd) such that each Jun is closed and included in a finite union of closed connected pieces of C1 hypersurfaces, each un belongs to C(¯ΩJun;Rd)Wm,(ΩJun;Rd) for every mN, and the following properties hold:

    (i)  unuin measure on Ω,(ii)  e(un)e(u)L2(Ω)0,(iii)  Hd1(JunJu)0.

    Proof. The result follows by combining [25,Theorem 1.1] and [28,Theorem 1.1]. First, [25,Theorem 1.1] yields an approximation un satisfying u_n \in SBV^2(\Omega; \Bbb R^d) \cap W^{1, \infty}({\Omega} \setminus J_{u_n}; \Bbb R^d) , and then [28,Theorem 1.1] gives the higher regularity.

    An adaption of the proof allows to impose boundary conditions on the approximating sequence. Suppose that the Lipschitz domains \Omega \subset \Omega' satisfy the conditions introduced in (2.15)–(2.16). By \mathcal{W}(\Omega; \Bbb R^d) we denote the space of all functions u \in SBV(\Omega; \Bbb R^d) such that J_u is a finite union of disjoint (d-1) -simplices and u \in W^{k, \infty}(\Omega \setminus J_u; \Bbb R^d) for every k \in \Bbb N .

    Theorem 3.6 (Density with boundary data). Let \Omega \subset \Omega' \subset \Bbb R^d be bounded Lipschitz domains satisfying (2.15)–(2.16). Let g \in W^{r, \infty}(\Omega') for r \in \Bbb N . Let u \in GSBD^2(\Omega') with u = g on \Omega' \setminus \overline{\Omega} . Then there exists a sequence of functions (u_n)_n \subset SBV^2(\Omega; \Bbb R^d) , a sequence of neighborhoods (U_n)_n \subset \Omega' of \Omega' \setminus \Omega , and a sequence of neighborhoods (\Omega_n)_n \subset \Omega of \Omega \setminus U_n such that u_n = g on \Omega' \setminus \overline{\Omega} , u_n|_{U_n} \in W^{r, \infty}(U_n; \Bbb R^d) , and u_n|_{\Omega_n} \in \mathcal{W}(\Omega_n; \Bbb R^d) , and the following properties hold:

    \begin{align} {\rm (i)} & \ \ u_n \to u \mathit{\text{in measure on }} \Omega', \\ {\rm(ii)} & \ \ \Vert e(u_n) - e(u) \Vert_{L^2(\Omega')} \to 0, \\ {\rm (iii)} & \ \ \mathcal{H}^{d-1}(J_{u_n}) \to \mathcal{H}^{d-1}(J_u). \end{align} (3.10)

    In particular, u_n \in W^{r, \infty}(\Omega \setminus J_{u_n}; \Bbb R^d) .

    Proof. The fact that u can be approximated by a sequence (u_n)_n \subset SBV^2(\Omega'; \Bbb R^d) \cap L^\infty(\Omega; \Bbb R^d) satisfying (3.10) and u_n = g in a neighborhood of \Omega' \setminus {\Omega} has been addressed in [25,Proof of Theorem 5.4]. Here, also the necessity of the geometric assumptions (2.15)–(2.16) is discussed, see [25,Remark 5.6]. The fact that the approximating sequence can be chosen as in the statement then follows by applying on each u_n a construction very similar to the one of [47,Proposition 2.5] along with a diagonal argument. This construction consists in a suitable cut-off construction and the application of the density result [29]. We also refer to [56,Theorem 3.5] for a similar statement.

    This section contains the proofs of our results.

    In this subsection we prove Proposition 2.1 and Theorem 2.2.

    Proof of Proposition 2.1. For y \in GSBV_2^2(\Omega; \Bbb R^d) we define

    \begin{align} \mathcal{E}'_\varepsilon(y) = \inf \big\{ \liminf\nolimits_{n \to \infty} \mathcal{E}_\varepsilon(y_n, \Omega): y_n \to y \ \text{ in measure on $\Omega$}\big\}, \end{align} (4.1)

    and define \overline{\mathcal{E}}_\varepsilon(\cdot, \Omega) as in (2.5). We need to check that \mathcal{E}'_\varepsilon = \overline{\mathcal{E}}_\varepsilon(\cdot, \Omega) . In the proof, we write \tilde{\subset} and \tilde{ = } for brevity if the inclusion or the identity holds up to an \mathcal{H}^{d-1} -negligible set, respectively.

    Step 1: \mathcal{E}'_\varepsilon \ge \overline{\mathcal{E}}_\varepsilon(\cdot, \Omega) . Since by definition \overline{\mathcal{E}}_\varepsilon(y, \Omega) \le {\mathcal{E}}_\varepsilon(y, \Omega) for all y \in GSBV^2_2(\Omega; \Bbb R^d) , see (2.4), it suffices to confirm that \overline{\mathcal{E}}_\varepsilon(\cdot, \Omega) is lower semicontinous with respect to the convergence in measure. To see this, consider (y_n)_n \subset GSBV^2_2(\Omega; \Bbb R^d) with y_n \to y in measure \Omega and \sup_{n\in \Bbb N}\overline{\mathcal{E}}_\varepsilon(y_n, \Omega) < + \infty . By using [45,Remark 2.2], there exists a continuous, increasing function \psi: [0, \infty) \to [0, \infty) with \lim_{t \to \infty} \psi(t) = + \infty such that up to subsequence (not relabeled) \sup_{n \in \Bbb N}\int_{\Omega} \psi(|y_n|)\, dx < + \infty . Then from Theorem 3.2 we obtain

    \overline{\mathcal{E}}_\varepsilon(y, \Omega) \le \liminf\limits_{n \to \infty} \overline{\mathcal{E}}_\varepsilon(y_n, \Omega).

    In fact, for the second and the third term in (2.5) we use (3.1)(ⅲ) and (ⅳ) for \gamma_1 = \gamma_2 , respectively. The first term in (2.5) is lower semicontinuous by the continuity of W , (3.1)(ⅱ), and Fatou's lemma. This shows that \overline{\mathcal{E}}_\varepsilon(\cdot, \Omega) is lower semicontinous and concludes the proof of \mathcal{E}'_\varepsilon \ge \overline{\mathcal{E}}_\varepsilon(\cdot, \Omega) .

    Step 2: \mathcal{E}'_\varepsilon \le \overline{\mathcal{E}}_\varepsilon(\cdot, \Omega) . In the proof, we will use the following argument several times: If y_1, y_2 \in GSBV^2(\Omega; \Bbb R^d) , then for a.e. t \in \Bbb R there holds that z: = y_1 + ty_2\in GSBV^2(\Omega; \Bbb R^d) satisfies J_z = J_{y_1} \cup J_{y_2} , see [38,Proof of Lemma 3.1] or [32,Proof of Lemma 4.5] for such an argument. We point out that here we exploit the fact that GSBV^2(\Omega; \Bbb R^d) is a vector space.

    Observe that for each y \in GSBV^2_2(\Omega; \Bbb R^d) and each \nu \in S^{d-1} , the function v : = \nabla y \cdot \nu lies in GSBV^2(\Omega; \Bbb R^d) \subset GSBD^2(\Omega) . We can choose \nu \in S^{d-1} such that there holds \mathcal{H}^{d-1}(J_v \triangle J_{\nabla y}) = 0 . We apply Theorem 3.5 to approximate v \in GSBD^2(\Omega) by a sequence (v_n)_n \subset SBV^2(\Omega; \Bbb R^d) such that v_n \in W^{2, \infty}({\Omega} \setminus J_{v_n}; \Bbb R^d) and

    \begin{align} \mathcal{H}^{d-1}(J_{v_n} \triangle J_{\nabla y} ) = \mathcal{H}^{d-1}(J_{v_n} \triangle J_{v} ) \to 0 \end{align} (4.2)

    as n \to \infty . We point out that J_{\nabla v_n} \tilde{\subset} J_{v_n} since v_n \in W^{2, \infty}({\Omega} \setminus J_{v_n}; \Bbb R^d) . Using v_n \in W^{2, \infty}({\Omega} \setminus J_{v_n}; \Bbb R^d) we can choose a sequence (\eta_n)_n with \eta_n \to 0 such that z_n : = y + \eta_n v_n \in GSBV^2_2(\Omega; \Bbb R^d) satisfies J_{z_n} \tilde{ = } J_y \cup J_{v_n} and there holds z_n \to y in measure on \Omega . By (4.2), the continuity of W , J_{z_n} \tilde{ = } J_y \cup J_{v_n} , and J_{\nabla z_n} \tilde{\subset} J_{\nabla y} \cup J_{v_n} we get

    \begin{align} \limsup\nolimits_{n\to \infty}\overline{\mathcal{E}}_\varepsilon(z_n, \Omega) \le \overline{\mathcal{E}}_\varepsilon(y, \Omega). \end{align} (4.3)

    As J_{z_n} \tilde{ = } J_y \cup J_{v_n} , J_{\nabla y} \tilde{ = } J_{v} , and J_{\nabla v_n} \tilde{\subset} J_{v_n} , we also get

    \begin{align} J_{\nabla z_n} \setminus J_{z_n} \, \tilde{\subset}\, (J_{\nabla y} \cup J_{\nabla v_n}) \setminus (J_y \cup J_{v_n}) \, \tilde{\subset} \, J_v \setminus J_{v_n}. \end{align} (4.4)

    In view of (4.2), by a Besicovitch covering argument we can cover the rectifiable sets J_v \setminus J_{v_n} by sets of finite perimeter (E_n)_n \subset \subset \Omega , each of which being a countable union of balls with radii smaller than \frac{1}{n} , such that

    \begin{align} \mathcal{L}^d(E_n) + \mathcal{H}^{d-1}(\partial^* E_n) \to 0. \end{align} (4.5)

    We finally define the sequence y_n \in GSBV^2_2(\Omega; \Bbb R^d) by y_n = z_n \chi_{\Omega \setminus E_n} + (\mathbf{id} + b_n) \chi_{E_n} for suitable constants (b_n)_n \subset \Bbb R^d which are chosen such that J_{y_n} \tilde{ = } (J_{z_n} \setminus E_n) \cup \partial^* E_n . Now in view of (4.4) and J_v \setminus J_{v_n} \tilde{\subset} E_n , we get J_{\nabla y_n}\tilde{\subset} J_{y_n} . By (4.5) and z_n \to y in measure on \Omega we get y_n \to y in measure on \Omega . By (2.3)(ⅲ) we obtain W(\nabla {y}_n) = 0 , \nabla^2 {y}_n = 0 on E_n . Then by (2.5), (4.3), (4.5), and the fact that J_{\nabla y_n}\tilde{\subset} J_{y_n} \tilde{ = } (J_{z_n} \setminus E_n) \cup \partial^* E_n we get

    \limsup\limits_{n\to \infty} \overline{\mathcal{E}}_\varepsilon(y_n, \Omega) \le \limsup\limits_{n\to\infty} \big(\overline{\mathcal{E}}_\varepsilon(z_n, \Omega) + \kappa\mathcal{H}^{d-1}(\partial^* E_n) \big) \le \overline{\mathcal{E}}_\varepsilon(y, \Omega).

    Since \overline{\mathcal{E}}_\varepsilon(y_n, \Omega) = \mathcal{E}_\varepsilon(y_n, \Omega) for all n \in \Bbb N by J_{\nabla y_n}\tilde{\subset} J_{y_n} , (4.1) implies \mathcal{E}'_\varepsilon(y) \le \overline{\mathcal{E}}_\varepsilon(y, \Omega) . This concludes the proof.

    Proof of Theorem 2.2. We prove the existence of minimizers via the direct method. Let (y_n)_n \subset GSBV^2_2(\Omega'; \Bbb R^d) with y_n = g on \Omega' \setminus\overline{\Omega} be a minimizing sequence for the minimization problem (2.6). By (2.3) we find W(F)\ge c_1|F|^2 - c_2 for c_1, c_2 > 0 . Thus, \sup_{n\in \Bbb N} \overline{\mathcal{E}}_\varepsilon(y_n) < +\infty also implies \sup_{n \in \Bbb N} \Vert \nabla y_n \Vert_{L^2(\Omega')} < + \infty , and we can apply Theorem 3.3. We obtain a sequence (z_n)_n \subset GSBV^2_2(\Omega'; \Bbb R^d) satisfying z_n = g on \Omega' \setminus \overline{\Omega} and a limiting function y \in GSBV^2_2(\Omega'; \Bbb R^d) with y = g on \Omega' \setminus \overline{\Omega} such that z_n \to y in measure on \Omega' . Using (2.5), (3.5), and g \in W^{2, \infty}(\Omega'; \Bbb R^d) we calculate

    \begin{align*} \limsup\limits_{n \to \infty} \big(\overline{\mathcal{E}}_\varepsilon(z_n) -\overline{\mathcal{E}}_\varepsilon(y_n) \big)& \le \limsup\limits_{n \to \infty} \Big( \varepsilon^{-2} C_{W, g} \mathcal{L}^d(S_n) + \varepsilon^{-2\beta} \Vert \nabla^2 g\Vert^2_{L^2(S_n)} \\ & \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ + \kappa \big( \mathcal{H}^{d-1}(J_{z_n} \cup J_{\nabla z_n} )- \mathcal{H}^{d-1}(J_{y_n} \cup J_{\nabla y_n} ) \big) \Big) \le 0, \end{align*}

    where the constant C_{W, g} depends on W and \Vert \nabla g \Vert_{L^\infty(\Omega')} . I.e., (z_n)_n is also a minimizing sequence. By z_n \to y in measure on \Omega' and the fact that \overline{\mathcal{E}}_\varepsilon is lower semicontinuous with respect to the convergence in measure on \Omega' , see Proposition 2.1, we get

    \overline{\mathcal{E}}_\varepsilon(y) \le \liminf\limits_{n \to \infty} \overline{\mathcal{E}}_\varepsilon(z_n) \le \liminf\limits_{n \to \infty} \overline{\mathcal{E}}_\varepsilon(y_n) = \inf\limits_{{\bar{y} \in GSBV^2_2(\Omega'; \Bbb R^d)}} \Big\{ \overline{\mathcal{E}}_\varepsilon(\bar{y}): \ \bar{y} = g \text{ on } \Omega' \setminus \overline{\Omega} \Big\}.

    This shows that y is a minimizer.

    This subsection is devoted to the proof of Theorem 2.3.

    Proof of Theorem 2.3(a). Consider a sequence (y_\varepsilon)_\varepsilon with y_\varepsilon \in \mathcal{S}_{\varepsilon, h} , i.e., y_\varepsilon = \mathbf{id} + {h}_{\varepsilon} on \Omega' \setminus \overline{\Omega} . Suppose that M: = \sup\nolimits_\varepsilon \overline{\mathcal{E}}_\varepsilon(y_\varepsilon) < +\infty . We first construct Caccioppoli partitions (Step 1) and the corresponding rotations (Step 2) in order to define y^{\rm rot}_\varepsilon . Then we confirm (2.10) (Step 3).

    Step 1: Definition of the Caccioppoli partitions. First, we apply the BV coarea formula (see [6,Theorem 3.40 or Theorem 4.34]) on each component (\nabla y_\varepsilon)_{ij} \in GSBV^2(\Omega') , 1 \le i, j\le d , to write

    \begin{align*} \int_{-\infty}^\infty \mathcal{H}^{d-1}\big( (\Omega' \setminus J_{\nabla y_\varepsilon}) \cap \partial^* \lbrace (\nabla y_\varepsilon)_{ij} \gt t \rbrace \big) \, dt & = |D (\nabla y_\varepsilon)_{ij}|(\Omega' \setminus J_{ \nabla y_\varepsilon}) \le \Vert \nabla^2 y_\varepsilon \Vert_{L^1(\Omega')}. \end{align*}

    Using Hölder's inequality and (2.5) along with \overline{\mathcal{E}}_\varepsilon(y_\varepsilon) \le M , we then get

    \begin{align} \int_{-\infty}^\infty \mathcal{H}^{d-1}\big( (\Omega' \setminus J_{\nabla y_\varepsilon}) \cap \partial^* \lbrace (\nabla y_\varepsilon)_{ij} \gt t \rbrace \big) \, dt \le (\mathcal{L}^d(\Omega'))^{1/2} \Vert \nabla^2 y_\varepsilon \Vert_{L^2(\Omega')} \le (\mathcal{L}^d(\Omega') M)^{1/2} \varepsilon^\beta. \end{align} (4.6)

    Fix \gamma \in (\frac{2}{3}, \beta) and define T_\varepsilon = \varepsilon^{\gamma} . For all k \in \Bbb Z we find t^{ij}_k \in (kT_\varepsilon, (k+1)T_\varepsilon] such that

    \begin{align} \mathcal{H}^{d-1}\big( (\Omega' \setminus J_{\nabla y_\varepsilon}) \cap \partial^*\lbrace (\nabla y_\varepsilon)_{ij} \gt t^{ij}_k \rbrace \big) \le \frac{1}{T_\varepsilon} \int_{kT_\varepsilon}^{{(k+1)T_\varepsilon}} \mathcal{H}^{d-1}\big( (\Omega' \setminus J_{\nabla y_\varepsilon}) \cap \partial^*\lbrace (\nabla y_\varepsilon)_{ij} \gt t \rbrace \big)\, dt. \end{align} (4.7)

    Let G_k^{\varepsilon, ij} = \lbrace (\nabla y_\varepsilon)_{ij} > t^{ij}_{k} \rbrace \setminus \lbrace (\nabla y_\varepsilon)_{ij} > t^{ij}_{k+1} \rbrace and note that each set has finite perimeter in \Omega' since it is the difference of two sets of finite perimeter. Now (4.6) and (4.7) imply

    \begin{align} \sum\nolimits_{k\in \Bbb Z} \mathcal{H}^{d-1}\big( (\Omega' \setminus J_{\nabla y_\varepsilon}) \cap \partial^* G_k^{\varepsilon, ij} \big) \le 2T^{-1}_\varepsilon (\mathcal{L}^d(\Omega') M)^{1/2} \varepsilon^\beta \le C \varepsilon^{\beta-\gamma} \end{align} (4.8)

    for a sufficiently large constant C > 0 independent of \varepsilon . Since \mathcal{L}^d(\Omega' \setminus \bigcup_{k \in \Bbb Z} G_k^{\varepsilon, ij}) = 0 , (G_k^{\varepsilon, ij})_{ k\in \Bbb Z } is a Caccioppoli partition of \Omega' . We let (P^{_j}_{\varepsilon})_{j\in \Bbb N} be the Caccioppoli partition of \Omega' consisting of the nonempty sets of

    \big\{ G_{k_{11}}^{\varepsilon, 11} \cap G_{k_{12}}^{\varepsilon, 12} \cap \ldots \cap G_{k_{dd}}^{\varepsilon, dd}: \ k_{ij} \in \Bbb Z\text{ for } i, j = 1, \ldots, d\big\}.

    Then (4.8) implies

    \begin{align} \sum\nolimits_{j = 1}^\infty\mathcal{H}^{d-1}\big(\partial^*P_j^\varepsilon \cap (\Omega' \setminus J_{\nabla y_\varepsilon}) \big) \le C\varepsilon^{\beta-\gamma} \end{align} (4.9)

    for a constant C > 0 independent of \varepsilon .

    Step 2: Definition of the rotations. We now define corresponding rotations. Recalling T_\varepsilon = \varepsilon^{\gamma} we get |t_k^{ij} -t_{k+1}^{ij}| \le 2T_\varepsilon = 2\varepsilon^\gamma for all k \in \Bbb Z , i, j = 1, \ldots, n . Then by the definition of G_k^{\varepsilon, ij} , for each component P^{_j}_{\varepsilon} of the Caccioppoli partition, we find a matrix F_j^\varepsilon \in \Bbb R^{d\times d} such that

    \begin{align} \Vert \nabla y_\varepsilon - F_j^\varepsilon \Vert_{L^\infty(P^{_j}_{\varepsilon})} \le c\varepsilon^{\gamma}, \end{align} (4.10)

    where c depends only on d . For each j \in \Bbb N with P^{_j}_{\varepsilon} \subset \Omega up to an \mathcal{L}^d -negligible set, we denote by \bar{R}_j^\varepsilon the nearest point projection of F_j^\varepsilon onto SO(d) . For all other components P^{_j}_{\varepsilon} , i.e., the components intersecting \Omega' \setminus \overline{\Omega} , we set \bar{R}_j^\varepsilon = \mathbf{Id} . We now show that for all j \in \Bbb N and for \mathcal{L}^d -a.e. x \in P_j^\varepsilon there holds

    \begin{align} |\nabla y_\varepsilon(x) - \bar{R}_j^\varepsilon| \le \max\big\{ C\varepsilon^{\gamma}, \ 2\operatorname{dist}(\nabla y_\varepsilon(x), SO(d)) \big\} \end{align} (4.11)

    for a constant C > 0 independent of \varepsilon .

    First, we consider components P^{_j}_{\varepsilon} which are contained in \Omega up to an \mathcal{L}^d -negligible set. Recall that \bar{R}_j^\varepsilon is defined as the nearest point projection of F_j^\varepsilon onto SO(d) . If |\bar{R}_j^\varepsilon- F_j^\varepsilon| \le 3c\varepsilon^{\gamma} , where c is the constant of (4.10), (4.11) follows from (4.10) and the triangle inequality. Otherwise, by (4.10) we get for \mathcal{L}^d -a.e. x \in P_j^\varepsilon

    \begin{align*} \operatorname{dist}(\nabla y_\varepsilon(x), SO(d)) & \ge \operatorname{dist}(F^{_j}_{\varepsilon}, SO(d)) - c\varepsilon^{\gamma} = |\bar{R}_j^\varepsilon- F_j^\varepsilon| - c\varepsilon^{\gamma} \\ & \ge \tfrac{1}{2} \big(|\bar{R}_j^\varepsilon- F_j^\varepsilon| + c\varepsilon^{\gamma} \big) \ge \tfrac{1}{2}|\bar{R}_j^\varepsilon- \nabla y_\varepsilon(x)|. \end{align*}

    This implies (4.11). Now consider a component P^{_j}_{\varepsilon} which intersects \Omega' \setminus \overline{\Omega} . Then by (4.10) and the fact that y_\varepsilon = \mathbf{id} + {h}_{\varepsilon} on \Omega' \setminus \overline{\Omega} there holds

    \begin{align*} \Vert \mathbf{Id} + \varepsilon \nabla h - F_j^\varepsilon \Vert_{L^\infty(P^{_j}_{\varepsilon} \setminus \Omega)} \le \Vert \nabla y_\varepsilon - F_j^\varepsilon \Vert_{L^\infty(P^{_j}_{\varepsilon})} \le c\varepsilon^{\gamma}. \end{align*}

    Since 0 < \gamma < 1 , this yields |F_j^\varepsilon - \mathbf{Id}| \le C\varepsilon^{\gamma} for a constant C depending also on \Vert \nabla h \Vert_{L^\infty(\Omega')} . This along with (4.10) implies (4.11) (for \bar{R}_j^\varepsilon = \mathbf{Id} ). We define the rotations in the statement by R_j^\varepsilon : = (\bar{R}^{_j}_{\varepsilon})^{-1} .

    Step 3: Proof of (2.10). We are now in a position to prove (2.10). We define y^{\rm rot}_{\varepsilon} as in (2.9), i.e., y^{\rm rot}_\varepsilon = \sum\nolimits_{j = 1}^\infty R_j^{\varepsilon}y_{\varepsilon}{\chi}_{P_j^{\varepsilon}} . Then (2.10)(ⅰ) follows from the fact that y_\varepsilon = \mathbf{id} + {h}_{\varepsilon} on \Omega' \setminus \overline{\Omega} and y_{\varepsilon}^{\rm rot} = y_\varepsilon on \Omega' \setminus \overline{\Omega} , where the latter holds due to R^{_j}_{\varepsilon} = \mathbf{Id} for all P_j^\varepsilon intersecting \Omega' \setminus \overline{\Omega} . Property (2.10)(ⅱ) is a direct consequence of the definition of y_\varepsilon^{\rm rot} and (4.9). To see (2.10)(ⅳ), we use (4.11) and R_j^\varepsilon = (\bar{R}^{_j}_{\varepsilon})^{-1} to get

    \begin{align*} \Vert \nabla y^{\rm rot}_\varepsilon - \mathbf{Id}\Vert^2_{L^2(\Omega')} & = \sum\nolimits_{j = 1}^\infty \Vert \nabla y_\varepsilon - \bar{R}^{_j}_{\varepsilon}\Vert^2_{L^2(P^{_j}_{\varepsilon})} \le C\varepsilon^{2\gamma} \mathcal{L}^d(\Omega') + 4 \Vert \operatorname{dist}(\nabla y_\varepsilon, SO(d)) \Vert^2_{L^2(\Omega')} \\& \le C(\varepsilon^{2\gamma} + \varepsilon^2) \end{align*}

    for a constant depending on M , where the last step follows from (2.3)(ⅲ), (2.5), and \overline{\mathcal{E}}_\varepsilon(y_\varepsilon) \le M . Since 0 < \gamma < 1 , (2.10)(ⅳ) is proved. It remains to show (2.10)(ⅲ). We recall the linearization formula (see [46,(3.20)])

    \begin{align} |{\rm sym}(F -\mathbf{Id})| = \operatorname{dist}(F, SO(d)) + {\rm O} (|F- \mathbf{Id}|^2) \end{align} (4.12)

    for F \in \Bbb R^{d \times d} . By Young's inequality and |{\rm sym}(F -\mathbf{Id})| \le |F -\mathbf{Id}| this implies

    \begin{align*} |{\rm sym}(F -\mathbf{Id})|^2& \le \min\big\{ |F- \mathbf{Id}|^2, \ C \operatorname{dist}^2(F, SO(d))+ C|F-\mathbf{Id}|^4 \big\}\\ & \le C \operatorname{dist}^2(F, SO(d)) + C\min\big\{ |F- \mathbf{Id}|^2, \ |F-\mathbf{Id}|^4 \big\}. \end{align*}

    Then we calculate

    \begin{align*} \int_{\Omega'} |{\rm sym}( \nabla y^{\rm rot}_\varepsilon - \mathbf{Id})|^2 &\le C\int_{\Omega'} \Big( \operatorname{dist}^2(\nabla y_\varepsilon^{\rm rot}, SO(d)) + \min\big\{ |\nabla y_\varepsilon^{\rm rot}- \mathbf{Id}|^2, \ |\nabla y_\varepsilon^{\rm rot}-\mathbf{Id}|^4 \big\}\Big) \\ &\le C\sum\nolimits_{j = 1}^\infty\int_{P^{_j}_{\varepsilon}} \Big(\operatorname{dist}^2(\nabla y_\varepsilon, SO(d)) + |\nabla y_\varepsilon- \bar{R}^{_j}_{\varepsilon}|^2 \, \min\big\{ 1, \, |\nabla y_\varepsilon-\bar{R}_j^\varepsilon|^2 \big\}\Big). \end{align*}

    By (4.11) we note that for a.e. x \in P_j^\varepsilon there holds

    |\nabla y_\varepsilon(x)- \bar{R}^{_j}_{\varepsilon}|^2 \, \min\big\{ 1 , \, |\nabla y_\varepsilon(x)-\bar{R}_j^\varepsilon|^2 \big\} \le C\varepsilon^{4\gamma} + C\operatorname{dist}^2(\nabla y_\varepsilon(x), SO(d)).

    Here, we used that, if |\nabla y_\varepsilon(x)-\bar{R}_j^\varepsilon|^2 > 1 , the maximum in (4.11) is attained for \operatorname{dist}(\nabla y_\varepsilon(x), SO(d)) , provided that \varepsilon is small enough. Therefore, we get

    \begin{align*} \int_{\Omega'} |{\rm sym}( \nabla y^{\rm rot}_\varepsilon - \mathbf{Id})|^2 &\le C\int_{ \Omega' } \operatorname{dist}^2(\nabla y_\varepsilon, SO(d)) + C\mathcal{L}^d(\Omega') \varepsilon^{4\gamma} \le C\varepsilon^2 + C\varepsilon^{4\gamma}, \end{align*}

    where in the last step we have again used (2.3)(ⅲ), (2.5), and \overline{\mathcal{E}}_\varepsilon(y_\varepsilon) \le M . Since \gamma > \frac{2}{3} \ge \frac{1}{2} , we obtain (2.10)(ⅲ). This concludes the proof of Theorem 2.3(a).

    Remark 4.1. For later purposes, we point out that the construction shows y_\varepsilon^{\rm rot} = y_{\varepsilon} on all P_j^\varepsilon intersecting \Omega'\setminus \overline{\Omega} .

    Proof of Theorem 2.3(b). We define the rescaled displacment fields u_\varepsilon : = \frac{1}{\varepsilon}(y^{\rm rot}_\varepsilon - \mathbf{id}) as in (2.11). Clearly, there holds u_\varepsilon \in GSBV^2(\Omega'; \Bbb R^d) \subset GSBD^2(\Omega') . Note that by (2.10)(ⅲ) we obtain \sup_\varepsilon \Vert e(u_\varepsilon) \Vert_{L^2(\Omega')} < + \infty , where for shorthand we again write e(u_\varepsilon) = \frac{1}{2}(\nabla u_\varepsilon^T + \nabla u_\varepsilon) . Moreover, in view of (2.10)(ⅱ) and \beta > \gamma , we get

    \begin{align} \limsup\nolimits_{\varepsilon \to 0} \mathcal{H}^{d-1}(J_{u_\varepsilon}) \le \limsup\nolimits_{\varepsilon \to 0} \mathcal{H}^{d-1}(J_{y_\varepsilon} \cup J_{\nabla y_\varepsilon}) \lt +\infty. \end{align} (4.13)

    Therefore, we can apply Theorem 3.4 on the sequence (u_\varepsilon)_\varepsilon to obtain A and u' \in GSBD^2(\Omega') such that (3.9) holds (up to passing to a subsequence). We first observe that E_u = A , where E_u : = \lbrace x\in \Omega: \, |u_\varepsilon(x)| \to \infty \rbrace and A : = \lbrace x\in \Omega': \, |u_\varepsilon(x)| \to \infty \rbrace . To see this, we have to check that A \subset \Omega . This follows from the fact that u_\varepsilon = h on \Omega' \setminus \overline{\Omega} for all \varepsilon , see (2.10)(ⅰ) and (2.11).

    We define u : = u' \chi_{\Omega' \setminus E_u} + \lambda \chi_{E_u} for some \lambda \in \Bbb R^d such that \partial^* E_u \cap \Omega' \subset J_u up to an \mathcal{H}^{d-1} -negligible set. Since J_u \subset J_{u'} \cup (\partial^*E_u\cap \Omega') , (3.9) then implies (2.12), where the last inequality in (2.12)(ⅲ) follows from (4.13). Finally, u \in GSBD^2_h follows from u_\varepsilon = h on \Omega' \setminus \overline{\Omega} and (2.12)(ⅰ).

    We now give the proof of Theorem 2.7.

    Proof of Theorem 2.7. Since \overline{\mathcal{E}}_\varepsilon \le \mathcal{E}_\varepsilon , see (2.4) and (2.5), the compactness result follows immediately from Theorem 2.3. It suffices to show the \Gamma -liminf inequality for \overline{\mathcal{E}}_\varepsilon and the \Gamma -limsup inequality for \mathcal{E}_\varepsilon .

    Step 1: \Gamma -liminf inequality. Consider u \in GSBD^2_h and (y_\varepsilon)_\varepsilon , y_\varepsilon \in \mathcal{S}_{\varepsilon, h} , such that y_\varepsilon \rightsquigarrow u , i.e., by Definition 2.4 there exist y^{\rm rot}_\varepsilon = \sum\nolimits_{j = 1}^\infty R^{_j}_{\varepsilon} \, y_\varepsilon \, \chi_{P^{_j}_{\varepsilon}} and u_\varepsilon: = \frac{1}{\varepsilon} (y^{\rm rot}_\varepsilon - \mathbf{id}) such that (2.10) and (2.12) hold for some fixed \gamma \in (\frac{2}{3}, \beta) . The essential step is to prove

    \begin{align} \liminf\limits_{\varepsilon \to 0} \frac{1}{\varepsilon^2}\int_{\Omega'} W(\nabla y_\varepsilon) \ge \int_{\Omega'} \frac{1}{2} Q(e( u) ). \end{align} (4.14)

    Once (4.14) is shown, we conclude by (2.5) and (2.12)(ⅲ) that

    \begin{align*} \liminf\limits_{\varepsilon \to 0} \overline{\mathcal{E}}_\varepsilon (y_\varepsilon) \ge \liminf\limits_{\varepsilon \to 0} \Big( \frac{1}{\varepsilon^2}\int_{\Omega'} W(\nabla y_\varepsilon) + \kappa\mathcal{H}^{d-1}(J_{y_\varepsilon} \cup J_{\nabla y_\varepsilon}) \Big) \ge \int_{\Omega'} \frac{1}{2} Q(e(u)) + \kappa\mathcal{H}^{d-1}(J_u). \end{align*}

    In view of (2.14), this shows \liminf_{\varepsilon \to 0} \overline{\mathcal{E}}_\varepsilon (y_\varepsilon) \ge \mathcal{E}(u) . To see (4.14), we first note that the frame indifference of W (see (2.3)(ⅱ)) and the definitions of y^{\rm rot}_\varepsilon and u_\varepsilon (see (2.9) and (2.11)) imply

    \begin{align} W(\nabla y_\varepsilon) = W(\nabla y_\varepsilon^{\rm rot}) = W(\mathbf{Id} + \varepsilon \nabla u_\varepsilon). \end{align} (4.15)

    In view of \gamma > 2/3 , we can choose \eta_\varepsilon \to +\infty such that

    \begin{align} \varepsilon^{1-\gamma} \eta_\varepsilon \to +\infty \ \ \ \text{and} \ \ \ \varepsilon{\eta}_{\varepsilon}^3 \to 0. \end{align} (4.16)

    We define \chi_\varepsilon \in L^\infty(\Omega') by \chi_\varepsilon(x) = \chi_{[0, \eta_\varepsilon)}(|\nabla u_\varepsilon(x)|) . Note that \mathcal{L}^d(\lbrace |\nabla u_\varepsilon(x)| > \eta_\varepsilon \rbrace) \le C(\varepsilon^{\gamma-1}/\eta_\varepsilon)^2 by (2.10)(ⅳ) and the fact that u_\varepsilon = \frac{1}{\varepsilon} (y^{\rm rot}_\varepsilon - \mathbf{id}) . Thus, (4.16) implies \chi_\varepsilon \to 1 boundedly in measure on \Omega' . The regularity of W implies W(\mathbf{Id} + F) = \frac{1}{2}Q({\rm sym}(F)) + \omega(F) , where Q is defined in (2.14) and \omega: \Bbb R^{d \times d}\to \Bbb R is a function satisfying |\omega(F)| \le C|F|^3 for all F \in \Bbb R^{d\times d} with |F| \le 1 . Then by (4.15) and W \ge 0 we get

    \begin{align*} \liminf\limits_{\varepsilon \to 0} \frac{1}{\varepsilon^2}\int_{\Omega'} W(\nabla y_\varepsilon) & \ge \liminf\limits_{\varepsilon\to 0} \frac{1}{\varepsilon^2}\int_{\Omega'} \chi_{\varepsilon}{W}(\mathbf{Id} + \varepsilon \nabla u_\varepsilon) \\ & = \liminf\limits_{\varepsilon\to 0} \int_{\Omega'} \chi_\varepsilon \Big( \frac{1}{2}Q(e(u_\varepsilon)) + \frac{1}{\varepsilon^2} \omega(\varepsilon \nabla u_\varepsilon) \Big) \\ &\ge \liminf\limits_{\varepsilon\to 0} \Big(\int_{\Omega'\setminus E_u} \chi_\varepsilon \frac{1}{2}Q(e(u_\varepsilon)) + \int_{\Omega'} \chi_\varepsilon |\nabla u_\varepsilon|^3 \varepsilon \frac{\omega(\varepsilon \nabla u_\varepsilon)}{|\varepsilon \nabla u_\varepsilon|^3} \Big), \end{align*}

    where E_u = \lbrace x\in \Omega: \, |u_\varepsilon(x)| \to \infty \rbrace . The second term converges to zero. Indeed, \chi_\varepsilon\frac{|\omega(\varepsilon \nabla u_\varepsilon)|}{|\varepsilon \nabla u_\varepsilon|^3} is uniformly controlled by C and \chi_\varepsilon |\nabla u_\varepsilon|^3 \varepsilon is uniformly controlled by \eta_\varepsilon^3 \varepsilon , where \eta_\varepsilon^3 \varepsilon \to 0 by (4.16). As e(u_\varepsilon) \rightharpoonup e(u) weakly in L^2(\Omega'\setminus E_u, \Bbb R^{d\times d}_{\rm sym}) by (2.12)(ⅱ), Q is convex, and \chi_\varepsilon converges to 1 boundedly in measure on \Omega' \setminus E_u , we conclude

    \begin{align*} \liminf\limits_{\varepsilon \to 0} \frac{1}{\varepsilon^2}\int_{\Omega'} W(\nabla y_\varepsilon) \ge \int_{\Omega' \setminus E_u} \frac{1}{2}Q(e(u)) = \int_{\Omega'} \frac{1}{2}Q(e(u)), \end{align*}

    where the last step follows from the fact that e(u) = 0 on E_u , see (2.12)(ⅳ). This shows (4.14) and concludes the proof of the \Gamma -liminf inequality.

    Step 2: \Gamma -limsup inequality. Consider u \in GSBD^2_h with h \in W^{2, \infty}(\Omega'; \Bbb R^d) . Let \gamma \in (\frac{2}{3}, \beta) . By Theorem 3.6 we can find a sequence (v_\varepsilon)_\varepsilon \in GSBV^2_2(\Omega'; \Bbb R^d) with v_\varepsilon = h on \Omega' \setminus \overline{\Omega} , v_\varepsilon \in W^{2, \infty}(\Omega' \setminus J_{v_\varepsilon}; \Bbb R^d) , and

    \begin{align} {\rm (i)} & \ \ v_\varepsilon \to u \text{ in measure on } \Omega', \\ {\rm (ii)} & \ \ \Vert e(v_\varepsilon) - e(u) \Vert_{L^2(\Omega')} \to 0, \\ {\rm (iii)} & \ \ \mathcal{H}^{d-1}(J_{v_\varepsilon}) \to \mathcal{H}^{d-1}(J_{u}), \\ {\rm (iv)} & \ \ \Vert \nabla v_\varepsilon \Vert_{L^\infty(\Omega')} + \Vert \nabla^2 v_\varepsilon \Vert_{L^\infty(\Omega')} \le \varepsilon^{(\beta-1)/2} \le \varepsilon^{\gamma-1}. \end{align} (4.17)

    Note that property (ⅳ) can be achieved since the approximations satisfy v_\varepsilon \in W^{2, \infty}(\Omega' \setminus J_{v_\varepsilon}; \Bbb R^d) . (Recall \gamma < \beta < 1 .) Moreover, v_\varepsilon \in W^{2, \infty}(\Omega' \setminus J_{v_\varepsilon}; \Bbb R^d) also implies J_{\nabla v_\varepsilon} \subset J_{v_\varepsilon} .

    We define the sequence y_\varepsilon = \mathbf{id} + {\varepsilon v}_{\varepsilon} . As v_\varepsilon \in GSBV_2^2(\Omega'; \Bbb R^d) and v_\varepsilon = h on \Omega' \setminus \overline{\Omega} , we get y_\varepsilon \in \mathcal{S}_{\varepsilon, h} , see (2.8). We now check that y_\varepsilon \rightsquigarrow u in the sense of Definition 2.4.

    We define y_\varepsilon^{\rm rot} = y_\varepsilon , i.e., the Caccioppoli partition in (2.9) consists of the set \Omega' only with corresponding rotation \mathbf{Id} . Then (2.10)(ⅰ), (ⅱ) are trivially satisfied. As \nabla y_\varepsilon^{\rm rot} - \mathbf{Id} = \varepsilon \nabla v_\varepsilon , (2.10)(ⅲ), (ⅳ) follow from (4.17)(ⅱ), (ⅳ). The rescaled displacement fields u_\varepsilon defined in (2.11) satisfy u_\varepsilon = v_\varepsilon . Then (2.12) for E_u = \emptyset follows from (4.17)(ⅰ)–(ⅲ) and J_{y_\varepsilon} = J_{v_\varepsilon} .

    Finally, we confirm \lim_{\varepsilon \to 0} \mathcal{E}_{\varepsilon}(y_\varepsilon) = \mathcal{E}(u) . In view of J_{y_\varepsilon} = J_{v_\varepsilon} , J_{\nabla y_\varepsilon} \subset J_{y_\varepsilon} , (4.17)(ⅲ), and the definition of the energies in (2.4), (2.14), it suffices to show

    \lim\limits_{\varepsilon \to 0} \Big( \frac{1}{\varepsilon^2}\int_{\Omega'} W(\nabla y_\varepsilon) + \frac{1}{\varepsilon^{2\beta}} \int_{\Omega'} |\nabla^2 y_\varepsilon|^2 \Big) = \int_{\Omega'} \frac{1}{2} Q(e(u)).

    The second term vanishes by (4.17)(ⅳ), \beta < 1 , and the fact that \nabla^2 y_\varepsilon = \varepsilon \nabla^2 v_\varepsilon . For the first term, we again use that W(\mathbf{Id} + F) = \frac{1}{2}Q({\rm sym}(F)) + \omega(F) with |\omega(F)|\le C|F|^3 for |F| \le 1 , and compute by (4.17)(ⅱ), (ⅳ)

    \begin{align*} \lim\limits_{\varepsilon \to 0}\frac{1}{\varepsilon^2} \int_{\Omega'} W(\nabla y_\varepsilon) & = \lim\limits_{\varepsilon \to 0} \frac{1}{\varepsilon^2} \int_{\Omega'} W(\mathbf{Id} + \varepsilon \nabla v_\varepsilon) = \lim\limits_{\varepsilon \to 0} \int_{\Omega'} \Big( \frac{1}{2} Q(e(v_\varepsilon)) + \frac{1}{\varepsilon^2} \omega(\varepsilon \nabla v_\varepsilon) \Big) \\ & = \int_{\Omega'} \frac{1}{2} Q(e(u)) + \lim\limits_{\varepsilon \to 0}\int_{\Omega'} {\rm O}\big( \varepsilon|\nabla v_\varepsilon|^3 \big) = \int_{\Omega'} \frac{1}{2} Q(e(u)), \end{align*}

    where in the last step we have used that \Vert \nabla v_\varepsilon \Vert_{L^\infty(\Omega')} \le C\varepsilon^{\gamma - 1} for some \gamma > 2/3 . This concludes the proof.

    Remark 4.2. The proof shows that one can readily incorporate a dependence on the material point x in the density W, as long as (2.3) still holds. We also point out that it suffices to suppose that W is C^{2, \alpha} in a neighborhood of SO(d) , provided that 1 > \beta > \gamma > \frac{2}{2+\alpha} . In fact, in that case, one has |\omega(F)| \le C|F|^{2+\alpha} for all |F|\le 1 , and all estimates remain true, where in (4.16) one chooses \eta_\varepsilon with \varepsilon^{1-\gamma} \eta_\varepsilon \to +\infty and \varepsilon^\alpha \eta^{2+\alpha}_\varepsilon \to 0 .

    We close this subsection with the proof of Corollary 2.8.

    Proof of Corollary 2.8. The statement follows in the spirit of the fundamental theorem of \Gamma -convergence, see, e.g., [12,Theorem 1.21]. We repeat the argument here for the reader's convenience. We observe that \inf_{\bar{y} \in \mathcal{S}_{\varepsilon, h}} \mathcal{E}_\varepsilon(\bar{y}) is uniformly bounded by choosing \mathbf{id} + {h}_{\varepsilon} as competitor. Given (y_\varepsilon)_\varepsilon , y_\varepsilon \in \mathcal{S}_{\varepsilon, h} , satisfying (2.18), we apply Theorem 2.7(a) to find a subsequence (not relabeled), and u \in GSBD^2_h such that y_\varepsilon \rightsquigarrow u in the sense of Definition 2.4. Thus, by Theorem 2.7(b) we obtain

    \begin{align} \mathcal{E}(u) \le \liminf\limits_{\varepsilon \to 0} \mathcal{E}_\varepsilon(y_\varepsilon) \le \liminf\limits_{\varepsilon \to 0} \inf\limits_{\bar{y} \in \mathcal{S}_{\varepsilon, h}} \mathcal{E}_\varepsilon(\bar{y}). \end{align} (4.18)

    By Theorem 2.7(c), for each v \in GSBD^2_h , there exists a sequence (w_\varepsilon)_\varepsilon with w_\varepsilon \rightsquigarrow v and \lim_{\varepsilon\to 0} \mathcal{E}_\varepsilon(w_\varepsilon) = \mathcal{E}(v) . This implies

    \begin{align} \limsup\limits_{\varepsilon \to 0} \inf\limits_{\bar{y} \in \mathcal{S}_{\varepsilon, h}} \mathcal{E}_\varepsilon(\bar{y}) \le \lim\limits_{\varepsilon \to 0} \mathcal{E}_\varepsilon(w_\varepsilon) = \mathcal{E}(v). \end{align} (4.19)

    By combining (4.18)–(4.19) we find

    \begin{align} \mathcal{E}(u) \le \liminf\limits_{\varepsilon \to 0} \inf\limits_{\bar{y} \in \mathcal{S}_{\varepsilon, h}} \mathcal{E}_\varepsilon(\bar{y}) \le \limsup\limits_{\varepsilon \to 0} \inf\limits_{\bar{y} \in \mathcal{S}_{\varepsilon, h}} \mathcal{E}_\varepsilon(\bar{y}) \le \mathcal{E}(v). \end{align} (4.20)

    Since v \in GSBD^2_h was arbitrary, we get that u is a minimizer of \mathcal{E} . Property (2.17) follows from (4.20) with v = u . In particular, the limit in (2.17) does not depend on the specific choice of the subsequence and thus (2.17) holds for the whole sequence.

    This final subsection is devoted to the proof of Lemma 2.6.

    Proof of Lemma 2.6. Proof of (a). As a preparation, we observe that for two given rotations R_1, R_2 \in SO(d) there holds

    \begin{align} |{\rm sym}(R_2 R_1^T -\mathbf{Id})| \le C |R_1- R_2|^2. \end{align} (4.21)

    This follows from formula (4.12) applied for F = R_2 R_1^T .

    Consider a sequence (y_\varepsilon)_\varepsilon . Let

    \begin{align} y^{{\rm rot}, i}_\varepsilon = \sum\nolimits_{j = 1}^\infty {R_j^{\varepsilon, i}} \, y_\varepsilon \, \chi_{P_j^{\varepsilon, i}}, \ \ \ i = 1, 2, \end{align} (4.22)

    be two sequences such that the corresponding rescaled displacement fields u_\varepsilon^i = \varepsilon^{-1}(y_\varepsilon^{{\rm rot}, i} - \mathbf{id}) , i = 1, 2 , converge to u_1 and u_2 , respectively, in the sense of (2.12), where the exceptional sets are denoted by E_{u_1} and E_{u_2} , respectively. In particular, pointwise \mathcal{L}^d -a.e. in \Omega' there holds

    \begin{align} e(u_\varepsilon^1) - e(u_\varepsilon^2) & = \varepsilon^{-1} {\rm sym} \Big( \sum\nolimits_j {R_j^{\varepsilon, 1}} \, \nabla y_\varepsilon \, \chi_{P_j^{\varepsilon, 1}} - \sum\nolimits_j {R_j^{\varepsilon, 2}} \, \nabla y_\varepsilon \, \chi_{P_j^{\varepsilon, 2}} \Big) \\ & = \varepsilon^{-1} {\rm sym} \Big( \sum\nolimits_{j, k} \big( {R_j^{\varepsilon, 1}} - {R_k^{\varepsilon, 2}} \big) \, \chi_{P_j^{\varepsilon, 1} \cap P_k^{\varepsilon, 2}} \, \nabla y_\varepsilon \Big) \\ & = \varepsilon^{-1} {\rm sym} \Big( \sum\nolimits_{j, k} \big( \mathbf{Id} - {R_k^{\varepsilon, 2} (R_j^{\varepsilon, 1})^T } \big) \, \chi_{P_j^{\varepsilon, 1} \cap P_k^{\varepsilon, 2}} \, \nabla y^{{\rm rot}, 1}_\varepsilon \Big). \end{align} (4.23)

    For brevity, we define Z_\varepsilon \in L^\infty(\Omega'; \Bbb R^{d \times d}) by

    \begin{align} Z_\varepsilon : = \sum\nolimits_{j, k} \big( \mathbf{Id} - {R_k^{\varepsilon, 2} (R_j^{\varepsilon, 1})^T } \big) \, \chi_{P_j^{\varepsilon, 1} \cap P_k^{\varepsilon, 2}}. \end{align} (4.24)

    By (2.10)(ⅳ) and the triangle inequality we get

    \begin{align*} \sum\limits_{j, k} \big\| {R_j^{\varepsilon, 1}} - {R_k^{\varepsilon, 2} }\big\|^2_{L^2(P_j^{\varepsilon, 1} \cap P_k^{\varepsilon, 2})} &\le C\sum\limits_{j = 1}^\infty \Vert (\nabla y_\varepsilon)^T - R_j^{\varepsilon, 1} \Vert^2_{L^2(P_j^{\varepsilon, 1})} + C\sum\limits_{k = 1}^\infty \Vert (\nabla y_\varepsilon)^T - R_k^{\varepsilon, 2} \Vert^2_{L^2(P_k^{\varepsilon, 2})} \\ & = C\Vert \nabla y_\varepsilon^{{\rm rot}, 1}- \mathbf{Id} \Vert_{L^2(\Omega')}^2 + C\Vert \nabla y_\varepsilon^{{\rm rot}, 2}- \mathbf{Id} \Vert_{L^2(\Omega')}^2 \le C\varepsilon^{2\gamma} \end{align*}

    for some given \gamma \in (\frac{2}{3}, \beta) , and C > 0 independent of \varepsilon . Equivalently, this means

    \begin{align*} \sum\nolimits_{j, k} \mathcal{L}^d\big( P_j^{\varepsilon, 1} \cap P_k^{\varepsilon, 2} \big) \big| R_j^{\varepsilon, 1} - {R_k^{\varepsilon, 2}} \big|^2 \le C\varepsilon^{2\gamma}. \end{align*}

    By recalling (4.21) and (4.24) we then get

    \begin{align*} \Vert {\rm sym} (Z_\varepsilon) \Vert_{L^1(\Omega')} \le C\varepsilon^{2\gamma}, \ \ \ \ \ \ \ \ \ \ \ \Vert Z_\varepsilon \Vert_{L^2(\Omega')} \le C\varepsilon^{\gamma}. \end{align*}

    This along with Hölder's inequality, (2.10)(ⅳ) for y_\varepsilon^{{\rm rot}, 1} , and (4.23) yields

    \begin{align} \Vert e(u_\varepsilon^1) - e(u_\varepsilon^2) \Vert_{L^1(\Omega')} & = \frac{1}{\varepsilon} \Vert {\rm sym} \big( Z_\varepsilon\, \nabla y_\varepsilon^{{\rm rot}, 1} \big)\Vert_{L^1(\Omega')} \\& \le \frac{1}{\varepsilon} \Vert {\rm sym} \big( Z_\varepsilon\, \big(\nabla y_\varepsilon^{{\rm rot}, 1} - \mathbf{Id} \big) \big)\Vert_{L^1(\Omega')} + \frac{1}{\varepsilon} \Vert {\rm sym} \big( Z_\varepsilon \big)\Vert_{L^1(\Omega')}\\ & \le \frac{1}{\varepsilon} \Vert Z_\varepsilon \Vert_{L^2(\Omega')} \Vert \nabla y_\varepsilon^{{\rm rot}, 1} - \mathbf{Id} \Vert_{L^2(\Omega')} + \frac{1}{\varepsilon} \Vert {\rm sym} \big( Z_\varepsilon \big)\Vert_{L^1(\Omega')} \le C\varepsilon^{2\gamma - 1}. \end{align} (4.25)

    We have that e(u_\varepsilon^1) - e(u_\varepsilon^2) converges to e(u_1) - e(u_2) weakly in L^2(\Omega' \setminus (E_{u_1} \cup E_{u_2}); \Bbb R^{d\times d}_{\rm sym}) , see (2.12)(ⅱ). Then (4.25) and the fact that \gamma > \frac{2}{3} > \frac{1}{2} imply that e(u_1) - e(u_2) = 0 on \Omega' \setminus (E_{u_1} \cup E_{u_2}) . This shows part (a) of the statement.

    Proof of (b). Let (y_\varepsilon)_\varepsilon be a sequence satisfying (2.13). Consider two piecewise rotated functions y^{{\rm rot}, i}_\varepsilon as given in (4.22) and let u_1, u_2 be the limits identified in (2.12), where the corresponding exceptional sets are denoted by E_{u_1}, E_{u_2} . We let \mathcal{J}^i = \lbrace j \in \Bbb N: \, P_j^{\varepsilon, i} \subset \Omega \text{ up to an $\mathcal{L}^d$ -negligible set} \rbrace for i = 1, 2 , and set D_\varepsilon : = \bigcup_{i = 1, 2}\bigcup_{j \in \mathcal{J}^i} P_j^{\varepsilon, i} . By (2.10)(ⅱ) and \gamma < \beta we obtain

    \begin{align} \limsup\nolimits_{\varepsilon \to 0} \mathcal{H}^{d-1}\big( \big(\partial^*D_\varepsilon \cap \Omega' \big) \setminus \big(J_{y_\varepsilon}\cup J_{\nabla y_\varepsilon} \big) \big) = 0. \end{align} (4.26)

    As also \sup_\varepsilon\mathcal{H}^{d-1}(J_{y_\varepsilon} \cup J_{\nabla y_{\varepsilon}}) < + \infty , we get that \mathcal{H}^{d-1}(\partial^* D_\varepsilon) is uniformly controlled. Therefore, we may suppose that D_\varepsilon \to D in measure for a set of finite perimter D \subset \Omega , see [6,Theorem 3.39]. We observe that y^{{\rm rot}, i}_\varepsilon = y_\varepsilon on \Omega' \setminus D_\varepsilon for i = 1, 2 by Remark 4.1. Therefore, (2.11) implies that E_{u_1} \setminus D = E_{u_2} \setminus D . In the following, we denote this set by \hat{E} . Then, (2.11) and (2.12)(ⅰ) also yield

    \begin{align} u_1 = u_2 \ \ \ \text{ a.e. on } \Omega' \setminus (D \cup \hat{E}). \end{align} (4.27)

    To compare u_1 and u_2 inside D \cup \hat{E} , we introduce modifications: For i = 1, 2 and sequences (\lambda_\varepsilon)_\varepsilon \subset \Bbb R^d , let

    \begin{align} y^{\lambda_\varepsilon, i}_\varepsilon : = y^{{\rm rot}, i}_\varepsilon + \lambda_\varepsilon \, \chi_{D_\varepsilon}. \end{align} (4.28)

    By definition, D_\varepsilon does not intersect \Omega' \setminus \overline{\Omega} and has finite perimeter by (4.26). Thus, we get y^{\lambda_\varepsilon, i}_\varepsilon \in \mathcal{S}_{\varepsilon, h} , see (2.8) and (2.10)(ⅰ). By (2.10)(ⅱ), (4.26), and the fact that the elastic energy is frame indifferent we also observe that (y^{\lambda_\varepsilon, i}_\varepsilon)_\varepsilon is a minimizing sequence for i = 1, 2 and all (\lambda_\varepsilon)_\varepsilon \subset \Bbb R^d . We obtain

    \begin{align} y_\varepsilon = y^{{\rm rot}, i}_\varepsilon = y^{\lambda_\varepsilon, i}_\varepsilon \ \ \ \ \text{ on $\Omega' \setminus D_\varepsilon $ for all $(\lambda_\varepsilon)_\varepsilon \subset \Bbb R^d$, $i = 1, 2$}. \end{align} (4.29)

    This follows from (4.28) and y^{{\rm rot}, i}_\varepsilon = y_\varepsilon on \Omega' \setminus D_\varepsilon for i = 1, 2 , see Remark 4.1. We now consider two different cases:

    (1) Fix i = 1, 2 , \lambda \in \Bbb R^d , and consider \lambda_\varepsilon = \lambda \varepsilon . In view of (2.11), (2.12)(ⅰ), and (4.28), we get that \varepsilon^{-1}(y_\varepsilon^{\lambda_\varepsilon, i} - \mathbf{id}) \to u_i + \lambda \chi_D in measure on \Omega' \setminus E_{u_i} . Thus, one can check that y^{\lambda_\varepsilon, i}_\varepsilon\rightsquigarrow u^{\lambda}_i for some u^\lambda_i \in GSBD^2_h satisfying

    \begin{align} u^\lambda_i = u_i + \lambda \chi_D \text{ on } \Omega'\setminus E_{u_i}. \end{align} (4.30)

    (2) Recall that \hat{E} = E_{u_1} \setminus D = E_{u_2} \setminus D = \lbrace x \in \Omega \setminus D: \, |\varepsilon^{-1}(y_\varepsilon^{{\rm rot}, i} - \mathbf{id})| \to \infty\rbrace for i = 1, 2 . In view of (4.28), we can choose a suitable sequence (\lambda_\varepsilon)_\varepsilon such that |\varepsilon^{-1}(y_\varepsilon^{\lambda_\varepsilon, i} - \mathbf{id})| \to \infty on \hat{E} \cup D for i = 1, 2 . This along with (4.29) and (2.12)(ⅰ), (ⅳ) implies that for i = 1, 2 we have {y}^{\lambda_\varepsilon, i}_\varepsilon \rightsquigarrow \hat{u} for some \hat{u} \in GSBD_h^2 satisfying

    \begin{align} {\rm (i)} & \ \ \hat{u} = u_1 = u_2 \ \ \ \text{ a.e. on } \Omega'\setminus (\hat{E} \cup D), \\ {\rm (ii)} & \ \ e(\hat{u}) = 0 \ \ \text{ a.e. on } \ \hat{E} \cup D, \ \ \ \mathcal{H}^{d-1}(J_{\hat{u}} \cap (\hat{E} \cup D)^1) = 0, \end{align} (4.31)

    where (\cdot)^1 denotes the set of points with density 1 .

    We now combine the cases (1) and (2) to obtain the statement: since (y_\varepsilon^{\lambda_\varepsilon, i})_\varepsilon are minimizing sequences, Corollary 2.8 implies that each u^\lambda_i , \lambda \in \Bbb R^d , i = 1, 2 , and \hat{u} are minimizers of the problem \min_{v \in GSBD^2_h} \mathcal{E}(v) . In particular, as e(u_i^\lambda) = e(u_i) for all \lambda \in \Bbb R^d for both i = 1, 2 , the jump sets of u^\lambda_1 , u^\lambda_2 have to be independent of \lambda , i.e., \mathcal{H}^{d-1}(J_{u_i} \triangle J_{u_i^\lambda}) = 0 for all \lambda \in \Bbb R^d and i = 1, 2 . In view of (4.30) and (2.12)(ⅳ), this yields \partial^*E_{u_i} \cap \Omega', \partial^* (D \setminus E_{u_i}) \cap \Omega' \subset J_{u_i} up to \mathcal{H}^{d-1} -negligigble sets. Since \hat{E} = E_{u_i}\setminus D , this implies for i = 1, 2 that

    \begin{align} \partial^* (\hat{E} \cup D) \cap \Omega' \subset J_{u_i} \ \ \ \ \ \text{ up to $\mathcal{H}^{d-1}$-negligigble sets.} \end{align} (4.32)

    Recall that u_1, u_2 are both minimizers, that also \hat{u} is a minimzer, and that there holds \hat{u} = u_1 = u_2 on \Omega' \setminus (\hat{E} \cup D) , see (4.31)(ⅰ). This along with (4.31)(ⅱ) and (4.32) yields e({u}_i) = 0 on \hat{E} \cup D and \mathcal{H}^{d-1}(J_{{u}_i} \cap (\hat{E} \cup D)^1) = 0 for i = 1, 2 . Then (4.27) and (4.32) show that e(u_1) = e(u_2) \mathcal{L}^d -a.e. on \Omega' , and J_{u_1} = J_{u_2} up to an \mathcal{H}^{d-1} -negligible set.

    We finally provide an example that in case (a) the strains cannot be compared inside E_{u_1} \cup E_{u_2} .

    Example 4.3 Similar to Example 2.5, we consider \Omega' = (0, 3) \times (0, 1) , \Omega = (1, 3) \times (0, 1) , \Omega_1 = (0, 2) \times (0, 1) , \Omega_2 = (2, 3) \times (0, 1) , and h \equiv 0 . Let z \in W^{2, \infty}(\Omega'; \Bbb R^d) with \lbrace z = 0 \rbrace = \emptyset , and define

    y_\varepsilon(x) = x \chi_{\Omega_1}(x) + \big(x + {z}_{\varepsilon}(x)\big) \chi_{\Omega_2}(x) \ \ \ \ \text{for} \ x \in \Omega'.

    Note that J_{y_\varepsilon} = \partial\Omega_1 \cap\Omega' = \partial\Omega_2 \cap\Omega' . Then two possible alternatives are

    \begin{align*} (1)& \ \ P^{\varepsilon}_1 = \Omega_1, \ \ P^{\varepsilon}_2 = \Omega_2, \ \ R_1^{\varepsilon} = \mathbf{Id}, \ \ R_2^{\varepsilon} = \bar{R}_\varepsilon, \\ (2)& \ \ \tilde{P}^{\varepsilon}_1 = \Omega', \ \ \tilde{R}_1^{\varepsilon} = \mathbf{Id}, \end{align*}

    where \bar{R}_\varepsilon \in SO(2) satisfies \bar{R}_\varepsilon = \mathbf{Id} + \varepsilon^\gamma A + {\rm O}(\varepsilon^{2\gamma}) for some A \in \Bbb R^{2 \times 2}_{\rm skew} , \gamma \in (\frac{2}{3}, \beta) . Let u_{\varepsilon} = \varepsilon^{-1} (\sum_{j = 1}^2 R_j^{\varepsilon}y_{\varepsilon} \chi_{P_j^{\varepsilon}} -\mathbf{id}) and \tilde{u}_{\varepsilon} = \varepsilon^{-1} (y_{\varepsilon} -\mathbf{id}) , We observe that |u_\varepsilon| \to \infty on \Omega_2 . Possible limits identified in (2.12) are u = \lambda \chi_{\Omega_2} for some \lambda \in \Bbb R^{d} , \lambda \neq 0 , with E_u = \Omega_2 , and \tilde{u}(x) = z(x) \, \chi_{\Omega_2}(x) with E_{\tilde{u}} = \emptyset . This shows that in general there holds e(u) \neq e(\tilde{u}) in E_{u} .

    This work was supported by the DFG project FR 4083/1-1 and by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany's Excellence Strategy EXC 2044 -390685587, Mathematics Münster: Dynamics–Geometry–Structure.

    The author declares no conflict of interest.



    [1] Agostiniani V, Dal Maso G, DeSimone A (2012) Linear elasticity obtained from finite elasticity by Γ-convergence under weak coerciveness conditions. Ann Inst H Poincaré Anal Non Linéaire 29: 715-735.
    [2] Alicandro R, Dal Maso G, Lazzaroni G, et al. (2018) Derivation of a linearised elasticity model from singularly perturbed multiwell energy functionals. Arch Ration Mech Anal 230: 1-45. doi: 10.1007/s00205-018-1240-6
    [3] Ambrosio L (1990) Existence theory for a new class of variational problems. Arch Ration Mech Anal 111: 291-322. doi: 10.1007/BF00376024
    [4] Ambrosio L (1994) On the lower semicontinuity of quasi-convex integrals in S BV(\Omega; \R^k). Nonlinear Anal Theor 23: 405-425. doi: 10.1016/0362-546X(94)90180-5
    [5] Ambrosio L, Coscia A, Dal Maso G (1997) Fine properties of functions with bounded deformation. Arch Ration Mech Anal 139: 201-238. doi: 10.1007/s002050050051
    [6] Ambrosio L, Fusco N, Pallara D (2000) Functions of Bounded Variation and Free Discontinuity Problems, Oxford: Oxford University Press.
    [7] Babadjian JF, Giacomini A (2014) Existence of strong solutions for quasi-static evolution in brittle fracture. Ann Sc Norm Super Pisa Cl Sci 13: 925-974.
    [8] Ball JM, Currie JC, Olver PL (1981) Null Lagrangians, weak continuity, and variational problems of arbitrary order. J Funct Anal 41: 135-174. doi: 10.1016/0022-1236(81)90085-9
    [9] Batra RC (1976) Thermodynamics of non-simple elastic materials. J Elasticity 6: 451-456. doi: 10.1007/BF00040904
    [10] Blake A, Zisserman A (1987) Visual Reconstruction, Cambridge: The MIT Press.
    [11] Bourdin B, Francfort GA, Marigo JJ (2008) The variational approach to fracture. J Elasticity 91: 5-148. doi: 10.1007/s10659-007-9107-3
    [12] Braides A (2002) Γ-Convergence for Beginners, Oxford: Oxford University Press.
    [13] Braides A, Solci M, Vitali E (2007) A derivation of linear elastic energies from pair-interaction atomistic systems. Netw Heterog Media 2: 551-567. doi: 10.3934/nhm.2007.2.551
    [14] Capriz G (1985) Continua with latent microstructure. Arch Ration Mech Anal 90: 43-56. doi: 10.1007/BF00281586
    [15] Carriero M, Leaci A, Tomarelli F (1996) A second order model in image segmentation: Blake & Zisserman functional, In: Variational Methods for Discontinuous Structures, Basel: Birkhäuser, 57-72.
    [16] Carriero M, Leaci A, Tomarelli F (2004) Second order variational problems with free discontinuity and free gradient discontinuity, In: Calculus of Variations: Topics from the Mathematical Heritage of Ennio De Giorgi, Quad Mat, 14: 135-186.
    [17] Carriero M, Leaci A, Tomarelli F (1992) Plastic free discontinuities and special bounded hessian. C R Acad Sci 314: 595-600.
    [18] Chambolle A (2003) A density result in two-dimensional linearized elasticity, and applications. Arch Ration Mech Anal 167: 167-211.
    [19] Chambolle A, Conti S, Francfort G (2016) Korn-Poincaré inequalities for functions with a small jump set. Indiana Univ Math J 65: 1373-1399. doi: 10.1512/iumj.2016.65.5852
    [20] Chambolle A, Conti S, Francfort G (2018) Approximation of a brittle fracture energy with a constraint of non-interpenetration. Arch Ration Mech Anal 228: 867-889. doi: 10.1007/s00205-017-1207-z
    [21] Chambolle A, Conti S, Francfort G (2019) Approximation of functions with small jump sets and existence of strong minimizers of Griffith's energy. J Math Pures Appl 128: 119-139. doi: 10.1016/j.matpur.2019.02.001
    [22] Chambolle A, Giacomini A, Ponsiglione M (2007) Piecewise rigidity. J Funct Anal 244: 134-153. doi: 10.1016/j.jfa.2006.11.006
    [23] Conti S, Focardi M, Iurlano F (2018) Which special functions of bounded deformation have bounded variation? P Roy Soc Edinb A 148: 33-50. doi: 10.1017/S030821051700004X
    [24] Conti S, Focardi M, Iurlano F (2017) Integral representation for functionals defined on S BDp in dimension two. Arch Rat Mech Anal 223: 1337-1374. doi: 10.1007/s00205-016-1059-y
    [25] Chambolle A, Crismale V (2019) A density result in GS BDp with applications to the approximation of brittle fracture energies. Arch Ration Mech Anal 232: 1329-1378. doi: 10.1007/s00205-018-01344-7
    [26] Chambolle A, Crismale V (2018) Compactness and lower semicontinuity in GS BD. J Eur Math Soc http://arxiv.org/abs/1802.03302.
    [27] Chambolle A, Crismale V (2018) Existence of strong solutions to the Dirichlet problem for Griffith energy. Calc Var Partial Dif 58: 136.
    [28] Crismale V (2018) On the approximation of S BD functions and some applications. http://arxiv.org/abs/1806.03076.
    [29] Cortesani G, Toader R (1999) A density result in SBV with respect to non-isotropic energies. Nonlinear Anal 38: 585-604. doi: 10.1016/S0362-546X(98)00132-1
    [30] Dal Maso G (1993) An Introduction to Γ-Convergence, Basel: Birkhäuser.
    [31] Dal Maso G (2013) Generalized functions of bounded deformation. J Eur Math Soc 15: 1943-1997. doi: 10.4171/JEMS/410
    [32] Dal Maso G, Francfort GA, Toader R (2005) Quasistatic crack growth in nonlinear elasticity. Arch Ration Mech Anal 176: 165-225. doi: 10.1007/s00205-004-0351-4
    [33] Dal Maso G, Negri M, Percivale D (2002) Linearized elasticity as Γ-limit of finite elasticity. Set-valued Anal 10: 165-183. doi: 10.1023/A:1016577431636
    [34] Dal Maso G, Lazzaroni G (2010) Quasistatic crack growth in finite elasticity with noninterpenetration. Ann Inst H Poincaré Anal Non Linéaire 27: 257-290.
    [35] De Giorgi E, Ambrosio L (1988) Un nuovo tipo di funzionale del calcolo delle variazioni. Acc Naz Lincei, Rend Cl Sci Fis Mat Natur 82: 199-210.
    [36] Dunn JE, Serrin J (1985) On the thermomechanics of interstitial working. Arch Ration Mech Anal 88: 95-133. doi: 10.1007/BF00250907
    [37] Francfort GA, Marigo JJ (1998) Revisiting brittle fracture as an energy minimization problem. J Mech Phys Solids 46: 1319-1342. doi: 10.1016/S0022-5096(98)00034-9
    [38] Francfort GA, Larsen CJ (2003) Existence and convergence for quasi-static evolution in brittle fracture. Comm Pure Appl Math 56: 1465-1500. doi: 10.1002/cpa.3039
    [39] Friedrich M (2017) A derivation of linearized Griffith energies from nonlinear models. Arch Ration Mech Anal 225: 425-467. doi: 10.1007/s00205-017-1108-1
    [40] Friedrich M (2017) A Korn-type inequality in SBD for functions with small jump sets. Math Mod Meth Appl Sci 27: 2461-2484. doi: 10.1142/S021820251750049X
    [41] Friedrich M (2018) A piecewise Korn inequality in SBD and applications to embedding and density results. SIAM J Math Anal 50: 3842-3918. doi: 10.1137/17M1129982
    [42] Friedrich M (2019) A compactness result in GS BVp and applications to Γ-convergence for free discontinuity problems. Calc Var Partial Dif 58: 86. doi: 10.1007/s00526-019-1530-3
    [43] Friedrich M, Kružík M (2018) On the passage from nonlinear to linearized viscoelasticity. SIAM J Math Anal 50: 4426-4456. doi: 10.1137/17M1131428
    [44] Friedrich M, Schmidt B (2015) On a discrete-to-continuum convergence result for a two dimensional brittle material in the small displacement regime. Netw Heterog Media 10: 321-342. doi: 10.3934/nhm.2015.10.321
    [45] Friedrich M, Solombrino F (2018) Quasistatic crack growth in 2d-linearized elasticity. Ann Inst H Poincaré Anal Non Linéaire 35: 27-64.
    [46] Friesecke G, James RD, MÜller S (2002) A theorem on geometric rigidity and the derivation of nonlinear plate theory from three-dimensional elasticity. Comm Pure Appl Math 55: 1461-1506. doi: 10.1002/cpa.10048
    [47] Giacomini A (2005) Ambrosio-Tortorelli approximation of quasi-static evolution of brittle fractures. Calc Var Partial Dif 22: 129-172. doi: 10.1007/s00526-004-0269-6
    [48] Giacomini A, Ponsiglione M (2006) A Γ-convergence approach to stability of unilateral minimality properties. Arch Ration Mech Anal 180: 399-447. doi: 10.1007/s00205-005-0392-3
    [49] Iurlano F (2014) A density result for GSBD and its application to the approximation of brittle fracture energies. Calc Var Partial Dif 51: 315-342. doi: 10.1007/s00526-013-0676-7
    [50] Mielke A, Roubíček T (2015) Rate-Independent Systems: Theory and Application, New York: Springer.
    [51] Mielke A, Stefanelli U (2013) Linearized plasticity is the evolutionary Γ-limit of finite plasticity. J Eur Math Soc 15: 923-948. doi: 10.4171/JEMS/381
    [52] Negri M, Toader R (2015) Scaling in fracture mechanics by Bažant's law: from finite to linearized elasticity. Math Mod Meth Appl Sci 25: 1389-1420. doi: 10.1142/S0218202515500360
    [53] Negri M, Zanini C (2014) From finite to linear elastic fracture mechanics by scaling. Calc Var Partial Dif 50: 525-548. doi: 10.1007/s00526-013-0645-1
    [54] Schmidt B (2008) Linear Γ-limits of multiwell energies in nonlinear elasticity theory. Continuum Mech Thermodyn 20: 375-396. doi: 10.1007/s00161-008-0087-8
    [55] Schmidt B (2009) On the derivation of linear elasticity from atomistic models. Netw Heterog Media 4: 789-812. doi: 10.3934/nhm.2009.4.789
    [56] Schmidt B, Fraternali F, Ortiz M (2009) Eigenfracture: An eigendeformation approach to variational fracture. SIAM Mult Model Simul 7: 1237-1266. doi: 10.1137/080712568
    [57] Toupin RA (1962) Elastic materials with couple stresses. Arch Ration Mech Anal 11: 385-414. doi: 10.1007/BF00253945
    [58] Toupin RA (1964) Theory of elasticity with couple stress. Arch Ration Mech Anal 17: 85-112. doi: 10.1007/BF00253050
  • This article has been cited by:

    1. L. De Luca, M. Ponsiglione, Variational models in elasticity, 2021, 3, 2640-3501, 1, 10.3934/mine.2021015
    2. Rufat Badal, Manuel Friedrich, Martin Kružík, Nonlinear and Linearized Models in Thermoviscoelasticity, 2023, 247, 0003-9527, 10.1007/s00205-022-01834-9
    3. Maria Giovanna Mora, Filippo Riva, Pressure live loads and the variational derivation of linear elasticity, 2022, 0308-2105, 1, 10.1017/prm.2022.79
    4. Manuel Friedrich, Leonard Kreutz, Konstantinos Zemas, Derivation of effective theories for thin 3D nonlinearly elastic rods with voids, 2024, 34, 0218-2025, 723, 10.1142/S0218202524500131
    5. Stefano Almi, Elisa Davoli, Manuel Friedrich, Non-interpenetration conditions in the passage from nonlinear to linearized Griffith fracture, 2023, 175, 00217824, 1, 10.1016/j.matpur.2023.05.001
    6. Stefano Almi, Maicol Caponi, Manuel Friedrich, Francesco Solombrino, Geometric rigidity on Sobolev spaces with variable exponent and applications, 2025, 32, 1021-9722, 10.1007/s00030-024-01016-4
    7. Antonio Flavio Donnarumma, Manuel Friedrich, Stochastic homogenisation for functionals defined on asymptotically piecewise rigid functions, 2025, 64, 0944-2669, 10.1007/s00526-025-02930-w
    8. Rita Ferreira, José Matias, Elvira Zappale, Junction in a thin multi-domain for nonsimple grade two materials in BH, 2025, 84, 14681218, 104322, 10.1016/j.nonrwa.2025.104322
  • Reader Comments
  • © 2020 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(3856) PDF downloads(419) Cited by(8)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog