Loading [MathJax]/jax/element/mml/optable/GeneralPunctuation.js
Research article

Signorini problem as a variational limit of obstacle problems in nonlinear elasticity

  • An energy functional for the obstacle problem in linear elasticity is obtained as a variational limit of nonlinear elastic energy functionals describing a material body subject to pure traction load under a unilateral constraint representing the rigid obstacle. There exist loads pushing the body against the obstacle, but unfit for the geometry of the whole system body-obstacle, so that the corresponding variational limit turns out to be different from the classical Signorini problem in linear elasticity. However, if the force field acting on the body fulfils an appropriate geometric admissibility condition, we can show coincidence of minima. The analysis developed here provides a rigorous variational justification of the Signorini problem in linear elasticity, together with an accurate analysis of the unilateral constraint.

    Citation: Francesco Maddalena, Danilo Percivale, Franco Tomarelli. Signorini problem as a variational limit of obstacle problems in nonlinear elasticity[J]. Mathematics in Engineering, 2024, 6(2): 261-304. doi: 10.3934/mine.2024012

    Related Papers:

    [1] 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
    [2] Michiaki Onodera . Linear stability analysis of overdetermined problems with non-constant data. Mathematics in Engineering, 2023, 5(3): 1-18. doi: 10.3934/mine.2023048
    [3] Manuel Friedrich . Griffith energies as small strain limit of nonlinear models for nonsimple brittle materials. Mathematics in Engineering, 2020, 2(1): 75-100. doi: 10.3934/mine.2020005
    [4] L. De Luca, M. Ponsiglione . Variational models in elasticity. Mathematics in Engineering, 2021, 3(2): 1-4. doi: 10.3934/mine.2021015
    [5] Matteo Novaga, Marco Pozzetta . Connected surfaces with boundary minimizing the Willmore energy. Mathematics in Engineering, 2020, 2(3): 527-556. doi: 10.3934/mine.2020024
    [6] Jason Howell, Katelynn Huneycutt, Justin T. Webster, Spencer Wilder . A thorough look at the (in)stability of piston-theoretic beams. Mathematics in Engineering, 2019, 1(3): 614-647. doi: 10.3934/mine.2019.3.614
    [7] Matteo Fogato . Asymptotic finite-dimensional approximations for a class of extensible elastic systems. Mathematics in Engineering, 2022, 4(4): 1-36. doi: 10.3934/mine.2022025
    [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] Sergio Conti, Patrick Dondl, Julia Orlik . Variational modeling of paperboard delamination under bending. Mathematics in Engineering, 2023, 5(2): 1-28. doi: 10.3934/mine.2023039
    [10] Biagio Cassano, Lucrezia Cossetti, Luca Fanelli . Spectral enclosures for the damped elastic wave equation. Mathematics in Engineering, 2022, 4(6): 1-10. doi: 10.3934/mine.2022052
  • An energy functional for the obstacle problem in linear elasticity is obtained as a variational limit of nonlinear elastic energy functionals describing a material body subject to pure traction load under a unilateral constraint representing the rigid obstacle. There exist loads pushing the body against the obstacle, but unfit for the geometry of the whole system body-obstacle, so that the corresponding variational limit turns out to be different from the classical Signorini problem in linear elasticity. However, if the force field acting on the body fulfils an appropriate geometric admissibility condition, we can show coincidence of minima. The analysis developed here provides a rigorous variational justification of the Signorini problem in linear elasticity, together with an accurate analysis of the unilateral constraint.



    In its original formulation (see [54]) the Signorini problem in linear elastostatics consists in finding the equilibrium configuration of an elastic body Ω resting on a frictionless rigid support EΩ in its natural configuration and subject to body forces and surface forces acting on ΩE; precisely, if u:ΩR3 denotes the displacement field of the body, C represents the classical linear elasticity tensor and E denotes the linear strain tensor, we assume that

    Q(x,E):=12ETC(x)E

    is the corresponding strain energy density (see [27]) and that the body is subject to a load system of forces f:ΩR3 and g:ΩER3 such that

    L(u):=Ωfu dx+ΩEgu dH2 (1.1)

    is the load potential, where H2 is the two-dimensional Hausdorff measure. Assuming that H2(E)>0, the variational formulation of the Signorini problem consists in finding a minimizer of the functional

    E(u):=ΩQ(x,E(u))dxL(u) (1.2)

    among all u in the Sobolev space H1(Ω;R3) such that un0 H2-a.e. on E, where n is the inward unit vector normal to Ω. A classical result (see [22]) states that a solution of (1.2) exists if the following condition is verified: Every infinitesimal rigid displacement v fulfills L(v)0 if vn0 H2-a.e. on E and L(v)=0 if and only if vn0 H2-a.e. on E. Moreover if E is planar, that is EΩ{x3=0}, and if L(e3)<0, fC0,α(¯Ω;R3) and gL2(ΩE;R3) then a minimizer of (1.2) exists if and only if the above condition holds (see [22, Theorem XXXII] and [10]): In particular if Ω is the cylinder

    Ω:={x:(x1ax3)2+x22<R2, 0<x3<H},
    E:={x:x21+x22<R2, x3=0},

    with a0,R>0,H>0, and f=e3, g=0, then a minimum is attained if and only if aH<2R that is 0ϑ:=arctana<arctan2R/H where ϑ is the inclination of the cylinder with respect to the x3-axis.

    More recent formulations of constrained problems in the calculus of variations use the notion of capacity (see Section 2 for details) leading to consider more general geometries since any set of null capacity has null H2 measure (see [57, Theorem 4]) but there exist sets of null H2 measure and strictly positive capacity (see [1, Theorem 5.4.1]). Indeed, a proper generalization of the latter case is to assume that the set E{x30} has positive capacity and accordingly modify the obstacle condition by requiring x3+u3(x)0 on E up to a set of null capacity (shortly, q.e. on E): The existence of minimizers for this general setting was proved by [12, Theorem 4.5]. Although the original obstacle formulation given in [54] may look different from the generalized notion exploited in this work, it can be shown (see Remark 2.3) that if the set EΩ is regular in an appropriate sense (see Remark 2.3) then the two frameworks coincide.

    Like in the analysis of many problems in linear elastostatics, it is quite natural to ask whether there exists a sequence of functionals in finite elasticity whose minimizing sequences converge to a minimizer of (1.2), possibly under suitable compatibility conditions on the functional L: Such an approach provides a variational justification of the linearized theory and could help finding other reasonable models rigorously deduced.

    In this paper we show sharp conditions on L entailing that a wide class of energy functionals in finite elasticity fulfill this variational property in the context of obstacle problems; in addition we also show examples of loads leading to the failure of this convergence. In this perspective, denoting by y:ΩR3 the deformation field and by h>0 an adimensional parameter, we introduce a family of energy functionals defined by

    Fh(y):=h2ΩW(x,y(x))dxh1L(yx) (1.3)

    where L is defined as in (1.1) and W:Ω×R3×3[0,+] is the strain energy density. For every xΩ, the function W(x,) is assumed to be frame indifferent and attaining its minimum value 0 at rigid deformations only. We also assume that W is C2-regular in a neighborhood of rigid deformations and satisfies a natural coercivity condition, see (2.22).

    According to a standard approach in the deduction of linearized theories in continuum mechanics, if yh is a minimizing sequence of Fh (see (2.40)) in a class Ah of deformations satisfying a suitable obstacle constraint, we aim to investigate whether F(yh) converges, as h goes to 0, to the minimum of E (with C=D2W(x,I)) among displacements fulfilling u3(x)+x30 q.e. on E. Since here the aim is the deduction of the Signorini problem in linear elasticity, it is natural to assume that the unilateral constraint in nonlinear approximating problems takes the form x3+h1(yh,3x3)0 on E that is

    yh,3(1h)x3  on E. (1.4)

    We define the functionals Gh coupling the energies Fh with the unilateral constraint due to rigid obstacle:

    Gh(y)={Fh(y),  if  y3(1h)x3 on E, +,  else, (1.5)

    where E, the portion of the elastic body sensitive to the obstacle, has an horizontal projection with non negligible capacity. Moreover we have to assume that

    L(yx)0 (1.6)

    for every deformation y=y(x) fulfilling (1.4) and such that

    ΩW(x,y)dx=0. (1.7)

    On the contrary if y satisfied (1.4) and (1.7) but L(yx)>0 then

    Gh(y)=h1L(yx) as h0+.

    Under our assumptions on W, Eq (1.7) holds true if and only if y is rigid deformation, i.e., y(x)=Rx+c for some RSO(3) and cR3, while (1.4) is fulfilled by these y if and only if (Rx)3+c3(1h)x3 on E, a condition which is satisfied for every h>0 if and only if

    c3((RI)x)3on E. (1.8)

    Thus, due to (1.6) we have to assume this geometrical compatibility between load and obstacle

    L((RI)x+c)0,RSO(3), cR3 verifying (1.8). (1.9)

    Nevertheless though (1.9) entails the compatibility assumptions of Theorem 4.5 in [12] and though compatibility assumptions of Theorem 4.5 in [12] entail the existence of minimizers of the Signorini functional in linear elasticity, these compatibility assumptions alone do not warrant the equiboundedness from below of approximating functionals Gh (see Example 3.6 below). In the main result of this paper (Theorem 2.4) we show that if L satisfies the necessary condition (1.9) together with L(e3)<0 and

    L((Rxx)1e1+(Rxx)2e2)0,RSO(3), (1.10)

    then, under some capacitary assumptions on E (see (2.13)), we have

    limh0(infGh)=minG, (1.11)

    where

    G(u):={ΩQ(x,E(u))dxmaxSL,EL(Ru),  if  x3+u30 on E,  +,  else,  (1.12)
    Q(x,F):=12FTD2W(x,I)F,FR3×3,xΩ,
    SL,E={RSO(3): L((RI)x)minxEess((Rx)3x3)L(e3)=0}.

    Along this paper Eess denotes the essential part of E with respect to the capacity (see (2.9)), a closed canonical subset of ¯E such that EEess has null capacity.

    Under the hypotheses detailed previously we will show (see Lemma 3.8) that either SL,E{I} or SL,E={RSO(3):Re3=e3}. If SL,E{I} then clearly GE, hence in this case the minimum of Signorini problem in linearized elasticity is the limit of the infGh but, quite surprisingly, the second alternative is much more subtle and indeed we are able to exhibit examples such that

    minG<minE, (1.13)

    namely a gap between limh0(infGh) and minE may appear (see Section 5). However the coincidence of minimizers of G and E may hold true even if SL,E is not reduced to the identity matrix: In particular, if Ω is contained in the upper half-space, E is either ¯Ω or Ω, the load

    L(v):=Ωfv3dx+Ωgv3dH2

    satisfies condition (1.9) and L(e3)<0, then L(v)=L(Rv) for every RSL,E hence minG=minE, say the energy of minimizing sequences for Gh converges to the minimum energy of E. On the other hand, it is always possible to rotate the external forces in such a way that GR and ER (the functionals obtained replacing the load functional L with LR defined by LR(v):=L(Rv)) have the same minimum as shown in Theorem 5.5.

    For several contributions facing issues strictly connected with the context of the present paper we refer to [3,4,5,6,8,9,11,13,14,15,16,17,18,19,21,23,25,26,27,28,29,32,33,34,35,36,37,38,39,40,41,42,43,46,47,48,49,50,51,52,53,55,56].

    We set a+:=max{a,0}, a:=max{a,0} for every aR; notations x=(x1,x2,x3) and y=(y1,y2,y3) represent generic points in R3; ej, j=1,2,3 denote the unitary basis vectors of R3, R3×3 is the set of 3×3 real matrices, endowed with the Euclidean norm |F|=Tr(FTF). R3×3sym (resp. R3×3skew) denotes the subset of symmetric (resp. skew-symmetric) matrices. For every FR3×3 we define symF:=12(F+FT), SO(3) will denote the special orthogonal group and for every RSO(3) there exist ϑ[0,2π] and aR3, |a|=1 such that the following Euler-Rodrigues representation formula holds

    Rx=x+sinϑ(ax)+(1cosϑ)(a(ax)),xR3. (2.1)

    For every compact set KRN we define the capacity of K by setting (see [1, Definition 2.2.1])

    capK=inf{ (2.2)

    If G\subset {{\mathbb R}}^N is open we define (see [1, Definition 2.2.2])

    \begin{equation} {\mathop{{{\rm{cap}}}}\nolimits} G: = \sup\{{\mathop{{{\rm{cap}}}}\nolimits} K: K\ \text{compact},\ K\subset G\} \end{equation} (2.3)

    and, since (see [1, Proposition 2.2.3])

    \begin{equation} {\mathop{{{\rm{cap}}}}\nolimits} K = \inf\{{\mathop{{{\rm{cap}}}}\nolimits} G: G\ \text{open},\ K\subset G\}, \qquad \forall\,K\text{ compact} , \end{equation} (2.4)

    we may extend the above definitions to an arbitrary set by setting (see [1, Definition 2.2.4])

    \begin{equation} {\mathop{{{\rm{cap}}}}\nolimits} E: = \inf\{{\mathop{{{\rm{cap}}}}\nolimits} G: G\ \text{open},\ E\subset G\}, \qquad \forall\,E\subset {{\mathbb R}}^N . \end{equation} (2.5)

    A straightforward consequence of (2.3) and (2.5) is that

    \begin{equation} {\mathop{{{\rm{cap}}}}\nolimits} E_1\le {\mathop{{{\rm{cap}}}}\nolimits} E_2,\qquad \forall\ E_1\subset E_2\subset {{\mathbb R}}^N. \end{equation} (2.6)

    On the other hand, for every E\subset {{\mathbb R}}^N the Bessel capacity is defined as (see [1, Definition 2.3.3])

    \begin{equation} {\mathop{{{\rm{Cap}}}}\nolimits} E: = \inf \left\{\|f\|^2_{L^2({{\mathbb R}}^N)} : \,f\ge 0,\ \text{a.e. on}\,{{\mathbb R}}^N, \,\,(f*g_1)({{\bf x}})\ge 1\ \forall\,{{\bf x}}\in E\right\}, \end{equation} (2.7)

    where g_1\!\in\! L^1({{\mathbb R}}^3) is the first-order Bessel kernel in {{\mathbb R}}^N defined as the inverse Fourier transform of (1+|\boldsymbol\xi|^2)^{-1/2} , say

    g_1({{\bf x}}): = (2\pi)^{-N}\int_{{{\mathbb R}}^3}(1+|{\boldsymbol{\xi}}|^2)^{-1/2}e^{i{{\bf x}}\cdot{\boldsymbol{\xi}}}\,d{\boldsymbol{\xi}} = \frac 1 {2\pi}\int_0^\infty t^{-(N+1)/2} e^{-\pi|{\mathbf{x}}|^2/t} e^{-t/(4\pi)}\,dt.

    Notice that since f\ge 0 a.e. we have that f*g_1 is everywhere defined if we allow it to take the value +\infty (see [1, Definition 2.3.1]) and that f*g_1 is l.s.c. by Proposition 2.3.2 of [1]. Thus inequality (f*g_1)({{\bf x}})\ge 1 for every {{\bf x}}\in E appearing in formula (2.7) has a precise meaning.

    In addition it is possible to show that there exist two constants \alpha, \beta > 0 such that

    \begin{equation} \alpha {\mathop{{{\rm{Cap}}}}\nolimits} E\le {\mathop{{{\rm{cap}}}}\nolimits} E\le \beta {\mathop{{{\rm{Cap}}}}\nolimits} E,\quad \forall E\subset{{\mathbb R}}^N, \end{equation} (2.8)

    (see [1, Definition 2.2.6 and Proposition 2.3.13]).

    A property is said to hold quasi-everywhere (q.e. for short) if it holds true outside a set of zero capacity. It is convenient to introduce (see [12]) a canonical representative of the set E , called the essential part of E and denoted by E_{ess} , which nevertheless coincides with E itself whenever it is a smooth closed manifold or the closure of an open subset of \mathbb R^N .

    For every set E\subset{{\mathbb R}}^3 we define the essential part E_{ess} of E (with respect to the capacity) by

    \begin{equation} E_{ess}\ : = \ \bigcap \,\{\, C:\ C \hbox{ is closed and } {\mathop{{{\rm{cap}}}}\nolimits}(E\!\setminus\! C) = 0\,\}. \end{equation} (2.9)

    It has been shown in [12] that

    \begin{equation} E_{ess}\ \text{ is a closed subset of}\ \overline E, \end{equation} (2.10)
    \begin{equation} {\mathop{{{\rm{cap}}}}\nolimits}(E\!\setminus\! E_{ess}) = 0, \end{equation} (2.11)
    \begin{equation} {\mathop{{{\rm{cap}}}}\nolimits} E = 0\ \text{ if and only if}\ E_{ess} = \emptyset. \end{equation} (2.12)

    In the following {\mathop{{{\rm{co}}}}\nolimits} A , {\mathop{{{\rm{aff}}}}\nolimits} A , {\mathop{{{\rm{ri}}}}\nolimits} A , {\rm r}\partial A and {\mathop{{{\rm{proj}}}}\nolimits} A denote respectively, the closed convex hull of the set A\!\subset\! {{\mathbb R}}^3 (say, the intersection of all convex sets containing A ), the affine hull of the set A (say, the smallest affine space containing A ), the relative interior of A (say, the interior part of A with respect to the affine hull of A ), the relative boundary of A (say, the boundary of A with respect to the affine hull of A : {\mathop{{{\rm{ri}}}}\nolimits}\partial A = \overline A\setminus{\mathop{{{\rm{ri}}}}\nolimits} A ) and the projection of A onto the horizontal plane \{x_3 = 0\} .

    Throughout the paper we will assume that

    \begin{equation} {\mathop{{{\rm{cap}}}}\nolimits} ({\mathop{{{\rm{proj}}}}\nolimits} ({\mathop{{{\rm{co}}}}\nolimits} E_{ess})) > 0 . \end{equation} (2.13)

    Notice that {\mathop{{{\rm{cap}}}}\nolimits} E\! > \!0 does not imply (2.13) whereas the converse is true: Indeed, if {\mathop{{{\rm{cap}}}}\nolimits} E\! = \!0 then by (2.12) we get E_{ess}\! = \!\emptyset so {\mathop{{{\rm{proj}}}}\nolimits} ({\mathop{{{\rm{co}}}}\nolimits} E_{ess})\! = \emptyset and {\mathop{{{\rm{cap}}}}\nolimits} ({\mathop{{{\rm{proj}}}}\nolimits} ({\mathop{{{\rm{co}}}}\nolimits} E_{ess}))\! = \! 0 , a contradiction to (2.13).

    In the following \Omega will denote the reference configuration of an elastic body and it is always assumed to be a nonempty, bounded, connected, Lipschitz open set in \mathbb R^3 . We need to show that any function in the Sobolev space H^{1}(\Omega; {{\mathbb R}}^3) actually has a precise representative defined quasi-everywhere on the whole \overline\Omega with respect to the capacity. Indeed, if {\mathbf{uu}}\!\in\! H^{1}(\Omega; {{\mathbb R}}^3) and {\mathbf{v}}\!\in\! H^{1}({{\mathbb R}}^3;{{\mathbb R}}^3) is a Sobolev extension of {\mathbf{uu}} , it is well known (see [1, Proposition 6.1.3]) that the limit

    \begin{equation} {{\mathbf{v}}}^* ({\mathbf{x}})\ : = \ \lim\limits_{r \downarrow 0}\ \frac{1}{|B_r({\mathbf{x}})|}\,\int_{ B_r({\mathbf{x}})}\! {\mathbf{v}}(\mathbf \xi)\,d\mathbf \xi \end{equation} (2.14)

    exists for q.e. {{\bf x}}\!\in\!{{\mathbb R}}^3. The function {{\mathbf{v}}}^* is called the precise representative of {\mathbf{v}} and is a quasicontinuous function in {{\mathbb R}}^3 , that is to say, for every {\varepsilon} > 0 there exists an open set V\subset {{\mathbb R}}^3 such that {\mathop{{{\rm{cap}}}}\nolimits} V < {\varepsilon} and {{\mathbf{v}}}^* is continuous in {{\mathbb R}}^3\!\setminus\! V . We claim that if {\mathbf{v}}_1, \ {\mathbf{v}}_2 are two distinct Sobolev extensions of {\mathbf{uu}} then

    \begin{equation} {\mathbf{v}}_1^*({\mathbf{x}})\ = \ {\mathbf{v}}_2^*({\mathbf{x}}), \qquad \text{q.e.}\ {\mathbf{x}}\in \overline \Omega. \end{equation} (2.15)

    The claim is trivial for q.e. {\mathbf{x}}\in \Omega , thus we are left to show (2.15) for q.e. {\mathbf{x}}\in \partial \Omega .

    Ler R > 0 such that \overline \Omega\subset B_R(0) and let \Omega_R: = B_R(0)\setminus \overline \Omega . Since \Omega has Lipschitz boundary it is well known (see [2]) that

    \begin{equation} \lim\limits_{r\downarrow 0}\frac{|B_r({\mathbf{x}})\cap \Omega_R|}{|B_r({\mathbf{x}})|} = \frac{1}{2} \end{equation} (2.16)

    and for \mathcal H^2 a.e. {\mathbf{x}}\in \partial \Omega ,

    \begin{equation} {\mathbf{uu}}({\mathbf{x}}) = \lim\limits_{r \downarrow 0}\frac{2}{|B_r({\mathbf{x}})|} \!\int_{ B_r({\mathbf{x}})\cap \Omega_R} \!\!\!\! {\mathbf{v}}_1(\mathbf \xi)\,d\mathbf \xi = \lim\limits_{r \downarrow 0}\ \frac{2}{|B_r({\mathbf{x}})|} \!\int_{ B_r({\mathbf{x}})\cap \Omega_R}\! \!\!\!\! {\mathbf{v}}_2(\mathbf \xi)\,d\mathbf \xi, \quad \mathcal H^2\,\hbox{a.e.}\,{\mathbf{x}}\in \!\partial \Omega, \end{equation} (2.17)

    where we have denoted again with {\mathbf{uu}} the trace of {\mathbf{uu}} on \partial \Omega . Hence

    \frac{1}{|B_r({\mathbf{x}})|}\,\int_{ B_r({\mathbf{x}})\cap \Omega_R}\! ({\mathbf{v}}_1(\mathbf \xi)-{\mathbf{v}}_2(\mathbf \xi))\,d\mathbf \xi = 0,\quad \mathcal H^2\, \text{a.e.}\ {\mathbf{x}}\in \partial \Omega,

    so, by taking account \partial \Omega\subset\partial \Omega_R and by recalling that \partial \Omega is Lipschitz, we may apply Theorem 2.1 of [20] to {\mathbf{v}}_1-{\mathbf{v}}_2\in H^1(\Omega_R; {{\mathbb R}}^3) and we get

    \frac{1}{|B_r({\mathbf{x}})|}\,\int_{ B_r({\mathbf{x}})\cap \Omega_R}\! ({\mathbf{v}}_1(\mathbf \xi)-{\mathbf{v}}_2(\mathbf \xi))\,d\mathbf \xi = 0, \qquad \hbox{ q.e. }{\mathbf{x}}\in \partial \Omega.

    Since {\mathbf{v}}_1 = {\mathbf{v}}_2 = {\mathbf{uu}} a.e. in B_r({\mathbf{x}})\!\setminus\! \Omega_R the claim follows easily by (2.14). Therefore if {\mathbf{uu}}\in H^{1}(\Omega; {{\mathbb R}}^3) we may define its precise representative for quasi-every {\mathbf{x}} on \overline \Omega by

    \begin{equation} {{\mathbf{uu}}}^* ({\mathbf{x}}) = \lim\limits_{r \downarrow 0}\ \frac{1}{|B_r({\mathbf{x}})|}\,\int_{ B_r({\mathbf{x}})}\! {\mathbf{v}}(\mathbf \xi)\,d\mathbf \xi, \qquad \text{q.e. }{\mathbf{x}}\in \overline \Omega, \end{equation} (2.18)

    where {\mathbf{v}} is any Sobolev extension of {\mathbf{uu}} .

    The function {{\mathbf{uu}}}^* is pointwise quasi-everywhere defined by (2.18) and is quasicontinuous on \overline \Omega i.e., for every {\varepsilon} > 0 there exists a relatively open set V\subset \overline \Omega such that {\mathop{{{\rm{cap}}}}\nolimits} V < {\varepsilon} and {{\mathbf{uu}}}^* is continuous in \overline \Omega\setminus V .

    Let {\mathcal L}^3 and {\mathcal B}^3 denote respectively the \sigma\mbox{-algebras} of Lebesgue measurable and Borel measurable subsets of {{\mathbb R}}^3 and let \mathcal W : \Omega \times \mathbb R^{3 \times 3} \to [0, +\infty ] be {\mathcal L}^3\! \times\! {\mathcal B}^{9} -measurable satisfying the following assumptions, see also [3,45]:

    \begin{equation} {\mathcal W}({{\bf x}}, \mathbf R\mathbf F) = {\mathcal W}({{\bf x}}, \mathbf F), \qquad \ \forall \ \mathbf\! \mathbf R\!\in\! SO(3), \ \ \forall\ \mathbf F\!\in\! \mathbb R^{3 \times 3},\ \ \mbox{for a.e. } {{\bf x}}\!\in\!\Omega, \end{equation} (2.19)
    \begin{equation} \min\limits_{\mathbf F} {\mathcal W}({{\bf x}},\mathbf F) = {\mathcal W}({{\bf x}},\mathbf I) = 0, \quad \mbox{for a.e. } {{\bf x}}\in\Omega \end{equation} (2.20)

    and as far as it concerns the regularity of \mathcal W , we assume that there exist an open neighborhood \mathcal U of SO(3) in {{\mathbb R}}^{3\times3} , an increasing function \omega:\mathbb R_+\to\mathbb R satisfying \lim_{t\to0^+}\omega(t) = 0 and a constant K > 0 such that for a.e. {{\bf x}}\in \Omega

    \begin{equation} \begin{array}{ll} & \mathcal W({{\bf x}},\cdot)\in C^{2}(\mathcal U),\;\;\; |D^2 \mathcal W({{\bf x}},\mathbf I)|\le K \;\;\hbox{and}\\ & |D^2 {\mathcal W}({{\bf x}},\mathbf F)-D^2 {\mathcal W}({{\bf x}},\mathbf G)|\le\omega(|\mathbf F-\mathbf G|),\quad\forall\; \mathbf F,\mathbf G\in\mathcal U. \end{array} \end{equation} (2.21)

    Moreover we assume that there exists C > 0 such that for a.e. {{\bf x}}\in\Omega

    \begin{equation} \begin{array}{ll} {\mathcal W}({{\bf x}},\mathbf F)\ge C (d(\mathbf F, SO(3)))^2,\qquad \forall\, \mathbf F\in \mathbb R^{3 \times 3}, \end{array} \end{equation} (2.22)

    where d(\, \cdot\, ,SO(3)) denotes the Euclidean distance function from the set of rotations.

    The frame indifference assumption (2.19) implies that there exists a function \mathcal V such that

    \begin{equation} {\mathcal W}({{\bf x}},\mathbf F) = {\mathcal V}({{\bf x}}, {\frac{1}{2}}( \mathbf F^T \mathbf F - \mathbf I)), \qquad \hbox{for a.e. } {{\bf x}}\in \Omega , \ \forall\, \mathbf F\in \mathbb R^{3\times 3}. \end{equation} (2.23)

    By (2.19)–(2.21), for a.e. {{\bf x}}\in\Omega , we have \mathcal W({{\bf x}}, \mathbf R) = D\mathcal W({{\bf x}}, \mathbf R) = \mathbf 0 for any \mathbf R\in SO(3). By (2.23), for a.e. {{\bf x}}\in \Omega , given \mathbf B\in\mathbb R^{3\times 3} and h > 0 , we have

    \mathcal W({{\bf x}},{\mathbf{I}}+h\mathbf B) = {\mathcal V}({{\bf x}},h\,{\rm sym}\mathbf B+\tfrac12h^{2}\mathbf B^{T}\mathbf B)

    and (2.20), (2.21) together imply

    \lim\limits_{h\to 0} h^{-2}\mathcal W({{\bf x}},\mathbf I+h\mathbf B) = \frac{1}{2} \,{\rm sym} \mathbf B\, D^2 {\mathcal V} ({{\bf x}}, \mathbf 0) \ {\rm sym}\mathbf B = \frac12\, \mathbf B^T D^2\mathcal W({{\bf x}},\mathbf I)\,\mathbf B, \qquad\forall \,\mathbf B\in\mathbb R^{3\times3}.

    Hence, by (2.22) and polar decomposition [27], we obtain, for a.e. {{\bf x}}\!\in\! \Omega and every \mathbf B\!\in\!\mathbb R^{3\times3} , eventually as h\to 0_+ (since \det(\mathbf I+h\mathbf B)\! > \!0 for small h )

    \begin{equation*} \label{ellipticity}\begin{aligned} \frac12\, \mathbf B^T D^2\mathcal W({{\bf x}},\mathbf I)\,\mathbf B& = \lim\limits_{h\to 0} h^{-2}\mathcal W({{\bf x}},\mathbf I+h\mathbf B) \ge \limsup\limits_{h\to 0}Ch^{-2}\,d^2(\mathbf I+h\mathbf B,SO(3)) \\& = \limsup\limits_{h\to 0} C h^{-2}\left|\sqrt{(\mathbf I+h\mathbf B)^T(\mathbf I+h\mathbf B)}-\mathbf I\right|^2 = C|\mathrm{sym}\mathbf B|^2.\end{aligned} \end{equation*}

    Moreover, as noticed also in [44], by expressing the remainder of Taylor's expansion in terms of the {{\bf x}} -independent modulus of continuity \omega of D^2 {\mathcal W}({{\bf x}}, \cdot) on the set \mathcal U from (2.21), we have

    \begin{equation} \left|\mathcal W({{\bf x}}, \mathbf I+h\mathbf B)- \frac{h^2}{2} \,{\rm sym} \mathbf B\, D^2\mathcal W ({{\bf x}}, \mathbf I) \ {\rm sym}\mathbf B\right|\le h^2\omega(h|\mathbf B|)|\mathbf B|^2 \end{equation} (2.24)

    for any small enough h (such that h\mathbf B\in\mathcal U ). Similarly, \mathcal V({{\bf x}}, \cdot) is C^2 in a neighborhood of the origin in \mathbb R^{3\times 3} , with an {{\bf x}} -independent modulus of continuity \eta:\mathbb R_+\to \mathbb R , which is increasing and such that \lim_{t\to0^+}\eta(t) = 0 , and we have

    \begin{equation} \left|\mathcal V({{\bf x}}, h\mathbf B)- \frac{h^2}{2} \,{\rm sym} \mathbf B\, D^2 {\mathcal V} ({{\bf x}}, \mathbf 0) \ {\rm sym}\mathbf B\right|\le h^2\eta(h|\mathbf B|)|\mathbf B|^2 \end{equation} (2.25)

    for any small enough h .

    We mention a general class of energy densities {\mathcal W} (the so called Yeoh materials) fulfilling the assumptions above (2.19)–(2.22) and for which the main result of the present paper (see Theorem 2.4 below) applies.

    Example 2.1. For simplicity, we consider the homogeneous case and assume that a standard isochoric-volumetric decomposition of elastic energy density by setting

    \begin{equation} \mathcal{W}(\mathbf{F}): = \left\{\begin{array}{ll} \mathcal{W}_{ \rm{iso}}\left(\frac{\mathbf{F}}{(\det\mathbf{F})^{1/3}}\right)+\mathcal{W}_{ \rm{vol}}(\mathbf{F}),\qquad&\mbox{if $\det\mathbf F > 0$}, \\ +\infty,\qquad&\mbox{if $\det\mathbf F\le 0$}, \end{array}\right. \end{equation} (2.26)

    where {\mathcal W}_{ \rm{iso}} is an energy density of Yeoh type which is defined by choosing

    \begin{equation} {\mathcal W}_{ \rm{iso}}(\mathbf{F}): = \sum\limits_{k = 1}^3 c_k(|\mathbf{F}|^2-3)^k \end{equation} (2.27)

    with coefficients c_k > 0 and \mathcal{W}_{ \rm{vol}}(\mathbf{F}) = g(\det\mathbf{F}) for some convex g\in C^2({{\mathbb R}}_{+}) such that

    \begin{equation} \left\{\begin{array}{ll} &g(t)\geq 0\text{ for all }t > 0,\quad g(t) = 0 \text{ if and only if }t = 1 , \\ &g''(1) > 0,\quad \lim\limits_{t\to 0^+}g(t) = +\infty , \\ &\hbox{there exists } \ C' > 0\text{ and } r\ge 2\ \hbox{such that } \ g(t)\geq C' t^r,\quad\text{for }t > 0\text{ sufficiently large}.\\ \end{array}\right. \end{equation} (2.28)

    It is easy to check that with this choice the energy density satisfies all assumptions from (2.19) to (2.21) while inequality (2.22) has been proven in [43].

    It is worth noticing that when material constants are suitably chosen then also Ogden-type energies may fulfil assumptions (2.19)–(2.22) and we refer to [43] for all details.

    We introduce a body force field \mathbf f\in L^{6/5}(\Omega, {{\mathbb R}}^3) and a surface force field \mathbf g\in L^{4/3}(\partial\Omega, {{\mathbb R}}^3) . From now on, \mathbf f and \mathbf g will always be understood to satisfy these summability assumptions. The load functional is the following linear functional

    \begin{equation} \mathcal L({\mathbf{v}}): = \int_ \Omega\mathbf f\cdot{\mathbf{v}}\,d{{\bf x}}+\int_{\partial \Omega}\mathbf g\cdot{\mathbf{v}}\,d\mathcal H^2({{\bf x}}),\qquad {\mathbf{v}}\in H^1( \Omega,\mathbb R^3). \end{equation} (2.29)

    We note that since \Omega is a bounded Lipschitz domain, the Sobolev embedding H^{1}(\Omega, \mathbb R^3)\hookrightarrow L^{6}(\Omega, \mathbb R^3) and the Sobolev trace embedding H^{1}(\Omega, \mathbb R^3)\hookrightarrow L^{4}(\partial\Omega, \mathbb R^3) imply that \mathcal L is a bounded functional over H^{1}(\Omega, \mathbb R^3) and throughout the paper we denote its norm with \|\mathcal L\|_* .

    For every \mathbf R\!\in\! SO(3) we set

    \begin{equation} \mathcal C_{\mathbf R}: = \{\mathbf c: c_3\ge -\min\limits_{{{\bf x}}\in E_{ess}}((\mathbf R{{\bf x}})_3-x_3)\} \end{equation} (2.30)

    and, as we have observed in the Introduction, we must assume the following geometrical compatibility between load and obstacle

    \begin{equation} \mathcal L((\mathbf R-\mathbf I){{\bf x}}+\mathbf c)\le 0,\qquad \forall \mathbf R\in SO(3),\ \forall \mathbf c\,\in \mathcal C_{\mathbf R} \end{equation} (2.31)

    together with

    \begin{equation} \mathcal L\left ((\mathbf R{{\bf x}}-{{\bf x}}\right)_\alpha\mathbf e_\alpha)\le 0,\quad \forall\,\mathbf R\in SO(3), \end{equation} (2.32)

    the summation convention over repeated index \alpha = 1, 2 being understood all along this paper. It can be shown that condition (2.31) is equivalent to (see Remark 3.4 below)

    \begin{equation} \mathcal L(\mathbf e_3)\le 0 = \mathcal L(\mathbf e_1) = \mathcal L(\mathbf e_2), \qquad \Phi(\mathbf R, E, \mathcal L)\le 0, \quad \forall\,\mathbf R\!\in\! SO(3) , \end{equation} (2.33)

    where we have set

    \Phi(\mathbf R, E, \mathcal L): = \mathcal L((\mathbf R-\mathbf I){{\bf x}})\,-\,\mathcal L(\mathbf e_3)\min\limits_{{{\bf x}}\in E_{ess}}\left\{((\mathbf R{{\bf x}})_3-x_3)\right\}

    and from now on we will use (2.33) in place of (2.31). On the other hand Remark 4.5 below will show that also condition (2.32) is in fact unavoidable.

    If E\!\subset\!\overline \Omega\!\subset\! \{\mathbf x\!: x_3\!\ge\! 0\} , the classical Signorini problem in linear elasticity can be described as the minimization of the functional \mathcal E: H^1(\Omega, \mathbb R^3)\to\mathbb R\cup\{+\infty\} defined by

    \begin{equation} {\mathcal E}({\mathbf{u}}): = \left\{\begin{array}{ll} \int_ \Omega \mathcal Q({{\bf x}}, \mathbb E({\mathbf{u}}))\,d{{\bf x}}-\mathcal L({\mathbf{u}}),\quad &\hbox{if} \ {\mathbf{u}}\in \mathcal A, \\ \ \!\!+\infty,\quad &\hbox{otherwise in} \ H^1( \Omega,\mathbb R^3), \end{array}\right. \end{equation} (2.34)

    where \mathbb E({\mathbf{u}}): = \! {\mathop{{{\rm{sym}}}}\nolimits} \nabla{\mathbf{u}} , {\mathcal Q} ({{\bf x}}, \mathbf F)\! = \tfrac12\, \mathbf F^T\mathbb C\, \mathbf F with \mathbb C = D^2\mathcal W({{\bf x}}, \mathbf I) and \mathcal A denotes the set of admissible displacements, defined by

    \begin{equation} \mathcal A\ : = \ \left\{ \,{\mathbf{uu}}\in H^1(\Omega;{{\mathbb R}}^3):\ \, u_3^*({\mathbf{x}})+x_3\ge 0 \ \,\hbox{q.e}\ {\mathbf{x}}\!\in\! E \,\right\} . \end{equation} (2.35)

    The meaning of such constraint is that, if the portion E of the elastic body is contained in \{x_3\!\ge\! 0\} in the reference configuration, then the deformed configuration of E , namely \{\mathbf y({\mathbf{x}})\!: = {\mathbf{x}}+{\mathbf{u}}({\mathbf{x}}), \, {\mathbf{x}}\!\in\! E\} , is constrained to remain in \{y_3\ge 0\} .

    For every \mathbf y\in H^1(\Omega, \mathbb R^3) we introduce the set

    \begin{equation} \mathcal M(\mathbf y)\,: = \, {\mathop{{{\rm{argmin}}}}\nolimits}\left\{ \,\int_ \Omega |\nabla\mathbf y-\mathbf R|^2\,d{{\bf x}}: \ \mathbf R\in SO(3)\right\} . \end{equation} (2.36)

    Thus, due to the rigidity inequality of [24], there exists a constant C = C(\Omega) > 0 such that for every {\mathbf{y}}\in H^1(\Omega, \mathbb R^{3}) and every \mathbf R\in \mathcal M(\mathbf y)

    \begin{equation} \int_{ \Omega}\Big(d\big(\nabla {\mathbf{y}}, SO(3)\big)\Big)^2\,d{{\bf x}} \,\ge\,C\int_{ \Omega}|\nabla{\mathbf{y}}-\mathbf R|^2\,d{{\bf x}}\ , \end{equation} (2.37)

    where d\big(\mathbf F, SO(3)\big): = \min\{|\mathbf F-\mathbf R|:\mathbf R\in SO(3)\} .

    We introduce the set of admissible deformations \mathcal A_h as

    \begin{equation} \ \mathcal A_h:\, = \,\{\,\mathbf y\!\in\! H^{1}(\Omega,{{\mathbb R}}^3)\!: \ y^{*}_{3}({\mathbf{x}})-x_3\ge -hx_3\ \hbox{ q.e.}\ {{\bf x}}\!\in\! E\,\} \end{equation} (2.38)

    and the rescaled finite elasticity functionals \mathcal G_h: H^1(\Omega, {{\mathbb R}}^3)\to {{\mathbb R}}\cup\{+\infty\} by setting

    \begin{equation} \mathcal G_h(\mathbf y) = \left\{\begin{array}{lr} & h^{-2}\int_ \Omega\mathcal W({{\bf x}},\nabla\mathbf y)\,d{{\bf x}} -h^{-1}\mathcal L(\mathbf y-\mathbf x),\ \qquad \hbox{if}\ \mathbf y\in \mathcal A_h,\\ &\\ & +\infty, \ \qquad \hbox{otherwise.}\\ \end{array}\right. \end{equation} (2.39)

    It is readily seen that, for every \mathbf R\in SO(3) and for every \mathbf c\in {{\mathbb R}}^3 such that

    c_3\ge -\min\limits_{E_{ess}}((\mathbf R{\mathbf{x}})_3-x_3)

    the map {\mathbf{y}} ({\mathbf{x}}): = \mathbf R{\mathbf{x}}+\mathbf c belongs to \mathcal A_h for every h > 0 . In the sequel we use the short notations \mathcal G_j: = \mathcal G_{h_j} and \mathcal A_j: = \mathcal A_{h_j} whenever \{h_j\}_{j\in \mathbb N} is a sequence of strictly positive real numbers such that h_j\to 0^+ as j\to +\infty .

    We say that (\mathbf y_j)_{j\in\mathbb N}\!\subset\! H^1(\Omega, \mathbb R^3) is a minimizing sequence of the sequence of functionals \mathcal G_j if

    \begin{equation} \lim\limits_{j\to+\infty}\left(\,\mathcal G_j(\mathbf y_j)\,-\inf\limits_{H^{1}(\Omega,\mathbb R^3)}\mathcal G_j\right) = 0. \end{equation} (2.40)

    The main focus of the paper is to investigate whether minimizers of (2.34) can be approximated by minimizing sequences of the sequence of functionals \mathcal G_j , as defined by (1.5) and (2.40).

    To this end we introduce the functionals \mathcal I, \ \widetilde{\mathcal G}, \ \mathcal G :H^1(\Omega, \mathbb R^3)\to\mathbb R\cup\{+\infty\} defined by

    \begin{equation} \mathcal I({\mathbf{uu}}): = \min\limits_{\mathbf b\in \mathbb R^2}\int_ \Omega \mathcal Q({{\bf x}}, \mathbb E({\mathbf{u}})+\frac{1}{2}b_{\alpha}(\mathbf e_{\alpha}\otimes\mathbf e_3+\mathbf e_{3}\otimes\mathbf e_\alpha))\,d{{\bf x}}, \end{equation} (2.41)
    \begin{equation} {\widetilde{\mathcal G}}({\mathbf{u}}): = \left\{\begin{array}{ll} \mathcal I({\mathbf{uu}})-\max\limits_{\mathbf R\in \mathcal S_{\mathcal L, E}} \mathcal L(\mathbf R{\mathbf{u}}),\quad &\hbox{if} \ {\mathbf{u}}\in \mathcal A, \\ \ \!\!+\infty,\quad &\hbox{otherwise in} \ H^1( \Omega,\mathbb R^3), \end{array}\right. \end{equation} (2.42)

    and

    \begin{equation} {\mathcal G}({\mathbf{u}}): = \left\{\begin{array}{ll} \int_ \Omega \mathcal Q({{\bf x}}, \mathbb E({\mathbf{u}})\,d{{\bf x}}-\max\limits_{\mathbf R\in \mathcal S_{\mathcal L, E}} \mathcal L(\mathbf R{\mathbf{u}}),\quad &\hbox{if} \ {\mathbf{u}}\in \mathcal A, \\ \ \!\!+\infty,\quad &\hbox{otherwise in} \ H^1( \Omega,\mathbb R^3), \end{array}\right. \end{equation} (2.43)

    where

    \begin{equation} \mathcal S_{\mathcal L,E} = \left\{\, \mathbf R\in SO(3): \Phi(\mathbf R, E, \mathcal L) = 0 \right\}. \end{equation} (2.44)

    Remark 2.2. It is worth noticing that \mathcal G\le \mathcal E , since \mathbf I\in\mathcal S_{\mathcal L, E} and it is straightforward checking that \mathcal I, \ \widetilde{\mathcal G}, \ \mathcal G are all continuous with respect to the strong convergence in H^1(\Omega; \mathbb R^3) .

    Before stating the main result in Theorem 2.4, we show the next Remark with some insight on technicalities implied by precise obstacle formulation in the Sobolev space H^1(\Omega) .

    Remark 2.3. If w\in H^1(\Omega) then w^-\!\in\! H^1(\Omega) too. Moreover, both (w^-)^* and (w^*)^- are quasicontinuous in \overline \Omega and (w^-)^* = (w^*)^- = w^- a.e. in \Omega . Then, by [30], (w^-)^* = (w^*)^- q.e. in \overline \Omega . Therefore the condition (w^-)^* = 0 q.e. in E_{ess} is equivalent to w^*\ge 0 q.e. in E_{ess} .

    In particular we claim that

    \begin{equation} (w^-)^* = 0\ \text{ q.e. in}\ \overline\Omega \end{equation} (2.45)

    is equivalent to

    \begin{equation} w\ge 0\ \text{ a.e. in}\ \Omega\ \text{ and }\ w\ge 0\ \ \mathcal H^2\,\text{a.e. on}\ \partial \Omega. \end{equation} (2.46)

    Indeed if w\ge 0 a.e. in \Omega then (w^-)^* = 0 a.e. in \Omega and hence (w^-)^* = 0 q.e. in \Omega .

    If w\ge 0\ \mathcal H^2\, a.e. on \partial \Omega and v is a Sobolev extension of w^- then

    \begin{equation*} \lim\limits_{r \downarrow 0}\frac{1}{|B_r({\mathbf{x}})\!\cap\!\Omega|}\,\int_{ B_r({\mathbf{x}})\cap \Omega}\! v(\mathbf \xi)\,d\mathbf \xi = \lim\limits_{r \downarrow 0}\frac{1}{|B_r({\mathbf{x}})\!\setminus\!\Omega|}\,\int_{ B_r({\mathbf{x}})\setminus \Omega}\! v(\mathbf \xi)\,d\mathbf \xi = 0, \quad\ \mathcal H^2\,\text{a.e.}\,{\mathbf{x}}\!\in\!\partial \Omega. \end{equation*}

    and by taking (2.16) into account we get

    \begin{equation*} \lim\limits_{r \downarrow 0}\frac{2}{|B_r({\mathbf{x}})|}\,\int_{ B_r({\mathbf{x}})\cap \Omega}\! v(\mathbf \xi)\,d\mathbf \xi = \lim\limits_{r \downarrow 0}\frac{2}{|B_r({\mathbf{x}})|}\,\int_{ B_r({\mathbf{x}})\setminus \Omega}\! v(\mathbf \xi)\,d\mathbf \xi = 0, \qquad\quad \mathcal H^2\,\text{a.e.}\,{\mathbf{x}}\!\in\!\partial \Omega. \end{equation*}

    By recalling that \Omega is a Lipschitz set, it is easily checked that \partial \Omega is Ahlfors 2 -regular, that is there are constants c_1, c_2 > 0 such that

    \begin{equation} c_1r^2\le \mathcal H^2( \partial \Omega \cap B_r({\mathbf{x}}))\le c_2 r^2 \end{equation} (2.47)

    for every 0\! < \! r \! < \! \hbox{diam}(\partial\Omega) and for every {\mathbf{x}}\in\partial \Omega . Therefore if we choose R\! > \! 0 such that \overline \Omega\!\subset\! B_R(0) we may apply Proposition 6.1.3. of [1] and Theorem 2.1 of [20] both in H^1(\Omega) and in H^1(B_R(0)\setminus\overline \Omega) and we get

    \lim\limits_{r \downarrow 0}\frac{2}{|B_r({\mathbf{x}})|}\,\int_{ B_r({\mathbf{x}})\cap \Omega}\! v(\mathbf \xi)\,d\mathbf \xi = \lim\limits_{r \downarrow 0}\frac{2}{|B_r({\mathbf{x}})|}\,\int_{ B_r({\mathbf{x}})\setminus \Omega}\! v(\mathbf \xi)\,d\mathbf \xi = 0, \qquad \text{ q.e. }{\mathbf{x}}\in \partial \Omega,

    that is (2.46) implies (2.45).

    Conversely if (2.45) holds then (w^-)^* = 0 a.e. in \Omega and \mathcal H^2 a.e. on \partial \Omega . Therefore w\ge 0 a.e. in \Omega and by recalling again that the negative part of the trace of w and the trace of its negative part coincide \mathcal H^2 a.e. on \partial \Omega we get w\ge 0\, \mathcal H^2 a.e. on \partial \Omega thus proving (2.46) and the claim.

    Similarly, again by Theorem 2.1 of [20], if E_{ess}\subset \partial \Omega is Ahlfors 2 -regular then the condition w\ge 0\ q.e. on E is equivalent to w\ge 0\ \ \mathcal H^2 a.e. on E, so the classical framework of [12,31,54] is equivalent to ours in this case as it was claimed in the Introduction.

    The convergence result is stated in the next theorem, referring to (2.36) and (2.40).

    Theorem 2.4. Assume (2.13), (2.19)–(2.22), (2.32)–(2.33) and \mathcal L(\mathbf e_3) < 0 . Let h_j\to 0^+ as j\to+\infty and let (\overline{\mathbf y}_j)_{j\in\mathbb N}\!\subset\! H^1(\Omega, \mathbb R^3) be a minimizing sequence of \mathcal G_j . If \mathbf R_j\!\in\!\mathcal M(\overline{\mathbf y}_j) for every j\in\mathbb N , then there are \overline{\mathbf c}_j\!\in\! \mathbb R^3 such that the sequence

    \begin{equation} \overline{{\mathbf{u}}}_j({{\bf x}})\,: = \,{h_j}^{-1}\mathbf R_J^T\!\left\{ \big(\overline{\mathbf y}_j-\overline{\mathbf c}_j-\mathbf R_j{{\bf x}}\big)_\alpha\,\mathbf e_\alpha\,+\,(\overline{ y}_{j,3}-x_3)\mathbf e_3\,\right\} \end{equation} (2.48)

    is weakly compact in H^1(\Omega, \mathbb R^3) . Therefore up to subsequences, \overline{{\mathbf{uu}}}_j{\rightharpoonup} \overline{\mathbf{u}} in H^1(\Omega, \mathbb R^3) and also

    \begin{equation} \mathcal G_j(\mathbf y_{j})\to \widetilde{\mathcal G}(\overline{\mathbf{u}}) = \min\limits_{H^1( \Omega,{{\mathbb R}}^3)}\widetilde{\mathcal G} = \min\limits_{H^1( \Omega,{{\mathbb R}}^3)}\mathcal G,\quad\qquad\mathit{\mbox{as}}\ j\to+\infty. \end{equation} (2.49)

    Remark 2.5. Since \widetilde{\mathcal G}\le \mathcal G then equality \min\widetilde{\mathcal G} = \min\mathcal G is equivalent to {\mathop{{{\rm{argmin}}}}\nolimits} \mathcal G\subset {\mathop{{{\rm{argmin}}}}\nolimits} \widetilde{\mathcal G} with possible strict inclusion (see Remark 5.6), thus in general \overline{\mathbf{u}} may not belong to {\mathop{{{\rm{argmin}}}}\nolimits} \mathcal G .

    Remark 2.6. Conditions (2.32) and (2.33) are compatible. Indeed set

    \begin{equation} \begin{array}{ll}& \Omega: = \{{{\bf x}}: x_1^2+x_2^2 < 1,\ 0 < x_3 < 1 \},\\ &\\ &E: = \{{{\bf x}}: x_{1}^{2}+x_{2}^{2} < 1,\ x_3 = 0\}, \\ \end{array} \end{equation} (2.50)

    and \mathbf f\! = {\bf 0}, \ \mathbf g\! = \!-\mathbf e_3 {\mathbf 1}_{E}. It is readily seen that \mathcal L(\mathbf e_3) < 0 = \mathcal L(\mathbf e_1) = \mathcal L(\mathbf e_2) and

    \mathcal L((\mathbf R{{\bf x}}-{{\bf x}})_{\alpha}\mathbf e_{\alpha}) = 0,\ \ \Phi( {\mathbf R}, E, \mathcal L) = -\pi\sqrt{1-R_{33}^2}\le 0, \qquad \forall \, \mathbf R\in SO(3),

    thus both (2.32) and (2.33) are fulfilled.

    On the other hand condition (2.32) does not entail (2.33). Indeed if \Omega and E are as in (2.50), \mathbf f\! = -\mathbf e_3\ \mathbf g\! = 0 then is not since if \overline{\mathbf R} = \mathbf e_1\otimes \mathbf e_1-\mathbf e_2\otimes \mathbf e_2-\mathbf e_3\otimes \mathbf e_3 a direct calculation yields

    \begin{equation} \ \Phi( {\overline{\mathbf R}}, E, \mathcal L) = 2\,\int_\Omega x_3\,d{\mathbf{x}} +| \Omega|\min\limits_{E_{ess}}(-2x_3) = \pi\, > \, 0. \end{equation} (2.51)

    Eventually (2.33) does not imply (2.32), see Remark 4.5.

    Example 2.7. Here we show an example where the all the assumptions in Theorem 2.4 concerning the geometry of the material body \Omega and its portion E sensitive to the constraint and their compatibility with the loads are fulfilled. Set

    \begin{equation} \Omega: = \{{{\bf x}}\!\in\!\mathbb R^3\!: x_1^2+x_2^2\! < \!1,\,0\! < \!x_3\! < \!1\}, \quad E\!: = \! \overline \Omega,\quad \end{equation} (2.52)
    \begin{equation} \mathbf f: = p\,{\mathbf e}_3,\quad \mathbf g\equiv 0 ,\quad p < 0. \end{equation} (2.53)

    Then \mathcal L({\mathbf{u}}) = \int_\Omega \mathbf f\cdot {\mathbf{u}}\, d{{\bf x}} , condition (2.33) is satisfied and \mathcal S_{\mathcal L, E} = \{\mathbf R\in SO(3): \mathbf R\mathbf e_3 = \mathbf e_3\} .

    Indeed it is readily seen that \mathcal L(\mathbf e_3) = p| \Omega| < 0 = \mathcal L(\mathbf e_1) = \mathcal L(\mathbf e_2) ; moreover if \mathbf R\in SO(3) and we denote its entries as R_{ij}\ i, j = 1, 2, 3 then, taking into account p < 0 , we get

    \begin{equation} \begin{array}{ll} \Phi(\mathbf R, E, \mathcal L)& = \pi\,(R_{11}+R_{22}-2) -\,p\,\pi\,\min\limits_{\overline \Omega}\big\{R_{31}x_1+R_{32}x_2+(R_{33}-1)x_3\big\} \\ &\\ & = \frac{\pi p}{2}(R_{33}-1)+p\pi\sqrt{R_{31}^2+R_{32}^2}+p\pi (1-R_{33})^+\\ &\\ & = \frac{\pi p}{2}(1-R_{33})+p\pi\sqrt{1-R_{33}^2}\le 0 \end{array} \end{equation} (2.54)

    and \Phi(\mathbf R, E, \mathcal L) = 0 if and only if R_{33} = 1 that is \mathbf R\mathbf e_3 = \mathbf e_3 as claimed.

    Both conditions (2.13) and (2.32) are trivially fulfilled.

    We emphasize that the above claims still hold true if the assumption on E in (2.52) is weakened by allowing any E\subset\Omega such that E fulfills {\mathop{{{\rm{co}}}}\nolimits} E_{ess} = \overline\Omega.

    This section makes explicit the properties of admissible loads by exploiting the conditions stated by (2.32) and (2.33).

    Lemma 3.1. Assume that (2.32) holds. Then

    \begin{equation} \mathcal L ((\mathbf a\wedge {{\bf x}})_{\alpha}\,\mathbf e_\alpha) = 0\ \ \mathit{\text{and}}\ \ \mathcal L ((\mathbf a\wedge (\mathbf a\wedge {{\bf x}}))_{\alpha}\,\mathbf e_\alpha)\le 0\ \ \ \forall\, \mathbf a\in \mathbb R^3. \end{equation} (3.1)

    Proof. By the Euler-Rodrigues formula (2.32) entails

    \begin{equation} \mathcal L \left(\frac{\sin \vartheta}{\vartheta}\, (\mathbf a \wedge {\mathbf{x}})_{\alpha}\mathbf e_\alpha+ \frac{1-\cos \vartheta}{\vartheta}\,(\mathbf a \wedge (\mathbf a \wedge {\mathbf{x}}))_{\alpha}\mathbf e_\alpha\right)\le 0 \end{equation} (3.2)

    for every \vartheta \in (0, 2\pi) and by letting \vartheta\to 0^+ we get \mathcal L ((\mathbf a\wedge {{\bf x}})_{\alpha}\mathbf e_\alpha)\le 0 for every \mathbf a\in \mathbb R^3 hence \mathcal L ((\mathbf a\wedge {{\bf x}})_{\alpha}\mathbf e_\alpha) = 0 for every \mathbf a\in \mathbb R^3 . The second inequality in (3.1) follows by the previous one.

    Remark 3.2. It is worth noticing that, by inserting \mathbf a = \mathbf e_1 or \mathbf a = \mathbf e_2 , the condition (3.1) entails \mathcal L(x_3\mathbf e_2) = 0 and \mathcal L(x_3\mathbf e_1) = 0 respectively.

    Lemma 3.3. Assume (2.13) and (2.31). Then

    (1) \, \mathcal L(\mathbf e_1) = \mathcal L(\mathbf e_2) = 0 and \mathcal L(\mathbf e_3) \le 0;

    (2) \, \mathcal L(\mathbf e_3\wedge {\mathbf{x}}) = 0;

    (3) \, \mathcal L\big(\mathbf e_3\wedge (\mathbf e_3\wedge{\mathbf{x}})\big) \le 0;

    (4) there exists {\mathbf{x}}_{\mathcal L} \in {\mathop{{{\rm{ri}}}}\nolimits} {\mathop{{{\rm{co}}}}\nolimits} E_{ess} such that \ \mathcal L\big(\, \mathbf a \wedge ({\mathbf{x}}-{\mathbf{x}}_{\mathcal L})\big) = 0 \ \forall\mathbf a\in{{\mathbb R}}^3 .

    Proof. By choosing \mathbf R = \mathbf I in (2.31) we get \mathcal L(\mathbf c)\le 0 for every \mathbf c\in \mathcal C_{\mathbf I} = \{\mathbf c\!\in\! {{\mathbb R}}^3\!: c_3\ge 0\} . Since c_1\, \mathbf e_1+c_2\, \mathbf e_2\in C_{\mathbf I} for every c_1, c_2\in {{\mathbb R}} , we get \mathcal L(\mathbf e_1) = \mathcal L(\mathbf e_2) = 0 . Moreover c_3\, \mathbf e_3\in C_{\mathbf I} for c_3\ge 0 entails \mathcal L(\mathbf e_3) \le 0 . Thus (1) is proved and (2.31) entails

    \begin{equation} \Phi(\mathbf R, E, \mathcal L): = \mathcal L((\mathbf R-\mathbf I){{\bf x}})\,-\,\mathcal L(\mathbf e_3)\min\limits_{{{\bf x}}\in E_{ess}}\left\{((\mathbf R{{\bf x}})_3-x_3)\right\}\le 0, \qquad \forall\,\mathbf R\!\in\! SO(3), \end{equation} (3.3)

    that is by the Euler-Rodrigues formula

    \begin{eqnarray} && \varphi_{\mathbf a}(\vartheta): = \mathcal L \big(\sin \vartheta\, (\mathbf a \wedge {\mathbf{x}})+ (1-\cos \vartheta )\,\mathbf a \wedge (\mathbf a \wedge {\mathbf{x}})\big) \\ && \qquad\qquad - \min\limits_{{\mathbf{x}}\in E_{ess}} \Big( \sin \vartheta\, (\mathbf a \wedge {\mathbf{x}})_3+ (1-\cos \vartheta )\,\mathbf a \wedge (\mathbf a \wedge {\mathbf{x}})_3\Big)\mathcal L(\mathbf e_3)\ \le \ 0, \\ && \qquad\qquad\qquad\qquad\qquad \forall\mathbf a\in{{\mathbb R}}^3, |\mathbf a| = 1, \ \forall\,\vartheta \in [0,2\pi]. \end{eqnarray} (3.4)

    If \mathbf a\! = \!\mathbf e_3 then \mathbf R\mathbf e_3\! = \!\mathbf e_3 and (3.4) reads

    \begin{equation} \varphi(\vartheta)\,: = \,\mathcal L \big(\sin \vartheta\, (\mathbf e_3 \wedge {\mathbf{x}})+ (1-\cos \vartheta )\,\mathbf e_3 \wedge (\mathbf e_3 \wedge {\mathbf{x}})\big) \le 0, \qquad \forall\,\vartheta\in [0,2\pi] . \end{equation} (3.5)

    By \varphi(0) = \varphi(2\pi) = 0 and \varphi(\vartheta)\le 0 we get 0\ge \varphi'(0) = \mathcal L(\mathbf e_3\wedge {\mathbf{x}}) = \varphi'(2\pi)\ge 0 . Thus (2) is proved. By (2) and (3.5) we get \mathcal L\big(\mathbf e_3 \wedge (\mathbf e_3 \wedge {\mathbf{x}})\big)\le 0 , say (3). In order to show (4), first we notice that \mathcal L(\mathbf e_1) = \mathcal L(\mathbf e_2) = 0 entail for every {\boldsymbol{\xi}}\in{{\mathbb R}}^3

    \begin{equation} \begin{array}{ll} \mathcal L(\mathbf a\wedge {\boldsymbol{\xi}})& = \mathcal L\big(\sum\limits_{j = 1}^3 a_j\mathbf e_j\wedge {\boldsymbol{\xi}}\big)\, = \, \sum\limits_{j = 1}^3 a_j \,\mathcal L (\mathbf e_j\wedge {\boldsymbol{\xi}})\, \\ &\\ & = a_1\,\mathcal L (\,-\xi_3 \mathbf e_2 + \xi_2 \mathbf e_3\,) + a_2 \,\mathcal L (\,\xi_3 \mathbf e_1 - \xi_1 \mathbf e_3\,) + a_3\, \mathcal L (\,-\xi_2 \mathbf e_1 + \xi_1 \mathbf e_2\,)\, \\ & \\ & = \mathbf a \cdot (\,\xi_2\mathbf e_1-\xi_1\mathbf e_2\,) \mathcal L(\mathbf e_3) = (\mathbf a\wedge{\boldsymbol{\xi}})_3 \mathcal L(\mathbf e_3) , \end{array} \end{equation} (3.6)

    moreover, \mathcal L(\mathbf e_3\wedge {\mathbf{x}}) = 0 entails

    \begin{equation} \mathcal L(\mathbf a\wedge {\mathbf{x}}) = a_1\,\mathcal L(\mathbf e _1\wedge {\mathbf{x}}) + a_2\,\mathcal L(\mathbf e _2\wedge {\mathbf{x}}), \qquad \forall\, \mathbf a\in{{\mathbb R}}^3\,. \end{equation} (3.7)

    Let us assume first that \mathcal L(\mathbf e_3) < 0 . In this case we can set

    \begin{equation} \widetilde x_1\ = \ -\,\frac{\mathcal L(\mathbf e_2\wedge {\mathbf{x}})}{\mathcal L(\mathbf e_3)}, \qquad \widetilde x_2\ = \ \frac{\mathcal L(\mathbf e_1\wedge {\mathbf{x}})}{\mathcal L(\mathbf e_3)} , \end{equation} (3.8)

    hence, by (3.6)–(3.8),

    \begin{equation*} \mathcal L(\mathbf a \wedge \widetilde{{\mathbf{x}}}) = (\mathbf a\wedge \widetilde {\mathbf{x}})_3\, \mathcal L(\mathbf e_3) = (a_1\widetilde x_2-a_2\widetilde x_1)\, \mathcal L(\mathbf e_3) = a_1\,\mathcal L(\mathbf e _1\wedge {\mathbf{x}}) + a_2\,\mathcal L(\mathbf e _2\wedge {\mathbf{x}}) = \mathcal L(\mathbf a \wedge {\mathbf{x}}) \end{equation*}

    say

    \begin{equation} \mathcal L(\mathbf a \wedge {\mathbf{x}}) = \mathcal L(\mathbf a \wedge \widetilde {\mathbf{x}}), \qquad \forall\,\mathbf a\in{{\mathbb R}}^3, \ \forall\,\widetilde {\mathbf{x}}\in \{(\widetilde x_1,\widetilde x_2,z): z\in \mathbb R\}. \end{equation} (3.9)

    Since \varphi_{\mathbf a}(0) = \varphi_{\mathbf a}(2\pi) = 0 and \varphi_{\mathbf a}(\vartheta)\le 0 for every \mathbf a\in{{\mathbb R}}^3, |\mathbf a| = 1 and for every \vartheta \in [0, 2\pi], (3.4) entails

    \begin{equation} 0\,\ge\,\limsup\limits_{\vartheta\to 0_+} \frac {\varphi_{\mathbf a}(\vartheta)}{\vartheta} = \mathcal L(\mathbf a \wedge {\mathbf{x}}) - \min\limits_{{\mathbf{x}}\in E_{ess}} (\mathbf a \wedge {\mathbf{x}})_3\,\mathcal L(\mathbf e_3),\qquad \forall\, |\mathbf a| = 1, \end{equation} (3.10)
    \begin{equation} 0\,\le\,\liminf\limits_{\vartheta\to 2\pi^-} \frac {\varphi_{\mathbf a}(\vartheta)}{\vartheta-2\pi} = \mathcal L(\mathbf a \wedge {\mathbf{x}}) - \max\limits_{{\mathbf{x}}\in E_{ess}} (\mathbf a \wedge {\mathbf{x}})_3\,\mathcal L(\mathbf e_3),\qquad\forall \, |\mathbf a| = 1. \end{equation} (3.11)

    Hence

    \begin{equation*} \max\limits_{{\mathbf{x}}\in E_{ess}}(\mathbf a \wedge {\mathbf{x}})_3\ \mathcal L(\mathbf e_3) \,\le\, \mathcal L(\mathbf a \wedge {\mathbf{x}}) \,\le\, \min\limits_{{\mathbf{x}}\in E_{ess}}(\mathbf a \wedge {\mathbf{x}})_3\ \mathcal L(\mathbf e_3), \end{equation*}

    so, by (3.6), (3.8) and (3.9),

    \begin{equation} \max\limits_{{\mathbf{x}}\in E_{ess}}(\mathbf a \wedge {\mathbf{x}})_3\ \mathcal L(\mathbf e_3) \,\le\, (\mathbf a\wedge \widetilde{\mathbf{x}})_3\,\mathcal L(\mathbf e_3) \,\le\, \min\limits_{{\mathbf{x}}\in E_{ess}}(\mathbf a \wedge {\mathbf{x}})_3\ \mathcal L(\mathbf e_3) . \end{equation} (3.12)

    By taking account of \mathcal L(\mathbf e_3) < 0 , we find

    \begin{equation*} \min\limits_{{\mathbf{x}}\in E_{ess}}(\mathbf a \wedge {\mathbf{x}})_3 \,\le\, (\mathbf a \wedge \widetilde{\mathbf{x}})_3 \,\le\, \max\limits_{{\mathbf{x}}\in E_{ess}}(\mathbf a \wedge {\mathbf{x}})_3 ,\qquad \forall\,\mathbf a\in {{\mathbb R}}^3: \ |\mathbf a| = 1, \end{equation*}

    hence, by linearity and by homogeneity,

    \begin{equation} \min\limits_{{\mathbf{x}}\in {\mathop{{{\rm{co}}}}\nolimits} E_{ess}}(\mathbf a \wedge {\mathbf{x}})_3 \,\le\, (\mathbf a \wedge \widetilde{\mathbf{x}})_3 \,\le\, \max\limits_{{\mathbf{x}}\in {\mathop{{{\rm{co}}}}\nolimits} E_{ess}}(\mathbf a \wedge {\mathbf{x}})_3, \qquad \forall\,\mathbf a\in {{\mathbb R}}^3 . \end{equation} (3.13)

    By subtracting (\mathbf a \wedge \widetilde{\mathbf{x}})_3 on each term of inequality (3.13) we get

    \begin{equation*} \min\limits_{{\mathbf{y}}\in\, {\mathop{{{\rm{co}}}}\nolimits} E_{ess}-\widetilde {\mathbf{x}}}\, (\mathbf a \wedge {\mathbf{y}})_3 \,\le\,0\,\le\, \max\limits_{{\mathbf{y}}\in\, {\mathop{{{\rm{co}}}}\nolimits} E_{ess}-\widetilde {\mathbf{x}}}\, (\mathbf a \wedge {\mathbf{y}})_3 \end{equation*}

    for every \mathbf a\in{{\mathbb R}}^3 and for every \widetilde {\mathbf{x}}\in \{(\widetilde x_1, \widetilde x_2, z): z\in \mathbb R\} .

    We claim that at least one of the above inequalities is strict for every \mathbf a\in \mathbb R^3 such that a_1^{\, 2}+a_2^{\, 2}\neq 0 . Indeed, if by contradiction there was \overline {\mathbf a} = (\overline a_1, \overline a_2, \overline a_3) with \overline a_1^{\, 2}+\overline a_2^{\, 2}\neq 0 such that

    \begin{equation*} \min\limits_{{\mathbf{x}}\in\, {\mathop{{{\rm{co}}}}\nolimits} E_{ess}-\widetilde {\mathbf{x}}}\, ( \overline {\mathbf a} \wedge {\mathbf{x}})_3 \, = \, \max\limits_{{\mathbf{x}}\in\, {\mathop{{{\rm{co}}}}\nolimits} E_{ess}-\widetilde {\mathbf{x}}}\, (\overline{\mathbf a} \wedge {\mathbf{x}})_3 = 0, \qquad \forall\, \widetilde{\mathbf{x}}\in \{(\widetilde x_1,\widetilde x_2,z),\, z\in{{\mathbb R}}\} , \end{equation*}

    then

    \begin{equation*} \big( {\mathop{{{\rm{co}}}}\nolimits} E_{ess}-\widetilde {\mathbf{x}} \,\big) \ \subset \ \big\{\, {\mathbf{x}}: \ \overline a_1x_2-\overline a_2x_1 = 0\,\big\} . \end{equation*}

    Since the plane \{\overline a_1x_2-\overline a_2x_1 = 0\} is orthogonal to \{x_3 = 0\} we obtain

    {\mathop{{{\rm{cap}}}}\nolimits} \big({\mathop{{{\rm{proj}}}}\nolimits}\{\,{\mathbf{x}}\in{{\mathbb R}}^3: \overline a_1x_2-\overline a_2x_1 = 0\,\}\big) = 0,

    hence

    0 = {\mathop{{{\rm{cap}}}}\nolimits}\Big({\mathop{{{\rm{proj}}}}\nolimits}\big( {\mathop{{{\rm{co}}}}\nolimits} E_{ess}-\widetilde {\mathbf{x}} \,\big)\Big) = {\mathop{{{\rm{cap}}}}\nolimits}\Big({\mathop{{{\rm{proj}}}}\nolimits} ({\mathop{{{\rm{co}}}}\nolimits} E_{ess})\Big)

    which contradicts (2.13).

    Without loss of generality we can proceed by assuming that the first inequality is strict, say

    \begin{equation*} \min\limits_{{\mathbf{x}}\in\, {\mathop{{{\rm{co}}}}\nolimits} E_{ess}-\widetilde {\mathbf{x}}}\, ( {\mathbf a} \wedge {\mathbf{x}})_3 \, < \, 0 \end{equation*}

    for every \mathbf a\in \mathbb R^3 such that a_1^{\, 2}+a_2^{\, 2}\neq 0 and for every \widetilde{\mathbf{x}}\in \{(\widetilde x_1, \widetilde x_2, z), \, z\in{{\mathbb R}}\} . Hence, by setting \, T: = {\mathop{{{\rm{proj}}}}\nolimits}({\mathop{{{\rm{co}}}}\nolimits} E_{ess}-\widetilde {\mathbf{x}}) , we get

    \begin{equation} \min\limits_{{\mathbf{x}}\in T}\, ({\mathbf a} \wedge {\mathbf{x}})_3 \, < \, 0 \end{equation} (3.14)

    for every \mathbf a\in \mathbb R^3 such that a_1^{\, 2}+a_2^{\, 2}\neq 0 . For every (a_1, a_2)\in \mathbb R^2 such that a_1^{\, 2}+a_2^{\, 2}\neq 0 we set now

    \begin{equation*} \Gamma(a_1,a_2)\,: = \, \big\{ \,(x_1,x_2)\in{{\mathbb R}}^2: \, a_1x_2-a_2x_1\ge \min\limits_{(x_1,x_2)\in T} (a_1x_2-a_2x_1)\,\big\}\, \end{equation*}

    then \big\{\, \Gamma(a_1, a_2):\ a_1^{\, 2}+a_2^{\, 2} = 1 \, \big\} is the set of half-planes supporting T . Since T is closed and convex, we get

    \begin{equation*} T\ \, = \, \bigcap\limits_{a_1^{\,2}+a_2^{\,2} = 1} \Gamma(a_1,a_2) . \end{equation*}

    By (3.14), we get

    {\mathop{{{\rm{dist}}}}\nolimits} \big(\,(0,0)\,,\,\partial \Gamma(a_1,a_2)\,\big) = \left|\,\min\limits_{{\mathbf{x}}\in T} \, (a_1x_2-a_2x_1) \,\right| > 0.

    Hence we deduce the existence of (\widetilde a_1, \widetilde a_2) :\ \widetilde a_1^{\, 2}+\widetilde a_2^{\, 2} = 1 such that

    \begin{equation*} \min\limits_{\widetilde a_1^{\,2}+\widetilde a_2^{\,2} = 1} {\mathop{{{\rm{dist}}}}\nolimits}\big(\,(0,0)\,,\,\partial \Gamma(a_1,a_2)) = |\min\limits_{{\mathbf{x}}\in T}\big(\widetilde a_1x_2-\widetilde a_2x_1\big)| > 0 , \end{equation*}

    so (0, 0)\, \in \, {\mathop{{{\rm{ri}}}}\nolimits} T that is (\widetilde x_1, \widetilde x_2, 0)\in {\mathop{{{\rm{ri}}}}\nolimits}{\mathop{{{\rm{proj}}}}\nolimits}({\mathop{{{\rm{co}}}}\nolimits} E_{ess}) .

    We are left to show that there exists \widetilde x_3 such that {\mathbf{x}}_{\mathcal L}: = (\widetilde x_1, \widetilde x_2, \widetilde x_3) \!\in\! {\mathop{{{\rm{ri}}}}\nolimits}{\mathop{{{\rm{co}}}}\nolimits} E_{ess} . To this aim it is readily seen that by taking account of {\mathop{{{\rm{cap}}}}\nolimits}({\mathop{{{\rm{proj}}}}\nolimits}({\mathop{{{\rm{co}}}}\nolimits} E_{ess})) > 0 , we get {\mathop{{{\rm{aff}}}}\nolimits}({\mathop{{{\rm{proj}}}}\nolimits}({\mathop{{{\rm{co}}}}\nolimits} E_{ess}))) = \{x_3 = 0\} so there exists \overline r > 0 such that

    \left\{(x_1,x_2): |x_1-\widetilde x_1|^2+|x_2-\widetilde x_2|^2 < \overline r^2\right\}\subset {\mathop{{{\rm{proj}}}}\nolimits}({\mathop{{{\rm{co}}}}\nolimits} E_{ess})).

    Let now

    J: = \{z: (x_1,x_2,z)\in {\mathop{{{\rm{co}}}}\nolimits} E_{ess}\}\neq \emptyset

    and assume that (x_1, x_2, z)\in \, ({\mathop{{{\rm{co}}}}\nolimits} E_{ess})\!\setminus ({\mathop{{{\rm{ri}}}}\nolimits}{\mathop{{{\rm{co}}}}\nolimits} E_{ess}) for every z\in J . Then

    B_r(x_1,x_2,z)\cap {\mathop{{{\rm{co}}}}\nolimits} E_{ess}\neq \emptyset,\qquad B_r(x_1,x_2,z)\cap \big(\,({\mathop{{{\rm{aff}}}}\nolimits}{\mathop{{{\rm{co}}}}\nolimits} E_{ess})\!\setminus\!({\mathop{{{\rm{co}}}}\nolimits} E_{ess})\,)\neq \emptyset

    for every z\in J and for every r > 0 , therefore by recalling that {\mathop{{{\rm{aff}}}}\nolimits}{\mathop{{{\rm{proj}}}}\nolimits}{\mathop{{{\rm{co}}}}\nolimits} E_{ess}\, = \, \{x_3 = 0\}

    {\mathop{{{\rm{proj}}}}\nolimits} B_r(x_1,x_2,z)\cap {\mathop{{{\rm{proj}}}}\nolimits}{\mathop{{{\rm{co}}}}\nolimits} E_{ess}\neq \emptyset,
    {\mathop{{{\rm{proj}}}}\nolimits} B_r(x_1,x_2,z)\cap (\{x_3 = 0\}\setminus{\mathop{{{\rm{proj}}}}\nolimits}{\mathop{{{\rm{co}}}}\nolimits} E_{ess})\neq \emptyset

    for every r > 0 . This is a contradiction since

    \begin{equation*} {\mathop{{{\rm{proj}}}}\nolimits} B_r(x_1,x_2,z)\subset \left\{(x_1,x_2): |x_1-\widetilde x_1|^2+|x_2-\widetilde x_2|^2 < \overline r^2\right\}\subset {\mathop{{{\rm{proj}}}}\nolimits}({\mathop{{{\rm{co}}}}\nolimits} E_{ess}) \end{equation*}

    for some r > 0 thus (4) is proved in this case since by construction \mathcal L(\mathbf a\wedge ({\mathbf{x}}-{\mathbf{x}}_{\mathcal L}) = 0 for every \mathbf a \in \mathbb R^3 . Eventually, if \mathcal L(\mathbf e_3) = 0 , (2.33) reduces to

    \begin{equation*} \mathcal L((\mathbf R-\mathbf I){{\bf x}})\le 0,\qquad \forall\, \mathbf R\!\in\! SO(3) , \end{equation*}

    say \sin\vartheta\, \mathcal L(\mathbf a \wedge {\mathbf{x}})+(1-\cos\vartheta) \mathcal L\big(\mathbf a\wedge(\mathbf a \wedge {\mathbf{x}})\big)\le 0 for all \mathbf a\in {{\mathbb R}}^3 , thus, by repeating the analysis made on (3.5), we get

    \begin{equation*} \mathcal L (\mathbf a\!\wedge\!{\mathbf{x}}) = 0 ,\qquad \forall\, \mathbf a \in {{\mathbb R}}^3, \end{equation*}

    and, since {\mathop{{{\rm{ri}}}}\nolimits} {\mathop{{{\rm{proj}}}}\nolimits} {\mathop{{{\rm{co}}}}\nolimits} E_{ess} \neq\emptyset due to (2.13), by exploiting identity (3.6) with \boldsymbol \xi = \widetilde {\mathbf{x}} we obtain, for whatever choice of \widetilde{\mathbf{x}}\in{\mathop{{{\rm{ri}}}}\nolimits} {\mathop{{{\rm{proj}}}}\nolimits} {\mathop{{{\rm{co}}}}\nolimits} E_{ess}

    \begin{equation*} \mathcal L (\mathbf a\wedge\big({\mathbf{x}}-\widetilde{\mathbf{x}})\big) = - \mathcal L (\mathbf a\wedge\widetilde{\mathbf{x}}) = - (\mathbf a\wedge\widetilde{\mathbf{x}})_3\,\mathcal L(\mathbf e_3) = 0, \qquad \forall\, \mathbf a \in {{\mathbb R}}^3, \end{equation*}

    that is (4) is proven also in this case.

    Remark 3.4. Conditions (2.31) and (2.33) are equivalent as claimed in Subsection 2.2.

    Indeed, as it has been pointed out in the proof of Lemma 3.3, condition (2.31) implies that \mathcal L(\mathbf e_3)\le 0, \ \mathcal L(\mathbf e_1) = \mathcal L(\mathbf e_2) = 0 and

    \begin{equation} \Phi(\mathbf R, E, \mathcal L): = \mathcal L((\mathbf R-\mathbf I){{\bf x}})\,-\,\mathcal L(\mathbf e_3)\min\limits_{{{\bf x}}\in E_{ess}}((\mathbf R{{\bf x}})_3-x_3)\le 0, \qquad \forall\,\mathbf R\!\in\! SO(3). \end{equation} (3.15)

    Conversely if the latter condition holds and \mathcal L(\mathbf e_3)\le 0, \ \ \mathcal L(\mathbf e_1) = \mathcal L(\mathbf e_2) = 0 , then

    \mathcal L((\mathbf R-\mathbf I){{\bf x}}+\mathbf c) = \mathcal L((\mathbf R-\mathbf I){{\bf x}}+c_3\mathbf e_3)\le 0

    for every \mathbf c\in {{\mathbb R}}^3 such that c_3\ge -\min_{{{\bf x}}\in E_{ess}}((\mathbf R{{\bf x}})_3-x_3)\} .

    Remark 3.5. We emphasize that conditions (1) and (4) in Lemma 3.3 together with (2.13) and \mathcal L(\mathbf e_3) < 0 coincide with conditions (4.9)–(4.11) of Theorem 4.5 of [12], which provides the solution to Signorini problem in linear elasticity.

    The whole set of conditions (1)–(4) appearing in the claim of Lemma 3.3 together with condition (2.13) on the set E is not equivalent to admissibility of the loads as expressed by (2.33): This phenomenon is made explicit by subsequent Example 3.6.

    Example 3.6. Let \Omega = \{{\mathbf{x}}:\, x_1^2+x_2^2 < 1, \, 0\! < \!x_3\! < \!H\} , E\! = \!E_{ess}\! = \!\{\, (x_1, x_2, 0)\!:\, x_1^2+x_2^2\!\le\! 1\} and \mathcal L({\mathbf{v}}) = \!\int_\Omega p\, v_3\, d{\mathbf{x}} with p < 0 , say \mathbf f = p\, \mathbf e_3, \ \mathbf g = \mathbf 0 .

    Then E fulfills (2.13), since {\mathop{{{\rm{cap}}}}\nolimits} E > 0 and {\mathop{{{\rm{proj}}}}\nolimits} (\, {\mathop{{{\rm{co}}}}\nolimits} E_{ess}) = E_{ess}\, \subset\, \overline{\Omega}\, \cap \{x_3 = 0\} ; moreover all claims (1)–(4) of Lemma 3.3 hold true: Indeed

    \mathcal L(\mathbf e_1) = \mathcal L(\mathbf e_2) = 0, \quad \mathcal L(\mathbf e_3) = p|\Omega| < 0,
    \mathcal L(\mathbf e_3\wedge {\mathbf{x}}) = \int_\Omega(-\mathbf e_3)\cdot(\mathbf e_3\wedge {\mathbf{x}})d{\mathbf{x}} = 0,
    \mathcal L(\mathbf e_3\wedge (\mathbf e_3\wedge {\mathbf{x}}) = \int_\Omega(-\mathbf e_3)\cdot\big(\mathbf e_3\wedge (\mathbf e_3\wedge {\mathbf{x}})\big)d{\mathbf{x}} = 0,

    eventually, by choosing {\mathbf{x}}_{\mathcal L} = (0, 0, 0)\in {\mathop{{{\rm{ri}}}}\nolimits}\ ({\mathop{{{\rm{co}}}}\nolimits} E_{ess}) = E and by taking the symmetry of \Omega into account, we get

    \begin{eqnarray*} \mathcal L\big(\mathbf a\wedge ({\mathbf{x}}- {\mathbf{x}}_{\mathcal L})\big) \!\!& = &\!\! \mathcal L(\mathbf a\wedge {\mathbf{x}}) \ = \ \int_\Omega \!(-\mathbf e_3)\cdot(\mathbf a\wedge {\mathbf{x}})\,d{\mathbf{x}} \ = \ \mathbf a\cdot\! \int_\Omega (-x_2\mathbf e_1+x_1\mathbf e_2)\,d{\mathbf{x}} \, = \, 0. \end{eqnarray*}

    Nevertheless, condition (2.33) is violated, since we can consider the \pi radians rotation around axis \mathbf e_1 which keeps E above the horizontal plane (obstacle boundary) but capsizes the body below the horizontal plane, namely \widetilde{\mathbf R}\in SO(3) defined by \widetilde{\mathbf R}{\mathbf{x}} = x_1\mathbf e_1-x_2\mathbf e_2-x_3\mathbf e_3 . Thus

    \begin{equation*} \mathcal L\big((\widetilde{\mathbf R}-\mathbf I){\mathbf{x}}\big)\, -\,\mathcal L(\mathbf e_3)\min\limits_{E_{ess}}\big((\widetilde{\mathbf R}{\mathbf{x}})_3-x_3\big) = \, -2p\,\int_\Omega x_3\,d{\mathbf{x}} -p| \Omega|\min\limits_{E_{ess}} (-2x_3) = -p\pi H^2\, > \, 0 . \end{equation*}

    The assumptions in Lemma 3.3 and in Theorem 2.4 cannot be weakened by assuming only {\mathop{{{\rm{cap}}}}\nolimits} E > 0 in place of (2.13) as it is shown in the next example, thus proving that Theorem 2.4 is a sharp result with respect to the sets E subject to the constraint that are admissible.

    Example 3.7. Choose \mathbf f = -\mathbf e_3, \ \mathbf g = \mathbf 0 and

    \Omega = \{{\mathbf{x}}: x_1^2+x_2^2 < 1,\ x_2 > 0,\ 0 < x_3 < 1\},
    E\, = \,E_{ess}\, = \,\{(x_1,x_2, x_3)\in \overline \Omega : \ x_2 = 0\} .

    It is readily seen that condition (2.32) is fulfilled since \mathcal L((\mathbf R{{\bf x}}-{{\bf x}})_{\alpha}\mathbf e_\alpha) = 0 moreover, since \mathcal L(\mathbf e_3) < 0 = \mathcal L(\mathbf e_1) = \mathcal L(\mathbf e_2),

    \begin{equation*} \begin{array}{ll} \Phi(\mathbf R,E,\mathcal L)& = -\frac{2}{3}R_{32}+\frac{\pi}{2}(1-R_{33}) +\,\min\limits_{E_{ess}}\left\{R_{31}x_1+R_{32}x_2+(R_{33}-1)x_3\right\}| \Omega|\\ &\\ & = -\pi |R_{31}|-\pi (R_{32})^{-}-\frac{2}{3}R_{32}+ \frac{\pi}{2}\,(R_{33}-1)\\ &\\ & = -\pi |R_{31}|-\frac{\pi}{2} |R_{32}|+\left (\frac{\pi}{2}-\frac{2}{3}\right)R_{32}+ \frac{\pi}{2}\,(R_{33}-1)\le 0 \end{array} \end{equation*}

    for every \mathbf R\in SO(3), then also condition (2.33) is satisfied. Nevertheless it can be easily checked that {\mathop{{{\rm{cap}}}}\nolimits} E > 0 but {\mathop{{{\rm{cap}}}}\nolimits} \big({\mathop{{{\rm{proj}}}}\nolimits} (\, {\mathop{{{\rm{co}}}}\nolimits} E_{ess}\, )\big) = 0 , thus E does not fulfil (2.13).

    If there was \overline{\mathbf{x}}\in {\mathop{{{\rm{ri}}}}\nolimits}{\mathop{{{\rm{co}}}}\nolimits} E_{ess} such that \mathcal L(\mathbf a\wedge({\mathbf{x}}-\overline {\mathbf{x}})) = 0 for every \mathbf a\in \mathbb R^3 then we get

    \begin{equation*} \mathcal L(\mathbf a\wedge\overline{\mathbf{x}}) = \mathcal L(\mathbf a\wedge{\mathbf{x}}) = -\int_\Omega \mathbf e_3\cdot(\mathbf a\wedge{\mathbf{x}})\,d{\mathbf{x}} = -\int_{ \Omega}(a_1x_2-a_2x_1)\,d{\mathbf{x}} = -\frac {2}{3} a_1 ; \end{equation*}

    then, since \mathcal L(\mathbf e_1) = \mathcal L(\mathbf e_2) = 0 , we could apply (3.6) and find

    -\frac {2}{3}a_1\, = \,\mathcal L(\mathbf a\wedge\overline{\mathbf{x}})\, = \,(\mathbf a\wedge\overline{\mathbf{x}})_3\,\mathcal L(\mathbf e_3) = -\pi(a_1\overline x_2-a_2\overline x_1), \qquad \forall\,\mathbf a\in \mathbb R^3 ,

    hence \overline x_1 = 0, \ \overline x_2 = \frac 2 {3\pi} , thus

    \overline{\mathbf{x}}\not\in {\mathop{{{\rm{ri}}}}\nolimits}{\mathop{{{\rm{co}}}}\nolimits} E_{ess} = \{(x_1,x_2, x_3)\!\in\! \overline \Omega : \,x_2 = 0\},

    a contradiction. Thus claim (4) of Lemma 3.3 cannot hold true in this case.

    Moreover if we set \mathbf w_k({{\bf x}}): = -k\, (\, x_2\, \mathbf e_3-x_3\, \mathbf e_2\, ), then \mathbb E(\mathbf w_k) = {\bf 0}, \ w_{k, 3}+x_3\equiv x_3\ge 0 on E whence \mathbf w_k\in \mathcal A and it is readily seen that

    \begin{equation} \mathcal L(\mathbf R{\mathbf{v}}) = \mathcal L({\mathbf{v}}),\qquad \forall\ \mathbf R\in \mathcal S_{\mathcal L,E}, \end{equation} (3.16)

    hence

    \mathcal G(\mathbf w_k) = -\mathcal L({\mathbf{w}}_k) = \int _ \Omega\mathbf e_3\cdot {\mathbf{w}}_k\,d{\mathbf{x}} = -k\!\int_ \Omega x_2\,d{{\bf x}} = -\frac{2}{3}k\to -\infty, \quad \hbox{as }k\!\to\!+\infty

    that is \inf_{\mathcal A}\mathcal G = -\infty so the convergence of the energies claimed in Theorem 2.4 fails to be true in this case thus showing sharpness of condition (2.13).

    Lemma 3.8. Assume that (2.13), (2.33) hold and that \mathcal L(\mathbf e_3) < 0 . Then

    \begin{equation} \hbox{either}\quad \mathcal S_{\mathcal L,E} = \{\,\mathbf I\,\} \quad \hbox{or} \quad \mathcal S_{\mathcal L,E} = \{\,\mathbf R\in SO(3): \mathbf R\mathbf e_3 = \mathbf e_3\,\}. \end{equation} (3.17)

    Proof. First, we prove the inclusion

    \begin{equation*} \mathcal S_{\mathcal L,E}\,\subset\, \{\,\mathbf R\in SO(3): \mathbf R\mathbf e_3 = \mathbf e_3\,\}. \end{equation*}

    Indeed if \mathbf R belongs to \mathcal S_{\mathcal L, E} and \mathbf a is a rotation axis of \mathbf R with |\mathbf a| = 1 , then

    \begin{equation} \begin{array}{ll} \varphi_{\mathbf a}(\vartheta)&: = \Phi(\mathbf R,E,\mathcal L) = \mathcal L\left(\sin \vartheta(\mathbf a\wedge {\mathbf{x}}) + (1-\cos \vartheta) \big(\mathbf a \wedge (\mathbf a\wedge {\mathbf{x}})\big) \right) \\ & -\min\limits_{E_{ess}} \left\{ \sin \vartheta(\mathbf a\wedge {\mathbf{x}})_3 + (1-\cos \vartheta) \big(\mathbf a \wedge (\mathbf a\wedge {\mathbf{x}})\big)_3 \right\}\mathcal L(\mathbf e_3) \\ & = 0 \end{array} \end{equation} (3.18)

    for every \vartheta\in[0, 2\pi] . By arguing now as in the proof of (4) of Lemma 3.3, we get

    \begin{equation} \begin{array}{lll} 0 &\ge& \lim\limits_{\vartheta\to 0^+} \frac {\varphi_{\mathbf a}(\vartheta)}{\vartheta} = \mathcal L(\mathbf a \wedge {\mathbf{x}}) - \min\limits_{{\mathbf{x}}\in E_{ess}} (\mathbf a \wedge {\mathbf{x}})_3\,\mathcal L(\mathbf e_3)\ \\ & \ge& \lim\limits_{\vartheta\to 2\pi^-} \frac {\varphi_{\mathbf a}(\vartheta)}{\vartheta- 2\pi} = \mathcal L(\mathbf a \wedge {\mathbf{x}}) - \max\limits_{{\mathbf{x}}\in E_{ess}} (\mathbf a \wedge {\mathbf{x}})_3\,\mathcal L(\mathbf e_3) \\ & \ge\ &0 \end{array} \end{equation} (3.19)

    and, since \mathcal L(\mathbf e_3) < 0, we get

    \begin{equation} \min\limits_{{\mathbf{x}}\in E_{ess}} (\mathbf a \wedge {\mathbf{x}})_3 = \max\limits_{{\mathbf{x}}\in E_{ess}} (\mathbf a \wedge {\mathbf{x}})_3 \end{equation} (3.20)

    that is the function

    {\mathbf{x}}\to (\mathbf a\wedge{\mathbf{x}})_3 = a_1x_2-a_2x_1

    is constant on E_{ess} hence it is constant in {\mathop{{{\rm{co}}}}\nolimits}(E_{ess}) thus {\mathop{{{\rm{cap}}}}\nolimits} {\mathop{{{\rm{proj}}}}\nolimits}{\mathop{{{\rm{co}}}}\nolimits} E_{ess} > 0 entails a_1 = a_2 = 0 that is \mathbf R\mathbf e_3 = \mathbf e_3 as claimed.

    We notice that \mathbf I\in \mathcal S_{\mathcal L, E} and, the other hand, if \mathcal S_{\mathcal L, E}\not\equiv\{\, \mathbf I\, \} then there is

    \widetilde {\mathbf R}\in \mathcal S_{\mathcal L,E}\subset \{\,\mathbf R\in SO(3): \mathbf R\mathbf e_3 = \mathbf e_3\,\}

    such that \widetilde {\mathbf R}\not = \mathbf I and

    \begin{equation*} \widetilde {\mathbf R}{\mathbf{x}}\, = \,{\mathbf{x}}\,+\,\sin\widetilde\vartheta\, (\mathbf e_3\wedge {\mathbf{x}})\,+\,\big(1-\cos\widetilde\vartheta\big)\, \big(\mathbf e_3\wedge (\mathbf e_3\wedge {\mathbf{x}})\big) \end{equation*}

    for every {\mathbf{x}}\in{{\mathbb R}}^3 and for some suitable \widetilde\vartheta\in(0, 2\pi) . By taking (2) of Lemma 3.3 into account, we get

    \begin{eqnarray*} 0 & = & \Phi(\widetilde{\mathbf R},E,\mathcal L) \ = \ \mathcal L\Big( (\widetilde{\mathbf R} - \mathbf I)\,{{\bf x}}\Big)-\,\mathcal L(\mathbf e_3) \,\min\limits_{{\mathbf{x}}\in E_{ess}}\big(\,(\widetilde{\mathbf R}-\mathbf I){\mathbf{x}})\big)_3\, \\ & = & \mathcal L\Big( (\widetilde{\mathbf R} - \mathbf I)\,{{\bf x}}\Big) \ = \ \sin\widetilde\vartheta\, \mathcal L (\mathbf e_3\wedge {\mathbf{x}})\,+\, \big(1-\cos\widetilde\vartheta\big)\, \mathcal L \big(\mathbf e_3\wedge (\mathbf e_3\wedge {\mathbf{x}})\big) \, \, \\ & = & \big(1-\cos\widetilde\vartheta\big)\, \mathcal L \big(\mathbf e_3\wedge (\mathbf e_3\wedge {\mathbf{x}})\big) , \end{eqnarray*}

    thus \ \mathcal L \big(\mathbf e_3\wedge (\mathbf e_3\wedge {\mathbf{x}})\big) = 0 .

    Therefore for any other \mathbf R\in SO(3), \ \mathbf R\neq\mathbf I such that \mathbf R\mathbf e_3 = \mathbf e_3 there is \vartheta\in(0, 2\pi) such that

    \begin{equation*} {\mathbf R}{\mathbf{x}}\, = \,{\mathbf{x}}\,+\,\sin\vartheta\, (\mathbf e_3\wedge {\mathbf{x}})\,+\,\big(1-\cos\vartheta\big)\, \big(\mathbf e_3\wedge (\mathbf e_3\wedge {\mathbf{x}})\big), \qquad\forall\,{\mathbf{x}}\in{{\mathbb R}}^3 , \end{equation*}

    thus, by taking again (2) of Lemma 3.3 into account, we get

    \begin{eqnarray*} \mathcal L\Big( ({\mathbf R} - \mathbf I)\,{{\bf x}}\Big)\!\!\!\!&-&\!\!\!\!\,\mathcal L(\mathbf e_3) \,\min\limits_{{\mathbf{x}}\in E_{ess}}\big(\,({\mathbf R}-\mathbf I){\mathbf{x}})\big)_3\, \\ & = & \mathcal L\Big( ({\mathbf R} - \mathbf I)\,{{\bf x}}\Big) \ = \ \sin\vartheta\, \mathcal L (\mathbf e_3\wedge {\mathbf{x}})\,+\, \big(1-\cos\vartheta\big)\, \mathcal L \big(\mathbf e_3\wedge (\mathbf e_3\wedge {\mathbf{x}})\big) \, \, \\ & = & \big(1-\cos\vartheta\big)\, \mathcal L \big(\mathbf e_3\wedge (\mathbf e_3\wedge {\mathbf{x}})\big) = 0 \end{eqnarray*}

    that is \mathbf R belongs to \mathcal S_{\mathcal L, E} thus concluding the proof of the lemma.

    Remark 3.9. It is possible to show that both alternatives in Lemma 3.8 can actually occur. Indeed in Example 2.7 we have exhibited an example in which \mathcal S_{\mathcal L, E} = \{\mathbf R\in SO(3): \mathbf R\mathbf e_3 = \mathbf e_3\} and we show here that also the other alternative may occur. Indeed set

    \begin{equation} \Omega: = \{{{\bf x}}\!\in\!\mathbb R^3\!: x_1^2+x_2^2\! < \!1,\,0\! < \!x_3\! < \!1\}, \quad E\!: = \! \overline \Omega,\quad \end{equation} (3.21)
    \begin{equation} \mathbf f: = -{\mathbf e}_3,\quad \mathbf g = {\mathbf 1}_{\partial_{l} \Omega}\mathbf n, \end{equation} (3.22)

    where \partial_{l} \Omega is the lateral boundary of \Omega and \mathbf n the unit outward vector normal to \partial_{l} \Omega . If \mathbf R\in SO(3) and we denote its entries as R_{ij}\ i, j = 1, 2, 3, then

    \mathcal L\left ((\mathbf R{{\bf x}}-{{\bf x}}\right)_\alpha\mathbf e_\alpha) = \sum\limits_{i = 1,2}(R_{ii}-1)\int_{\partial_{l} \Omega}x_i^2\,d\mathcal H^2\le 0

    that is condition (2.32) is satisfied. Moreover since

    \mathcal L(\mathbf e_3) = -| \Omega| < 0 = \mathcal L(\mathbf e_1) = \mathcal L(\mathbf e_2)

    and

    \begin{equation} \begin{array}{ll} \Phi(\mathbf R, E, \mathcal L)&=\mathcal L\left (\sum_{j=1}^3(\mathbf R{{\bf x}}-{{\bf x}})_j \mathbf e_j\right)+\pi\,(R_{11}+R_{22}-2) + \pi\,\min\limits_{\overline \Omega}\big\{R_{31}x_1+R_{32}x_2+(R_{33}-1)x_3\big\} \\ &\\ & = -\frac{\pi}{2}(R_{33}-1)-\pi\sqrt{R_{31}^2+R_{32}^2}-\pi (1-R_{33})^+ \sum\limits_{i=1,2}(R_{ii}-1)\int_{\partial_{l} \Omega}x_i^2\,d\mathcal H^2\\\ &\\ & = -\frac{\pi}{2}(1-R_{33})-\pi\sqrt{1-R_{33}^2}+ \sum\limits_{i=1,2}(R_{ii}-1)\int_{\partial_{l} \Omega}x_i^2\,d\mathcal H^2 \le 0 \end{array} \end{equation} (3.23)

    and equality holds if and only if R_{11} = R_{22} = R_{33} = 1 then condition (2.33) is satisfied and \mathcal S_{\mathcal L, E}\equiv \{\mathbf I\} as claimed.

    Lemma 3.10. Assume (2.13), (2.33) and \mathcal L(\mathbf e_3) < 0 . Let \mathbf R_j\in SO(3) be a sequence of rotations such that \, \mathbf R_j\mathbf e_3\neq\mathbf e_3 , \, \mathbf R_j\mathbf e_3\to \mathbf e_3 as j\to +\infty . Then

    \begin{equation} \limsup\limits_{j\to +\infty} \ \frac { \Phi (\mathbf R_j ,E, \mathcal L )} {|\mathbf R_j\mathbf e_3-\mathbf e_3 |\,|\mathcal L(\mathbf e_3)|} \ < \ 0 . \end{equation} (3.24)

    Proof. \Phi (\mathbf R_j, E, \mathcal L)\le 0 , by (2.33). Hence the \limsup in (3.24) cannot be strictly positive.

    We assume by contradiction

    \begin{equation} \limsup\limits_{j\to+\infty} \ \frac { \Phi (\mathbf R_j ,E, \mathcal L )} {|\mathbf R_j\mathbf e_3-\mathbf e_3 |\,|\mathcal L(\mathbf e_3)|} \ = \ 0 . \end{equation} (3.25)

    By Euler-Rodrigues formula there are sequences \mathbf a_j\in {{\mathbb R}}^3 and \vartheta_j\in [0, 2\pi] , such that |\mathbf a_j| = 1 and

    \begin{equation} \mathbf R_j{\mathbf{x}}\, = \, {\mathbf{x}}+ (\sin \vartheta_j)(\mathbf a_j\wedge {\mathbf{x}}) + (1-\cos \vartheta_j)\big((\mathbf a_j\wedge (\mathbf a_j\wedge {\mathbf{x}})\big), \qquad \forall {\mathbf{x}}\in{{\mathbb R}}^3, \end{equation} (3.26)

    thus a direct computation yields

    \begin{equation} |\mathbf R_j\mathbf e_3-\mathbf e_3 | \, = \, \sqrt {a_{1,j}^{\,2}+a_{2,j}^{\,2}}\,\sqrt{2\,(1-\cos \vartheta_j)}. \end{equation} (3.27)

    By taking account of \mathbf R_j\mathbf e_3\!\neq\!\mathbf e_3 and \mathbf R_j\mathbf e_3\to \mathbf e_3 as j\!\to\! +\infty , we get \mathbf a_j\neq \mathbf e_3 , \vartheta_j\!\in\! (0, 2\pi) and therefore, up to subsequences, we may assume: that \mathbf a_j\to \mathbf a, \ \vartheta_j\to \vartheta\in [0, 2\pi], that either \vartheta\in \{0, 2\pi\} or a_3 = 1 and that \mu_j\, {a_{i, j}}\to\, \alpha_i, \ i = 1, 2 with \alpha_1^2+\alpha_2^2 = 1 , where we have set

    \mu_j: = (a_{1,j}^{\,2}+a_{2,j}^{\,2})^{-\frac{1}{2}}.

    For every {\mathbf{x}}\in\Omega and {\mathbf{v}}\in \mathbb R^2, |{\mathbf{v}}| = 1, set

    \begin{equation} \begin{array}{ll} \mathbf p_{{\mathbf{v}}}({\mathbf{x}})& = \big(\, (x_1-\widetilde x_1)\mathbf e_1 +(x_2-\widetilde x_2)\mathbf e_2 +(x_3-\widetilde x_3)\mathbf e_3 \,\big) \wedge (-v_1\mathbf e_2+v_2\mathbf e_1) \\ &\\ & = (v_1(x_1-\widetilde x_1)+v_2(x_2-\widetilde x_2))\,\mathbf e_3\,+\, (x_3-\widetilde x_3) (v_1\mathbf e_1 +v_2\mathbf e_2), \\ \end{array} \end{equation} (3.28)

    where (\widetilde x_1, \widetilde x_2, \widetilde x_3) = {\mathbf{x}}_{\mathcal L}\in {\mathop{{{\rm{ri}}}}\nolimits}{\mathop{{{\rm{co}}}}\nolimits} E_{ess} is chosen as in the proof of (4) in Lemma 3.3. Hence, by taking account of (1) and (4) of Lemma 3.3, we have

    \begin{equation*} 0 = \mathcal L\big(\mathbf p_{{\mathbf{v}}}\big)\, = \, \mathcal L\big( (v_1 x_1 +v_2 x_2)\,\mathbf e_3 \big) - (v_1 \widetilde x_1 +v_2 \widetilde x_2) \, \mathcal L(\mathbf e_3) + \mathcal L\big(\, x_3(v_1 \mathbf e_1 +v_2 \mathbf e_2)\big) \end{equation*}

    that is

    \begin{equation} (v_1 \widetilde x_1 +v_2 \widetilde x_2) \,\mathcal L(\mathbf e_3)\, = \, \mathcal L\big( (v_1 x_1 +v_2 x_2)\mathbf e_3\big) + \mathcal L\big(\, x_3(v_1 \mathbf e_1 +v_2 \mathbf e_2)\big). \end{equation} (3.29)

    By (2) of Lemma 3.3 we know \mathcal L(\mathbf e_3\wedge {\mathbf{x}}) = 0 , then (3.8) entails

    \begin{equation} \mu_j\, \mathcal L(\mathbf a_j\wedge {\mathbf{x}}) = a_{1,j}\mu_j \,\mathcal L(\mathbf e_1\wedge {\mathbf{x}}) + a_{2,j}\mu_j \, \mathcal L(\mathbf e_2\wedge {\mathbf{x}}) \to (\alpha_1\widetilde x_2-\alpha_2\widetilde x_1) \,\mathcal L(\mathbf e_3) \end{equation} (3.30)

    and by (3) of Lemma 3.3 we have

    0\ge\mathcal L\big(\mathbf e_3\wedge (\mathbf e_3\wedge {\mathbf{x}})\big) = \mathcal L(x_3\mathbf e_3 -{\mathbf{x}}) = -\mathcal L(x_1\mathbf e_1 +x_2\mathbf e_2).

    By taking (3.29) into account and by recalling that either \vartheta\in \{0, 2\pi\} or a_3 = 1 , we get

    \begin{eqnarray*} &&\mu_j\,\mathcal L\big(\mathbf a_j\wedge (\mathbf a_j\wedge {\mathbf{x}})\big)\sin\frac{\vartheta_j}{2} = \mu_j\,\mathcal L\big((\mathbf a_j\cdot{\mathbf{x}})\,\mathbf a_j- {\mathbf{x}}\big)\sin\frac{\vartheta_j}{2} \\ && = \mu_j \sin\frac{\vartheta_j}{2}\Big(a_{1,j}\,\mathcal L\big((\mathbf a_j\cdot{\mathbf{x}})\,\mathbf e_1\big) + a_{2,j}\,\mathcal L\big((\mathbf a_j\cdot{\mathbf{x}})\,\mathbf e_2\big) + \mathcal L\big((a_{3,j}^{\,2}-1)\,x_3\,\mathbf e_3\big)\!-\! \mathcal L(x_1\mathbf e_1+x_2\mathbf e_2)\Big) \\ &&\le \ \mathcal L\big(\alpha_1x_3\mathbf e_1+\alpha_2x_3\mathbf e_2+ (\alpha_1x_1+\alpha_2x_2)\,\mathbf e_3 \big)\sin\frac{\vartheta}{2} + o(1) \\ && = (\alpha_1\widetilde x_1+\alpha_2\widetilde x_2)\,\sin\frac{\vartheta}{2}\,\mathcal L(\mathbf e_3) + o(1) , \end{eqnarray*}

    that is

    \begin{equation} \limsup\limits_{j\to + \infty}\ \mu_j\,\mathcal L\big(\mathbf a_j\wedge (\mathbf a_j\wedge {\mathbf{x}})\big)\sin\frac{\vartheta_j}{2} \ \le\ (\alpha_1\widetilde x_1+\alpha_2\widetilde x_2)\,\sin\frac{\vartheta}{2}\mathcal L(\mathbf e_3)\ . \end{equation} (3.31)

    Let now \eta\in C([0, 2\pi]) such that \eta(\vartheta) = \left (2(1-\cos\vartheta)\right)^{-\frac{1}{2}}\sin (\vartheta) for every \vartheta\in (0, 2\pi) .

    By recalling that either a_3 = 1 or \vartheta \in \{0, 2\pi\} we get

    \begin{equation} \mu_j(\mathbf a_j\wedge {\mathbf{x}})_3 \longrightarrow \alpha_1x_2-\alpha_2x_1 , \quad \mu_j(\mathbf a_j \wedge (\mathbf a_j\wedge {\mathbf{x}}))_3\,\sin\frac{\vartheta_j}{2} \longrightarrow (\alpha_1x_1+\alpha_2x_2)\sin\frac{\vartheta}{2}, \end{equation} (3.32)

    so, by taking (3.30)–(3.32) into account we obtain

    \begin{equation*} \begin{array}{ll} 0& = \limsup\limits_{j\to +\infty} \ \frac { \Phi (\mathbf R_j ,E, \mathcal L )} {|\mathbf R_j\mathbf e_3-\mathbf e_3 |\,|\mathcal L(\mathbf e_3)|}\\ &\\ & = \limsup\limits_{j\to +\infty} \frac {\mu_j} {|\mathcal L(\mathbf e_3)|} \, \Big\{ \eta(\vartheta_j)\mathcal L(\mathbf a_j \wedge \mathbf x) + \mathcal L\left (\mathbf a_j\wedge (\mathbf a_j\wedge {\mathbf{x}})\right )\sin\frac{\vartheta_j}{2}+\\ &\\ & -\min\limits_{{\mathbf{x}}\in E_{ess}}\left\{\eta(\vartheta_j) (\mathbf a_j\wedge{{\bf x}})_3+\sin\frac{\vartheta_j}{2}(\mathbf a_j\wedge (\mathbf a_h\wedge{\mathbf{x}}))_3\right\}\mathcal L(\mathbf e_3)\Big\} \\ &\\ & \le\min\limits_{{\mathbf{x}}\in E_{ess}}\Big\{ \eta(\vartheta)(\alpha_1x_2-\alpha_2x_1)+ (\alpha_1x_1+\alpha_2x_2)\sin \frac{\vartheta}{2}\Big\}+\\ &\\ & -\eta(\vartheta)(\alpha_1\widetilde x_2-\alpha_2\widetilde x_1)- (\alpha_1\widetilde x_1+\alpha_2 \widetilde x_2)\sin \frac{\vartheta}{2} \\ &\\ & = \min\limits_{{\mathbf{x}}\in co(E_{ess})} \Big\{ \eta(\vartheta)(\alpha_1(x_2-\widetilde x_2) -\alpha_2(x_1-\widetilde x_1))+ \big( \alpha_1(x_1-\widetilde x_1) + \alpha_2(x_2-\widetilde x_2) \big)\sin\frac{\vartheta}{2} \Big\} \le \ 0, \\ \end{array} \end{equation*}

    since (\widetilde x_1, \widetilde x_2, \widetilde x_3)\in {\mathop{{{\rm{ri}}}}\nolimits}{\mathop{{{\rm{co}}}}\nolimits} E_{ess} . Therefore the function

    g(x_1,x_2): = \eta(\vartheta)(\alpha_1(x_2-\widetilde x_2) -\alpha_2(x_1-\widetilde x_1))+ \big( \alpha_1(x_1-\widetilde x_1) + \alpha_2(x_2-\widetilde x_2) \big)\sin\frac{\vartheta}{2}

    attains its minimum on {\mathop{{{\rm{proj}}}}\nolimits}({\mathop{{{\rm{co}}}}\nolimits} E_{ess}) at (\widetilde x_1, \widetilde x_2)\in {\mathop{{{\rm{ri}}}}\nolimits} ({\mathop{{{\rm{proj}}}}\nolimits}({\mathop{{{\rm{co}}}}\nolimits} E_{ess})) hence, by taking into account of {\mathop{{{\rm{aff}}}}\nolimits} ({\mathop{{{\rm{proj}}}}\nolimits}({\mathop{{{\rm{co}}}}\nolimits} E_{ess})) = \{x_3 = 0\}, we get

    0 = |\nabla g(\widetilde x_1,\widetilde x_2)|^2 = \left(\eta(\vartheta)\alpha_1+\alpha_2\sin\frac{\vartheta}{2}\right)^2+\left(\alpha_1\sin\frac{\vartheta}{2}-\alpha_2\eta(\vartheta)\right)^2 = 2\left(\eta^2(\vartheta)+\sin^2\frac{\vartheta}{2}\right) > 0,

    a contradiction. Thus

    \begin{equation*} \limsup\limits_{j\to +\infty} \ \frac {\Phi (\mathbf R_j ,E, \mathcal L )} {|\mathbf R_j\mathbf e_3-\mathbf e_3|\,|\mathcal L(\mathbf e_3)|}\ < 0 \end{equation*}

    and the proof is achieved.

    Remark 3.11. If {\mathop{{{\rm{cap}}}}\nolimits} ({\mathop{{{\rm{proj}}}}\nolimits}({\mathop{{{\rm{co}}}}\nolimits} E_{ess})) = 0 then the claim of Lemma 3.10 may be false even if {\mathop{{{\rm{cap}}}}\nolimits} E > 0 . For instance, set for every j\in \mathbb N\setminus \{0\}

    \begin{equation*} \mathbf R_j: = \mathbf e_1\otimes\mathbf e_1+(1-j^{-1})(\mathbf e_2\otimes\mathbf e_2+\mathbf e_3\otimes\mathbf e_3)+\sqrt{2\,j^{-1}-j^{-2}}\,(\mathbf e_3\otimes\mathbf e_2-\mathbf e_2\otimes\mathbf e_3), \end{equation*}

    let

    \begin{equation*} \Omega: = \{{\mathbf{x}}: x_1^2+x_2^2 < 1,\ 0 < x_3 < 1\}, \end{equation*}
    \begin{equation*} E: = \overline \Omega\cap\{x_2 = 0,\ 0 < x_3 < \frac{1}{2}\}, \end{equation*}

    and \mathbf f: = -\mathbf e_3, \ \mathbf g = \mathbf 0. It is straightforward checking that \mathbf R_j\mathbf e_3\neq \mathbf e_3, \mathbf R_j \to \mathbf I , moreover since \mathcal L(\mathbf e_3) < 0 = \mathcal L(\mathbf e_1) = \mathcal L(\mathbf e_2),

    \begin{equation} \mathcal L((\mathbf R-\mathbf I){{\bf x}})-\min\limits_{{{\bf x}}\in E_{ess}}((\mathbf R{{\bf x}})_3-x_3)\}\,\mathcal L(\mathbf e_3) = \,-\,\pi \,|R_{31}|\,\le\, 0, \qquad \forall\,\mathbf R\in SO(3) \end{equation} (3.33)

    and

    \begin{equation} \mathcal L((\mathbf R-\mathbf I){{\bf x}})_\alpha\mathbf e_\alpha) = 0, \qquad \forall\,\mathbf R\in SO(3), \end{equation} (3.34)

    conditions (2.32) and (2.33) are satisfied. Nevertheless

    \begin{equation} \mathcal L((\mathbf R_j-\mathbf I){\mathbf{x}})-\min\limits_{E_{ess}}((\mathbf R_j-\mathbf I){\mathbf{x}})_3) \mathcal L(\mathbf e_3) = 0 \end{equation} (3.35)

    and the claim of Lemma 3.10 cannot be true in this case.

    This section contains the proof of our main result. We start by showing that sequences of deformations with equibounded energy correspond (up to suitably tuned rotations and translations of the horizontal components) to displacements that are equibounded in H^1 .

    Lemma 4.1. (compactness) Assume that E , \mathcal L and \mathcal W fulfil (2.13), (2.19)–(2.22), (2.33) and \mathcal L(\mathbf e_3) < 0 . If \, 0 < h_j\to 0^+ as j\to +\infty then for every {\mathbf{y}}_j\in H^1(\Omega; \mathbb R^3) with \mathcal G_j({\mathbf{y}}_j)\le M < +\infty there are \mathbf R_j\in SO(3), \ \mathbf c_j\in \mathcal C_{\mathbf R_j} such that

    \begin{equation} \mathbf R_j\to \mathbf R\in \mathcal S_{\mathcal L, E} \end{equation} (4.1)

    and the sequence

    {h_j}^{-1}\big(\,{\mathbf{y}}_j\,-\,\mathbf R_j{\mathbf{x}}\,-\mathbf c_j\,\big)_{\alpha}\mathbf e_{\alpha}+{h_j}^{-1}(y_{j,3}-x_3)\mathbf e_3

    is bounded in H^1(\Omega; \mathbb R^3) .

    Proof. Referring to (2.36), we can choose {\mathbf R}_j\in \mathcal M({\mathbf{y}}_j) in such a way that, up to subsequence and without relabelling, {\mathbf R}_j\to\mathbf R . Then we define \mathbf c_j = (c_{j, 1}, c_{j, 2}, c_{j, 3}) by

    \begin{equation} c_{j,\alpha} = \, |\Omega|^{-1}\!\!\int_\Omega ({\mathbf{y}}_j ({\mathbf{x}})-\mathbf R_j {\mathbf{x}})_\alpha\,d{\mathbf{x}}, \qquad \alpha = 1,2, \end{equation} (4.2)
    \begin{equation} c_{j,3} = -\min\limits_{{\mathbf{x}}\in E_{ess}} ((\mathbf R_j-\mathbf I){\mathbf{x}})_3. \end{equation} (4.3)

    By the rigidity inequality ([24]) there exists a constant C = C(\Omega) > 0 such that

    \begin{equation} \begin{array}{ll} M& \ge\ \mathcal G_j({\mathbf{y}}_j) \,\ge\, C\, h_j^{-2}\!\! \int_\Omega |\nabla {\mathbf{y}}_j- \mathbf R_j|^2 \,d{\mathbf{x}}- h_j^{-1} \mathcal L({\mathbf{y}}_j-{\mathbf{x}})\ \\ & = \, C\, h_j^{-2} \!\!\int_\Omega |\nabla {\mathbf{y}}_j- \mathbf R_j|^2 \,d{\mathbf{x}}- h_j^{-1} \mathcal L({\mathbf{y}}_j-\mathbf R_j{\mathbf{x}}-\mathbf c_j) - h_j^{-1} \mathcal L(\mathbf R_j{\mathbf{x}}-{\mathbf{x}}+\mathbf c_j). \end{array} \end{equation} (4.4)

    Thus, by (2.33), \mathcal L(\mathbf e_1)\! = \!\mathcal L(\mathbf e_2)\! = \!0 and the definition of c_{j, 3} , we get

    \begin{equation} M \,\ge\, C\, h_j^{-2} \int_\Omega |\nabla {\mathbf{y}}_j- \mathbf R_j|^2 \,d{\mathbf{x}}- h_j^{-1} \mathcal L({\mathbf{y}}_j-\mathbf R_j{\mathbf{x}}-\mathbf c_j)\, \end{equation} (4.5)

    and Poincaré inequality entails, for every \varepsilon\! > \!0 ,

    \begin{equation} \begin{array}{ll} h_j^{-1} \! \sum\limits_{\alpha = 1}^{2} \mathcal L\big(\,({\mathbf{y}}_j-{\mathbf R}_j{\mathbf{x}}-{\mathbf c}_j)_\alpha \,{\mathbf e}_{\alpha} \big) \, &\le\, h_j^{-1} C_P \|\mathcal L\|_{*} \Big ( \sum\limits_{\alpha = 1}^{2}\! \int_\Omega |\,(\nabla{{\mathbf{y}}_j}-{\mathbf R}_j)_\alpha\, |^2\,d{\mathbf{x}} \Big )^{1/2}\!\! \\ & \le\, \frac {C_P \|\mathcal L\|^2_{*}} {2\varepsilon} \,+\, \frac {\varepsilon\,h_j^{-2}\,C_P} {2}\, \sum\limits_{\alpha = 1}^{2}\int_\Omega |(\nabla{\mathbf{y}}_j-{\mathbf R}_j)_{\alpha}|^2\,d{\mathbf{x}}. \\ \end{array} \end{equation} (4.6)

    Estimates (4.5) and (4.6) together with Young inequality provide

    \begin{eqnarray} \quad M \!\!\!\!&\ge&\!\!\!\! h_j^{-2}\! \left(\!C-\frac {\varepsilon\, C_P}{2}\!\right) \!\! \int_\Omega \!|\nabla {\mathbf{y}}_j\!-\! \mathbf R_j|^2 d{\mathbf{x}} -\frac {C_P \,\|\mathcal L\|^2_{*}}{2\,\varepsilon} - h_j^{-1} \!\mathcal L\big(({\mathbf{y}}_j-\mathbf R_j{\mathbf{x}}-\mathbf c_j)_3 \,\mathbf e_3 \big)\! \\ \!\!\!\!&\ge&\!\!\!\! h_j^{-2}\! \left(\!C-\frac {\varepsilon\, C_P}{2}\right) \!\int_\Omega \!|\nabla {\mathbf{y}}_j\!-\! \mathbf R_j|^2 \,d{\mathbf{x}} -\frac {C_P \|\mathcal L\|^2_{*}} {2\,\varepsilon} \\ && \ \ - h_j^{-1} \,\|\mathcal L\|_{*} \, \big (\, \| ({\mathbf{y}}_j-\mathbf R_j{\mathbf{x}}-\mathbf c_j)_3 \|_{L^2(\Omega)} + \| \nabla ({\mathbf{y}}_j-\mathbf R_j{\mathbf{x}})_3 \|_{L^2(\Omega)} \,\big) \\ \!\!\!\!&\ge&\!\!\!\! h_j^{-2}\! \left(\!C-\frac {\varepsilon\,C_P}{2}-\frac {\varepsilon} 2\right) \!\int_\Omega \!|\nabla {\mathbf{y}}_j\!-\! \mathbf R_j|^2 \,d{\mathbf{x}} -\left(\frac {C_P } {2\,\varepsilon} + \frac 1 {2\,\varepsilon}\right) \|\mathcal L\|^2_{*} \\ && \ \ - h_j^{-1} \,\|\mathcal L\|^2_{*} \, \big (\, \| ({\mathbf{y}}_j-\mathbf R_h{\mathbf{x}}-\mathbf c_j)_3 \|_{L^2(\Omega)} . \end{eqnarray} (4.7)

    By choosing \varepsilon = C/ (C_P+1) , we get

    \begin{eqnarray} && h_j^{-2}\ \frac {C}{2} \int_\Omega \!|\nabla {\mathbf{y}}_j\!-\! \mathbf R_j|^2 \,d{\mathbf{x}} \ \\ && \le\, M\, +\, \frac {(C_P+1)^2}{2C} \|\mathcal L\|^2_{*} \,+\, h_j^{-1} \,\|\mathcal L\|^2_{*} \, \| ({\mathbf{y}}_j-\mathbf R_j{\mathbf{x}}-\mathbf c_j)_3 \|_{L^2(\Omega)} . \end{eqnarray} (4.8)

    Thus, if we show that \ h_j^{-1} \| ({\mathbf{y}}_j-\mathbf R_j{\mathbf{x}}-\mathbf c_j)_3 \|_{L^2(\Omega)}\, is uniformly bounded then, due to estimate (4.8), \|h_j^{-1}(\nabla{\mathbf{y}}_j-\mathbf R_j)\|_{L^2(\Omega)} is equibounded too. So \, h_j^{-1} \, ({\mathbf{y}}_j-\mathbf R_j{\mathbf{x}}-\mathbf c_j)\, is uniformly bounded in H^{1}(\Omega; {{\mathbb R}}^3) and we set

    \begin{equation} M_1: = \sup\limits_{j}\|\mathcal L\|_*\|{\mathbf{y}}_j-\mathbf R_j{{\bf x}}-\mathbf c_j\|_{H^1( \Omega;\mathbb R^3)} > 0. \end{equation} (4.9)

    To this aim we assume by contradiction that, up to subsequences,

    \begin{equation} t_j\ : = \ h_j^{-1} \, \| \,({\mathbf{y}}_j-\mathbf R_j{\mathbf{x}}-\mathbf c_j)_3\, \|_{L^2(\Omega)}\ \to\ +\infty\ \end{equation} (4.10)

    and set w_j \ : = t_j^{-1} h_j^{-1} \, ({\mathbf{y}}_j-\mathbf R_j{\mathbf{x}}-\mathbf c_j)_3. Then

    \begin{equation} \|w_j\|_{L^2(\Omega)} = 1,\qquad |\nabla {\mathbf{y}}_j\!-\! \mathbf R_j|^2\ = \ \sum\limits_{\alpha = 1}^2 \,|\,\nabla ({\mathbf{y}}_j\!-\! \mathbf R_j{\mathbf{x}})_\alpha\,|^2 \ +\ h_j^{\,2} t_j^{\,2}\,|\nabla w_j|^2 , \end{equation} (4.11)
    \begin{eqnarray} && C t_j^{\,2}\int_\Omega|\nabla w_j|^2\,d{\mathbf{x}} \,-\, t_j\,\mathcal L(w_j\,\mathbf e_3)\, \\ \le\, &&\!\! M-C\,h_j^{-2}\, \sum\limits_{\alpha = 1}^2 \,\int_\Omega|\, \nabla ({\mathbf{y}}_j\!-\! \mathbf R_j{\mathbf{x}})_\alpha\,|^2\,d{\mathbf{x}}\,+ \,h_j^{-1}\, \sum\limits_{\alpha = 1}^2 \, \mathcal L \big( \,({\mathbf{y}}_j\!-\! \mathbf R_j{\mathbf{x}}-\mathbf c_j)_\alpha\,\mathbf e_\alpha\,\big)\, \\ \le\, &&\!\! M-C\,h_j^{-2}\, \sum\limits_{\alpha = 1}^2 \,\int_\Omega|\, \nabla ({\mathbf{y}}_h\!-\! \mathbf R_j{\mathbf{x}})_\alpha\,|^2\,d{\mathbf{x}}\,+ \, \frac {C_P\,\|\mathcal L\|^2_{*} }{2\,\varepsilon} \, \\ +&&\!\! \frac {h_j^{-2}\, \varepsilon\,C_P} {2} \sum\limits_{\alpha = 1}^2\int_\Omega |\nabla ({\mathbf{y}}_j\!-\! \mathbf R_j{\mathbf{x}})_\alpha\,|^2\,d{\mathbf{x}}, \end{eqnarray} (4.12)

    and by choosing \varepsilon = 2C/C_P in (4.12) we get

    \begin{equation} C\, t_j^{\,2}\int_\Omega|\nabla w_j|^2\,d{\mathbf{x}} \,-\, t_j\,\mathcal L(w_j\,\mathbf e_3)\ \le\ \frac{C_P^2\,\|\mathcal L\|^2_{*}}{4\,C}+M \end{equation} (4.13)

    while, by choosing \varepsilon = C/C_P , (4.12) yields

    \begin{eqnarray} && \frac 1 2 \,C\,h_j^{-2}\, \sum\limits_{\alpha = 1}^2 \,\int_\Omega|\, \nabla ({\mathbf{y}}_h\!-\! \mathbf R_j{\mathbf{x}})_\alpha\,|^2\,d{\mathbf{x}}\,+\,C\, t_j^{\,2}\,|\nabla w_j|^2\,-t_j\, \mathcal L(w_h\,\mathbf e_3)\ \\ && \qquad \le \ \frac {C_P^2}{2C}\,\|\mathcal L\|^2_{*}+M. \end{eqnarray} (4.14)

    Thus

    \begin{equation} \frac 1 2 \,C\,\frac {h_j^{-2}}{t_j^{\,2}}\, \sum\limits_{\alpha = 1}^2 \,\int_\Omega|\, \nabla ({\mathbf{y}}_j\!-\! \mathbf R_j{\mathbf{x}})_\alpha\,|^2\,d{\mathbf{x}}\ \le\ \frac 1 {t_j}\, \mathcal L(w_j\,\mathbf e_3) \,+\, \frac 1 {{t_j}^{\,2}}\,\frac {C_P^2}{2C}\,\|\mathcal L\|^2_{*}+\frac{M}{t_j^2}. \end{equation} (4.15)

    Normalization \|w_j\|_{L^2} = 1 entails, for every \varepsilon > 0 ,

    \begin{eqnarray} \mathcal L \big( w_j\,\mathbf e_3\big) &\le& \|\mathcal L\|_{*} \big( \,\|w_j\|_{L^2}+ \| \nabla w_j \|_{L^2}\,\big) \, = \, \|\mathcal L\|_{*} \big( \,1+ \| \nabla w_j \|_{L^2}\,\big)\, \\ &\le& \|\mathcal L\|_{*}\,\,+\, \frac {\|\mathcal L\|^2_{*}}{2\,\varepsilon}\,+\, \frac{\varepsilon}2 \, \| \nabla w_j \|_{L^2}^2, \end{eqnarray} (4.16)

    and choosing \varepsilon = C\, t_j^2 therein we get, by (4.13),

    \begin{equation} \frac {C}2t_j^2\,\int_\Omega|\nabla w_j|^2\,\,d{\mathbf{x}}\ \le \ t_j\,\|\mathcal L\|_{*} + \frac {\|\mathcal L\|^2_{*}} {2C\,t_j}+M, \end{equation} (4.17)

    thus \int_\Omega |\nabla w_j|^2\, \, d{\mathbf{x}}\to 0 so by (4.11) w_j\to w in H^1(\Omega; {{\mathbb R}}^3) with \nabla w = 0 a.e. in \Omega that is w is a constant function since \Omega is a connected open set.

    Combining estimates (4.15)–(4.17), we get

    \begin{eqnarray} && \frac 1 2 C\,\frac {h_j^{-2}}{t_j^{\,2}}\, \sum\limits_{\alpha = 1}^2 \,\int_\Omega|\, \nabla ({\mathbf{y}}_j\!-\! \mathbf R_j{\mathbf{x}})_\alpha\,|^2\,d{\mathbf{x}} \ \\ && \qquad \le \ \frac 1 {t_j} \left( \|\mathcal L\|^2_{*}\,+\, \frac {\|\mathcal L\|^2_{*}}{2}\,+\, \frac {1}{2}\, \| \nabla w_j \|_{L^p}^2 \right) \,+\, \frac 1 {{t_j}^{\,2}}\,\frac {C_P^2}{2\,C_R\,C}\,\|\mathcal L\|^2_{*}+\frac{M}{t_j^2}, \end{eqnarray} (4.18)

    hence

    \begin{equation} \frac{1}{h_j\,t_j}\,\nabla \big( {\mathbf{y}}_j-\mathbf R_j{\mathbf{x}}\big)_\alpha\ \to \ 0, \qquad \hbox{in }L^2(\Omega), \qquad \hbox{if }\alpha = 1,2 \end{equation} (4.19)

    and

    \begin{equation} h_j^{-1}t_j^{-1}\,\left({\mathbf{y}}_j-\mathbf R_j{\mathbf{x}}-\mathbf c_j\right)\ \to\ w\,\mathbf e_3, \qquad \hbox{q.e. } {\mathbf{x}}\in E . \end{equation} (4.20)

    Moreover, by (4.13), we get

    \begin{equation*} \mathcal L ( w_j\,\mathbf e_3) \ \ge \ -\,\frac {C_P^{\,2}}{4\,C\, t_j}\, \|\mathcal L\|^2_{*}. \end{equation*}

    Hence, due to \mathcal L (w_j\, \mathbf e_3)\to \mathcal L (w\, \mathbf e_3) = w\, \mathcal L(\mathbf e_3) , we have w\, \mathcal L (\mathbf e_3)\!\ge\! 0, thus, by taking into account of \mathcal L (\mathbf e_3) < 0, we get w\le 0 and eventually, by \|w_j\|_{L^2} = 1 , we obtain w\ < \ 0 . Then, by (2.33), (4.4) and (4.9), we get

    \begin{equation} 0\ \le -\Phi(\mathbf R_j,E,\mathcal L) = \ \mathcal L \left(\,{\mathbf{x}}\,-\,\mathbf R_j{\mathbf{x}}\,-\,\mathbf c_j\,\right) \le \ (M+M_1)\,h_j . \end{equation} (4.21)

    Hence, due to \mathbf R_j\to \mathbf R, we have \Phi(\mathbf R, E, \mathcal L) = 0 thus \mathbf R\in \mathcal S_{\mathcal L, E} and (4.1) is proven.

    We notice that either \mathbf R_j\mathbf e_3\not = \mathbf e_3 for j large enough or \mathbf R_j\mathbf e_3 = \mathbf e_3 for infinitely many j . In the first case, by taking account of \mathcal L(\mathbf e_3) < 0, Lemma 3.10 entails

    \begin{equation} \limsup\limits_{j\to +\infty} \ \frac { \Phi(\mathbf R_j,E,\mathcal L)}{|\mathbf R_j\mathbf e_3-\mathbf e_3 |\,|\mathcal L(\mathbf e_3)|} \ < \ 0 . \end{equation} (4.22)

    By (4.21) we get \Phi(\mathbf R_j, E, \mathcal L)\ge - (M+M_1) h_j hence

    \begin{equation} \gamma: = \liminf\limits _{j\to +\infty}\frac {h_j}{|\mathbf R_j\mathbf e_3-\mathbf e_3 |} > 0 \end{equation} (4.23)

    and for large enough j (4.23) yields

    \begin{equation} |\mathbf R_j\mathbf e_3-\mathbf e_3 |\,\le\, \, \frac{2h_j}{\gamma}. \end{equation} (4.24)

    Therefore, for every {\mathbf{x}}\in\Omega

    \begin{equation} \frac {|(\mathbf R_j{\mathbf{x}})_3-{\mathbf{x}}_3|} {\ h_j\,t_j} \le \frac {|{{\bf x}}||\mathbf R_j^T\mathbf e_3-\mathbf e_3|} {\ h_j\,t_j}\mathop{\longrightarrow}^{j\to +\infty}\ 0, \end{equation} (4.25)

    hence, by taking into account of c_{j, 3}\! = \!-\min\, \left\{\, (\mathbf R_j{\mathbf{x}})_3-{\mathbf{x}}_3\, :\, {\mathbf{x}}\in E_{ess}\, \right\} , we get for q.e. {\mathbf{x}}\!\in\! E

    \begin{equation} \begin{array}{rl} \frac {{\mathbf{y}}_{j,3}^*-{\mathbf{x}}_3} {\ h_j\,t_j} = & \frac {{\mathbf{y}}_{j,3}^*-(\mathbf R_j{\mathbf{x}})_3-\mathbf c_{j,3}} {\ h_j\,t_j} \,+\, \frac {(\mathbf R_j{\mathbf{x}})_3-{\mathbf{x}}_3+\mathbf c_{j,3}} {\ h_j\,t_j} \\ = & \frac {{\mathbf{y}}_{h_j,3}^*-(\mathbf R_j{\mathbf{x}})_3-\mathbf c_{j,3}} {\ h_j\,t_j} \,+\, o(1) \ \mathop{\longrightarrow}^{j\to +\infty}\ w \ < \ 0,\\ \end{array} \end{equation} (4.26)

    a contradiction since {\mathbf{y}}_{j, 3}^*\ge (1-h_j)x_3 for q.e. {\mathbf{x}}\in E , that is (h_j\, t_j)^{-1}\big({\mathbf{y}}_{j, 3}^*-{\mathbf{x}}_3\big)\ge -\, x_3/t_j\longrightarrow 0 as j\to +\infty for q.e. {\mathbf{x}}\in E_{ess} . Therefore in this case the sequence t_j is bounded so h_j^{-1}\left({\mathbf{y}}_j-\mathbf R_j{\mathbf{x}}-\mathbf c_j\right) is equibounded in H^1(\Omega; {{\mathbb R}}^3) and in particular h_j^{-1}\left({\mathbf{y}}_j-\mathbf R_j{\mathbf{x}}-\mathbf c_j\right)_{\alpha}\mathbf e_{\alpha} is equibounded in H^1(\Omega; {{\mathbb R}}^3) .

    In the second case we may assume that \mathbf R_j\mathbf e_3 = \mathbf e_3 for every j so c_{j, 3} = 0 for every j . By arguing as in the previous case we may assume that

    \begin{equation*} \label{contradiction} t_j\ : = \ h_j^{-1} \, \| \,({\mathbf{y}}_j-\mathbf R_j{\mathbf{x}}-\mathbf c_j)_3\, \|_{L^2(\Omega)} = h_j^{-1}\| y_{j,3}- x_3\|_{L^2(\Omega)}\ \to\ +\infty \end{equation*}

    and by setting w_j \ : = t_j^{-1}h_j^{-1} \, (y_{j, 3}- x_3) we get w_j\to w < 0 as before which is again a contradiction, so t_j is a bounded sequence. Eventually we are left to show that, in the first case, h_j^{-1}(y_{j, 3}-x_3) is equibounded in H^1(\Omega; {{\mathbb R}}^3) . To this aim let C > 0 such that

    \left\|h_j^{-1}\left({\mathbf{y}}_{j,3}-(\mathbf R_j{\mathbf{x}})_3-\mathbf c_{j,3}\right)\right\|_{H^1( \Omega;\mathbb R^3)}\le C

    for every j\in \mathbb N and assume that for every n\in \mathbb N there exists j_n such that

    \begin{equation} \left\|h_{j_n}^{-1}({\mathbf{y}}_{j_n,3}-{\mathbf{x}}_3)\right\|_{H^1( \Omega;\mathbb R^3)}\ge n. \end{equation} (4.27)

    Then for every n > C we have \mathbf R_{j_n}\mathbf e_3\not = \mathbf e_3 otherwise c_{j_n, 3} = 0 and

    n\le \left\|h_{j_n}^{-1}({\mathbf{y}}_{j_n,3}-{\mathbf{x}}_3)\right\|_{H^1( \Omega;\mathbb R^3)} = \left\|h_{j_n}^{-1}\left({\mathbf{y}}_{j_n,3}-(\mathbf R_{j_n}{\mathbf{x}})_3-\mathbf c_{j_n,3}\right)\right\|_{H^1( \Omega;\mathbb R^3)} \le C,

    a contradiction. By taking account of (4.23) and (4.25) there exists \widetilde C > 0 such that

    |\mathbf R_{j_n}\mathbf e_3- \mathbf e_3|\le \widetilde C h_{j_n}

    for every n > C , hence

    \begin{equation} \begin{array}{ll} & \left |h_{j_n}^{-1}({\mathbf{y}}_{j_n,3}-{\mathbf{x}}_3)\right | \\ & \le \left |h_{j_n}^{-1}({\mathbf{y}}_{j_n,3}-(\mathbf R_{j_n}{\mathbf{x}})_3-\mathbf c_{j_n,3})\right |+h_{j_n}^{-1} |(\mathbf R_{j_n}{\mathbf{x}})_3-x_3| \\ & \le \left |h_{j_n}^{-1}({\mathbf{y}}_{j_n,3}-(\mathbf R_{j_n}{\mathbf{x}})_3-\mathbf c_{j_n,3})\right |+h_{j_n}^{-1}|\mathbf R_{j_n}\mathbf e_3- \mathbf e_3||{\mathbf{x}}| \\ & \le \left |h_{j_n}^{-1}({\mathbf{y}}_{j_n,3}-(\mathbf R_{j_n}{\mathbf{x}})_3-\mathbf c_{j_n,3})\right |+\widetilde C\sup\limits_{ \Omega}|{\mathbf{x}}|\\ \end{array} \end{equation} (4.28)

    thus showing that

    \left\|h_{j_n}^{-1}({\mathbf{y}}_{j_n,3}-{\mathbf{x}}_3)\right\|_{H^1( \Omega;\mathbb R^3)}\le C+\widetilde C\sup\limits_{ \Omega}|{\mathbf{x}}|| \Omega|

    which contradicts (4.27) and proves that h_j^{-1}(y_{j, 3}-x_3) is equibounded in H^1(\Omega; {{\mathbb R}}^3) .

    Remark 4.2. If {\mathop{{{\rm{cap}}}}\nolimits} ({\mathop{{{\rm{proj}}}}\nolimits}({\mathop{{{\rm{co}}}}\nolimits} E_{ess}))\! = \!0 , the claim of Lemma 4.1 may fail even if {\mathop{{{\rm{cap}}}}\nolimits} E > 0. Indeed, choose \Omega, E, \mathbf f, \mathbf g, \mathbf R_j as in Remark 3.11 and set h_j = j^{-1}. Thus both (2.32) and (2.33) are satisfied but (2.13) is not. It is readily seen that {\mathbf{y}}_j({\mathbf{x}}): = \mathbf R_j{\mathbf{x}} belongs to \mathcal A_j since y_{j, 3} = (1-j^{-1})x_3 on E and that \mathcal G_j({\mathbf{y}}_j) = \pi|\, \Omega|/2 for every j , but

    \begin{equation*} j\,(y_{j,3}-x_3) = j\,(x_2\sqrt{2j^{-1}-j^{-2}}-x_3 j^{-1}) \end{equation*}

    is not equibounded in H^1(\Omega; {{\mathbb R}}^3) as j\!\to\!+\infty : thus claim of Lemma 4.1 fails in this case.

    Lemma 4.3. Assume that E , \mathcal L and \mathcal W fulfil conditions (2.13), (2.19)–(2.22), (2.33) and \mathcal L(\mathbf e_3) < 0. Choose {\mathbf{y}}_j, \ \mathbf R_j as in Lemma 4.1 and set

    \begin{equation} {\mathbf{z}}_j({{\bf x}}): = h_j^{-1}\{(\mathbf R_j{\mathbf{x}})_3-x_3) - \min\limits_{{\mathbf{x}}\in E_{ess}} \big( (\mathbf R_j{\mathbf{x}})_3-x_3 \big)\}\mathbf e_3. \end{equation} (4.29)

    Then there exist b_1, \ b_2, \ b_3\in \mathbb R such that by setting

    \begin{equation} {\mathbf{z}}({{\bf x}}): = (b_{1}x_1+b_{2}x_2+b_{3})\mathbf e_3 \end{equation} (4.30)

    we have, up to subsequences, {\mathbf{z}}_j {\rightharpoonup} {\mathbf{z}} in w^*-W^{1, \infty}(\Omega; \mathbb R^3) .

    Proof. We may assume that \mathbf R_j\mathbf e_3\neq \mathbf e_3 for infinitely many j otherwise {\mathbf{z}}_j\equiv 0 for j large enough and thesis is obvious. Therefore by Euler-Rodrigues formula, there are sequences \mathbf a_j\in {{\mathbb R}}^3 and \vartheta_j\in (0, 2\pi) , s.t. |\mathbf a_j| = 1, \ \mathbf a_j\neq \mathbf e_3 and

    \begin{equation} \mathbf R_j{\mathbf{x}}\, = \, {\mathbf{x}}+ (\sin \vartheta_j)(\mathbf a_j\wedge {\mathbf{x}}) + (1-\cos \vartheta_j)\big((\mathbf a_j\wedge (\mathbf a_j\wedge {\mathbf{x}})\big), \qquad \forall {\mathbf{x}}\in{{\mathbb R}}^3. \end{equation} (4.31)

    By recalling (3.27) and (4.1) we have, up to subsequences, \mathbf R_j\mathbf e_3\to \mathbf e_3 . Then, up to subsequences, we may assume: that \mathbf a_j\to \mathbf a, \ \vartheta_j\to \vartheta\in [0, 2\pi], that either \vartheta\in \{0, 2\pi\} or a_3 = 1 and that \mu_j\, {a_{i, j}}\to\, \alpha_i, \ i = 1, 2 with \alpha_1^2+\alpha_2^2 = 1 , where we set \mu_j: = (a_{1, j}^{\, 2}+a_{2, j}^{\, 2})^{-\frac{1}{2}} . By recallling (4.24) we may assume that, up to subsequences,

    h_j^{-1} \min\limits_{{\mathbf{x}}\in E_{ess}} \big( (\mathbf R_j{\mathbf{x}})_3-x_3 \big)\to \beta

    for some \beta\in \mathbb R . Moreover by exploiting (3.27), (3.26), we get

    \begin{equation} \begin{array}{ll} & h_j^{-1}((\mathbf R_j{\mathbf{x}})_3-{\mathbf{x}}_3)\\ &\\ & = \frac{\mu_j}{\sqrt{2\,(1-\cos \vartheta_j)}}\big(\sin \vartheta_j(\mathbf a_j\wedge {\mathbf{x}})_3 + (1-\cos \vartheta_j)\big((\mathbf a_j\wedge (\mathbf a_j\wedge {\mathbf{x}})_3\big)\big)\frac{|\mathbf R_j\mathbf e_3-\mathbf e_3 |}{h_j}\\ &\\ & = \left (\mu_j\eta(\vartheta_j)(\mathbf a_j\wedge {\mathbf{x}})_3+\mu_j\big((\mathbf a_j\wedge (\mathbf a_j\wedge {\mathbf{x}})\big)_3\sin\frac{\vartheta_j}{2}\right)\frac{|\mathbf R_j\mathbf e_3-\mathbf e_3 |}{h_j} \end{array} \end{equation} (4.32)

    where \eta\in C([0, 2\pi]) is such that \eta(\vartheta) = \left (2(1-\cos\vartheta)\right)^{-\frac{1}{2}}\sin\vartheta for every \vartheta\in (0, 2\pi) .

    By arguing as in (3.32) and by taking (4.24) into account we get, up to subsequences,

    \begin{equation} h_j^{-1}((\mathbf R_j{\mathbf{x}})_3-x_3) \to \lambda\left (\eta(\vartheta)(\alpha_1x_2-\alpha_2x_1)+(\alpha_1x_1+\alpha_2x_2)\sin\frac{\vartheta}{2}\right ),\quad \forall\, {{\bf x}}\in \Omega \end{equation} (4.33)

    for some \lambda\ge 0 . On the other hand \nabla{\mathbf{z}}_j = h_j^{-1}(\mathbf R_j\mathbf e_3-\mathbf e_3) and (4.24) entail \|\nabla{\mathbf{z}}_j\|_{\infty}\le C for some C > 0 so {\mathbf{z}}_j{\rightharpoonup} {\mathbf{z}} in w^*-W^{1, \infty}(\Omega:\mathbb R^3) whenever b_1 = \lambda(-\alpha_{2}\eta(\vartheta)+\alpha_1\sin\frac{\vartheta}{2}), \ b_2 = \lambda(\alpha_{1}\eta(\vartheta)+\alpha_2\sin\frac{\vartheta}{2}), \ b_3 = -\beta.

    Lemma 4.4. (Lower bound) Assume that E , \mathcal L , \mathcal W fulfil the conditions (2.13)–(2.22), (2.32), (2.33) and \mathcal L(\mathbf e_3) < 0 . If h_j\to 0^+ as j\to +\infty then, for every sequence of deformations {\mathbf{y}}_j\in H^1(\Omega; \mathbb R^3) such that \mathcal G_j({\mathbf{y}}_j)\le M < +\infty and for every \mathbf R_j\in \mathcal M({\mathbf{y}}_j) there exist \mathbf c_j\in \mathcal C_{\mathbf R_j} such that by setting

    \begin{equation} {\mathbf{u}}_j({{\bf x}})\,: = \,{h_j}^{-1}\mathbf R_j^T\!\left\{ \big(\mathbf y_j-\mathbf c_j-\mathbf R_j{{\bf x}}\big)_\alpha\,\mathbf e_\alpha\,+\,(y_{j,3}-x_3)\mathbf e_3\,\right\}, \end{equation} (4.34)

    there is {\mathbf{u}}\in\mathcal A such that up to subsequences {\mathbf{u}}_j{\rightharpoonup} {\mathbf{u}} weakly in H^1(\Omega; {{\mathbb R}}^3) and

    \begin{equation} \liminf\limits_{j\to +\infty}\mathcal G_j({\mathbf{y}}_j)\ge \widetilde{\mathcal G}({\mathbf{u}}). \end{equation} (4.35)

    Proof. Due to Lemma 4.1, the sequence defined in (4.34) is equibounded in H^1(\Omega; \mathbb R^3) hence there exists {\mathbf{u}}\in H^1(\Omega; \mathbb R^3) such that up to subsequences {\mathbf{u}}_j{\rightharpoonup}{\mathbf{u}} in H^1(\Omega; \mathbb R^3) . By recalling Lemma A1 of [12] we get, again up to subsequences, {\mathbf{u}}_j^*({\mathbf{x}})\to {\mathbf{uu}}^*({\mathbf{x}}) for q.e. {\mathbf{x}}\in E hence by taking account of

    u_{j,3}^* = {h_j}^{-1}( y_{j,3}^*-x_3)\ge {h_j}^{-1}(x_3-h_jx_3-x_3) = -x_3

    for q.e. {\mathbf{x}}\in E we get u_{3}^*\ge -x_3 for q.e. {\mathbf{x}}\in E that is {\mathbf{uu}}\in \mathcal A .

    By taking account of \mathcal G_j({\mathbf{y}}_j)\le M and by arguing as in Lemma 4.1 the sequence h_j^{-1}({\mathbf{y}}_j-\mathbf R_j-\mathbf c_j) is bounded in H^1(\Omega; \mathbb R^3) hence (4.4) entails

    \begin{equation} 0\ \le \ \mathcal L \left({\mathbf{x}}-\mathbf R_j{\mathbf{x}}-\mathbf c_j\right)\ \le \ (M+M_1)\,h_j . \end{equation} (4.36)

    Therefore, by recalling (4.2), (4.3) and that, up to subsequences, \mathbf R_j\to \mathbf R we get \mathbf R\in \mathcal S_{\mathcal L, E} . By defining {\mathbf{z}}_j as in Lemma 4.3 and by setting

    \mathbf D_{j}: = \mathbb E({\mathbf{u}}_{j})+\frac{1}{2}h_{j}\nabla{\mathbf{u}}_{j}^{T}\nabla{\mathbf{u}}_{j},\ \mathbf F_j: = \mathbb E(\mathbf R_j^T{\mathbf{z}}_{j})+\frac{1}{2}h_{j}\nabla(R_j^T{\mathbf{z}}_{j})^{T}\nabla(R_j^T{\mathbf{z}}_{j})

    a straightforward calculation shows that

    \begin{equation} \nabla{\mathbf{y}}_j^T\nabla{\mathbf{y}}_j-\mathbf I = 2h_j(\mathbf D_{j}+\mathbf F_j). \end{equation} (4.37)

    If now

    \begin{equation*} \label{Bj} B_j: = \{x\in \Omega: \sqrt h_j|\nabla{\mathbf{u}}_j|\le 1\}, \end{equation*}

    we immediately notice that, by Tchebycheff inequality, | \Omega\setminus B_j|\to 0 as j\to +\infty and that for large enough j

    \begin{equation} h_j|\mathbf D_j|\le \sqrt{h_j}\left(\sqrt{h_j}|\nabla{\mathbf{v}}_j|+\tfrac12h_j^{3/2}|\nabla {\mathbf{v}}_j^T||\nabla {\mathbf{v}}_j|\right)\le 2\sqrt{h_j}, \qquad \hbox{on }B_j. \end{equation} (4.38)

    Moreover by Lemma 4.3 there exists C > 0 such that

    \begin{equation} h_j |\mathbf F_j|\le C h_j, \quad\text{in}\ \Omega \end{equation} (4.39)

    hence by defining {\mathbf{z}} as in (4.30) and by taking account of \mathbf R_j\to\mathbf R\in \mathcal S_{\mathcal L, E} , we get

    \begin{equation} \mathbf F_j{\rightharpoonup} \mathbb E({\mathbf{z}}),\quad w^*-L^{\infty}( \Omega;\mathbb R^{3\times 3}). \end{equation} (4.40)

    By taking account of \mathcal L(\mathbf e_\alpha) = 0 for \alpha = 1, 2 , (2.32) entails

    \begin{equation} \begin{array}{ll} &\mathcal L({\mathbf{y}}_j-{{\bf x}}) = \mathcal L((y_{j,3}-x_3)\mathbf e_3)+\mathcal L(({\mathbf{y}}_j-\mathbf R_j{{\bf x}}-\mathbf c_j)_\alpha\mathbf e_\alpha) \\ &\\ &+\mathcal L((\mathbf R_j{{\bf x}}-{{\bf x}})_\alpha\mathbf e_\alpha)\le h_j\mathcal L(\mathbf R_j{\mathbf{u}}_j) \end{array} \end{equation} (4.41)

    thus, since \eta is increasing, by (2.21)–(2.23), (2.25), (4.37)–(4.40) we get for large j

    \begin{equation} \begin{aligned} \mathcal G_j({\mathbf{y}}_j)&\ge\frac1{h_j^2}\int_{B_j}\mathcal V({{\bf x}},h_j \mathbf D_j+h_j \mathbf F_j)\,d{{\bf x}}-\mathcal L(\mathbf R_j{\mathbf{u}}_j) \\ & \ge \int_{B_j}\mathcal Q({{\bf x}},\mathbf D_j+\mathbf F_j)\,d{{\bf x}}-\int_{B_j} \eta(h_j\mathbf D_j+h_j\mathbf F_j)|\mathbf D_j+\mathbf F_j|^2\,d{{\bf x}}-\mathcal L(\mathbf R_j{\mathbf{u}}_j) \\&\ge \int_ \Omega \mathcal Q({{\bf x}},{\mathbf 1}_{B_j}(\mathbf D_j+\mathbf F_j))\,d{{\bf x}}- \eta(3\sqrt h_j)\int_ \Omega|{\mathbf 1}_{B_j}(\mathbf D_j+\mathbf F_j)|^2\,d{{\bf x}}-\mathcal L(\mathbf R_j{\mathbf{u}}_j). \end{aligned} \end{equation} (4.42)

    Since h_{j}\nabla{\mathbf{u}}_{j}^T\nabla{\mathbf{u}}_j\to 0 a.e. in \Omega and |{\mathbf 1}_{B_j}h_{j}\nabla{\mathbf{u}}_{j}^T\nabla{\mathbf{u}}_j| \le 1, by taking account of | \Omega\setminus B_j|\to 0 as j\to +\infty we get {\mathbf 1}_{B_j}h_{j}\nabla{\mathbf{u}}_{j}^T\nabla{\mathbf{u}}_j{\rightharpoonup} 0 weakly in L^2(\Omega, \mathbb R^{3\times3}) . By taking account of {\mathbf 1}_{B_j}\nabla{\mathbf{u}}_j{\rightharpoonup} \nabla{\mathbf{u}} weakly in L^2(\Omega, \mathbb R^{3\times3}) and (4.40), we then obtain

    \begin{equation} {\mathbf 1}_{B_j}(\mathbf D_j+\mathbf F_j){\rightharpoonup} \mathbb E({\mathbf{u}})+\mathbb E({\mathbf{z}}) = \mathbb E({\mathbf{u}})+\frac12 b_\alpha(\mathbf e_\alpha\otimes\mathbf e_3+\mathbf e_3\otimes\mathbf e_\alpha)\ \ (\alpha = 1,2) \end{equation} (4.43)

    weakly in L^2(\Omega, \mathbb R^{3\times 3}) . Since \mathbf R_j\to\mathbf R\in \mathcal S_{\mathcal L, E} , then

    \begin{equation*} \lim\limits_{j\to +\infty}\mathcal L(- \mathbf R_j {\mathbf{u}}_j) = -\mathcal L( \mathbf R {\mathbf{u}}) \end{equation*}

    and by (2.21) and (4.42), the weak L^2(\Omega, \mathbb R^{3\times3}) lower semicontinuity of the map \mathbf B\mapsto\int_\Omega \mathcal Q({{\bf x}}, \mathbf B)d{{\bf x}} entails

    \begin{equation*} \label{pi} \liminf\limits_{j\to +\infty}\mathcal G_j({\mathbf{y}}_j)\ge \int_ \Omega \mathcal Q({{\bf x}},\mathbb E({\mathbf{u}})+\mathbb E({\mathbf{z}}))\,d{{\bf x}}-\mathcal L(\mathbf R{\mathbf{u}})\ge \widetilde{\mathcal G}({\mathbf{u}}) \end{equation*}

    which, by recalling (4.30), ends the proof.

    Remark 4.5. If condition (2.32) is not satisfied then the thesis of Lemma 4.4 may fail. Indeed let \mathbf f: = -\mathbf e_3+6(x_3-\frac{1}{2})\mathbf e_1, \ \mathbf g = \mathbf 0 and

    \begin{equation*} E = \Omega: = \{{\mathbf{x}}: x_1^2+x_2^2 < 1,\ 0 < x_3 < 1\}. \end{equation*}

    It is straightforward checking that \mathcal L(\mathbf e_3) < 0 = \mathcal L(\mathbf e_1) = \mathcal L(\mathbf e_2) and

    \begin{equation} \begin{array}{ll} & \mathcal L((\mathbf R-\mathbf I){{\bf x}})-\min\limits_{{{\bf x}}\in E_{ess}}((\mathbf R{{\bf x}})_3-x_3)\}\,\mathcal L(\mathbf e_3) \\ &\\ & = \frac{\pi}{2}R_{13}-\frac{\pi}{2}(1-R_{33})-\pi\sqrt{1-R_{33}^2} \\ &\\ & \le\frac{\pi}{2}\sqrt{1-R_{33}^2} -\frac{\pi}{2}(1-R_{33})-\pi\sqrt{1-R_{33}^2} \\ &\\ & = -\frac{\pi}{2}\sqrt{1-R_{33}}\left\{\sqrt{1-R_{33}}+\sqrt{1+R_{33}}\right\}\le -\frac{\pi\sqrt 2}{2}\sqrt{1-R_{33}}\le 0 \end{array} \end{equation} (4.44)

    for every \mathbf R\in SO(3) . On the other hand if R_{13} > 0 we have

    \mathcal L((\mathbf R{{\bf x}}-{{\bf x}})_\alpha\mathbf e_\alpha) = \frac{\pi}{2}R_{13} > 0,

    so (2.33) is satisfied while (2.32) is not. Choose now h_j: = j^{-1},

    \begin{equation*} \mathbf R_j: = \mathbf e_2\otimes\mathbf e_2+(1-j^{-2})(\mathbf e_1\otimes\mathbf e_1+\mathbf e_3\otimes\mathbf e_3)+j^{-1}\sqrt{2 -j^{-2}}\,(-\mathbf e_3\otimes\mathbf e_1+\mathbf e_1\otimes\mathbf e_3) \end{equation*}

    and set {\mathbf{y}}_j: = \mathbf R_j{{\bf x}}+j^{-1}\sqrt{2 -j^{-2}}\mathbf e_3 . It is readily seen that

    \begin{equation} \begin{array}{ll} y_{j,3}& = -j^{-1}\sqrt{2 -j^{-2}}x_1+(1-j^{-2})x_3+j^{-1}\sqrt{2 -j^{-2}} \\ &\\ & \ge (1-j^{-2})x_3\ge x_3-j^{-1}x_3\\ \end{array} \end{equation} (4.45)

    hence {\mathbf{y}}_j\in \mathcal A_j and by taking (4.2) into account we get ({\mathbf{y}}_j-\mathbf R_j{{\bf x}}- \mathbf c_j)_{\alpha} \equiv 0, \ \alpha = 1, 2 . Therefore bearing in mind that \mathbf R_j^T\to \mathbf I we have

    {\mathbf{u}}_j = j\mathbf R_j^T((y_{j,3}-x_3)\mathbf e_3) = \mathbf R_j^T\left\{(\sqrt{2 -j^{-2}}(1-x_1)+j^{-1}x_3)\mathbf e_3\right\}\to {\mathbf{u}}: = \sqrt 2(1-x_1)\mathbf e_3

    and by Lemma 3.8 we get \mathbf R{\mathbf{u}} = {\mathbf{u}} for every \mathbf R\in \mathcal S_{\mathcal L, E}, hence

    \widetilde{\mathcal G}({\mathbf{u}})\ge -\max\limits_{\mathbf R\in \mathcal S_{\mathcal L, E}}\mathcal L(\mathbf R{\mathbf{u}}) = -\mathcal L({\mathbf{u}}) = \pi\sqrt 2.

    On the other hand by taking account of

    y_{j,1}-x_1 = -j^{-2}x_1+j^{-1}\sqrt{2-j^{-2}}x_3,\ y_{j,2}-x_2 = 0

    and

    y_{j,3}-x_3 = j^{-1}\sqrt{2 -j^{-2}}(1-x_1)+j^{-2}x_3

    it is straightforward checking that

    \mathcal G_j({\mathbf{y}}_j) = -j\mathcal L({\mathbf{y}}_j-{{\bf x}}) = \pi\sqrt{2-j^{-2}}+\pi j^{-1}-\frac{\pi}{2}\sqrt{2-j^{-2}}\to \frac{\pi\sqrt 2}{2} < \widetilde{\mathcal G}({\mathbf{u}})

    thus proving that the claim of Lemma 4.4 fails in this case.

    Lemma 4.6. (Upper bound) Assume (2.13), (2.19)–(2.22), (2.32), (2.33), \mathcal L(\mathbf e_3)\! < \!0 and let 0 < h_j\to 0^+ as j\to +\infty . For every {\mathbf{u}}\in C^1(\overline\Omega, \mathbb R^3) there exists \widetilde{{\mathbf{y}}}_j\in C^1(\overline\Omega, \mathbb R^3) such that

    \limsup\limits_{j\to +\infty} {\mathcal G}_j(\widetilde{\mathbf y}_j) \le \widetilde{\mathcal G}({\mathbf{u}}).

    Proof. We assume without loss of generality that {\mathbf{u}}\in \mathcal A and let

    \mathbf b^*\in {\mathop{{{\rm{argmin}}}}\nolimits} \left\{\int_ \Omega \mathcal Q({{\bf x}}, \mathbb E({\mathbf{u}})+\frac{1}{2}b_{\alpha}(\mathbf e_{\alpha}\otimes\mathbf e_3+\mathbf e_{3}\otimes\mathbf e_\alpha))\,d{{\bf x}}: \mathbf b\in \mathbb R^2\right\},
    \begin{equation} \widetilde{\mathbf{u}}({{\bf x}}): = {\mathbf{u}}({{\bf x}})+x_3(b^*_1\mathbf e_1+b^*_2\mathbf e_2). \end{equation} (4.46)

    It is readily seen that \widetilde{\mathbf{u}}\in \mathcal A , that \mathbb E(\widetilde{\mathbf{u}}) = \mathbb E({\mathbf{u}})+\frac{1}{2}b_{\alpha}^*(\mathbf e_{\alpha}\otimes\mathbf e_3+\mathbf e_{3}\otimes\mathbf e_\alpha) hence

    \begin{equation} \mathcal I({\mathbf{u}}) = \int_ \Omega \mathcal Q({{\bf x}}, \mathbb E(\widetilde{\mathbf{u}}))\,d{{\bf x}}. \end{equation} (4.47)

    Moreover, by (3.1) of Lemma 3.1 and Remark 3.2 we obtain

    \begin{equation} \mathcal L(\mathbf R \widetilde{\mathbf{u}}) = \mathcal L(\mathbf R{\mathbf{u}})+\mathcal L(x_3(b^*_1\mathbf R\mathbf e_1+b^*_2\mathbf R\mathbf e_2)) = \mathcal L(\mathbf R{\mathbf{u}}), \qquad \forall\, \mathbf R\in \mathcal S_{\mathcal L, E}. \end{equation} (4.48)

    Therefore by choosing

    \widetilde{\mathbf R}\in {\mathop{{{\rm{argmin}}}}\nolimits}\left\{ -\mathcal L(\mathbf R\widetilde{\mathbf{u}}): \mathbf R\in \mathcal S_{\mathcal L, E}\right\}

    we get

    \widetilde{\mathcal G}({\mathbf{u}}) = \int_ \Omega \mathcal Q({{\bf x}},\mathbb E(\widetilde{\mathbf{u}}))\,d{{\bf x}}-\mathcal L(\widetilde{\mathbf R}\widetilde{\mathbf{u}}).

    By setting \widetilde{\mathbf y}_j: = \widetilde{\mathbf R}({{\bf x}}+h_j\widetilde{\mathbf{u}}) , taking account \mathcal S_{\mathcal L, E}\subset \{\mathbf R: \mathbf R\mathbf e_3 = \mathbf e_3\} and Lemma 3.8, we get \mathcal L(\widetilde{\mathbf R}{{\bf x}}-{{\bf x}}) = 0 and for q.e. {\mathbf{x}}\in E

    \widetilde y_{j,3}^* = x_3+h_j\widetilde u_{3}\ge (1-h_j)x_3.

    Therefore \widetilde{{\mathbf{y}}}_j\in \mathcal A_j and by (2.24) we get

    \limsup\limits_{j\to+\infty} \mathcal |\mathcal G_{h_j}(\mathbf y_j)-\widetilde{\mathcal G}({\mathbf{u}})| \le \limsup\limits_{j\to+\infty}\int_{\Omega} \left|\frac{1}{h_j^2}\mathcal W({{\bf x}},\mathbf I+h_j\nabla\widetilde{\mathbf{u}})- \mathcal Q({{\bf x}},\mathbb E(\widetilde{\mathbf{uu}}))\right |\,d{{\bf x}} = 0

    which proves the lemma.

    We are now in a position to prove our main theorem.

    Proof of Theorem 2.4. If (\overline{\mathbf y}_j)_{j\in\mathbb N}\subset H^1(\Omega, \mathbb R^3) is a minimizing sequence for \mathcal G_j then \mathcal G_j(\overline{\mathbf y}_j)\le \mathcal G_j(\mathbf x) = 0, moreover if \mathbf R_j belong \mathcal A(\overline{\mathbf y}_j) and \overline{\mathbf c}_j is defined by (4.2) and (4.3), then Lemma 4.1 entails that the sequence

    \begin{equation*} {\overline{\mathbf{u}}}_j({{\bf x}}): = {h_j}^{-1}\mathbf R_j^T\left\{\big({\overline{\mathbf{y}}}_j\,-\,\mathbf R_j{\mathbf{x}}\,-{\overline{\mathbf c}}_j\,\big)_{\alpha}\mathbf e_{\alpha}+({\overline y}_{j,3}-x_3)\mathbf e_3\right\} \end{equation*}

    is bounded in H^1(\Omega; \mathbb R^3) . Therefore up to subsequences {\overline{\mathbf{u}}}_j\to \overline{\mathbf{u}} weakly in H^1(\Omega; {{\mathbb R}}^3), so, by Lemma 4.4, we have \overline{\mathbf{u}}\in \mathcal A and

    \begin{equation*} \label{liminf} \liminf\limits_{j\to +\infty}{\mathcal G}_j({\overline{\mathbf y}}_j)\ge \widetilde{\mathcal G}(\overline{\mathbf{u}}). \end{equation*}

    On the other hand, by Lemma 4.6, for every {\mathbf{u}}\in C^1(\overline\Omega, \mathbb R^3)\cap\mathcal A there exists a sequence \mathbf y_j\in C^1(\overline\Omega, \mathbb R^3) such that

    \limsup\limits_{j\to +\infty} {\mathcal G}_j(\mathbf y_j) \le \widetilde{\mathcal G}({\mathbf{u}}).

    Since

    \begin{equation} {\mathcal G}_j(\overline{\mathbf{y}}_j)+o(1) = \inf\limits_{H^1( \Omega,\mathbb R^3)}\mathcal G_{j}\le {\mathcal G}_j(\mathbf y_j),\qquad \mbox{as } j\to+\infty,\end{equation} (4.49)

    by passing to the limit as j\to +\infty , we get

    \begin{equation} \widetilde{\mathcal G}(\overline{\mathbf{u}})\le\widetilde {\mathcal G}({\mathbf{u}}), \qquad \forall {\mathbf{u}}\in C^1(\overline\Omega,\mathbb R^3)\cap\mathcal A. \end{equation} (4.50)

    Now fix a generic {\mathbf{u}}\!\in\! \mathcal A and denote again by {\mathbf{u}} a Sobolev extension of {\mathbf{u}} to the whole \mathbb R^3 . We claim that there exists {\mathbf{u}}_j\in C^1(\overline\Omega, \mathbb R^3)\cap\mathcal A such that {\mathbf{u}}_j\to {\mathbf{u}} in H^1(\Omega; \mathbb R^3) : Indeed, since \widetilde u_3({\mathbf{x}})+x_3\ge 0 for q.e. {\mathbf{x}}\in E , by Lemma A.4 it is enough to choose u_{3, j}: = \eta_j-x_3 where \eta_j\in C^1(\mathbb R^3), \ \eta_j\ge 0\ \text{q.e. in}\ E, \ \eta_j \to u_3+x_3 in H^1(\mathbb R^3) (here u_3+x_3 denotes also an extension to the whole H^1(\mathbb R^3) ) and u_{\alpha, j}: = u_\alpha*\rho_j, \ \alpha = 1, 2 where \rho_j is a sequence of smooth mollifiers. By (4.50) we have

    \widetilde{\mathcal G}(\overline{\mathbf{u}})\le \widetilde{\mathcal G}({\mathbf{u}}_j)

    whence by Remark 2.2,

    \widetilde{\mathcal G}(\overline{\mathbf{u}})\le \lim\limits_{j\to +\infty}\widetilde{\mathcal G}({\mathbf{u}}_j) = \widetilde{ \mathcal G}({\mathbf{u}}), \qquad \forall\, {\mathbf{u}}\in \mathcal A,

    that is \overline{\mathbf{u}}\in{\mathop{{{\rm{argmin}}}}\nolimits} \widetilde{\mathcal G} .

    We show that {\mathcal G}_j(\mathbf y_j)\to \widetilde{\mathcal G}(\overline{\mathbf{u}}) : By Lemma A.4 in the Appendix, for every {\varepsilon} > 0 there is \overline{\mathbf{u}}_{\varepsilon}\in C^1(\overline \Omega; \mathbb R^3))\cap\mathcal A such that

    \begin{equation*} \widetilde{\mathcal G}(\overline{\mathbf{u}}_{\varepsilon}) < \widetilde{\mathcal G}(\overline{\mathbf{u}})+{\varepsilon} \end{equation*}

    and by Lemma 4.6 there exists \mathbf y_{j, {\varepsilon}}\in C^1(\overline \Omega; \mathbb R^3) such that by taking account of (4.49) we have

    \limsup\limits_{j\to +\infty} {\mathcal G}_j(\overline{\mathbf y}_j) \le \limsup\limits_{j\to +\infty} {\mathcal G}_j(\mathbf y_{j,{\varepsilon}}) \le \widetilde{\mathcal G}({\mathbf{u}}_{\varepsilon}) < \widetilde{\mathcal G}(\overline{\mathbf{u}})+{\varepsilon}

    for every {\varepsilon} > 0 . Since by Lemma 4.4,

    \liminf\limits_{j\to +\infty}{\mathcal G}_j(\overline{\mathbf y}_j)\ge \widetilde{\mathcal G}(\overline{\mathbf{u}}),

    we get {\mathcal G}_j(\mathbf y_j)\to \widetilde{\mathcal G}(\overline{\mathbf{u}}) as claimed.

    We are only left to show that \min\widetilde{\mathcal G} = \min\mathcal G. To this aim we show first that for every {\mathbf{u}}\in \mathcal A there exists {\mathbf{u}}_*\in \mathcal A such that \mathcal G({\mathbf{u}}_*) = \widetilde{\mathcal G}({\mathbf{u}}) . Indeed if \widetilde{\mathbf{u}} is defined as in (4.46) then by (4.47) and (4.48) we get

    \begin{equation} \widetilde{\mathcal G}({\mathbf{u}}) = \mathcal I({\mathbf{u}})-\max\limits_{\mathbf R\in \mathcal S_{\mathcal L, E}}\mathcal L(\mathbf R{\mathbf{u}}) = \int_ \Omega \mathcal Q({{\bf x}},\mathbb E(\widetilde{\mathbf{u}})\,d{{\bf x}}-\max\limits_{\mathbf R\in \mathcal S_{\mathcal L, E}}\mathcal L(\mathbf R\widetilde{\mathbf{u}}) = \mathcal G(\widetilde{\mathbf{u}}) \end{equation} (4.51)

    as claimed. By recalling that \widetilde{\mathcal G}(\overline{\mathbf{u}}) = \min \widetilde{\mathcal G}\le \inf \mathcal G let us assume that inequality is strict. Then by (4.51) there exists {\overline{\mathbf{u}}}_*\in \mathcal A such that \mathcal G({\overline{\mathbf{u}}}_*) = \widetilde{\mathcal G}(\overline{\mathbf{u}}) < \inf \mathcal G , a contradiction. Thus again by (4.51) \mathcal G({\overline{\mathbf{u}}}_*) = \widetilde{\mathcal G}(\overline{\mathbf{u}}) = \min \mathcal G .

    In this section we will exhibit a choice of energy density \mathcal W , open set \Omega , dead loads \mathbf f, \mathbf g and set E\!\subset\! \overline \Omega fulfilling all the assumptions of Theorem 2.4 but such that the minimum of the limit functional \mathcal G is strictly less than the minimum of the Signorini functional (see [45] for a counterexample exhibiting an analogous gap for unconstrained pure traction problem).

    We shall consider the energy density already defined in (2.26) by setting

    \begin{equation} \mathcal{W}(\mathbf{F}): = \left\{\begin{array}{ll} \mathcal{W}_{ \rm{iso}}\left(\frac{\mathbf{F}}{(\det\mathbf{F})^{1/3}}\right)+\mathcal{W}_{ \rm{vol}}(\mathbf{F}),\qquad&\mbox{if } \det\mathbf F > 0, \\ +\infty,\qquad&\mbox{if } \det\mathbf F\le 0 , \end{array}\right. \end{equation} (5.1)

    where {\mathcal W}_{ \rm{iso}} is the energy density of Yeoh type defined in (2.27) with 2c_1 = \mu > 0 and \mathcal W_{ \rm{vol}}(\mathbf F) = g(\det\mathbf F) where g:\mathbb R_+\to\mathbb R is the convex C^2 function (satisfying (2.28) with r = 2 ) defined by

    \begin{equation*} g(t) = \frac{\mu}{6}(t^2-1-2\log t). \end{equation*}

    By recalling Example 2.1 it is readily seen that \mathcal W satisfies (2.14)–(2.17) and by taking into account of

    \det(\mathbf I+ h\mathbf B) = 1+h\, {{\rm{Tr}}} \mathbf B+(h^2/2)\big( ( {{\rm{Tr}}} \mathbf B)^2- {{\rm{Tr}}} \,\mathbf B^2\big) + h^3\det\mathbf B

    and {{\rm{Tr}}} (\mathbf B^T \mathbf B) \! = \! |\mathbf B|^2 for every \mathbf B\in \mathbb R^{3\times3} , we obtain as h\to 0

    \begin{equation*} \label{taylor} \frac{|\mathbf I+ h \mathbf B|^2}{\det (\mathbf I+h \mathbf B)^{2/3}}-3\, = \, h^2 \big(\,2\,|\mathbf B|^2-\tfrac{2}{3}| {{\rm{Tr}}} \mathbf B|^2\,\big)+o(h^2) \end{equation*}

    for every \mathbf B\in \mathbb R^{3\times3}_{\mathrm{sym}} . Moreover by recalling (2.27) (with 2c_1 = \mu )

    \begin{equation*} \mathcal W_{ \rm{vol}}(\mathbf I+ h \mathbf B) = g(\det(\mathbf I+ h\mathbf B)) = \frac{h^2}{2}| {{\rm{Tr}}} \mathbf B|^2+ o(h^2) = \frac{\mu}{3}\,| {{\rm{Tr}}} \mathbf B|^2\,h^2\,+ \,o(h^2) , \end{equation*}
    \begin{equation*} \mathcal W_{ \rm{iso}}(\mathbf I+ h \mathbf B)\, = \, \frac{\mu}{2}\,h^2\left( \,2\,|\mathbf B|^2-\frac 2 3 ( {{\rm{Tr}}} \mathbf B)^2 \,\right) \,+ \,o(h^2), \end{equation*}

    so

    \begin{equation} \frac{1}{2}\mathbf B\,D^2\mathcal W(\mathbf I)\,\mathbf B\, = \,\mu\,|\mathbf B|^2. \end{equation} (5.2)

    Let

    \begin{equation} \Omega: = \{{{\bf x}}\in\mathbb R^3: x_1^2+x_2^2 < 1,\,0 < x_3 < 1\}, \quad E: = \{{{\bf x}}\in\mathbb R^3: x_1^2+x_2^2 < 1,\ x_3 = 0\}\\ \end{equation} (5.3)

    and \varphi\in C^2(\overline E) such that

    \begin{equation} \begin{array}{ll} & \Delta\varphi\not\equiv 0\,,\quad \varphi(x_1,x_2) = \phi(r)\,,\quad r: = \sqrt{x_1^2+x_2^2}\,, \quad \phi(1)\! = \!\phi'(1)\! = \! \int_0^1 r^2\phi'(r)\,dr = 0\\ \end{array} \end{equation} (5.4)

    (for instance \phi(r): = 1-6r^2+9r^4-4r^6 fulfills (5.4)). It is readily seen that condition (2.13) is fulfilled and that E_{ess} = \overline E . We define

    \begin{equation} \begin{array}{ll} \mathbf R_*&: = -\mathbf e_1\otimes \mathbf e_2+ \mathbf e_2\otimes \mathbf e_1+\mathbf e_3\otimes \mathbf e_3,\\ &\\ \mathcal L({\mathbf{u}})\!&: = \!\int_\Omega u_{\alpha}\,\nabla_{\alpha}\varphi \,d{{\bf x}}- \int_E u_3(x_1,x_2,0)\,dx_1\,dx_2. \\ \end{array} \end{equation} (5.5)

    Condition (2.32) is satisfied since

    \mathcal L((\mathbf R{{\bf x}}-{{\bf x}})_{\alpha}\mathbf e_\alpha) = \pi\,(R_{11}+R_{22}-2)\int_0^1r^2\phi'(r)\,dr = 0,\qquad \forall\, \mathbf R\in SO(3).

    Moreover \mathcal L(\mathbf e_1) = \mathcal L(\mathbf e_2) = 0, \ \mathcal L(\mathbf e_3) < 0 and

    \begin{equation} \begin{array}{ll} \Phi(\mathbf R, E, \mathcal L)& = \pi\,(R_{11}+R_{22}-2)\int_0^1r^2\phi'(r)\,dr+ \,\pi\,\min\limits_{E_{ess}}\big\{R_{31}x_1+R_{32}x_2+(R_{33}-1)x_3\big\}\\ &\\ & = -\pi\sqrt{R_{31}^2+R_{32}^2}\le 0.\\ \end{array} \end{equation} (5.6)

    so (2.33) is fulfilled too. By taking account of \mathbf R_*\mathbf e_3 = \mathbf e_3 , we get

    \Phi(\mathbf R_*,E, \mathcal L) = -2\pi\int_0^1r^2\phi'(r)\,dr = 0

    whence \mathbf R_*\in \mathcal S_{\mathcal L, E} and Lemma 3.8 entails \mathcal S_{\mathcal L, E} = \{\mathbf R: \mathbf R\mathbf e_3 = \mathbf e_3\} . Since E fulfills (2.13), \mathcal W satisfies (2.14)–(2.17) and \mathcal L satisfies (2.27) together with \mathcal L(\mathbf e_3) < 0 then, by taking into account of (5.2), Theorem 2.4 entails

    \begin{equation} \inf\mathcal G_j\to \min\limits_{{\mathbf{u}}\in\mathcal A} \mathcal G = \min\left\{\mu\int_ \Omega|\mathbb E({\mathbf{u}})|^2\,d{\mathbf{x}}-\max\limits_{\mathbf R\in \mathcal S_{\mathcal L, E}}\mathcal L(\mathbf R{\mathbf{u}}) : {\mathbf{u}}\in\mathcal A\right\}. \end{equation} (5.7)

    We set

    \begin{equation} \mathcal E({\mathbf{uu}}): = \mu\int_ \Omega|\mathbb E({\mathbf{u}})|^2\,d{\mathbf{x}}-\mathcal L({\mathbf{u}}) \end{equation} (5.8)

    and, for every \mathbf R\in \mathcal S_{\mathcal L, E},

    \begin{equation} \mathcal E_{\mathbf R}({\mathbf{uu}}): = \mu\int_ \Omega|\mathbb E({\mathbf{u}})|^2\,d{\mathbf{x}}-\mathcal L(\mathbf R{\mathbf{u}}). \end{equation} (5.9)

    We aim to show

    \begin{equation} \min\{\mathcal E_{\mathbf R_*}({\mathbf{uu}}): {\mathbf{u}}\in \mathcal A\} < \min\{\mathcal E({\mathbf{uu}}): {\mathbf{u}}\in \mathcal A\} \end{equation} (5.10)

    so that, once (5.10) is proved, we deduce

    \begin{equation} \min\limits_{{\mathbf{u}}\in\mathcal A} \mathcal G < \min\limits_{{\mathbf{u}}\in\mathcal A} \mathcal E. \end{equation} (5.11)

    In order to show inequality (5.10) we need some properties of minimizers of \mathcal E which have been essentially proven in [45]. In the following \overline{\mathbb E}(\cdot) will denote the upper-left 2\times2 submatrix of \mathbb E(\cdot) and \overline{\mathbf R}\in SO(2) the upper-left 2\times2 submatrix of any \mathbf R\in \mathcal S_{\mathcal L, E} .

    Lemma 5.1. Let {\mathbf{u}}\in \mathcal A and let

    \begin{equation} {\mathbf{v}}({{\bf x}})\!: = v_\alpha(x_1,x_2)\mathbf e_\alpha+v_3(x_3)\mathbf e_3,\ \alpha\! = \!1,2 \end{equation} (5.12)

    where

    v_\alpha(x_1,x_2): = \int_0^1\!u_\alpha({\mathbf{x}}) \,dx_3,\ \alpha\! = \!1,2,\quad v_3(x_3): = \pi^{-1}\!\!\!\int_{E} \!u_3({\mathbf{x}}) dx_1dx_2.

    Then {\mathbf{v}}\in \mathcal A and

    \begin{equation} {\mathcal E}_{\mathbf R}({\mathbf{u}})\ge \mathcal J_{\overline{\mathbf R}}(\overline{\mathbf{v}}),\qquad \forall\, \mathbf R\in. \mathcal S_{\mathcal L, E}, \end{equation} (5.13)

    where \overline{\mathbf{v}}: = v_\alpha\mathbf e_\alpha , and

    \begin{equation} {\mathcal J}_{\overline{\mathbf R}}(\overline{\mathbf{v}}): = \mu\int_E|\overline{\mathbb E}(\overline{\mathbf{v}})|^2\,dx_1dx_2-\int_E\overline{\mathbf R}^{T}\nabla\varphi\cdot \overline{\mathbf{v}}\ \,dx_1dx_2. \end{equation} (5.14)

    In particular if \mathbf R = \mathbf I , then (5.13) reduces to \mathcal E({\mathbf{u}})\ge \mathcal J(\overline{\mathbf{v}}) having set \mathcal J: = \mathcal J_{\overline{\mathbf I}} .

    Proof. Since u_{3}^*\ge 0 q.e. on E = \partial \Omega\cap \{x_3 = 0\} then by Remark 2.3 we get u_3\ge 0\ \ \mathcal H^2 - q.e. in E that is v_3(0)\ge 0 hence, again by Remark 2.3, {\mathbf{v}}\in \mathcal A . Moreover, by using the notation u_{3, 3}: = \partial_3u_3 , Jensen inequality entails

    \begin{equation} \begin{array}{ll} {\mathcal E}_{\mathbf R}({\mathbf{u}}) \!\!&\ge\ \mu \int_E\left|\int_0^1\overline{\mathbb E}({\mathbf{u}})\,dx_3\right|^2\,dx_1\,dx_2 \,+\,\mu\,\pi\int_0^1\left|\frac1\pi\int_E u_{3,3}\,dx_1\,dx_2\right|^2\,dx_3\\ &\\ & \ \ \ \ -\int_E {\overline R}_{\beta\alpha}\nabla_\beta \varphi\left(\int_0^1u_\alpha\,dz\right)\,dx_1\,dx_2\,+\int_E u_3(x_1,x_2,0)\,dx_1\,dx_2\\ &\\ &\ge \ \ \mathcal J_{\overline{\mathbf R}}(\overline v)+\mu\,\pi\int_0^1|\dot v_3|{\,^2}\,dx_3+\pi v_3(0)\ge \mathcal J_{\overline{\mathbf R}}(\overline v), \end{array} \end{equation} (5.15)

    thus proving the lemma.

    We need now the following characterization of minimizers of \mathcal J which has been given in [45].

    Lemma 5.2. There exists \overline{\Phi}\in H^2(E) such that

    \begin{equation} \min\limits_{{\mathbf{u}}\in H^1(E)} \mathcal J({\mathbf{u}}) = \mathcal J(\nabla\overline{\Phi})\ge \min\limits_{\Phi\in H^2(E)}\int_E(2\,\mu\,\Phi_{,12}^2+\mu\,\Phi_{,11}^2+\mu\,\Phi_{,22}^2+\Phi\Delta\varphi)\,dx_1dx_2, \end{equation} (5.16)

    where we have used the notation \Phi_{, \alpha\beta}: = \partial^2_{\alpha\beta}\Phi .

    A straightforward application of Lemma 5.1 (with \mathbf R = \mathbf I ) and Lemma 5.2 yields the following precise calculation of the energy level of {\mathbf{u}}\in {\mathop{{{\rm{argmin}}}}\nolimits}_{\mathcal A}\mathcal E .

    Lemma 5.3. There holds

    \begin{equation} \min\limits_{{\mathbf{u}}\in \mathcal A}\mathcal E({\mathbf{u}}) = \min\limits_{\Phi\in H^2(E)}\int_E(2\mu\Phi_{,12}^2+\mu\Phi_{,11}^2+ \mu\Phi_{,22}^2+\Phi\Delta\varphi)\,dx_1dx_2. \end{equation} (5.17)

    Proof. It is readily seen that any displacement of the kind (\nabla\Phi(x_1, x_2), v_3(x_3))\in \mathcal A if and only if \Phi\in H^2(E), \ v_3\in H^1(0, 1) and v_3(0)\ge 0 . Therefore, by Lemmas 5.1 and 5.2, we get

    \begin{equation*} \begin{array}{ll} \min\limits_{{\mathbf{u}}\in \mathcal A}\mathcal E({\mathbf{u}})&\ge \min\limits_{\Phi\in H^2(E)}\int_E(2\mu\Phi_{,12}^2+\mu\Phi_{,11}^2+ \mu\Phi_{,22}^2+\Phi\Delta\varphi)\,dx_1dx_2 \\ &\\ & +\inf\left\{\mu\,\pi\!\int_0^1 \! |\dot v_3|^2\,dx_3+\pi \,v_3(0)\,: v_3\in H^1(0,1),\ v_3(0)\ge 0\right\}\\ &\\ & = \min\limits_{\Phi\in H^2(E)}\int_E(2\mu\Phi_{,12}^2+\mu\Phi_{,11}^2+ \mu\Phi_{,22}^2+\Phi\Delta\varphi)\,dx_1dx_2. \end{array} \end{equation*}

    The opposite inequality follows by choosing {\mathbf{v}}: = (\nabla\Phi, 0) with \Phi\in H^2(B) and by taking into account of

    \min\limits_{{\mathbf{u}}\in \mathcal A}\mathcal E({\mathbf{u}})\le \mathcal E({\mathbf{v}})

    for every choice of \Phi\in H^2(E) . Let now \Phi\in H^2(E) . Then {\mathbf{v}}: = \Phi_{, 2}\mathbf e_1 -\Phi_{, 1}\mathbf e_2\in \mathcal A and a direct computation shows that

    \begin{equation} \min\limits_{\mathcal A}{\mathcal E}_{{\mathbf R}_*}\le \min\limits_{\Phi\in H^2(E)} \int_{E}(2\mu{\Phi_{,12}}^2+\frac{\mu}{2}(\Phi_{,22}-\Phi_{,11})^2+ \Phi \Delta\varphi)\,dx_1dx_2. \end{equation} (5.18)

    Therefore inequality (5.10) is an immediate consequence of the next proposition.

    Proposition 5.4. There holds

    \begin{equation} \begin{array}{ll} & \min\limits_{\Phi\in H^2(E)} \int_{E}(2\mu{\Phi_{,12}}^2+\frac{\mu}{2}(\Phi_{,22}-\Phi_{,11})^2+ \Phi \Delta\varphi) \,dx_1dx_2 \\ & < \min\limits_{\Phi\in H^2(E)}\int_E(2\mu\Phi_{,12}^2+\mu\Phi_{,11}^2+ \mu\Phi_{,22}^2+\Phi\Delta\varphi)\,dx_1dx_2.\end{array} \end{equation} (5.19)

    Proof. The proof is the same of formula (5.14) of [45].

    The previous explicit example shows that a gap phenomenon may actually develop. Nevertheless one can prove that whenever \mathbf f, \ \mathbf g satisfy (2.33) then they can be suitable rotated in order to avoid the gap. In order to state such result, we introduce suitable notation: Set

    \begin{equation} \mathcal L_{\mathbf R}({\mathbf{v}}): = \mathcal L(\mathbf R{\mathbf{v}}) = \int_ \Omega\mathbf R^T\mathbf f\cdot{\mathbf{v}}\,d{{\bf x}}+\int_{\partial \Omega}\mathbf R^T\mathbf g\cdot{\mathbf{v}}\,d\mathcal H^2, \qquad \forall\, \mathbf R\in SO(3) , \end{equation} (5.20)

    say \mathcal L_{\mathbf R} is the load functional associated to the external forces \mathbf R^T\mathbf f, \mathbf R^T\mathbf g and let \mathcal E_{\mathbf R} be the functional defined by replacing \mathcal L with \mathcal L_{\mathbf R} in the definition of \mathcal E .

    Theorem 5.5. Assume (2.13), (2.33), \mathcal L(\mathbf e_3) < 0 and \mathbf R\in \mathcal S_{\mathcal L, E} . Then the functional \mathcal L_{\mathbf R} fulfills (2.33) and \mathcal S_{\mathcal L, E}\equiv\mathcal S_{\mathcal L_{\mathbf R}, E} . Moreover, if {\mathbf{u}} minimizes \mathcal G over H^1 \Omega, {{\mathbb R}}^3), \mathbf R\in\mathcal S_{\mathcal L, E} attains the maximum in definition (1.12) of \mathcal G({\mathbf{u}}) then {\mathbf{u}}\in {\mathop{{{\rm{argmin}}}}\nolimits} {\mathcal E}_{\mathbf R} and

    \begin{equation} \min\limits_{H^1( \Omega,{{\mathbb R}}^3)}\mathcal G\ \, = \,\min\limits_{H^1( \Omega,{{\mathbb R}}^3)}\mathcal E_{\mathbf R}. \end{equation} (5.21)

    Proof. By Lemma 3.8 we have either \mathcal S_{\mathcal L, E}\equiv \{\mathbf I\} or \mathcal S_{\mathcal L, E} = \{ \mathbf R: \mathbf R\mathbf e_3 = \mathbf e_3\} : In the first case there is nothing to prove, in the second one by (2.33) we get

    \begin{equation} 0 = \Phi(\mathbf R, E, \mathcal L) = \mathcal L((\mathbf R-\mathbf I){\mathbf{x}}). \end{equation} (5.22)

    Therefore for any other \mathbf S\in SO(3) by taking account of \mathbf R\mathbf e_3 = \mathbf e_3 and (5.22) we have

    \begin{equation} \Phi(\mathbf S, E, \mathcal L_{\mathbf R}) = \Phi({\mathbf R}\mathbf S, E, \mathcal L)\le 0 \end{equation} (5.23)

    that is \mathcal L_{\mathbf R} satisfies (2.33). By Remark 3.5 conditions (4.9)–(4.11) of Theorem 4.5 in [12] are fulfilled hence \mathcal E_{\mathbf R} achieves a finite minimum. Moreover since \mathbf R\mathbf e_3 = \mathbf e_3 implies \mathbf R^2\mathbf e_3 = \mathbf e_3 , (5.23) together with Lemma 3.8 entails

    \begin{equation} \Phi(\mathbf R, E, \mathcal L_{\mathbf R}) = \Phi({\mathbf R}^2, E, \mathcal L) = 0, \end{equation} (5.24)

    whence \mathbf R\in \mathcal S_{\mathcal L_{\mathbf R}, E} whenever \mathbf R\in \mathcal S_{\mathcal L, E} . Then since \mathcal L_{\mathbf R}(\mathbf e_3) = \mathcal L(\mathbf R\mathbf e_3) = \mathcal L(\mathbf e_3) < 0 we get, again by Lemma 3.8, \mathcal S_{\mathcal L_{\mathbf R}, E}\not\equiv \{\mathbf I\} hence \mathcal S_{\mathcal L_{\mathbf R}, E} = \{ \mathbf R: \mathbf R\mathbf e_3 = \mathbf e_3\} = \mathcal S_{\mathcal L, E} as claimed.

    We conclude by checking that if {\mathbf{u}} minimizes \mathcal G then it is also a minimizer of \mathcal E_{\mathbf R} over H^1(\Omega, {{\mathbb R}}^3) . If {\mathbf{u}}\in H^1(\Omega, \mathbb R^3) minimizes \mathcal G and \mathbf R attains the maximum then

    \begin{aligned}\min\limits_{H^1( \Omega,\mathbb R^3)}\mathcal G& = \mathcal G({\mathbf{u}}) = \int_ \Omega\mathcal Q({{\bf x}},\mathbb E({\mathbf{u}}))\,d{{\bf x}}-\mathcal L({\mathbf R} {\mathbf{u}}) = \mathcal E_{\mathbf R}({\mathbf{u}}). \end{aligned}

    Thus since \mathcal G\le \mathcal E_{\mathbf R} then (5.21) is proven.

    Remark 5.6. By choosing \mathcal W as in (5.1), \Omega, E as in (5.3) we provide an example where the inclusion {\mathop{{{\rm{argmin}}}}\nolimits} \mathcal G \subset {\mathop{{{\rm{argmin}}}}\nolimits} \widetilde{\mathcal G} is strict. Indeed Lemma 5.1 shows that for every \mathbf R\in \mathcal S_{\mathcal L, E} there exists \mathbf w\in{\mathop{{{\rm{argmin}}}}\nolimits}\mathcal E_{\mathbf R} such that \mathbf w({{\bf x}}) = w_{\alpha}(x_1, x_2)\mathbf e_{\alpha} . By Theorem 5.5 there exists {\mathbf{u}}\in {\mathop{{{\rm{argmin}}}}\nolimits} \mathcal G\subset {\mathop{{{\rm{argmin}}}}\nolimits} \widetilde{\mathcal G} such that {\mathbf{u}}({{\bf x}}) = u_{\alpha}(x_1, x_2)\mathbf e_{\alpha} . Then we can set {\mathbf{u}}_*({{\bf x}}): = {\mathbf{u}}({{\bf x}})+x_3\mathbf e_1 . By taking account of Lemma 3.1 and Remark 3.2, we get \mathcal L(\mathbf R{\mathbf{u}}) = \mathcal L(\mathbf R{\mathbf{u}}_*) for every \mathbf R\in \mathcal S_{\mathcal L, E} hence \widetilde {\mathcal G}({\mathbf{u}}) = \widetilde {\mathcal G}({\mathbf{u}}_*) and {\mathbf{u}}_*\in {\mathop{{{\rm{argmin}}}}\nolimits} \widetilde{\mathcal G} . Moreover, again by taking account of Lemma 3.1 and Remark 3.2, we have

    \mathcal G({\mathbf{u}}_*)- \mathcal G({\mathbf{u}}) = \mu\int_ \Omega |\mathbb E({\mathbf{u}}_*)|^2\,d{{\bf x}}-\mu\int_ \Omega |\mathbb E({\mathbf{u}})|^2\,d{{\bf x}} = \frac{\mu}{2}| \Omega| > 0,

    thus {\mathbf{u}}_*\not\in {\mathop{{{\rm{argmin}}}}\nolimits} \mathcal G and the inclusion is strict in this case.

    We have showed a rigorous variational linearization for a classical obstacle problem in nonlinear elasticity, namely an elastic body subject to pure traction load, supported on a unilateral rigid plane. Under suitable geometric admissibility conditions on the loads we obtain coincidence of minima with the classical Signorini problem in linear elasticity. On the other hand, we have shown the existence of loads violating such admissibility condition and entailing a gap between the minimum of the two problems.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    Francesco Maddalena and Franco Tomarelli are members of the Unione Matematica Italiana and the Gruppo Nazionale per l'Analisi Matematica, la Probabilità e le loro Applicazioni (GNAMPA) of the Istituto Nazionale di Alta Matematica (INdAM). The authors are partially funded by INdAM–GNAMPA 2023 Project Codice CUP–E53C22001930001, "Modelli variazionali ed evolutivi per problemi di adesione e di contatto", Italian M.U.R. PRIN grant number 2022J4FYNJ "Variational Methods for Stationary and Evolution Problems with Singularities and Interfaces", INdAM–GNAMPA 2023 Project Codice CUP E53C22001930001, "Modelli variazionali ed evolutivi per problemi di adesione e di contatto".

    Research Project of National Relevance "Evolution problems involving interacting scales" granted by the Italian Ministry of Education, University and Research (MIUR Prin 2022, project code 2022M9BKBC, Grant No. CUP D53D23005880006).

    All authors declare no conflicts of interest in this paper.

    For reader's convenience and aiming to the precise formulation of unilateral constraint, in this section we encompass some results about capacity theory which are essential to achieve the results of present paper and somehow present in the literature, though they are spread in several different contexts and not easy to find as stated in this form: In particular Proposition A.1 and Eq (A.2) can be proven as like as Propositions 5.8.3 and 5.8.4 in [7] although the results seem slightly different.

    Proposition A.1. Let G an open bounded subset of {{\mathbb R}}^N . Then

    \begin{equation} {\mathop{{{\rm{cap}}}}\nolimits} G = \inf \left\{\|w\|_{H^1({{\mathbb R}}^N)}^2: w\!\in\! C^{\infty}_0(\mathbb R^N),\, w\ge 1\,{{on}}\ G\right\}. \end{equation} (A.1)

    The above property can be generalized to every bounded subset of {{\mathbb R}}^N by the following:

    Proposition A.2. Let E a bounded subset of {{\mathbb R}}^N . Then

    \begin{equation} \begin{array}{ll} {\mathop{{{\rm{cap}}}}\nolimits} E\!\!& = \,\inf \left\{\|w\|_{H^1({{\mathbb R}}^N)}^2: w\!\in\! C^{\infty}_0(\mathbb R^N),\, w\ge 1\,\;{{on \;a\; neighborhood\; of}}\, E\right\} \\ &\\ & = \, \inf \left\{\|w\|_{H^1({{\mathbb R}}^N)}^2: w\!\in\! C^{\infty}_0(\mathbb R^N; [0,1]),\, w\equiv 1\,{{on\; a \;neighborhood \;of}}\, E\right\}. \end{array} \end{equation} (A.2)

    We state and prove some results which play a crucial role in the proof of our main theorem.

    In the sequel \Omega will denote an open bounded subset of {{\mathbb R}}^N with Lipschitz boundary and E will denote a subset of \overline \Omega such that {\mathop{{{\rm{cap}}}}\nolimits} E > 0 .

    Lemma A.3. Let u\in H^1(\Omega), \ u\ge 0 a.e. in \Omega such that u^*({\mathbf{x}}) = 0 for q.e. {\mathbf{x}}\in D , where D is a closed subset of \overline \Omega . Then there is an extension v\in H^1({{\mathbb R}}^N) of u such that {\mathop{{{\rm{spt}}}}\nolimits} v is compact, v\ge 0\ {{a.e.\; in}}\ {{\mathbb R}}^N and

    \begin{equation*} \lim\limits_{r\to 0^+}\frac{1}{\left |B_r({\mathbf{x}})\right |}\int_{B_r({\mathbf{x}})}v({\boldsymbol{\xi}})\,d{\boldsymbol{\xi}} = 0 \end{equation*}

    for q.e. {\mathbf{x}}\in D .

    Proof. We recall that u^* is defined as

    \begin{equation} u^*({\mathbf{x}}) = \lim\limits_{r\to 0^+}\frac{1}{\left |B_r({\mathbf{x}})\right |}\int_{B_r({\mathbf{x}})}w({\boldsymbol{\xi}})\,d{\boldsymbol{\xi}} \end{equation} (A.3)

    for q.e. {\mathbf{x}}\in \overline \Omega , where w is any Sobolev extension of u . Therefore the claim follows easily by choosing a cut off function \varphi\in C^{\infty}_0({{\mathbb R}}^N), \ \varphi\equiv 1 on \Omega and by setting v: = w^+\varphi which is a Sobolev extension of u with compact support since v = u a.e. in \Omega , and {\mathop{{{\rm{spt}}}}\nolimits} v\subset {\mathop{{{\rm{spt}}}}\nolimits} \varphi .

    Lemma A.4. Let u\in H^1(\mathbb R^N) with compact support such that u\ge 0\ \mathit{\text{a.e. in}}\ {{\mathbb R}}^N and u^*({\mathbf{x}}) = 0 for q.e. {\mathbf{x}}\in E . Then there exists a sequence u_j\in C^1(\mathbb R^N)\cap H^1(\mathbb R^N), \ u_j\ge 0\ \mathit{\text{in}}\ {{\mathbb R}}^N such that u_j({\mathbf{x}}) = 0 for q.e. {\mathbf{x}}\in E and u_j\to u in H^1(\mathbb R^3) .

    Proof. For every j\in \mathbb N, \ j\ge 1 let \overline u_j: = \min\{u, j^{\frac{1}{4}}\}\in H^1(\mathbb R^N) . By Theorem 3.11.6 and Remark 3.11.7 of [57], there exists v_j\in C^1(\mathbb R^N)\cap H^1(\mathbb R^N) with {\mathop{{{\rm{spt}}}}\nolimits} v_j\subset \{{{\bf x}}: d({{\bf x}}, {\mathop{{{\rm{spt}}}}\nolimits} {\overline u}_j) \le j^{-1}\} such that if F_j: = \{{\mathbf{x}}: v_j({\mathbf{x}})\neq {\overline u}_j({\mathbf{x}})\} then

    \begin{equation} {\mathop{{{\rm{Cap}}}}\nolimits} F_j < \frac{1}{j},\quad \|v_j-{\overline u}_j\|_{H^1(\mathbb R^N)} < \frac{1}{j}, \end{equation} (A.4)

    so (2.8) entails

    \begin{equation} {\mathop{{{\rm{cap}}}}\nolimits} F_j < \beta j^{-1}. \end{equation} (A.5)

    By recalling that {\mathop{{{\rm{spt}}}}\nolimits} {\overline u}_j is compact we get that F_j is bounded so, by taking account of (A.2), there exists w_j\in C^{\infty}_0(\mathbb R^N; [0, 1]), \ w_j\equiv 1 in a neighbourhood U_j of F_j such that

    \begin{equation} \| w_j\|_{H^1({{\mathbb R}}^N)}^2 < \beta j^{-1}. \end{equation} (A.6)

    We define u_j: = (1-w_j)v_j : it is readily seen that u_j\in C^1(\mathbb R^N)\cap H^1(\mathbb R^N), \ u_j\ge 0\ \text{in}\ {{\mathbb R}}^N , that u_j({\mathbf{x}}) = 0 for q.e. {\mathbf{x}}\in E\cap U_j and that u_j\equiv \overline u_j(1-w_j) outside U_j , hence, by recalling that {\overline u_j}^*({\mathbf{x}}) = u^*({\mathbf{x}}) = 0 for q.e. {\mathbf{x}}\in E , we get u_j({\mathbf{x}}) = 0 for q.e. {\mathbf{x}}\in E\setminus U_j that is u_j({\mathbf{x}}) = 0 for q.e. {\mathbf{x}}\in E . We claim that u_j\to u in H^1(\mathbb R^N) : to this aim, by noticing that u_j - u = v_j- \overline u_j+\overline u_j-u-v_jw_j and that \overline u_j\to u in H^1(\mathbb R^N) , thanks to (A.4) we have only to show that v_jw_j\to 0 in H^1(\mathbb R^N) . We first notice that by setting

    B_j: = \{{\mathbf{x}}: |w_j({\mathbf{x}})|\ge j^{-\frac{1}{4}}\},

    then by (A.6) and Tchebichev inequality we get |B_j|\le \beta j^{-\frac{1}{2}} , therefore

    \begin{equation} \begin{array}{ll} \int_{\mathbb R^N} |w_j|^2|v_j|^2\,d{\mathbf{x}} & = \int_{B_j}|w_j|^2|v_j|^2\,d{\mathbf{x}} +\int_{\mathbb R^3\setminus B_j} |w_j|^2|v_j|^2\,d{\mathbf{x}}\\ & \le \int_{B_j}|v_j|^2\,d{\mathbf{x}}+j^{-\frac{1}{2}} \int_{\mathbb R^N\setminus B_j}|v_j|^2\,d{\mathbf{x}} \\ & \le 2\int_{B_j}|v_j-\overline u_j|^2\,d{\mathbf{x}}+2\int_{B_j}|\overline u_j|^2\,d{\mathbf{x}} \\ & +2j^{-\frac{1}{2}} \int_{\mathbb R^N}|v_j-\overline u_j|^2\,d{\mathbf{x}}+2j^{-\frac{1}{2}} \int_{\mathbb R^N}|\overline u_j|^2\,d{\mathbf{x}}\to 0,\\ \end{array} \end{equation} (A.7)

    since

    \begin{equation*} \|v_j-\overline u_j\|_{H^1(\mathbb R^N)} < \frac{1}{j},\quad \overline u_j\to u\ \text{in}\ H^1(\mathbb R^N),\quad |B_j|\to 0. \end{equation*}

    Analogously by recalling (A.6), that w_j\equiv 1 on U_j , that v_j\equiv \overline u_j outside U_j and \|\overline u_j\|_{\infty}^2\le \sqrt j , we get

    \begin{equation} \begin{array}{ll} \int_{\mathbb R^N} |\nabla(w_jv_j)|^2\,d{\mathbf{x}}&\le 2\int_{\mathbb R^N} |w_j|^2|\nabla v_j|^2\,d{\mathbf{x}}+2\int_{\mathbb R^N\setminus U_j} |v_j|^2|\nabla w_j|^2\,d{\mathbf{x}} \\ & \le 2\int_{\mathbb R^N} |w_j|^2|\nabla v_j|^2\,d{\mathbf{x}}+2\|{\overline u}_j\|_{\infty}^2\int_{\mathbb R^N\setminus U_j} |\nabla w_j|^2\,d{\mathbf{x}} \\ \\ & \le 2\int_{\mathbb R^N\setminus B_j} |w_j|^2|\nabla v_j|^2\,d{\mathbf{x}} +2\int_{B_j} |w_j|^2|\nabla v_j|^2\,d{\mathbf{x}}+2j^{-\frac{1}{2}}\to 0,\\ \end{array} \end{equation} (A.8)

    as in (A.7) thus proving the lemma.

    Lemma A.5. Let u\in H^1(\Omega) such that u^*({\mathbf{x}})\ge 0 for q.e. {\mathbf{x}}\in E . Then there exists a sequence u_j\in C^1(\overline \Omega) such that u_j({\mathbf{x}})\ge 0 for q.e. {\mathbf{x}}\in E and u_j\to u in H^1(\Omega) .

    Proof. We recall that by Remark 2.3 u^*({\mathbf{x}})\ge 0 for q.e. {\mathbf{x}}\in E if and only if (u^-)^*({\mathbf{x}}) = 0 for q.e. {\mathbf{x}}\in E_{ess} and that E_{ess} is a closed subset of \overline \Omega . By Lemma A.3 there exists a Sobolev extension v of u^- such that {\mathop{{{\rm{spt}}}}\nolimits} v is compact, v\ge 0\ \text{a.e. in}\ {{\mathbb R}}^3 and

    \begin{equation*} \lim\limits_{r\to 0^+}\frac{1}{\left |B_r({\mathbf{x}})\right |}\int_{B_r({\mathbf{x}})}v({\boldsymbol{\xi}})\,d{\boldsymbol{\xi}} = 0 \end{equation*}

    for q.e. {\mathbf{x}}\in E_{ess} , so by Lemma A.4, there exists a sequence v_j\in C^1(\mathbb R^N)\cap H^1(\mathbb R^N), \ v_j\ge 0\ \text{in}\ {{\mathbb R}}^N such that v_j({\mathbf{x}}) = 0 for q.e. {\mathbf{x}}\in E_{ess} and v_j\to v in H^1(\mathbb R^N) . Let now w be a Sobolev extension of u^+ . We may assume without loss of generality that w\ge 0 a.e. in {{\mathbb R}}^N and if \rho_j is a sequence of smooth mollifiers then w_j: = w*\rho_j\ge 0 and w_j\to w in H^1(\mathbb R^N) . Therefore by setting u_j: = w_j-v_j , we have u_j\in C^1(\overline \Omega) , u_j({\mathbf{x}})\ge 0 for q.e. {\mathbf{x}}\in E and u_j\to u in H^1(\Omega) thus proving the lemma.

    Remark A.6. If E is a non empty subset of \overline \Omega and u\in H^1(\Omega) we say that u\ge 0 on E in the sense of H^1(\Omega) if there exists a sequence u_j\in C^1(\overline \Omega) such that u_j\ge 0 on E and u_j\to u in H^1(\Omega) (according to [31, Definition 5.1]). We claim that (u^-)^* = 0 q.e. in E (or equivalently u^*\ge 0 q.e. in E ) if and only if u\ge 0 on E in the sense of H^1(\Omega) : Indeed if (u^-)_* = 0 q.e. in E then Lemma A.5 provides a sequence u_j\in C^1(\overline \Omega) such that u_j\to u in H^1(\Omega), \ u_j\ge 0 on E , while the converse follows easily by recalling that if u_j\to u in H^1(\Omega) then, up to subsequences, u_j({{\bf x}})\to u^{*}({{\bf x}}) for q.e. {\mathbf{x}}\in E . $



    [1] D. R. Adams, L. I. Hedberg, Function spaces and potential theory, Springer, 1999.
    [2] L. Ambrosio, N. Fusco, D. Pallara, Functions of bounded variation and free discontinuity problems, Oxford Academic, 2000.
    [3] V. Agostiniani, G. Dal Maso, A. De Simone, , Linear elasticity obtained from finite elasticity by \Gamma-convergence under weak coerciveness conditions, Ann. Inst. H. Poincaré Anal. Non Linéaire, 29 (2012), 715–735. https://doi.org/10.1016/j.anihpc.2012.04.001 doi: 10.1016/j.anihpc.2012.04.001
    [4] R. Alicandro, G. Dal Maso, G. Lazzaroni, M. Palombaro, Derivation of a linearised elasticity model from singularly perturbed multiwell energy functionals, Arch. Rational Mech. Anal., 230 (2018), 1–45. https://doi.org/10.1007/s00205-018-1240-6 doi: 10.1007/s00205-018-1240-6
    [5] R. Alicandro, G. Lazzaroni, M. Palombaro, Derivation of linear elasticity for a general class of atomistic energies, SIAM J. Math. Anal., 53 (2021), 5060–5093. https://doi.org/10.1137/21M1397179 doi: 10.1137/21M1397179
    [6] G. Anzellotti, S. Baldo, D. Percivale, Dimension reduction in variational problems, asymptotic development in \Gamma-convergence and thin structures in elasticity, Asymptotic Anal., 9 (1994), 61–100. https://doi.org/10.3233/ASY-1994-9105 doi: 10.3233/ASY-1994-9105
    [7] H. Attouch, G. Buttazzo, G. Michaille, Variational analysis in Sobolev and BV spaces: applications to PDEs and optimization, 2 Eds., Society for Industrial and Applied Mathematics, 2014. https://doi.org/10.1137/1.9781611973488
    [8] B. Audoly, Y. Pomeau, Elasticity and geometry, Oxford University Press, 2010.
    [9] C. Baiocchi, F. Gastaldi, F. Tomarelli, Inéquations variationnelles non coercives, C. R. Acad. Sci. Paris, 299 (1984), 647–650.
    [10] C. Baiocchi, F. Gastaldi, F. Tomarelli, Some existence results on noncoercive variational inequalities, Ann. Scuola Normale Sup. Pisa., Cl.Sci., 13 (1986), 617–659.
    [11] T. Bagby, Quasi topologies and rational approximation, J. Funct. Anal., 10 (1972), 259–268. https://doi.org/10.1016/0022-1236(72)90025-0 doi: 10.1016/0022-1236(72)90025-0
    [12] C. Baiocchi, G. Buttazzo, F. Gastaldi, F. Tomarelli, General existence theorems for unilateral problems in continuum mechanics, Arch. Rational Mech. Anal., 100 (1988), 149–189. https://doi.org/10.1007/BF00282202 doi: 10.1007/BF00282202
    [13] P. Bella, R. V. Kohn, Wrinkles as the result of compressive stresses in an annular thin film, Commun. Pure Appl. Math., 67 (2014), 693–747. https://doi.org/10.1002/cpa.21471 doi: 10.1002/cpa.21471
    [14] G. Buttazzo, F. Tomarelli, Compatibility conditions for nonlinear Neumann problems, Adv. Math., 89 (1991), 127–143. https://doi.org/10.1016/0001-8708(91)90076-J doi: 10.1016/0001-8708(91)90076-J
    [15] M. Carriero, A. Leaci, F. Tomarelli, Strong solution for an elastic-plastic plate, Calc. Var. Partial Differ. Equ., 2 (1994), 219–240. https://doi.org/10.1007/BF01191343 doi: 10.1007/BF01191343
    [16] G. Dal Maso, An introduction to \Gamma-convergence, Boston: Birkhäuser Boston Inc., 1993.
    [17] G. Dal Maso, P. Longo, \Gamma-limits of obstacles, Ann. Mat. Pura Appl., 128 (1981), 1–50. https://doi.org/10.1007/BF01789466 doi: 10.1007/BF01789466
    [18] G. Dal Maso, M. Negri, D. Percivale, Linearized elasticity as \Gamma-limit of finite elasticity, Set-Valued Anal., 10 (2002), 165–183. https://doi.org/10.1023/A:1016577431636 doi: 10.1023/A:1016577431636
    [19] L. C. Evans, R. F. Gariepy, Measure theory and fine properties of functions, New York: Chapman and Hall/CRC, 2015. https://doi.org/10.1201/b18333
    [20] M. Egert, P. Tolksdorf, Characterizations of Sobolev functions that vanish on a part of the boundary, Discrete Cont. Dyn. Syst.-Ser. S, 10 (2016), 729–743. https://doi.org/10.3934/dcdss.2017037 doi: 10.3934/dcdss.2017037
    [21] G. Fichera, Sul problema elastostatico di Signorini con ambigue condizioni al contorno, Atti Accad. Naz. Lincei, Ⅷ. Ser., Rend., Cl. Sci. Fis. Mat. Nat., 8 (1963), 138–142.
    [22] G. Fichera, Problemi elastostatici con vincoli unilaterali: il problema di Signorini con ambigue condizioni al contorno, Atti. Acad. Naz. Lincei. Mem. Cl. Sci. Nat., 8 (1964), 91–140.
    [23] G. Fichera, Boundary value problems of elasticity with unilateral constraints, In: C. Truesdell, Linear theories of elasticity and thermoelasticity, Springer, 1972,391–424. https://doi.org/10.1007/978-3-662-39776-3_4
    [24] G. Frieseke, R. D. James, S. Müller, A theorem on geometric rigidity and the derivation of nonlinear plate theory from three-dimensional elasticity, Commun. Pure Appl. Math., 55 (2002), 1461–1506. https://doi.org/10.1002/cpa.10048 doi: 10.1002/cpa.10048
    [25] G. Frieseke, R. D. James, S. Müller, A hierarchy of plate models derived from nonlinear elasticity by Gamma-convergence, Arch. Rational Mech. Anal., 180 (2006), 183–236. https://doi.org/10.1007/s00205-005-0400-7 doi: 10.1007/s00205-005-0400-7
    [26] D. Grandi, M. Kru\check{\hbox{z}}ik, E. Mainini, U. Stefanelli, Equilibrium for multiphase solids with Eulerian interfaces, J. Elast., 142 (2020), 409–431. https://doi.org/10.1007/s10659-020-09800-w doi: 10.1007/s10659-020-09800-w
    [27] M. E. Gurtin, An introduction to continuum mechanics, Academic Press, 1981.
    [28] V. P. Havin, Approximation in the mean by analytic functions, Dokl. Akad. Nauk SSSR, 178 (1968), 1025–1028.
    [29] R. Haller-Dintelmann, J. Rehberg, M. Egert, Hardy's inequality for functions vanishing on a part of the boundary, Potential Anal., 43 (2015), 49–78. https://doi.org/10.1007/s11118-015-9463-8 doi: 10.1007/s11118-015-9463-8
    [30] T. Kilpelainen, A remark on the uniqueness of quasicontinuous functions, Ann. Acad. Sci. Fenn. Math., 23 (1998), 261–262.
    [31] D. Kinderlehrer, G. Stampacchia, An introduction to variational inequalities and their applications, New York: Academic Press, 1980.
    [32] A. E. Love, A treatise on the mathematical theory of elasticity, Dover, 1944.
    [33] J. L. Lions, G. Stampacchia, Variational inequalities, Commun. Pure Appl. Math., 20 (1967), 493–519. https://doi.org/10.1002/cpa.3160200302 doi: 10.1002/cpa.3160200302
    [34] F. Maddalena, D. Percivale, Variational models for peeling problems, Interfaces Free Bound., 10 (2008), 503–516. https://doi.org/10.4171/ifb/199 doi: 10.4171/ifb/199
    [35] F. Maddalena, D. Percivale, G. Puglisi, L. Truskinowsky, Mechanics of reversible unzipping, Continuum Mech. Thermodyn., 21 (2009), 251–268. https://doi.org/10.1007/s00161-009-0108-2 doi: 10.1007/s00161-009-0108-2
    [36] F. Maddalena, D. Percivale, F. Tomarelli, Adhesive flexible material structures, Discrete Cont. Dyn. Syst.-Ser. S, 17 (2012), 553–574. https://doi.org/10.3934/dcdsb.2012.17.553 doi: 10.3934/dcdsb.2012.17.553
    [37] F. Maddalena, D. Percivale, F. Tomarelli, Elastic structures in adhesion interaction, In: G. Buttazzo, A. Frediani, Variational analysis and aerospace engineering: mathematical challenges for aerospace design, Springer, 66 (2012), 289–304. https://doi.org/10.1007/978-1-4614-2435-2_12
    [38] F. Maddalena, D. Percivale, F. Tomarelli, Local and nonlocal energies in adhesive interaction, IMA J. Appl. Math., 81 (2016), 1051–1075. https://doi.org/10.1093/imamat/hxw044 doi: 10.1093/imamat/hxw044
    [39] F. Maddalena, D. Percivale, F. Tomarelli, Variational problems for Föppl-von Kármán plates, SIAM J. Math. Anal., 50 (2018), 251–282. https://doi.org/10.1137/17M1115502 doi: 10.1137/17M1115502
    [40] F. Maddalena, D. Percivale, F. Tomarelli, The gap between linear elasticity and the variational limit of finite elasticity in pure traction problems, Arch. Rational Mech. Anal., 234 (2019), 1091–1120. https://doi.org/10.1007/s00205-019-01408-2 doi: 10.1007/s00205-019-01408-2
    [41] F. Maddalena, D. Percivale, F. Tomarelli, A new variational approach to linearization of traction problems in elasticity, J. Optim. Theory Appl., 182 (2019), 383–403, https://doi.org/10.1007/s10957-019-01533-8 doi: 10.1007/s10957-019-01533-8
    [42] F. Maddalena, D. Percivale, F. Tomarelli, Elastic-brittle reinforcement of flexural structures, Atti Accad. Naz. Lincei Cl. Sci. Fis. Mat. Natur., 32 (2021), 691–724. https://doi.org/10.4171/RLM/954 doi: 10.4171/RLM/954
    [43] E. Mainini, R. Ognibene, D. Percivale, Asymptotic behavior of constrained local minimizers in finite elasticity, J. Elast., 152 (2022), 1–27. https://doi.org/10.1007/s10659-022-09946-9 doi: 10.1007/s10659-022-09946-9
    [44] E. Mainini, D. Percivale, Variational linearization of pure traction problems in incompressible elasticity, Z. Angew. Math. Phys., 71 (2020), 146. https://doi.org/10.1007/s00033-020-01377-7 doi: 10.1007/s00033-020-01377-7
    [45] E. Mainini, D. Percivale, Sharp conditions for the linearization of finite elasticity, Calc. Var. Partial Differ. Equ., 60 (2021), 164. https://doi.org/10.1007/s00526-021-02037-y doi: 10.1007/s00526-021-02037-y
    [46] C. Maor, M. G. Mora, Reference configurations versus optimal rotations: a derivation of linear elasticity from finite elasticity for all traction forces, J. Nonlinear Sci., 31 (2021), 62. https://doi.org/10.1007/s00332-021-09716-2 doi: 10.1007/s00332-021-09716-2
    [47] M. G. Mora, F. Riva, Pressure live loads and the variational derivation of linear elasticity, Proc. Roy. Soc. Edinb., 153 (2022), 1929–1964. https://doi.org/10.1017/prm.2022.79 doi: 10.1017/prm.2022.79
    [48] D. Percivale, F. Tomarelli, Scaled Korn-Poincaré inequality in BD and a model of elastic plastic cantilever, Asymptotic Anal., 23 (2000), 291–311.
    [49] D. Percivale, F. Tomarelli, From SBD to SBH: the elastic-plastic plate, Interfaces Free Bound., 4 (2002), 137–165. https://doi.org/10.4171/ifb/56 doi: 10.4171/ifb/56
    [50] D. Percivale, F. Tomarelli, A variational principle for plastic hinges in a beam, Math. Mod. Meth. Appl. Sci., 19 (2009), 2263–2297. https://doi.org/10.1142/S021820250900411X doi: 10.1142/S021820250900411X
    [51] D. Percivale, F. Tomarelli, Smooth and broken minimizers of some free discontinuity problems, In: P. Colli, A. Favini, E. Rocca, G. Schimperna, J. Sprekels, Solvability, regularity, and optimal control of boundary value problems for PDEs, Springer INdAM Series, 22 (2017), 431–468, https://doi.org/10.1007/978-3-319-64489-9_17
    [52] P. Podio-Guidugli, On the validation of theories of thin elastic structures, Meccanica, 49 (2014), 1343–1352. https://doi.org/10.1007/s11012-014-9901-5 doi: 10.1007/s11012-014-9901-5
    [53] B. D. Reddy, F. Tomarelli, The obstacle problem for an elastoplastic body, Appl. Math. Optim., 21 (1990), 89–110. https://doi.org/10.1007/BF01445159 doi: 10.1007/BF01445159
    [54] A. Signorini, Questioni di elasticit non linearizzata e semilinearizzata, Rend. Mat. Appl., 18 (1959), 95–139.
    [55] F. Tomarelli, Signorini problem in Hencky plasticity, Ann. Univ. Ferrara, 36 (1990), 73–84. https://doi.org/10.1007/BF02837208 doi: 10.1007/BF02837208
    [56] C. Truesdell, W. Noll, The non-linear field theories of mechanics, In: The non-linear field theories of mechanics/die nicht-linearen feldtheorien der mechanik, Springer, 1965, 1–541. https://doi.org/10.1007/978-3-642-46015-9_1
    [57] W. P. Ziemer, Weakly differentiable functions, Springer, 1989.
  • Reader Comments
  • © 2024 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(1202) PDF downloads(229) Cited by(0)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog