This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Discrete-to-continuum limits of planar disclinations

Pierluigi Cesana Patrick van Meurs
Abstract

In materials science, wedge disclinations are defects caused by angular mismatches in the crystallographic lattice. To describe such disclinations, we introduce an atomistic model in planar domains. This model is given by a nearest-neighbor-type energy for the atomic bonds with an additional term to penalize change in volume. We enforce the appearance of disclinations by means of a special boundary condition.

Our main result is the discrete-to-continuum limit of this energy as the lattice size tends to zero. Our proof method is relaxation of the energy. The main mathematical novelty of our proof is a density theorem for the special boundary condition. In addition to our limit theorem, we construct examples of planar disclinations as solutions to numerical minimization of the model and show that classical results for wedge disclinations are recovered by our analysis.

1 Introduction

This paper is devoted to the mathematical analysis of a discrete model that describes frustrations in atomistic lattices induced by rotational mismatches. Such configurations are called wedge disclinations, which are angular defects. Disclinations are observed in solids in situations where the rotational symmetry is violated at the level of the crystal lattice. Figure 1 illustrates classical, simple examples of disclinations.

Refer to caption1-10111-1011ϕ=2π5\phi=\tfrac{2\pi}{5}Refer to caption1-10111-1011ϕ=2π7\phi=\tfrac{2\pi}{7}
Figure 1: Simple examples of two wedge disclinations in planar triangular lattices. The left figure shows a 5-type or positive disclination and the right figure a 7-type or negative disclination. The atom in the center has a different number of bonds than the other atoms. ϕ\phi is the angle between two such consecutive bonds. The lattice consists of rotated copies of the single wedge indicated by the thick dots.

Historically, the existence of disclinations was predicted by Volterra alongside dislocations (translational defects) in a celebrated paper [Vol07]. However, it was not until the late 1960s that disclinations saw a systematical investigation both from an experimental and theoretical perspective. First examples of disclinations over planar lattices have been discovered in superconductors and reported in [TE68] and [ET67]. While a dislocation is a singularity of the deformation field which may be described by a lattice-valued vector, called Burgers vector, a disclination, as stated in [AEST68] (see also [Nab67]), is characterized by a closure failure of rotation … for a closed circuit round the disclination centre. A continuum theory for disclinations in linearized elasticity has been first derived by de Wit in [dW70] based on the idea of compatible elasticity and later elaborated by the same author in a series of articles [dW73a], [dW73b], [dW73c]. A comprehensive theory for disclinations (alongside dislocations) in non-linear elasticity has been developed by Zubov [Zub97]. We refer to [ZA18] for more recent developments on unified approaches to treat dislocations and disclinations as well as for a review of mechanical models of disclinations.

Our work is inspired by a number of experimental observations which have revealed formation of disclinations in metallic structures under a variety of mechanical loading and forces, geometrical regimes and kinematical constraints. Here, we elaborate on two such observations.

In austenite-to-martensite transformations, disclinations may emerge during the formation of rotated (and constant-strain) shear-bands [Bha03]. Such transformations are purely elastic and of the type solid-to-solid. They appear in a class of metals; in particular, in shape-memory alloys. Upon symmetry-break (typically triggered and driven by a negative temperature gradient), austenite, the high-symmetry and highly homogeneous crystal phase, turns into martensite, an anisotropic crystal phase with lower symmetry. The crystallographic lattice accommodates this phase change by forming a complicated microstructure composed of a mixture of thin plates and needle-shaped regions exhibiting differently rotated copied of martensitic phases. In a zero stress-microstructure, martensite can be described by a piecewise constant deformation gradient. Constant-strain regions, that is, regions of constant crystal orientation, are separated by sharp planar interfaces according to kinematical compatibility. This compatibility is called a rank-1 connection of the corresponding deformation tensors. However, such configurations are ideal and typically atomistic non-idealities such as dislocations and disclinations appear in large numbers. Outstanding examples of disclinations in martensite are represented by ”nested” star-shaped geometries observed in Pb3(VO4)2\textrm{Pb}_{3}(\textrm{VO}_{4})_{2}. Experiments are described in [KK91, MA80]; modeling work as well as numerical and exact constructions are described in [PL13] and [CPL14] respectively, and mathematical theories are developed in [CDPR+20]. Significantly more complicated microstructures rich in defects are described in [ILTH17] for Ti-Nb-Al-based alloys. Here, martensite nucleates and evolves in the form of thin plates embedded in an austenitic lattice. The evolution of these plates is complex due to plates colliding against surrounding structures, undergoing further branching into additional martensitic sub-plates or being reflected after hitting a grain boundary. Such evolution results in self-similar patterns resembling fractal structure which are rich in disclinations and dislocations [ILTH17]. The available models of such microstructures are essentially based on statistical analysis [IHM13] and probability [BCH15, CH], and thus the consistency with atomistic models remains elusive. This is where we aim to contribute.

The second experimental observation is the discovery of a superior (and, yet to date, largely unexplained) structural reinforcement mechanism which has recently spurred scientific interest on the experimental as well as on the theoretical investigation of kink formation in certain classes of metal alloys. Formation of kinks consisting of approximately constant-strain bands accompanied by high rotational stretches of the lattice are observed in classes of laminate ”mille-feuille” structures under uniaxial compression [HOI+16, HMH+16]. In one of their most typical morphologies, bands manifest themselves in the form of planar and sharp ridge-shaped regions which appear at various length-scales and are accompanied by localized plastic stresses and formation of disclinations [LN15]. Kinks of various morphologies have been described in [Ina19] where constructions of piecewise affine deformation maps based on the rank-1 connection rule and incorporating angular mismatches are presented. While [Ina19]’s analysis sheds light on the kinematics of the disclination-kinking mechanism for various morphologies of kinks, there is no available model based on atomistic descriptions which describes the energy of such disclination-kinking mechanisms.

A common aspect on both martensitic microstructures and kink formation is that planar regions of approximately constant strains and constant-orientation lattices need to rotate in order to preserve the continuity of the deformation field across their common border. As a result, these materials necessarily develop angular lattice misfits which are striking examples of wedge disclinations. This motivates the modeling assumption of this paper that wedge disclinations are caused by large (non-linear) rotational stretches in planar-confined geometries.

Our aim is to take the first step in the direction of a comprehensive variational theory that is suitable to simultaneously treat microscale and localized defects and to predict their effect on large (non-linear) elastic and plastic deformations including kinks and shear-bands. By designing a simple, nearest-neighbor-type interaction mechanism, we construct a model which we apply to describe a single disclination in a planar lattice and which is at the same time potentially adaptable to describe multi-disclination systems and to incorporate other lattice defects such as voids, dislocations and grain boundaries.

By pursuing this aim, we also fill a gap in the literature on atomistic modeling of planar lattice defects and the limit thereof as the lattice spacing tends to 0. To identify this gap, we review related literature. First, in the framework of linearized elasticity, the formal asymptotic expansions in [SN88] provide a continuum model for lattice defects. Second, in [LPS13] two triangular lattices with different lattice spacings are attached together, which forces dislocations to form at the interface. The main result is a continuum limit of this model as the lattice spacing tends to 0. Third, based on the discrete calculus of lattices constructed in [AO05], the stress in a periodic crystal induced by parallel screw dislocations is computed in [Pon07, HO14, HO15, EOS16, BBO19]. These results provide quantitative estimates between the displacement in the lattice and the displacement computed from linear elasticity in the continuum counter part. In [BHO19] these techniques are extended to capture cracks. All these works deal with either localized or small deformation to the underlying two dimensional lattice, which is unfit for disclinations (see, e.g., the relatively large deformations in Figure 1).

For large deformations, the recent work in [KM18] presents and anlyzes an atomistic model of a stretchable hexagonal lattice defined over a smooth manifold, in which the main result is the continuum limit in the form of Γ\Gamma-convergence. Since the approach in [KM18] is designed to describe the energetics of highly distorted membranes, we follow a similar approach. However, the choice in [KM18] that the number of bonds per atom is always 66 makes the result not applicable to disclinations.

As in [KM18], we assign an energy EεE_{\varepsilon} to a deformed lattice, where ε\varepsilon is the lattice spacing. Two examples of such deformed lattices are illustrated by the dotted nodes in Figure 1. Given a deformed lattice, the energy EεE_{\varepsilon} penalizes stretching and compression of the atomic bonds and changes in volume of the triangles formed by three neighboring atoms. To enforce the appearance of a 55- or 77-type disclination, we require the deformed lattice to satisfy a special boundary condition such that 55 or 77 rotated copies of it fit together such as in Figure 1.

Our main result, Theorem 3.1, is the exact derivation of EE, the macroscale energy obtained in the limit as ε0\varepsilon\to 0, that is,

E=Γ-limε0EεE=\Gamma\hbox{-}\lim_{\varepsilon\to 0}E_{\varepsilon}

in a suitable topology.

From the mathematical perspective, the interest of our analysis lies in the treatment of the special boundary condition. Conceptually, this boundary conditions requires us to incorporate a pointwise constraint in the space of traces on W1,pW^{1,p}-Sobolev vector maps which are characterized by means of a nonlocal norm. Our main contribution on this is Proposition 4.4.

The paper is organized as follows. In Section 2 we present the discrete mechanical model and the mathematical setting required for the related variational analysis. In Section 3 we state and prove our main result, Theorem 3.1, on the Γ\Gamma-convergence of the lattice energy EεE_{\varepsilon}. In this proof, we postpone the proofs of technical lemmas to Section 4, which includes the proof of Proposition 4.4 on density in the space of admissible lattice displacements. In Section 5 we explore the physical implications of our Γ\Gamma-convergence analysis in terms of energy and stress states of the minimizers of the continuum model. Finally in Section 6 we present several numerical realizations of both positive and negative disclinations.

2 The lattice energy EεE_{\varepsilon}

Here we define the atomic lattice energy EεE_{\varepsilon} as briefly introduced in the introduction. We start with the kinematics. Inspired by Figure 1, we consider a two-dimensional model. This corresponds conceptually to the mid-section of a 3-dimensional body. Furthermore, we impose rotational symmetry so that we may confine the domain to a single wedge in Figure 1 indicated by the black dots. The reference domain is given by the equilateral open triangle Ω2\Omega\subset\mathbb{R}^{2} of size 1 with boundary Γ=i=13Γi\Gamma=\cup_{i=1}^{3}\Gamma_{i} as depicted in Figure 2. We take the reference positions of the atoms as a triangular lattice ε\mathcal{L}_{\varepsilon}, where ε\mathcal{L}_{\varepsilon} is such that it fits on top of Ω\Omega as in Figure 2, i.e., the lattice spacing ε(0,1)\varepsilon\in(0,1) is such that 1ε\frac{1}{\varepsilon}\in\mathbb{N} and ε\mathcal{L}_{\varepsilon} is positioned such that each closed line segment Γi\Gamma_{i} fits on top of the atoms in one of the lattice directions. We denote by

Bε:={εRθe1:θπ3}B^{\varepsilon}:=\Big{\{}\varepsilon R_{\theta}e_{1}:\theta\in\frac{\pi}{3}\mathbb{Z}\Big{\}} (1)

the set of the six outward-pointing bonds in ε\mathcal{L}_{\varepsilon} from any lattice point, where RθSO(2)R_{\theta}\in SO(2) is the counter-clockwise rotation matrix by angle θ\theta. For later use, we sometimes interpret ε\mathcal{L}_{\varepsilon} as a planar graph 𝒢ε=(𝒱ε,ε,𝒯ε)\mathcal{G}_{\varepsilon}=(\mathcal{V}_{\varepsilon},\mathcal{E}_{\varepsilon},\mathcal{T}_{\varepsilon}), where 𝒱ε\mathcal{V}_{\varepsilon} is the set of all vertices xix_{i}, ε\mathcal{E}_{\varepsilon} is the set of all edges eije_{ij} between neighboring vertices xi,xj𝒱εx_{i},x_{j}\in\mathcal{V}_{\varepsilon}, and 𝒯ε\mathcal{T}_{\varepsilon} is the set of all open triangles TijkΩT_{ijk}\in\Omega with sides given by the three edges eij,eik,ejkεe_{ij},e_{ik},e_{jk}\in\mathcal{E}_{\varepsilon}.

00Ω\OmegaΓ1\Gamma_{1}Γ2\Gamma_{2}Γ3\Gamma_{3}ϕ\phie1e_{1}Rπ/3e1R_{\pi/3}e_{1}ε\varepsilon11ε𝒢ε=(𝒱ε,ε,𝒯ε)\mathcal{L}_{\varepsilon}\cong\mathcal{G}_{\varepsilon}=(\mathcal{V}_{\varepsilon},\mathcal{E}_{\varepsilon},\mathcal{T}_{\varepsilon})uεu_{\varepsilon}uε(ε)u_{\varepsilon}(\mathcal{L}_{\varepsilon})
Figure 2: The reference lattice ε\mathcal{L}_{\varepsilon} and an admissible deformation uεu_{\varepsilon} (see (2)). The dotted triangle on the right is a visual aid to see that uεu_{\varepsilon} satisfies the boundary condition.

The set 𝒜ε\mathcal{A}_{\varepsilon} of admissible displacements is given by

𝒜ε={uε:ε2:uε(Rπ/3x)=Rϕuε(x) for all xΓ1𝒱ε},\displaystyle\mathcal{A}_{\varepsilon}=\{u_{\varepsilon}:\mathcal{L}_{\varepsilon}\to\mathbb{R}^{2}:u_{\varepsilon}(R_{\pi/3}x)=R_{\phi}u_{\varepsilon}(x)\ \text{ for all }x\in\Gamma_{1}\cap\mathcal{V}_{\varepsilon}\}, (2)

where ϕ{2π5,2π7}\phi\in\{\frac{2\pi}{5},\frac{2\pi}{7}\} is the angle associated with a 55- or 77-type disclination (see Figure 1). Since 𝒱ε\mathcal{V}_{\varepsilon} is a finite set, the map uεu_{\varepsilon} can be identified with a vector in 2|𝒱ε|\mathbb{R}^{2|\mathcal{V}_{\varepsilon}|}. The boundary condition in (2) is such that the rotated copy Rϕuε(ε)R_{\phi}u_{\varepsilon}(\mathcal{L}_{\varepsilon}) of uε(ε)u_{\varepsilon}(\mathcal{L}_{\varepsilon}) fits seamlessly to uε(ε)u_{\varepsilon}(\mathcal{L}_{\varepsilon}). Adding more rotated copies, we obtain a deformed lattice with rotational symmetry such as that in Figure 1, which has a 55- or 77-type disclination at the origin.

Next we define the lattice energy on 𝒜ε\mathcal{A}_{\varepsilon} as inspired by [KM18]. We start with a formal description. Given uε𝒜εu_{\varepsilon}\in\mathcal{A}_{\varepsilon}, we set the lattice energy formally as

E~ε(uε)=ε2eijεw(eij)Φ(|uε(xi)uε(xj)|ε1)+ε22T𝒯εΨ(sgn(uε(T))|uε(T)||T|).\displaystyle\tilde{E}_{\varepsilon}(u_{\varepsilon})=\varepsilon^{2}\sum_{e_{ij}\in\mathcal{E}_{\varepsilon}}w(e_{ij})\Phi\left(\frac{|u_{\varepsilon}(x_{i})-u_{\varepsilon}(x_{j})|}{\varepsilon}-1\right)+\frac{\varepsilon^{2}}{2}\sum_{T\in\mathcal{T}_{\varepsilon}}\Psi\Big{(}\operatorname{sgn}(u_{\varepsilon}(T))\frac{|u_{\varepsilon}(T)|}{|T|}\Big{)}. (3)

where the potential

Φ(r):=1p|r|p,p2\displaystyle\Phi(r):=\frac{1}{p}|r|^{p},\quad p\geq 2

penalizes atomic bonds which are not of length ε\varepsilon (the parameter value p=2p=2 corresponds to linear elasticity), the weight function

w:ε{12,1},w(eij)={12if eijΓ1otherwisew:\mathcal{E}_{\varepsilon}\to\{\tfrac{1}{2},1\},\qquad w(e_{ij})=\left\{\begin{aligned} &\tfrac{1}{2}&&\text{if }e_{ij}\subset\Gamma\\ &1&&\text{otherwise}\end{aligned}\right. (4)

counts the outer edges as half111we model edges as part of the volume of the medium around them., and the potential Ψ\Psi penalizes change in volume of the triangles T𝒯εT\in\mathcal{T}_{\varepsilon} (here, |T||T| is the volume of TT, and |uε(T)||u_{\varepsilon}(T)| is the volume of the triangle TT after applying the displacement uεu_{\varepsilon}), especially if the volume of TT gets inverted under uεu_{\varepsilon}. Note that while the identity map minimizes E~ε\tilde{E}_{\varepsilon}, it is not in 𝒜ε\mathcal{A}_{\varepsilon} because of the boundary condition. Hence, the boundary condition enforces mechanical frustration. For p=2p=2 and Ψ0\Psi\equiv 0, Figure 1 illustrates the deformed lattice of a local minimizer in 𝒜ε\mathcal{A}_{\varepsilon} of E~ε\tilde{E}_{\varepsilon} which does not contain negative change in volume. In this paper we always assume p2p\geq 2 unless specified otherwise.

Since the term involving Ψ\Psi is not derived from physical principles, we elaborate on this modelling choice. It is well-known that when Ψ0\Psi\equiv 0, lattice energies with only nearest-neighbor interactions such as E~ε\tilde{E}_{\varepsilon} do not penalize folding or any other negative change in volume, and as a result, the related continuum energy may not penalize compression. While our boundary condition in (2) is not standard, we show in Section 6 by means of numerical simulations that folded patterns appear as local minimizers of E~ε\tilde{E}_{\varepsilon} when Ψ0\Psi\equiv 0.

To penalize folding, we pose the following minimal requirements on Ψ\Psi:

  • (Ψ1\Psi 1)

    C>0a,b:|Ψ(a)Ψ(b)|C(|a|p21+|b|p21+1)|ab|\exists\,C>0\ \forall\,a,b\in\mathbb{R}:|\Psi(a)-\Psi(b)|\leq C\Big{(}|a|^{\tfrac{p}{2}-1}+|b|^{\tfrac{p}{2}-1}+1\Big{)}|a-b|;

  • (Ψ2\Psi 2)

    Ψ0\Psi\geq 0, and Ψ(a)=0a=1\Psi(a)=0\Longleftrightarrow a=1.

Condition (Ψ1\Psi 1) is a continuity estimate, which implies both local Lipschitz continuity and p2\frac{p}{2}-growth. Condition (Ψ2\Psi 2) ensures that any change in volume is penalized. A simple example of Ψ\Psi is Ψ(a)=|a1|\Psi(a)=|a-1|. These conditions are more general than those in [KM18, Section 3.3].

The term in E~ε\tilde{E}_{\varepsilon} related to Ψ\Psi has an inconvenient form. We fix this by changing variables. The result is a rigorously defined energy functional EεE_{\varepsilon}, which we will use in the remainder of the paper.

We change variables by describing the state uεu_{\varepsilon} as the deformation of a displacement UεU_{\varepsilon} defined on Ω\Omega. Given uε𝒜εu_{\varepsilon}\in\mathcal{A}_{\varepsilon}, we set Uε:Ω2U_{\varepsilon}:\Omega\to\mathbb{R}^{2} as the piecewise linear extension of uεu_{\varepsilon} to Ω\Omega, i.e.,

Uε(αxi+βxj+(1αβ)xk):=αuε(xi)+βuε(xj)+(1αβ)uε(xk)Tijk𝒯ε, 0α1, 0β1α.U_{\varepsilon}(\alpha x_{i}+\beta x_{j}+(1-\alpha-\beta)x_{k}):=\alpha u_{\varepsilon}(x_{i})+\beta u_{\varepsilon}(x_{j})+(1-\alpha-\beta)u_{\varepsilon}(x_{k})\\ \forall\,T_{ijk}\in\mathcal{T}_{\varepsilon},\ 0\leq\alpha\leq 1,\ 0\leq\beta\leq 1-\alpha.

We note that

[Uε(x)]Te=uε(x)uε(xm)εTijk𝒯ε,xTijk,\big{[}\nabla U_{\varepsilon}(x)\big{]}^{T}e=\frac{u_{\varepsilon}(x_{\ell})-u_{\varepsilon}(x_{m})}{\varepsilon}\quad\forall\,T_{ijk}\in\mathcal{T}_{\varepsilon},\ x\in T_{ijk}, (5)

where eB1e\in B^{1} is any lattice bond of unit direction (see (1)), and the indices ,m{i,j,k}\ell,m\in\{i,j,k\} depend on ee. In particular, Uε\nabla U_{\varepsilon} is constant on each T𝒯εT\in\mathcal{T}_{\varepsilon},

|detUε|=|Uε(T)||T|on T|\det\nabla U_{\varepsilon}|=\frac{|U_{\varepsilon}(T)|}{|T|}\quad\text{on }T

is the relative change in volume of TT under UεU_{\varepsilon}, and the sign of the determinant determines whether the volume of TT is inverted under UεU_{\varepsilon}. From these observations, we define the second term in (3) rigorously by

Ψ(sgn(uε(T))|uε(T)||T|):=Ψ(detUε)on T.\Psi\Big{(}\operatorname{sgn}(u_{\varepsilon}(T))\frac{|u_{\varepsilon}(T)|}{|T|}\Big{)}:=\Psi(\det\nabla U_{\varepsilon})\quad\text{on }T. (6)

Next we rewrite 𝒜ε\mathcal{A}_{\varepsilon} and E~ε\tilde{E}_{\varepsilon} in terms of UεU_{\varepsilon}. This yields

εϕ:={Uε:Ω¯2Uε is ε-piecewise linear and xΓ1:RϕUε(x)=Uε(Rπ/3x)}.\mathcal{B}_{\varepsilon}^{\phi}:=\big{\{}U_{\varepsilon}:\overline{\Omega}\to\mathbb{R}^{2}\mid U_{\varepsilon}\text{ is }\mathcal{L}_{\varepsilon}\text{-piecewise linear and }\forall\,x\in\Gamma_{1}:R_{\phi}U_{\varepsilon}(x)=U_{\varepsilon}(R_{\pi/3}x)\big{\}}.

Using (5) and (6), we rewrite (3) as

32E~ε(uε)\displaystyle\frac{\sqrt{3}}{2}\tilde{E}_{\varepsilon}(u_{\varepsilon}) =32(ε2T𝒯ε12eB11|T|TΦ(|eUε(x)|1)dx\displaystyle=\frac{\sqrt{3}}{2}\bigg{(}\varepsilon^{2}\sum_{T\in\mathcal{T}_{\varepsilon}}\frac{1}{2}\sum_{e\in B^{1}}\frac{1}{|T|}\int_{T}\Phi\left(\big{|}e\nabla U_{\varepsilon}(x)\big{|}-1\right)\,dx
+ε22T𝒯ε1|T|TΨ(detUε(x))dx)\displaystyle\qquad\qquad+\frac{\varepsilon^{2}}{2}\sum_{T\in\mathcal{T}_{\varepsilon}}\frac{1}{|T|}\int_{T}\Psi(\det\nabla U_{\varepsilon}(x))\,dx\bigg{)}
=ΩeB1Φ(|eUε(x)|1)+Ψ(detUε(x))dx\displaystyle=\int_{\Omega}\sum_{e\in B^{1}}\Phi\left(\big{|}e\nabla U_{\varepsilon}(x)\big{|}-1\right)+\Psi(\det\nabla U_{\varepsilon}(x))\,dx
=ΩW(Uε(x))𝑑x,\displaystyle=\int_{\Omega}W(\nabla U_{\varepsilon}(x))\,dx, (7)

where

W(A):=eB1Φ(|eA|1)+Ψ(detA).W(A):=\sum_{e\in B^{1}}\Phi\left(\big{|}eA\big{|}-1\right)+\Psi(\det A). (8)

In the computation above, the weight function w(eij)w(e_{ij}) turns into the factor 12\frac{1}{2} due to the fact that each edge in the interior of Ω\Omega borders two triangles in 𝒯ε\mathcal{T}_{\varepsilon}. Eq. (7) motivates us to define

Eε:εϕ[0,),Eε(Uε):=ΩW(Uε(x))𝑑x.E_{\varepsilon}:\mathcal{B}_{\varepsilon}^{\phi}\to[0,\infty),\qquad E_{\varepsilon}(U_{\varepsilon}):=\int_{\Omega}W(\nabla U_{\varepsilon}(x))\,dx. (9)
Remark 2.1 (Properties of εϕ\mathcal{B}_{\varepsilon}^{\phi}).

We note that εϕLip(Ω¯;2)\mathcal{B}_{\varepsilon}^{\phi}\subset\operatorname{Lip}(\overline{\Omega};\mathbb{R}^{2}) and that any UεεϕU_{\varepsilon}\in\mathcal{B}_{\varepsilon}^{\phi} satisfies Uε(0)=0U_{\varepsilon}(0)=0. While εϕ\mathcal{B}_{\varepsilon}^{\phi} does not contain a subspace of constants, it does contain a subspace of linear maps Uε(x)=AxU_{\varepsilon}(x)=Ax. A possible choice for A2×2A\in\mathbb{R}^{2\times 2} is the one that satisfies Ae1=e1Ae_{1}=e_{1} and A(Rπ/3e1)=Rϕe1A(R_{\pi/3}e_{1})=R_{\phi}e_{1}.

3 Continuum limit

Having introduced the lattice energy EεE_{\varepsilon} on the triangular lattice, we are now in a position to discuss the continuum limit as ε0\varepsilon\to 0. To keep track of the asymptotic behavior of minima and minimizers of EεE_{\varepsilon} we characterize the continuum model with Γ\Gamma-convergence [DM93].

Let p2p\geq 2. The domain of the continuum energy is

Wϕ1,p(Ω;2):={UW1,p(Ω;2):RϕUε(x)=Uε(Rπ/3x) for a.e. xΓ1}.W_{\phi}^{1,p}(\Omega;\mathbb{R}^{2}):=\bigg{\{}U\in W^{1,p}(\Omega;\mathbb{R}^{2}):R_{\phi}U_{\varepsilon}(x)=U_{\varepsilon}(R_{\pi/3}x)\text{ for a.e.\ }x\in\Gamma_{1}\bigg{\}}.

Observe that Wϕ1,p(Ω;2)W_{\phi}^{1,p}(\Omega;\mathbb{R}^{2}) is linear (and, in particular, convex), and non-empty since εϕW01,p(Ω;2)Wϕ1,p(Ω;2)\mathcal{B}_{\varepsilon}^{\phi}\cup W_{0}^{1,p}(\Omega;\mathbb{R}^{2})\subset W_{\phi}^{1,p}(\Omega;\mathbb{R}^{2}). Moreover, if UWϕ1,p(Ω;2)U\in W_{\phi}^{1,p}(\Omega;\mathbb{R}^{2}), there holds RUWϕ1,p(Ω;2)RU\in W_{\phi}^{1,p}(\Omega;\mathbb{R}^{2}) for any RSO(2)R\in SO(2). Thanks to the properties of traces (see Lemma 4.6 below), we have that Wϕ1,p(Ω;2)W_{\phi}^{1,p}(\Omega;\mathbb{R}^{2}) is strongly closed and, therefore, weakly closed as well thanks to convexity.

The continuum energy functional is given by

E:Lp(Ω;2)[0,],E(U)={ΩQW(U)UWϕ1,p(Ω;2)UWϕ1,p(Ω;2),E:L^{p}(\Omega;\mathbb{R}^{2})\to[0,\infty],\qquad E(U)=\left\{\begin{aligned} &\int_{\Omega}QW(\nabla U)&&U\in W_{\phi}^{1,p}(\Omega;\mathbb{R}^{2})\\ &\infty&&U\notin W_{\phi}^{1,p}(\Omega;\mathbb{R}^{2}),\end{aligned}\right. (10)

where WW is as in (8) and QWQW is the quasiconvex envelop of WW defined by

QW(A):=sup{f(A):VW,f quasiconvex},QW(A):=\sup\{f(A):V\leq W,\ f\text{ quasiconvex}\}, (11)

where ff being quasiconvex means that f:2×2f:\mathbb{R}^{2\times 2}\to\mathbb{R} is Borel measurable, locally bounded and satisfies

f(A)1|ω|ωf(A+ϑ(x))𝑑xf(A)\leq\frac{1}{|\omega|}\int_{\omega}f(A+\nabla\vartheta(x))\,dx

for any bounded open set ω2\omega\subset\mathbb{R}^{2}, any A2×2A\in\mathbb{R}^{2\times 2} and any ϑW01,(ω,2)\vartheta\in W^{1,\infty}_{0}(\omega,\mathbb{R}^{2}).

Theorem 3.1.

For ϕ{2π5,2π7}\phi\in\{\frac{2\pi}{5},\frac{2\pi}{7}\} and 2p<2\leq p<\infty, EεE_{\varepsilon} (see (9)) Γ\Gamma-converges as ε0\varepsilon\to 0 to EE in the strong Lp(Ω)L^{p}(\Omega) topology.

We prove Theorem 3.1 in Section 3.2. Since the ε\varepsilon-dependence of EεE_{\varepsilon} appears only in its domain ϕε\mathcal{B}_{\phi}^{\varepsilon}, proving a Γ\Gamma-limit result reduces to proving a relaxation result, i.e., finding the lower semi-continuous envelope of

UΩW(U)U\mapsto\int_{\Omega}W(\nabla U)

in the right functional framework. There is a large literature on such relaxation problems; in Section 3.1 we cite the relevant classical theory. While this theory gives a useful roadmap for proving Theorem 3.1, it does not capture Theorem 3.1 because of the periodic boundary condition in εϕ\mathcal{B}_{\varepsilon}^{\phi}. Therefore, in Section 3.2 we give the skeleton of the proof of Theorem 3.1 based on the classical theory, and identify the missing steps as technical lemmas which we prove in Section 4.

3.1 Classical relaxation result on \mathcal{E}

We review some classical relaxation results as preparation for proving Theorem 3.1. All theorem references below in this section refer to Dacorogna’s book [Dac08]. Another relevant reference is [AF84].

We recall that Ω\Omega is a bounded Lipschitz domain. Here and in what follows we adopt the Frobenius norm for matrices. Let 𝒲:2×2\mathcal{W}:\mathbb{R}^{2\times 2}\to\mathbb{R} satisfy

  • (W1W1)

    pp-growth. C,C>0A2×2:1C|A|pC𝒲(A)C(|A|p+1)\exists\,C,C^{\prime}>0\ \forall\,A\in\mathbb{R}^{2\times 2}:\dfrac{1}{C}|A|^{p}-C\leq\mathcal{W}(A)\leq C^{\prime}(|A|^{p}+1);

  • (W2W2)

    Continuity estimate. C>0A,B2×2:|𝒲(A)𝒲(B)||AB|C(|A|p1+|B|p1+1)\exists\,C>0\ \forall\,A,B\in\mathbb{R}^{2\times 2}:\dfrac{|\mathcal{W}(A)-\mathcal{W}(B)|}{|A-B|}\leq C(|A|^{p-1}+|B|^{p-1}+1).

Note that (W1W1) includes a uniform bound from below, and that (W2W2) provides a local Lipschitz estimate. We set

:Lp(Ω;2),(U)={Ω𝒲(U)UW1,p(Ω)UW1,p(Ω).\mathcal{E}:L^{p}(\Omega;\mathbb{R}^{2})\to\mathbb{R},\qquad\mathcal{E}(U)=\left\{\begin{aligned} &\int_{\Omega}\mathcal{W}(\nabla U)&&U\in W^{1,p}(\Omega)\\ &\infty&&U\notin W^{1,p}(\Omega).\end{aligned}\right.

Here and henceforth, we remove the range 2\mathbb{R}^{2} from the notation of the function space if there is no danger for confusion.

As preparation, we cite Theorem 6.9 for the alternative characterization of (11)(\ref{2004262117}) as the quasiconvexification of 𝒲\mathcal{W} given by

Q𝒲(A)=inf{|ω|1ω𝒲(A+ϑ):ϑW01,(ω,2)},Q\mathcal{W}(A)=\inf\left\{|\omega|^{-1}\int_{\omega}\mathcal{W}(A+\nabla\vartheta):\vartheta\in W^{1,\infty}_{0}(\omega,\mathbb{R}^{2})\right\}, (12)

where ω\omega is a subset of 2\mathbb{R}^{2} with |ω|=0|\partial\omega|=0 (i.e., ω\partial\omega has zero two-dimensional volume).

Since 𝒲\mathcal{W} satisfies (W1W1), Theorem 9.1 implies that for all UW1,p(Ω)U\in W^{1,p}(\Omega) there exists a sequence (Uk)kW01,p(Ω)+{U}(U_{k})_{k}\subset W_{0}^{1,p}(\Omega)+\{U\} such that

UkUin Lp(Ω)\displaystyle U_{k}\to U\quad\text{in }L^{p}(\Omega)

and

(Uk)=Ω𝒲(Uk(x))𝑑xΩQ𝒲(U(x))𝑑x as k.\displaystyle\mathcal{E}(U_{k})=\int_{\Omega}\mathcal{W}(\nabla U_{k}(x))\,dx\to\int_{\Omega}Q\mathcal{W}(\nabla U(x))\,dx\quad\text{ as }k\to\infty. (13)

By the two properties of 𝒲\mathcal{W}, it follows from Theorem 6.9 and Theorem 5.3(iv) that Q𝒲Q\mathcal{W} is continuous. Then, Theorem 1.13 implies that UQ𝒲(U)U\mapsto\int Q\mathcal{W}(\nabla U) is sequentially weakly lower semicontinuous in W1,p(Ω)W^{1,p}(\Omega). In particular, for any (Uk)kW1,p(Ω)(U_{k})_{k}\subset W^{1,p}(\Omega) converging weakly in W1,p(Ω)W^{1,p}(\Omega) to some UU, we have that

lim infk(Uk)lim infkΩQ𝒲(Uk(x))𝑑xΩQ𝒲(U(x))𝑑x.\displaystyle\liminf_{k\to\infty}\mathcal{E}(U_{k})\geq\liminf_{k\to\infty}\int_{\Omega}Q\mathcal{W}(\nabla U_{k}(x))\,dx\geq\int_{\Omega}Q\mathcal{W}(\nabla U(x))\,dx. (14)

3.2 Proof of Theorem 3.1

The proof below relies on the following three statements which we make precise and prove in Section 4:

  • Poincaré Inequality holds on Wϕ1,p(Ω)W^{1,p}_{\phi}(\Omega) (Lemma 4.1);

  • WW satisfies Properties (W1W1) and (W2W2) defined in Section 3.1 (Lemma 4.2);

  • ϕε\mathcal{B}_{\phi}^{\varepsilon} is dense in Wϕ1,p(Ω)W^{1,p}_{\phi}(\Omega) (Proposition 4.4).

The proof of Theorem 3.1 follows from matching a lower bound with an upper bound, which is the standard method for computing Γ\Gamma-limits [DM93]. To identify the set where the limit functional is finite, we first investigate the (equi-)compactness of minimizing sequences.

Proof of Theorem 3.1.

Compactness. By the lower bound in (W1W1),

Eε(Uε)1CUεLp(Ω)pCE_{\varepsilon}(U_{\varepsilon})\geq\frac{1}{C}\|\nabla U_{\varepsilon}\|_{L^{p}(\Omega)}^{p}-C

for some constants C>0C>0 independent of ε\varepsilon and UεU_{\varepsilon}, and thus any finite-energy sequence (Uε)ε(\nabla U_{\varepsilon})_{\varepsilon} is bounded in Lp(Ω)L^{p}(\Omega). By the Poincaré Inequality (Lemma 4.1), we then infer that UεU_{\varepsilon} is bounded in W1,p(Ω)W^{1,p}(\Omega), and thus strongly convergent (along a subsequence) in Lp(Ω)L^{p}(\Omega).

Γ\Gamma-liminf. Since we can focus on finite-energy sequences UεUU_{\varepsilon}\to U in Lp(Ω)L^{p}(\Omega), the compactness statement implies that UεUU_{\varepsilon}\rightharpoonup U in W1,p(Ω)W^{1,p}(\Omega) as ε0\varepsilon\to 0. Since (Uε)Wϕ1,p(Ω)(U_{\varepsilon})\subset W^{1,p}_{\phi}(\Omega), which is a closed subspace of W1,p(Ω)W^{1,p}(\Omega), we also have UWϕ1,p(Ω)U\in W^{1,p}_{\phi}(\Omega). Hence, by applying (14), we obtain the required Γ\Gamma-liminf estimate.

Γ\Gamma-limsup. Thanks to (13) it suffices to find, for any UWϕ1,p(Ω)U\in W_{\phi}^{1,p}(\Omega), a sequence (Uε)εϕε(U_{\varepsilon})_{\varepsilon}\subset\mathcal{B}_{\phi}^{\varepsilon} such that

Uεε0U in Lp(Ω)andEε(Uε)ε0ΩW(U)dx=:(U).U_{\varepsilon}\xrightarrow{\varepsilon\to 0}U\ \text{ in }L^{p}(\Omega)\quad\text{and}\quad E_{\varepsilon}(U_{\varepsilon})\xrightarrow{\varepsilon\to 0}\int_{\Omega}W(\nabla U)dx=:\mathcal{E}(U). (15)

To prove (15), we infer from Proposition 4.4 and Eε(Uε)=(Uε)E_{\varepsilon}(U_{\varepsilon})=\mathcal{E}(U_{\varepsilon}) that it is enough to show that \mathcal{E} is continuous in W1,p(Ω)W^{1,p}(\Omega) at UU. The continuity of \mathcal{E} follows by (W2W2), Lemma 4.1 and Hölder’s inequality from

|(U)(V)|\displaystyle|\mathcal{E}(U)-\mathcal{E}(V)| Ω|W(U)W(V)|𝑑x\displaystyle\leq\int_{\Omega}|W(\nabla U)-W(\nabla V)|\,dx
CΩ(|U|p1+|V|p1+1)|UV|𝑑x\displaystyle\leq C\int_{\Omega}\big{(}|\nabla U|^{p-1}+|\nabla V|^{p-1}+1\big{)}|\nabla U-\nabla V|\,dx
C(UW1,p(Ω)p1+VW1,p(Ω)p1+1)UVW1,p(Ω),\displaystyle\leq C\left(\|U\|_{W^{1,p}(\Omega)}^{p-1}+\|V\|_{W^{1,p}(\Omega)}^{p-1}+1\right)\|U-V\|_{W^{1,p}(\Omega)},

where VW1,p(Ω)V\in W^{1,p}(\Omega) is arbitrary. ∎

4 Technical steps in the proof of Theorem 3.1

Here we state rigorously and prove the three statements mentioned at the start of Section 3.2. We start with the Poincaré Inequality.

Lemma 4.1 (Poincaré Inequality on Wϕ1,p(Ω)W^{1,p}_{\phi}(\Omega)).

There exists C>0C>0 such that for all UWϕ1,p(Ω)U\in W^{1,p}_{\phi}(\Omega) it holds that

ULp(Ω)CULp(Ω).\|U\|_{L^{p}(\Omega)}\leq C\|\nabla U\|_{L^{p}(\Omega)}.
Proof.

We follow a standard proof by contradiction. Assuming that there exists (Un)Wϕ1,p(Ω)(U_{n})\subset W^{1,p}_{\phi}(\Omega) such that

1=UnLp(Ω)>nUnLp(Ω),1=\|U_{n}\|_{L^{p}(\Omega)}>n\|\nabla U_{n}\|_{L^{p}(\Omega)},

we obtain by compactness that UnU_{n} converges along a subsequence (not relabelled) to UU strongly in Lp(Ω)L^{p}(\Omega) and weakly in Wϕ1,p(Ω)W^{1,p}_{\phi}(\Omega). Hence, ULp(Ω)=1\|U\|_{L^{p}(\Omega)}=1 and ULp(Ω)=0\|\nabla U\|_{L^{p}(\Omega)}=0. Since Ω\Omega is connected, UV0U\equiv V_{0} for some constant and non-zero vector V02V_{0}\in\mathbb{R}^{2}. However, for a.e. xΓ1x\in\Gamma_{1},

RϕU(x)U(Rπ/3x)=(RϕI)V00.R_{\phi}U(x)-U(R_{\pi/3}x)=(R_{\phi}-I)V_{0}\neq 0.

Hence, UWϕ1,p(Ω)U\notin W^{1,p}_{\phi}(\Omega), which completes the proof. ∎

4.1 Properties of WW

Lemma 4.2.

WW defined in (8) satisfies Properties (W1W1) and (W2W2) defined in Section 3.1.

Proof.

If Ψ0\Psi\equiv 0, then (W1W1) is obvious since Φ(r)=1p|r|p\Phi(r)=\frac{1}{p}{|}r{|}^{p}. Then, since Ψ0\Psi\geq 0 by (Ψ2\Psi 2), the lower bound in (W1W1) is immediate. The upper bound follows by (Ψ1\Psi 1) from

Ψ(detA)\displaystyle\Psi(\det A) =Ψ(detA)Ψ(detI)\displaystyle=\Psi(\det A)-\Psi(\det I)
C(|detA|p21+2)|detA1|\displaystyle\leq C\Big{(}|\det A|^{\tfrac{p}{2}-1}+2\Big{)}|\det A-1|
C(|A|p2+2)(|A|2+1).\displaystyle\leq C^{\prime}(|A|^{p-2}+2)(|A|^{2}+1).

(W2W2) follows by (Ψ1\Psi 1) from

|W(A)W(B)|\displaystyle|W(A)-W(B)|
eB1|Φ(|eA|1)Φ(|eB|1)|+|Ψ(detA)Ψ(detB)|\displaystyle\leq\sum_{e\in B^{1}}\big{|}\Phi\left(|eA|-1\right)-\Phi\left(|eB|-1\right)\big{|}+\big{|}\Psi(\det A)-\Psi(\det B)\big{|}
eB1Cp(|eA|p1+|eB|p1+1)||eA||eB||\displaystyle\leq\sum_{e\in B^{1}}C_{p}\big{(}|eA|^{p-1}+|eB|^{p-1}+1\big{)}\big{|}|eA|-|eB|\big{|}
+C(|detA|p21+|detB|p21+1)|detAdetB|\displaystyle\qquad+C\Big{(}|\det A|^{\tfrac{p}{2}-1}+|\det B|^{\tfrac{p}{2}-1}+1\Big{)}\big{|}\det A-\det B\big{|}
C(|A|p1+|B|p1+1)|AB|+C′′(|A|p2+|B|p2+1)(|A|+|B|)|AB|\displaystyle\leq C^{\prime}\big{(}|A|^{p-1}+|B|^{p-1}+1\big{)}|A-B|+C^{\prime\prime}\big{(}|A|^{p-2}+|B|^{p-2}+1\big{)}(|A|+|B|)|A-B|
C′′′(|A|p1+|B|p1+1)|AB|.\displaystyle\leq C^{\prime\prime\prime}\big{(}|A|^{p-1}+|B|^{p-1}+1\big{)}|A-B|.

For later use (although not necessary for the proof of Theorem 3.1) we elaborate on the rigidity properties of the energy density WW. The rigidity estimate is a direct consequence of Assumptions (W1), (W2). Lengthy computations are postponed to the Appendix.

We start with introducing the singular value decomposition

A=P1ΣP2,where Σ:=[σ100σ2] and P1,P2O(2).\displaystyle A=P_{1}\Sigma P_{2},\quad\text{where }\Sigma:=\begin{bmatrix}\sigma_{1}&0\\ 0&\sigma_{2}\end{bmatrix}\textrm{ and }P_{1},P_{2}\in O(2). (16)

In (16), 0σ1σ20\leq\sigma_{1}\leq\sigma_{2} are the ordered singular values of AA, that is, the square roots of the eigenvalues of the matrix AATAA^{T}. We also recall the definition of the distance function

distp(A,SO(2))=infQSO(2)|AQ|p.\displaystyle\operatorname{dist}^{p}(A,SO(2))=\inf_{Q\in SO(2)}|A-Q|^{p}.
Lemma 4.3.

For WW defined in (8), there exists c2>0c_{2}>0 such that for all A2×2A\in\mathbb{R}^{2\times 2}

W(A)c2distp(A,SO(2)).\displaystyle W(A)\geq c_{2}\operatorname{dist}^{p}(A,SO(2)). (17)
Proof.

Let A=P1ΣP2A=P_{1}\Sigma P_{2} be the singular decomposition as in (16). We split two cases; detA0\det A\geq 0 and detA<0\det A<0. We start with the first case. From the definition of WW we get

W(A)1peB1||Ae|1|p=1peB1||ΣP2e|1|p=1pk=05||ΣRθ+kπ/3e1|1|p,\displaystyle W(A)\geq\frac{1}{p}\sum_{e\in B_{1}}\bigl{|}|Ae|-1\bigr{|}^{p}=\frac{1}{p}\sum_{e\in B_{1}}\bigl{|}|\Sigma P_{2}e|-1\bigr{|}^{p}=\frac{1}{p}\sum_{k=0}^{5}\bigl{|}|\Sigma R_{\theta+k\pi/3}e_{1}|-1\bigr{|}^{p}, (18)

where e1e_{1} is the unit vector, RR is the rotation matrix, and θ[0,π/3)\theta\in[0,\pi/3) is fixed by P2P_{2}. We first consider the case p=2p=2. Thanks to Lemma A.1 we have

14k=05||ΣRθ+kπ/3e1|1|2(σ11)2+(σ21)2.\displaystyle 14\sum_{k=0}^{5}\bigl{|}|\Sigma R_{\theta+k\pi/3}e_{1}|-1\bigr{|}^{2}\geq(\sigma_{1}-1)^{2}+(\sigma_{2}-1)^{2}.

Recalling the well-known relation

(σ11)2+(σ21)2=dist2(A,O(2))(\sigma_{1}-1)^{2}+(\sigma_{2}-1)^{2}=\operatorname{dist}^{2}(A,O(2))

and noting that detA0\det A\geq 0 implies dist(A,O(2))=dist(A,SO(2))\operatorname{dist}(A,O(2))=\operatorname{dist}(A,SO(2)), we obtain (17). For the case p>2p>2, applying Jensen’s inequality in (18) yields

W(A)61p2p|k=05(|ΣRθ+kπ/3e1|1)2|p2.\displaystyle W(A)\geq\frac{6^{1-\frac{p}{2}}}{p}\Bigl{|}\sum_{k=0}^{5}(|\Sigma R_{\theta+k\pi/3}e_{1}|-1)^{2}\Bigr{|}^{\frac{p}{2}}.

Then, (17) follows from the argument above.

We continue with the second case, detA<0\det A<0. By (W1)(W1) in Section 3.1, there exist M>0M>0 independent of AA such that W(A)1M(|A|pM)W(A)\geq\frac{1}{M}(|A|^{p}-M). We separate two cases:

  1. 1.

    If |A|2M|A|\leq 2M, then detA2M2\det A\geq-2M^{2}, and thus

    W(A)Ψ(detA)min[2M2,0]Ψ,W(A)\geq\Psi(\det A)\geq\min_{[-2M^{2},0]}\Psi,

    which is positive by (Ψ2)(\Psi 2). On the other hand, for the right-hand side of (17), we obtain

    distp(A,SO(2))max|B|2Mdistp(B,SO(2))<.\operatorname{dist}^{p}(A,SO(2))\leq\max_{|B|\leq 2M}\operatorname{dist}^{p}(B,SO(2))<\infty.

    Hence, there exists c2>0c_{2}>0 independent of AA such that (17) holds.

  2. 2.

    If |A|>2M|A|>2M, we note from the triangle inequality

    dist(A,SO(2))|A|+dist(0,SO(2))=|A|+2\operatorname{dist}(A,SO(2))\leq|A|+\operatorname{dist}(0,SO(2))=|A|+\sqrt{2}

    that

    W(A)1M|A|pMc2(|A|+2)pc2distp(A,SO(2))W(A)\geq\frac{1}{M}|A|^{p}-M\geq c_{2}(|A|+\sqrt{2})^{p}\geq c_{2}\operatorname{dist}^{p}(A,SO(2))

    for some c2>0c_{2}>0 independent of AA.

4.2 Density of ϕε\mathcal{B}_{\phi}^{\varepsilon} in Wϕ1,p(Ω)W^{1,p}_{\phi}(\Omega)

Proposition 4.4 (Density of ϕε\mathcal{B}_{\phi}^{\varepsilon} in Wϕ1,p(Ω)W^{1,p}_{\phi}(\Omega)).

For all UWϕ1,p(Ω)U\in W^{1,p}_{\phi}(\Omega) there exists UεBϕεU_{\varepsilon}\in B_{\phi}^{\varepsilon} parametrized by 1ε\frac{1}{\varepsilon}\in\mathbb{N} such that UUεW1,p(Ω)0\|U-U_{\varepsilon}\|_{W^{1,p}(\Omega)}\to 0 as ε0\varepsilon\to 0.

Before giving the proof of Proposition 4.4 at the end of this section, we first outline the idea of the proof, and then establish some technical lemmas.

In order to explain the idea of the proof, we first recall two classical density results in the following lemma. To state it, we define Lipϕ(Ω¯):=Lip(Ω¯)Wϕ1,p(Ω)\operatorname{Lip}_{\phi}(\overline{\Omega}):=\operatorname{Lip}(\overline{\Omega})\cap W^{1,p}_{\phi}(\Omega).

Lemma 4.5 (Density of BϕεB_{\phi}^{\varepsilon}).

For any U:Ω2U:\Omega\to\mathbb{R}^{2} with either UW01,p(Ω)U\in W^{1,p}_{0}(\Omega) or UC1(Ω)Lipϕ(Ω¯)U\in C^{1}(\Omega)\cap\operatorname{Lip}_{\phi}(\overline{\Omega}) there exists UεBϕεU_{\varepsilon}\in B_{\phi}^{\varepsilon} parametrized by 1ε\frac{1}{\varepsilon}\in\mathbb{N} such that UUεW1,p(Ω)0\|U-U_{\varepsilon}\|_{W^{1,p}(\Omega)}\to 0 as ε0\varepsilon\to 0.

Proof.

This is a standard result in numerical analysis; see e.g. [ET99, Chap. X, Prop. 2.1, 2.6 and 2.9]. We give the details of the proof to show how our boundary condition fits in.

For UW01,p(Ω)U\in W^{1,p}_{0}(\Omega), it is not restrictive by density to assume that UCc(Ω)U\in C^{\infty}_{c}(\Omega). Then, setting Uε(xi):=U(xi)U_{\varepsilon}(x_{i}):=U(x_{i}) with piecewise affine continuation, it is obvious that UεBϕεU_{\varepsilon}\in B_{\phi}^{\varepsilon} and that UεU\nabla U_{\varepsilon}\to\nabla U uniformly as ε0\varepsilon\to 0.

For UC1(Ω)Lipϕ(Ω¯)U\in C^{1}(\Omega)\cap\operatorname{Lip}_{\phi}(\overline{\Omega}), we set again Uε(xi):=U(xi)U_{\varepsilon}(x_{i}):=U(x_{i}) with piecewise affine continuation, and define Vε:=UεUV_{\varepsilon}:=U_{\varepsilon}-U. Since UC(Ω¯)U\in C(\overline{\Omega}), we have Vε0\|V_{\varepsilon}\|_{\infty}\to 0, and thus VεLp(Ω)0\|V_{\varepsilon}\|_{L^{p}(\Omega)}\to 0. For the gradient, we split Ω=ΩδNδ\Omega=\Omega_{\delta}\cup N_{\delta}, where the disjoint sets Ωδ\Omega_{\delta} and NδN_{\delta} are such that Ωδ¯Ω\overline{\Omega_{\delta}}\subset\Omega and NδN_{\delta} has volume that vanishes as δ0\delta\to 0. Then,

VεLp(Ω)p=VεLp(Ωδ)p+VεLp(Nδ)p.\|\nabla V_{\varepsilon}\|_{L^{p}(\Omega)}^{p}=\|\nabla V_{\varepsilon}\|_{L^{p}(\Omega_{\delta})}^{p}+\|\nabla V_{\varepsilon}\|_{L^{p}(N_{\delta})}^{p}.

First, since UC(Ωδ¯)\nabla U\in C(\overline{\Omega_{\delta}}), we have by the argument above that VεC(Ωδ¯)0\|\nabla V_{\varepsilon}\|_{C(\overline{\Omega_{\delta}})}\to 0, and thus VεLp(Ωδ)p0\|\nabla V_{\varepsilon}\|_{L^{p}(\Omega_{\delta})}^{p}\to 0 as ε0\varepsilon\to 0. Second, since ULipϕ(Ω¯)U\in\operatorname{Lip}_{\phi}(\overline{\Omega}), UεU<\|\nabla U_{\varepsilon}\|_{\infty}\leq\|\nabla U\|_{\infty}<\infty, and thus VεLp(Nδ)pC|Nδ|\|\nabla V_{\varepsilon}\|_{L^{p}(N_{\delta})}^{p}\leq C|N_{\delta}| uniformly in ε\varepsilon. Hence, by taking ε\varepsilon small enough with respect to δ\delta and δ0\delta\to 0, we conclude that VεLp(Ω)0\|\nabla V_{\varepsilon}\|_{L^{p}(\Omega)}\to 0 as ε0\varepsilon\to 0. ∎

Thanks to Lemma 4.5, the proof of Proposition 4.4 narrows down to constructing a decomposition U=U1+U2+U3U=U_{1}+U_{2}+U_{3} where

{U1C1(Ω)Lipϕ(Ω¯),U2W01,p(Ω),U3W1,p(Ω) is small.\left\{\begin{aligned} &U_{1}\in C^{1}({\Omega})\cap\operatorname{Lip}_{\phi}(\overline{\Omega}),\\ &U_{2}\in W_{0}^{1,p}(\Omega),\\ &\|U_{3}\|_{W^{1,p}(\Omega)}\text{ is small.}\end{aligned}\right. (19)

Indeed, if such a decomposition exists, then Lemma 4.5 provides approximations in BϕεB_{\phi}^{\varepsilon} of U1U_{1} and U2U_{2}, and U3U_{3} can simply be approximated by 0. The difficulty in constructing such a decomposition is in finding a U1U_{1} for which U1|ΓU_{1}|_{\Gamma} is sufficiently close to U|ΓU|_{\Gamma}. Approximation by convolution does not work directly since U1U_{1} has to satisfy the boundary condition in Lipϕ(Ω¯)\operatorname{Lip}_{\phi}(\overline{\Omega}). Instead, we use the Trace Theorem to approximate U|ΓU|_{\Gamma} in an appropriate function space on Γ\Gamma. This approximation is based on convolution, but care is needed because of the boundary condition, the corners of Γ\Gamma and the fact that the norm of the function space on Γ\Gamma is nonlocal.

Next, we prepare for proving Proposition 4.4 by citing a Trace Theorem (Lemma 4.6) and proving a density result on Γ\Gamma (Lemma 4.9). To avoid technical difficulties with the corners in Γ\Gamma, we first transform Ω\Omega to the unit disc 𝔻={x2:|x|<1}\mathbb{D}=\{x\in\mathbb{R}^{2}:|x|<1\}. With this aim, let φ:Ω¯𝔻¯\varphi:\overline{\Omega}\to\overline{\mathbb{D}} be a related transformation (see Figure 3) such that

  • φ\varphi is bi-Lipschitz;

  • φC1(Ω)\varphi\in C^{1}(\Omega);

  • φ(Γi)=γi\varphi(\Gamma_{i})=\gamma_{i} for i=1,2,3i=1,2,3, where γi𝔻\gamma_{i}\subset\partial\mathbb{D} are given, in terms of the polar angle coordinate, by

    γ1:=[0,2π3],γ2:=[2π3,0],γ3:=[2π3,4π3];\gamma_{1}:=[0,\tfrac{2\pi}{3}],\quad\gamma_{2}:=[-\tfrac{2\pi}{3},0],\quad\gamma_{3}:=[\tfrac{2\pi}{3},\tfrac{4\pi}{3}];
  • if UWϕ1,p(Ω)U\in W_{\phi}^{1,p}(\Omega), then

    Uφ1Wϕ1,p(𝔻):={VW1,p(𝔻):V(θ)=RϕV(θ) for a.e.  0<θ<2π3}.U\circ\varphi^{-1}\in W_{\phi}^{1,p}(\mathbb{D}):=\{V\in W^{1,p}(\mathbb{D}):V(-\theta)=R_{\phi}V(\theta)\>\text{ for a.e.\ }\,0<\theta<\tfrac{2\pi}{3}\}.
Γ1\Gamma_{1}11Γ3\Gamma_{3}Γ2\Gamma_{2}Ω\Omegaφ\varphi𝔻\mathbb{D}γ3\gamma_{3}θ\theta11γ1\gamma_{1}γ2\gamma_{2}
Figure 3: The deformation φ\varphi and the related sections of the boundaries of Ω\Omega and 𝔻\mathbb{D}.

Above and in the following, we will often identify the unit circle 𝕊:=𝔻\mathbb{S}:=\partial\mathbb{D} with the periodic interval [0,2π)[0,2\pi). We also adopt the convention that a subscript ϕ\phi in a function space indicates the boundary condition. We note that there are constants 0<c,C0<c,C such that for all UWϕ1,p(Ω)U\in W_{\phi}^{1,p}(\Omega)

cUW1,p(Ω)Uφ1W1,p(𝔻)CUW1,p(Ω).c\|U\|_{W^{1,p}(\Omega)}\leq\|U\circ\varphi^{-1}\|_{W^{1,p}(\mathbb{D})}\leq C\|U\|_{W^{1,p}(\Omega)}.

Hence, it is sufficient to construct the decomposition of UU after transforming it to Wϕ1,p(𝔻)W_{\phi}^{1,p}(\mathbb{D}).

To cite the Trace Theorem, we fix s=11ps=1-\frac{1}{p} and recall the usual norm of the fractional Sobolev space Ws,p(𝕊)W^{s,p}(\mathbb{S}):

UWs,p(𝕊)p\displaystyle\|U\|_{W^{s,p}(\mathbb{S})}^{p} :=ULp(𝕊)p+[U]Ws,p(𝕊)p,\displaystyle:=\|U\|_{L^{p}(\mathbb{S})}^{p}+[U]_{W^{s,p}(\mathbb{S})}^{p},
[U]Ws,p(𝕊)p\displaystyle[U]_{W^{s,p}(\mathbb{S})}^{p} :=02π02π|U(θ)U(ρ)|p|θρ|𝕊p𝑑ρ𝑑θ,|θ|𝕊=mink|θ2πk|.\displaystyle:=\int_{0}^{2\pi}\int_{0}^{2\pi}\frac{|U(\theta)-U(\rho)|^{p}}{|\theta-\rho|_{\mathbb{S}}^{p}}\,d\rho d\theta,\qquad|\theta|_{\mathbb{S}}=\min_{k\in\mathbb{Z}}|\theta-2\pi k|.
Lemma 4.6 (Trace [Gag57, Thm. 1.I]).

There exists a C>0C>0 such that for all UW1,p(𝔻)U\in W^{1,p}(\mathbb{D})

U|𝕊Ws,p(𝕊)CUW1,p(𝔻).\big{\|}U|_{\mathbb{S}}\big{\|}_{W^{s,p}(\mathbb{S})}\leq C\|U\|_{W^{1,p}(\mathbb{D})}.

Conversely, there exists a C>0C>0 such that for all fWs,p(𝕊)f\in W^{s,p}(\mathbb{S}) there exists a UW1,p(𝔻)U\in W^{1,p}(\mathbb{D}) with U|𝕊=fU|_{\mathbb{S}}=f such that

UW1,p(𝔻)CfWs,p(𝕊).\|U\|_{W^{1,p}(\mathbb{D})}\leq C\|f\|_{W^{s,p}(\mathbb{S})}.

In order to prove a density result in Wϕs,p(𝕊)W_{\phi}^{s,p}(\mathbb{S}), we first recall two technical lemmas:

Lemma 4.7 ([DNPV12], Lem. 5.2).

If fWs,p(0,π)f\in W^{s,p}(0,\pi), then the even extension f~\tilde{f} satisfies

f~Ws,p(π,π)CfWs,p(0,π)\|\tilde{f}\|_{W^{s,p}(-\pi,\pi)}\leq C\|f\|_{W^{s,p}(0,\pi)}

for some universal constant C>0C>0.

Lemma 4.8 (Continuity of the translation operator).

The translation operator τaf(θ):=f(θ+a)\tau_{a}f(\theta):=f(\theta+a) is continuous in the strong Ws,p(𝕊)W^{s,p}(\mathbb{S}) topology, that is, for all fWs,p(𝕊)f\in W^{s,p}(\mathbb{S}),

lima0τaffWs,p(𝕊)=0.\lim_{a\to 0}\|\tau_{a}f-f\|_{W^{s,p}(\mathbb{S})}=0.
Proof.

The proof is standard; we give a sketch. The statement is obvious for fC1(𝕊)f\in C^{1}(\mathbb{S}). For general fWs,p(𝕊)f\in W^{s,p}(\mathbb{S}), it suffices to approximate it by ψC1(𝕊)\psi\in C^{1}(\mathbb{S}), and note that

τaψτafWs,p(𝕊)=τa(ψf)Ws,p(𝕊)=ψfWs,p(𝕊).\|\tau_{a}\psi-\tau_{a}f\|_{W^{s,p}(\mathbb{S})}=\|\tau_{a}(\psi-f)\|_{W^{s,p}(\mathbb{S})}=\|\psi-f\|_{W^{s,p}(\mathbb{S})}.

Lemma 4.9 (Approximation on 𝕊\mathbb{S}).

Cϕ(𝕊)C_{\phi}^{\infty}(\mathbb{S}) is dense in Wϕs,p(𝕊)W_{\phi}^{s,p}(\mathbb{S}).

Proof.

Let fWϕs,p(𝕊)f\in W_{\phi}^{s,p}(\mathbb{S}). We split f=f1+f2+f3f=f_{1}+f_{2}+f_{3} with fiWs,p(0,2π)f_{i}\in W^{s,p}(0,2\pi) such that

f1={fon (0,5π6)0on (7π6,2π),f2={0on (0,5π6)fon (7π6,2π).f_{1}=\left\{\begin{array}[]{ll}f&\text{on }(0,\tfrac{5\pi}{6})\\ 0&\text{on }(\tfrac{7\pi}{6},2\pi),\end{array}\right.\qquad f_{2}=\left\{\begin{array}[]{ll}0&\text{on }(0,\tfrac{5\pi}{6})\\ f&\text{on }(\tfrac{7\pi}{6},2\pi).\end{array}\right.

Note that f3Wϕs,p(𝕊)f_{3}\in W_{\phi}^{s,p}(\mathbb{S}) with suppf3[5π6,7π6]γ3\operatorname{supp}f_{3}\subset[\tfrac{5\pi}{6},\tfrac{7\pi}{6}]\subset\subset\gamma_{3}, and that it is not directly clear that f1,f2Ws,p(𝕊)f_{1},f_{2}\in W^{s,p}(\mathbb{S}). We will construct approximating sequences (ψ12k)k,(ψ3k)kC(𝕊)(\psi_{12}^{k})_{k},(\psi_{3}^{k})_{k}\subset C^{\infty}(\mathbb{S}) of f1+f2f_{1}+f_{2} and f3f_{3} respectively such that ψk:=ψ12k+ψ3kCϕ(𝕊)\psi^{k}:=\psi_{12}^{k}+\psi_{3}^{k}\in C_{\phi}^{\infty}(\mathbb{S}). By construction of f3f_{3}, ψ3k:=ηkf3\psi_{3}^{k}:=\eta_{k}*f_{3} is a suitable approximating sequence, where ηk\eta_{k} is the usual mollifier.

To construct ψ12k\psi_{12}^{k}, we first show that f1,f2Ws,p(𝕊)f_{1},f_{2}\in W^{s,p}(\mathbb{S}). Since by construction f1,f2Ws,p(0,2π)f_{1},f_{2}\in W^{s,p}(0,2\pi), it is sufficient to show that f1,f2Ws,p(r,r)f_{1},f_{2}\in W^{s,p}(-r,r) for some r>0r>0. We start with f1f_{1}. For any θ𝕊\theta\in\mathbb{S} and any parameter 0απ0\leq\alpha\leq\pi, let

f1α(θ):={Rαf1(θ)π<θ<0f1(θ)0<θ<π.f_{1}^{\alpha}(\theta):=\left\{\begin{array}[]{ll}R_{\alpha}f_{1}(-\theta)&-\pi<\theta<0\\ f_{1}(\theta)&0<\theta<\pi.\end{array}\right.

Clearly, f1αWs,p(0,π)Ws,p(π,2π)f_{1}^{\alpha}\in W^{s,p}(0,\pi)\cap W^{s,p}(\pi,2\pi) for any α\alpha. Since f1ϕ(θ)=f(θ)f_{1}^{\phi}(\theta)=f(\theta) for a.e. θ(5π6,5π6)\theta\in(-\tfrac{5\pi}{6},\tfrac{5\pi}{6}) and fWϕs,p(𝕊)f\in W_{\phi}^{s,p}(\mathbb{S}), we also have f1ϕWs,p(5π6,5π6)f_{1}^{\phi}\in W^{s,p}(-\tfrac{5\pi}{6},\tfrac{5\pi}{6}). Since f10f_{1}^{0} is the even extension of f1|(0,π)f_{1}|_{(0,\pi)}, we obtain from Lemma 4.7 that f10Ws,p(π,π)f_{1}^{0}\in W^{s,p}(-\pi,\pi). Hence, for the odd extension f1πf_{1}^{\pi}, we find from the linear relation

f1ϕ=I+Rϕ2f10+IRϕ2f1πf_{1}^{\phi}=\frac{I+R_{\phi}}{2}f_{1}^{0}+\frac{I-R_{\phi}}{2}f_{1}^{\pi}

that f1πWs,p(5π6,5π6)f_{1}^{\pi}\in W^{s,p}(-\tfrac{5\pi}{6},\tfrac{5\pi}{6}). Finally,

f1=f10+f1π2Ws,p(5π6,5π6).f_{1}=\frac{f_{1}^{0}+f_{1}^{\pi}}{2}\in W^{s,p}(-\tfrac{5\pi}{6},\tfrac{5\pi}{6}).

The proof of f2Ws,p(5π6,5π6)f_{2}\in W^{s,p}(-\tfrac{5\pi}{6},\tfrac{5\pi}{6}) is analogous.

Finally we construct the approximating sequences ψ12k\psi_{12}^{k}. Care is needed to ensure that ψ12k\psi_{12}^{k} satisfies the boundary condition. With this aim, we first approximate f1f_{1} and f2f_{2} by the translations τaf1\tau_{a}f_{1} and τaf2\tau_{-a}f_{2} with 0<a<π60<a<\frac{\pi}{6}; see Lemma 4.8. Note that τaf1+τaf2Wϕs,p(𝕊)\tau_{a}f_{1}+\tau_{-a}f_{2}\in W_{\phi}^{s,p}(\mathbb{S}) and (τaf1+τaf2)|(a,a)0(\tau_{a}f_{1}+\tau_{-a}f_{2})|_{(-a,a)}\equiv 0. Then, we define ψ12k:=ηk(τakf1+τakf2)\psi_{12}^{k}:=\eta_{k}*(\tau_{a_{k}}f_{1}+\tau_{-{a_{k}}}f_{2}) for some ak0a_{k}\to 0 such that suppηk(ak,ak)\operatorname{supp}\eta_{k}\subset(-a_{k},a_{k}), and note that ψ12kCϕ(𝕊)\psi_{12}^{k}\in C_{\phi}^{\infty}(\mathbb{S}). This completes the proof. ∎

We are ready to prove Proposition 4.4.

Proof of Proposition 4.4.

Step 1: decomposition of UU. Let δ>0\delta>0 and UW1,pϕ(Ω)U\in W^{1,p}_{\phi}(\Omega) be given. Set U~:=Uφ1W1,pϕ(𝔻)\tilde{U}:=U\circ\varphi^{-1}\in W^{1,p}_{\phi}(\mathbb{D}) and f~:=U~|𝕊\tilde{f}:=\tilde{U}|_{\mathbb{S}}. Then, by Lemma 4.6 we have f~Wϕs,p(𝕊)\tilde{f}\in W_{\phi}^{s,p}(\mathbb{S}). By Lemma 4.9 we find ψ~Cϕ(𝕊)\tilde{\psi}\in C_{\phi}^{\infty}(\mathbb{S}) with ψ~f~Ws,p(𝕊)<δ\|\tilde{\psi}-\tilde{f}\|_{W^{s,p}(\mathbb{S})}<\delta. By Lemma 4.6, there exists U~3W1,pϕ(𝔻)\tilde{U}_{3}\in W^{1,p}_{\phi}(\mathbb{D}) with U~3|𝕊=f~ψ~\tilde{U}_{3}|_{\mathbb{S}}=\tilde{f}-\tilde{\psi} and

U~3W1,p(𝔻)Cψ~f~Ws,p(𝕊)Cδ\|\tilde{U}_{3}\|_{W^{1,p}(\mathbb{D})}\leq C\|\tilde{\psi}-\tilde{f}\|_{W^{s,p}(\mathbb{S})}\leq C\delta

for some C>0C>0 independent of δ\delta.

Next, we take U~1Cϕ(𝔻)\tilde{U}_{1}\in C_{\phi}^{\infty}(\mathbb{D}) as the harmonic extension of ψ~\tilde{\psi}. Then, translating back to Ω\Omega, we set Ui:=U~iφU_{i}:=\tilde{U}_{i}\circ\varphi for i=1,3i=1,3. Taking U2:=UU1U3U_{2}:=U-U_{1}-U_{3}, we observe that UiU_{i} satisfy (19) for i=1,2,3i=1,2,3.

Step 2: Construction of UεU_{\varepsilon}. By Lemma 4.5 we find sequences (U1ε)ε,(U2ε)ε(U_{1}^{\varepsilon})_{\varepsilon},(U_{2}^{\varepsilon})_{\varepsilon} parametrized by 1ε\frac{1}{\varepsilon}\in\mathbb{N} such that UiεBϕεU_{i}^{\varepsilon}\in B_{\phi}^{\varepsilon} and

UiUiεW1,p(Ω)ε00\|U_{i}-U_{i}^{\varepsilon}\|_{W^{1,p}(\Omega)}\xrightarrow{\varepsilon\to 0}0

for i=1,2i=1,2. Hence, setting Uε:=U1ε+U2εU_{\varepsilon}:=U_{1}^{\varepsilon}+U_{2}^{\varepsilon} and taking ε\varepsilon small enough with respect to δ\delta, we obtain

UUεW1,p(Ω)U1U1εW1,p(Ω)+U2U2εW1,p(Ω)+U3W1,p(Ω)Cδ\|U-U_{\varepsilon}\|_{W^{1,p}(\Omega)}\leq\|U_{1}-U_{1}^{\varepsilon}\|_{W^{1,p}(\Omega)}+\|U_{2}-U_{2}^{\varepsilon}\|_{W^{1,p}(\Omega)}+\|U_{3}\|_{W^{1,p}(\Omega)}\leq C\delta

for some CC independent of UiU_{i}, δ\delta or ε\varepsilon. Since δ>0\delta>0 is arbitrary, we conclude that UUεW1,p(Ω)0\|U-U_{\varepsilon}\|_{W^{1,p}(\Omega)}\to 0 as ε0\varepsilon\to 0. ∎

5 Physical interpretation of Theorem 3.1

We show that, despite the relaxation process implied by Γ\Gamma-convergence, the equilibrium solution of the model are necessarily stressed. This is a consequence of the rotational boundary conditions incorporated into W1,pϕ(Ω)W^{1,p}_{\phi}(\Omega) and of the finite penalization to folding imposed by Ψ\Psi.

Proposition 5.1.
minLp(Ω)E(U)>0.\displaystyle\min_{L^{p}(\Omega)}E(U)>0. (20)

The proof of Proposition 5.1 follows the lines of [KM18, Sec. 5]. For the readers’ convenience, we display the main steps.

Proof.

First, we show

QW(F)cdistp(F,SO(2)),QW(F)\geq c\operatorname{dist}^{p}(F,SO(2)), (21)

for some c>0c>0. From Lemma 4.3, we have W(F)c2distp(F,SO(2))W(F)\geq c_{2}\operatorname{dist}^{p}(F,SO(2)) for every F2×2F\in\mathbb{R}^{2\times 2}. Now, thanks to (12), for every ε>0\varepsilon>0 and every smooth open ω2\omega\subset\mathbb{R}^{2} with |ω|=1|\omega|=1, there exists a map ϑεW1,0(ω,2)\vartheta_{\varepsilon}\in W^{1,\infty}_{0}(\omega,\mathbb{R}^{2}) (in fact even Cc(ω,2)C^{\infty}_{c}(\omega,\mathbb{R}^{2})) such that

QW(F)ωW(F+ϑε(x))dxεc2ωdistp(F+ϑε(x),SO(2))dxε.QW(F)\geq\int_{\omega}W(F+\nabla\vartheta_{\varepsilon}(x))dx-\varepsilon\geq c_{2}\int_{\omega}\operatorname{dist}^{p}(F+\nabla\vartheta_{\varepsilon}(x),SO(2))dx-\varepsilon.

We now invoke Rigidity Theorem 3.1 [FJM02] (which applies for p2p\geq 2, see [CS06, Sec. 2.4]) yielding the existence of a constant c>0c>0 and a matrix RSO(2)R\in SO(2) such that

c2ωdistp(F+ϑε(x),SO(2))dxcω|F+ϑε(x)R|pdx.\displaystyle c_{2}\int_{\omega}\operatorname{dist}^{p}(F+\nabla\vartheta_{\varepsilon}(x),SO(2))dx\geq c\int_{\omega}|F+\nabla\vartheta_{\varepsilon}(x)-R|^{p}dx. (22)

Now, interpreting ||p:2×2|\cdot|^{p}:\mathbb{R}^{2\times 2}\to\mathbb{R} as a convex function on 4\mathbb{R}^{4}, we obtain

|(FR)+ϑε(x)|p|FR|p+p|FR|p2(FR):ϑε(x)\displaystyle|(F-R)+\nabla\vartheta_{\varepsilon}(x)|^{p}\geq|F-R|^{p}+p|F-R|^{p-2}(F-R):\nabla\vartheta_{\varepsilon}(x) (23)

for all xωx\in\omega. By applying (23) to (22) we have

ωdistp(F+ϑε,SO(2))dxω|FR|pdx,\displaystyle\int_{\omega}\operatorname{dist}^{p}(F+\nabla\vartheta_{\varepsilon},SO(2))dx\geq\int_{\omega}|F-R|^{p}dx,

because the integral of (FR):ϑε(F-R):\nabla\vartheta_{\varepsilon} vanishes as ϑε\vartheta_{\varepsilon} has zero boundary datum. Putting our estimates together and minimising over RR, we get

QW(F)cωdistp(F,SO(2))dxεQW(F)\geq c\int_{\omega}\operatorname{dist}^{p}(F,SO(2))dx-\varepsilon

and the desired result in (21) follows by sending ε0\varepsilon\to 0.

Applying (21) we obtain

minW1,pϕ(Ω)(U)=(U¯)Ωdistp(U¯,SO(2))dx.\displaystyle\min_{W^{1,p}_{\phi}(\Omega)}\mathcal{E}(U)=\mathcal{E}(\overline{U})\geq\int_{\Omega}\operatorname{dist}^{p}(\nabla\overline{U},SO(2))dx.

It is left to prove that the right-hand side is positive. Suppose instead that it is 0. Then, U¯SO(2)\nabla\overline{U}\in SO(2) a.e. in Ω\Omega. Hence, U¯\nabla\overline{U} is a constant rotation (see [BJ87]). This contradicts with U¯W1,pϕ(Ω)\overline{U}\in W^{1,p}_{\phi}(\Omega). ∎

6 Numerical computations

In this section we explore numerically several energy wells of EεE_{\varepsilon} and inspect whether the obtained minimizers satisfies detUε>0\det\nabla U_{\varepsilon}>0. We do this for the simplified setting given by Ψ0\Psi\equiv 0, which does not penalize folded patterns. Fortunately, this simplified setting turns out to be stable enough to reproduce configurations with detUε>0\det\nabla U_{\varepsilon}>0 (see, e.g., Figure 1) as local minima. Yet, local minimizers are of limited use to our theoretical result Theorem 3.1, because Γ\Gamma-convergence only guarantees the convergence of global minimizers in the limit ε0\varepsilon\to 0. Instead, we show by direct inspection of the energy values at the computed local minimizers that there is a good agreement with available continuum models for disclinations for decreasing values of ε\varepsilon. We compute local minimizers by employing Newton’s method for different initial conditions. In all computations below, Newton’s method converges quadratically.

6.1 The limit ε0\varepsilon\to 0

Here we verify and quantify some of the computations in [SN88], in which disclinations satisfying detUε>0\det\nabla U_{\varepsilon}>0 are obtained. The only difference with [SN88] is in the definition of ww in (4); in [SN88] w1w\equiv 1 is taken instead.

For both ϕ=2π5\phi=\frac{2\pi}{5} and ϕ=2π7\phi=\frac{2\pi}{7} we chose the initial condition such that the local minimizer of EεE_{\varepsilon} satisfies detUε>0\det\nabla U_{\varepsilon}>0. With this aim, for ε=21\varepsilon=2^{-1} we took as initial condition one of the linear deformations Uε(x)=AxU_{\varepsilon}(x)=Ax in BϕεB_{\phi}^{\varepsilon} with detA=1\det A=1 (see Remark 2.1). For the subsequent values ε=22,23,,28\varepsilon=2^{-2},2^{-3},\ldots,2^{-8}, we constructed the initial condition from the minimizer obtained for the previous value of ε\varepsilon by linear interpolation. The resulting local minimizers U¯ε\overline{U}_{\varepsilon} turn out to satisfy detU¯ε>0\det\nabla\overline{U}_{\varepsilon}>0. For ε=23\varepsilon=2^{-3}, U¯ε\overline{U}_{\varepsilon} is illustrated in Figure 1. The energy values eε:=Eε(U¯ε)e_{\varepsilon}:=E_{\varepsilon}(\overline{U}_{\varepsilon}) are plotted in Figure 4. The behavior of eεe_{\varepsilon} as ε0\varepsilon\to 0 is qualitatively similar to the computations of [SN88] which are based on a model in linearized elasticity.

Next, we test the convergence of eεe_{\varepsilon} as ε0\varepsilon\to 0. Indeed, since U¯ε\overline{U}_{\varepsilon} need not be global minimizers, the Γ\Gamma-convergence result in Theorem 3.1 does not guarantee convergence. To test the convergence, we impose the power law ansatz

eεC1C22p,e_{\varepsilon}\sim C_{1}-C_{2}2^{p}, (24)

where the constants C1,C2,p>0C_{1},C_{2},p>0 need to be fitted from the data. Without computing C1C_{1} and C2C_{2}, we obtain the exponent pp of the power law numerically from

pε:=1log2loge4εe2εe2εeε.p_{\varepsilon}:=\frac{1}{\log 2}\log\frac{e_{4\varepsilon}-e_{2\varepsilon}}{e_{2\varepsilon}-e_{\varepsilon}}. (25)

Table 1 lists the values of pεp_{\varepsilon}. Since the values of pεp_{\varepsilon} are positive and increasing with logε-\log\varepsilon, eεe_{\varepsilon} seems to converge as ε0\varepsilon\to 0 implying a power law behavior.

ϕ\phi
ε\varepsilon 2π5\tfrac{2\pi}{5} 2π7\tfrac{2\pi}{7}
232^{-3} 1.552 1.276
242^{-4} 1.712 1.488
252^{-5} 1.812 1.595
262^{-6} 1.866 1.645
272^{-7} 1.898 1.671
282^{-8} 1.918 1.686
Table 1: The values of pεp_{\varepsilon} (see (25) as computed from the numerical data to test the power law in the ansatz in (24)).
Refer to caption212^{-1}232^{-3}252^{-5}272^{-7}ε\varepsilon112233eε×103e_{\varepsilon}\times 10^{3}ϕ=2π5\phi=\tfrac{2\pi}{5}ϕ=2π7\phi=\tfrac{2\pi}{7}
Figure 4: The energy values eε=Eε(U¯ε)e_{\varepsilon}=E_{\varepsilon}(\overline{U}_{\varepsilon}) obtained from the simulations. Here, U¯ε\overline{U}_{\varepsilon} is verified to satisfy detU¯ε>0\det\nabla\overline{U}_{\varepsilon}>0.

6.2 Folded configurations

In the next simulations we explore other local minimizers of the energy by starting Newton’s method from other initial conditions. In particular, since we take Ψ0\Psi\equiv 0, there is no penalisation on detUε0\det\nabla U_{\varepsilon}\leq 0, and thus local minimizers with folded patterns may occur.

With this aim, we set ε=22\varepsilon=2^{-2} and fold the reference lattice as illustrated in Figure 5. This folding procedure is done such that the following property used in the boundary condition in (2)

Γ2𝒱ε={Rπ/3x:xΓ1𝒱ε}\Gamma_{2}\cap\mathcal{V}_{\varepsilon}=\{R_{\pi/3}x:x\in\Gamma_{1}\cap\mathcal{V}_{\varepsilon}\}

is conserved during the folding process. After folding, we construct the initial condition for Newton’s method by deforming the folded reference domain by the linear map defined in Remark 2.1.

Ω\Omegafold=1\ell=1fold=2\ell=2fold=3\ell=3
Figure 5: The folding procedure of the reference domain to construct the initial conditions. Lines along which we fold are drawn in boldface.

Figures 7 and 8 illustrate the local minimizers obtained for each folded pattern shown in Figure 5 for ϕ=2π5\phi=\frac{2\pi}{5} and ϕ=2π7\phi=\frac{2\pi}{7} respectively. Figure 6 shows the related energy values. It is clear that the folded local minimizers have lower energy. In fact, it seems that the energy values decrease in an affine manner with the number of folds, and that, after linear extrapolation, 0 energy would be obtained around 4 folds.

The observation that EεE_{\varepsilon} has folded patterns as local minima with low energy is not unexpected from the expression of the continuum energy EE. Indeed, Ψ=0\Psi=0 implies that the minimization problem of EE lacks rigidity; Lemma 4.3 and Proposition 5.1 are not valid anymore. In fact, if we substitute Ψ=0\Psi=0 into (8), we obtain (see Lemma A.2) QW(0)=0QW(0)=0, which implies that 0Lp(Ω)0\in L^{p}(\Omega) is a global minimizer of EE. This is consistent with the observation in Figure 6, where the energy values decay with the number of folds.

Refer to caption0112233number of folds01122eε×103e_{\varepsilon}\times 10^{3}ϕ=2π5\phi=\tfrac{2\pi}{5}ϕ=2π7\phi=\tfrac{2\pi}{7}
Figure 6: The energy values eε=Eε(U¯ε)e_{\varepsilon}=E_{\varepsilon}(\overline{U}_{\varepsilon}) for ε=22\varepsilon=2^{-2} computed from folded initial conditions (see Figure 5).
Refer to caption1-10111-10110 foldsRefer to caption1-10111-101111 foldRefer to caption1-10111-101122 foldsRefer to caption1-10111-101133 folds
Figure 7: Minimizers of EεE_{\varepsilon} obtained from the folded initial conditions shown in Figure 5.
Refer to caption1-10111-10110 foldsRefer to caption1-10111-101111 foldRefer to caption1-10111-101122 foldsRefer to caption1-10111-101133 folds
Figure 8: Minimizers of EεE_{\varepsilon} obtained from the folded initial conditions shown in Figure 5.

7 Future perspectives

We develop a continuum model (the energy EE in (10)) for the energy of a disclination in a planar geometry. We construct EE rigorously from an atomistic model (Theorem 3.1), and show that any minimizer of EE corresponds to a stressed configuration of the medium (Proposition 5.1). We intend this first rigorous study as a stepping stone towards a general variational theory capable of describing both morphology and energetics of disclinations and their effect on the macrosocopic properties of a material. By analyzing the interaction of mismatches and distortions on the hexagonal lattice, we take a first step towards the investigation of the interaction of defects with the lattice kinematics, thus opening a possible path to modeling more complex systems such as austenite-martensite microstructures in Shape-Memory Alloys and kink formations as mentioned in the introduction.

Acknowledgements

P.C. is supported by JSPS Grant-in-Aid for Young Scientists (B) 16K21213 and by JSPS Innovative Area Grant 19H05131. P.C. holds an honorary appointment at La Trobe University and is a member of GNAMPA. The authors acknowledge the Research Institute for Mathematical Sciences, an International Joint Usage and Research Center located in Kyoto University, where part of the work contained in this paper was carried out.

Appendix A Appendix

Lemma A.1.

For all 0σ1σ20\leq\sigma_{1}\leq\sigma_{2} and all θ\theta\in\mathbb{R},

14k=05(|ΣRθ+kπ/3e1|1)2(σ11)2+(σ21)2,where Σ:=[σ100σ2].\displaystyle 14\sum_{k=0}^{5}(|\Sigma R_{\theta+k\pi/3}e_{1}|-1)^{2}\geq(\sigma_{1}-1)^{2}+(\sigma_{2}-1)^{2},\quad\text{where }\Sigma:=\begin{bmatrix}\sigma_{1}&0\\ 0&\sigma_{2}\end{bmatrix}. (26)
Proof.

We separate 2 cases: σ12+σ222\sigma_{1}^{2}+\sigma_{2}^{2}\geq 2 and σ12+σ222\sigma_{1}^{2}+\sigma_{2}^{2}\leq 2. In the first case, we start with bounding the right-hand side of (26) from above. If σ11\sigma_{1}\geq 1, then clearly (σ11)2+(σ21)22(σ21)2(\sigma_{1}-1)^{2}+(\sigma_{2}-1)^{2}\leq 2(\sigma_{2}-1)^{2}. If σ11\sigma_{1}\leq 1, then from the observation that the line segment between the points (0,2)(0,\sqrt{2}) and (1,1)(1,1) on the circle 2𝕊\sqrt{2}\mathbb{S} is below the arc of 2𝕊\sqrt{2}\mathbb{S} between the same two points, we obtain that σ21+(21)(1σ1)\sigma_{2}\geq 1+(\sqrt{2}-1)(1-\sigma_{1}). Hence,

01σ1σ2121=(2+1)(σ21).0\leq 1-\sigma_{1}\leq\frac{\sigma_{2}-1}{\sqrt{2}-1}=(\sqrt{2}+1)(\sigma_{2}-1).

This together with the bound for σ11\sigma_{1}\geq 1 we get

(σ11)2+(σ21)2<7(σ21)2.(\sigma_{1}-1)^{2}+(\sigma_{2}-1)^{2}<7(\sigma_{2}-1)^{2}. (27)

We continue by bounding the left-hand side in (26) from below. Writing

|ΣRθ+kπ/3e1|2=σ12cos2(θ+kπ/3)+σ22sin2(θ+kπ/3),|\Sigma R_{\theta+k\pi/3}e_{1}|^{2}=\sigma_{1}^{2}\cos^{2}(\theta+k\pi/3)+\sigma_{2}^{2}\sin^{2}(\theta+k\pi/3),

we note from the facts that αsin2α\alpha\mapsto\sin^{2}\alpha is π\pi-periodic and {α[0,π]:sin2α34}=[π/3,2π/3]\{\alpha\in[0,\pi]:\sin^{2}\alpha\geq\frac{3}{4}\}=[\pi/3,2\pi/3] that there are at least two values for k{0,,5}k\in\{0,\ldots,5\} for which sin2(θ+kπ/3)34\sin^{2}(\theta+k\pi/3)\geq\frac{3}{4}. For these values of kk, we have by σ2max{1,σ1}\sigma_{2}\geq\max\{1,\sigma_{1}\} and σ12+σ222\sigma_{1}^{2}+\sigma_{2}^{2}\geq 2 that

|ΣRθ+kπ/3e1|214σ12+34σ2212σ22+121,|\Sigma R_{\theta+k\pi/3}e_{1}|^{2}\geq\frac{1}{4}\sigma_{1}^{2}+\frac{3}{4}\sigma_{2}^{2}\geq\frac{1}{2}\sigma_{2}^{2}+\frac{1}{2}\geq 1,

and thus

14k=05(|ΣRθ+kπ/3e1|1)228(12σ22+11)2.14\sum_{k=0}^{5}(|\Sigma R_{\theta+k\pi/3}e_{1}|-1)^{2}\geq 28\Big{(}\frac{1}{\sqrt{2}}\sqrt{\sigma_{2}^{2}+1}-1\Big{)}^{2}. (28)

Estimating the convex function f(x)=x2+1f(x)=\sqrt{x^{2}+1} from below by its tangent at x=1x=1 yields

σ22+12+(σ21)/2.\displaystyle\sqrt{\sigma_{2}^{2}+1}\geq\sqrt{2}+(\sigma_{2}-1)/\sqrt{2}.

Plugging this estimate into (28), we observe that the resulting lower bound equals the upper bound in (27). This completes the proof in the case σ12+σ222\sigma_{1}^{2}+\sigma_{2}^{2}\geq 2.

In the second case, σ12+σ222\sigma_{1}^{2}+\sigma_{2}^{2}\leq 2, we follow a similar procedure. We claim that (σ21)2(σ11)2(\sigma_{2}-1)^{2}\leq(\sigma_{1}-1)^{2}. Then, instead of (27), we get

(σ11)2+(σ21)22(1σ1)2.(\sigma_{1}-1)^{2}+(\sigma_{2}-1)^{2}\leq 2(1-\sigma_{1})^{2}. (29)

To prove this claim, we note from σ1σ2\sigma_{1}\leq\sigma_{2} that σ11σ21\sigma_{1}-1\leq\sigma_{2}-1. To get an upper bound for σ21\sigma_{2}-1, we obtain from σ12+σ222\sigma_{1}^{2}+\sigma_{2}^{2}\leq 2 that

σ22σ12.\sigma_{2}\leq\sqrt{2-\sigma_{1}^{2}}.

Since the right-hand side is concave in σ1\sigma_{1}, we can bound it from above by its tangent at σ1=1\sigma_{1}=1, this yields σ22σ1\sigma_{2}\leq 2-\sigma_{1}, and thus σ211σ1\sigma_{2}-1\leq 1-\sigma_{1}. The claim follows.

For the left-hand side of (26), similar to the previous case, there are at least two values for k{0,,5}k\in\{0,\ldots,5\} for which cos2(θ+kπ/3)34\cos^{2}(\theta+k\pi/3)\geq\frac{3}{4}. For these values of kk, we have by σ1σ2\sigma_{1}\leq\sigma_{2}, σ12+σ222\sigma_{1}^{2}+\sigma_{2}^{2}\leq 2 and σ11\sigma_{1}\leq 1 that

|ΣRθ+kπ/3e1|234σ12+14σ2212σ12+121,|\Sigma R_{\theta+k\pi/3}e_{1}|^{2}\leq\frac{3}{4}\sigma_{1}^{2}+\frac{1}{4}\sigma_{2}^{2}\leq\frac{1}{2}\sigma_{1}^{2}+\frac{1}{2}\leq 1,

and thus

14k=05(1|ΣRθ+kπ/3e1|)228(112σ12+1)2.14\sum_{k=0}^{5}(1-|\Sigma R_{\theta+k\pi/3}e_{1}|)^{2}\geq 28\Big{(}1-\frac{1}{\sqrt{2}}\sqrt{\sigma_{1}^{2}+1}\Big{)}^{2}. (30)

Using that f(x)=x2+1f(1)+(1x)(f(0)f(1))f(x)=\sqrt{x^{2}+1}\leq f(1)+(1-x)(f(0)-f(1)), we obtain

σ12+12(1σ1)(21).\sqrt{\sigma_{1}^{2}+1}\leq\sqrt{2}-(1-\sigma_{1})(\sqrt{2}-1).

Plugging this estimate into (30), we observe that the resulting lower bound is larger than the upper bound in (29). ∎

Lemma A.2.

Let WW be as in (8) with Ψ=0\Psi=0, that is,

W(A)=eB1Φ(|eA|1).W(A)=\sum_{e\in B^{1}}\Phi\left(\big{|}eA\big{|}-1\right).

Then,

QW(0)=0.QW(0)=0.
Proof.

Since 0W(A)0\leq W(A) for all A2×2A\in\mathbb{R}^{2\times 2}, it follows 0QW(0)0\leq QW(0). We are left to show the reverse inequality. To do this, we introduce RW:2RW:\mathbb{R}^{2}\to\mathbb{R}, the rank-1 convex envelope of WW (see [Dac08, Sec. 6.4]) which is characterized by the following formula ([Dac08, Thm. 6.10])

RW(A)=inf{i=1IλiW(Ai):I,λi0,Ii=1λi=1,i=1IλiAi=A,(λi,Ai) satisfy (HI)},RW(A)=\inf\bigg{\{}\sum_{i=1}^{I}\lambda_{i}W(A_{i}):\\ I\in\mathbb{N},\ \lambda_{i}\geq 0,\ \sum^{I}_{i=1}\lambda_{i}=1,\ \sum_{i=1}^{I}\lambda_{i}A_{i}=A,\ (\lambda_{i},A_{i})\textrm{ satisfy }(H_{I})\bigg{\}}, (31)

where (HI)(H_{I}) is the hierarchical compatibility constraint defined in [Dac08, Definition 5.14]. Recall also that for W:2W:\mathbb{R}^{2}\to\mathbb{R}, then QWRWQW\leq RW (see [Dac08, Sec. 6.1]). Hence, it remains to show that RW(0)0RW(0)\leq 0.

To show that RW(0)0RW(0)\leq 0, we take in (31) I=4I=4, λi=14\lambda_{i}=\frac{1}{4} and consider the family

A1:=[1001],A2:=[1001],A3=A1,A4=A2\displaystyle A_{1}:=\begin{bmatrix}1&0\\ 0&-1\end{bmatrix},A_{2}:=\begin{bmatrix}1&0\\ 0&1\end{bmatrix},A_{3}=-A_{1},A_{4}=-A_{2}

and observe that

  1. 1)

    14A1+14A2+14A3+14A4=0\frac{1}{4}A_{1}+\frac{1}{4}A_{2}+\frac{1}{4}A_{3}+\frac{1}{4}A_{4}=0;

  2. 2)

    rank(A1A2)=rank(A3A4)=1\textrm{rank}(A_{1}-A_{2})=\textrm{rank}(A_{3}-A_{4})=1;

  3. 3)

    rank(A1+A22A3+A42)=1\textrm{rank}(\frac{A_{1}+A_{2}}{2}-\frac{A_{3}+A_{4}}{2})=1;

  4. 4)

    |Aiej|2=ejTAiTAiej=1|A_{i}e_{j}|^{2}=e_{j}^{T}A_{i}^{T}A_{i}e_{j}=1, j=1,,6j=1,\dots,6 where ejB1e_{j}\in B^{1} and i=1,,4i=1,\dots,4.

By [Dac08, Example 5.15], Properties 2)3)2)-3) imply that (λi,Ai)(\lambda_{i},A_{i}) satisfy condition (HI)(H_{I}). Then, it is then easy to see from 1) that (λi,Ai)(\lambda_{i},A_{i}) constitute an admissible candidate in the minimization problem (31). Finally, condition 4)4) implies W(Ai)=0W(A_{i})=0 for i=1,,4.i=1,\dots,4.

Collecting all the results above, we have

0QW(0)RW(0)i=14λiW(Ai)=0.0\leq QW(0)\leq RW(0)\leq\sum_{i=1}^{4}\lambda_{i}W(A_{i})=0.

References

  • [AEST68] K. Anthony, U. Essmann, A. Seeger, and H. Trauble. Disclinations and the Cosserat-Continuum with Incompatible Rotations, volume Mechanics of Generalized Continua, Proceedings of the IUTAM-Symposium on The Generalized Cosserat Continuum and the Continuum Theory of Dislocations with Applications, Freudenstadt and Stuttgart (Germany) 1967, pages 355–358. Springer-Verlag Berlin Heidelberg, 1968.
  • [AF84] E. Acerbi and N. Fusco. Semicontinuity problems in the calculus of variations. Archive for Rational Mechanics and Analysis, 86(2):125–145, 1984.
  • [AO05] M. P. Ariza and M. Ortiz. Discrete crystal elasticity and discrete dislocations in crystals. Archive for Rational Mechanics and Analysis, 178(2):149–226, 2005.
  • [BBO19] J. Braun, M. Buze, and C. Ortner. The effect of crystal symmetries on the locality of screw dislocation cores. SIAM Journal on Mathematical Analysis, 51(2):1108–1136, 2019.
  • [BCH15] J. Ball, P. Cesana, and P. Hambly. A probabilistic model for martensitic avalanches. MATEC Web of Conferences, 33, 2015.
  • [Bha03] K. Bhattacharya. Microstructure of martensite: why it forms and how it gives rise to the shape-memory effect. Oxford University Press, 2003.
  • [BHO19] M. Buze, T. Hudson, and C. Ortner. Analysis of an atomistic model for anti-plane fracture. Mathematical Models and Methods in Applied Sciences, 29(13):2469–2521, 2019.
  • [BJ87] J. Ball and R. James. Fine phase mixtures as minimizers of eenergies. Arch. Ration. Mech. Anal., 100:13–52, 1987.
  • [CDPR+20] P. Cesana, F. Della Porta, A. Rueland, C. Zillinger, and B. Zwicknagl. Exact constructions in the (non-linear) planar theory of elasticity: From elastic crystals to nematic elastomers. Archive for Rational Mechanics and Analysis, 237(1):383–445, 2020.
  • [CH] P. Cesana and P. Hambly. A probabilistic model for interfaces in a martensitic phase transition. Preprint.
  • [CPL14] P. Cesana, M. Porta, and T. Lookman. Asymptotic analysis of hierarchical martensitic microstructure. Journal of the Mechanics and Physics of Solids, 72, 2014.
  • [CS06] S. Conti and B. Schweizer. Rigidity and gamma convergence for solid‐solid phase transitions with SO(2) invariance. Commun. Pure Appl. Math., 59(06):830–868, 2006.
  • [Dac08] B. Dacorogna. Direct Methods in the Calculus of Variations. Springer, Heidelberg, 2nd edition, 2008.
  • [DM93] G. Dal Maso. An Introduction to Γ\Gamma-Convergence. Birkhäuser Verlag, Boston, 1993.
  • [DNPV12] E. Di Nezza, G. Palatucci, and E. Valdinoci. Hitchhiker’s guide to the fractional Sobolev spaces. Bulletin des Sciences Mathématiques, 136(5):521–573, 2012.
  • [dW70] R. de Wit. Linear theory of static disclinations. In J. A. Simmons, R. de Wit, and R. Bullough (eds.) Fundamental Aspects of Dislocation Theory, volume 317, pages 651–673. Nat. Bur. Stand. (US), Spec. Publ., 1970.
  • [dW73a] R. de Wit. Theory of disclinations: II. continuous and discrete disclinations in anisotropic elasticity. Journal of Research of the National Bureau of Standards - A, Physics and Chemistry, 77A(1), 1973.
  • [dW73b] R. de Wit. Theory of disclinations: III. continuous and discrete disclinations in isotropic elasticity. Journal of Research of the National Bureau of Standards - A, Physics and Chemistry, 73A(3), 1973.
  • [dW73c] R. de Wit. Theory of disclinations: IV. straight disclinations. Journal of Research of the National Bureau of Standards - A Physics and Chemistry, 77A(5), 1973.
  • [EOS16] V. Ehrlacher, C. Ortner, and A. V. Shapeev. Analysis of boundary conditions for crystal defect atomistic simulations. Archive for Rational Mechanics and Analysis, 222(3):1217–1268, 2016.
  • [ET67] D. Essmann and H. Träuble. The direct observation of individual flux lines in type ii superconductors. Phys. Letters, 24A(526), 1967.
  • [ET99] I. Ekeland and R. Temam. Convex Analysis and Variational Problems, volume 28. SIAM, 1999.
  • [FJM02] G. Friesecke, R. James, and S. Müller. A theorem on geometric rigidity and the derivation of non linear plate theory from three dimensional elasticity. Commun. Pure Appl. Math., 55:1461—1506, 2002.
  • [Gag57] E. Gagliardo. Caratterizzazioni delle tracce sulla frontiera relative ad alcune classi di funzioni in n variabili. Rend. Sem. Mat. Univ. Padova, 27(405):284–305, 1957.
  • [HMH+16] K. Hagihara, T. Mayama, M. Honnami, M. Yamasaki, H. Izuno, T. Okamoto, T. Ohashi, T. Nakano, and Y. Kawamura. Orientation dependence of the deformation kink band formation behavior in Zn single crystal. International Journal of Plasticity, 77:174–191, 2016.
  • [HO14] T. Hudson and C. Ortner. Existence and stability of a screw dislocation under anti-plane deformation. Archive for Rational Mechanics and Analysis, 213(3):887–929, 2014.
  • [HO15] T. Hudson and C. Ortner. Analysis of stable screw dislocation configurations in an antiplane lattice model. SIAM Journal on Mathematical Analysis, 47(1):291–320, 2015.
  • [HOI+16] K. Hagihara, T. Okamoto, H. Izuno, M. Yamasaki, M. Matsushita, T. Nakano, and Y. Kawamura. Plastic deformation behavior of 10H10\textrm{H}-type synchronized LPSO phase in a Mg-Zn-Y system. Acta Materialia, 109:90–102, 2016.
  • [IHM13] T. Inamura, H. Hosoda, and S. Miyazaki. Incompatibility and preferred morphology in the self-accommodation microstructure of β\beta-titanium shape memory alloy. Philosophical Magazine, 93(6):618–634, 2013.
  • [ILTH17] T. Inamura, M. Li, M. Tahara, and H. Hosoda. Formation process of the incompatible martensite microstructure in a beta-titanium shape memory alloy. Acta Materialia, 124:351–359, 2017.
  • [Ina19] T. Inamura. Geometry of kink microstructure analysed by rank-1 connection. Acta Materialia, 173:270–280, 2019.
  • [KK91] Y. Kitano and K. Kifune. HREM study of disclinations in MgCd ordered alloy. Ultramicroscopy, 39:279–286, 1991.
  • [KM18] R. Kupferman and C. Maor. Variational convergence of discrete geometrically-incompatible elastic models. Calculus of Variations and Partial Differential Equations, 57, 2018.
  • [LN15] X.W. Lei and A. Nakatani. A deformation mechanism for ridge-shaped kink structure in layered solids. Journal of Applied Mechanics, 82:071016, 1–6, 2015.
  • [LPS13] G. Lazzaroni, M. Palombaro, and A. Schlömerkemper. Dislocations in nanowire heterostructures: from discrete to continuum. Proceedings in Applied Mathematics and Mechanics, 13(1):541–544, 2013.
  • [MA80] C. Manolikas and S. Amelinckx. Phase transitions in ferroelastic lead orthovanadate as observed by means of electron microscopy and electron diffraction. Phys. Stat. Sol., 60(2):607–617, 1980.
  • [Nab67] F.R.N. Nabarro. Theory of crystal dislocations. International Series of Monographs on Physics. Oxford: Clarendon Press, 1967.
  • [PL13] M. Porta and T. Lookman. Heterogeneity and phase transformation in materials: Energy minimization, iterative methods and geometric nonlinearity. Acta Materialia, 61(14):5311–5340, 2013.
  • [Pon07] M. Ponsiglione. Elastic energy stored in a crystal induced by screw dislocations: from discrete to continuous. SIAM Journal on Mathematical Analysis, 39(2):449–469, 2007.
  • [SN88] S. Seung, H. and R. Nelson, D. Defects in flexible membranes with crystalline order. Physical Review A, 38:1005, 1988.
  • [TE68] H. Träuble and D. Essmann. Fehler im flussliniengitter von supraleitern zweiter art. Phys. Stat. Sol., 25(373), 1968.
  • [Vol07] V. Volterra. Sur l’équilibre des corps élastiques multiplement connexes. Annales scientifiques de l’École Normale Supérieure, 24:401–517, 1907.
  • [ZA18] C. Zhang and A. Acharya. On the relevance of generalized disclinations in defect mechanics. Journal of the Mechanics and Physics of Solids, 119:188–223, 2018.
  • [Zub97] L. M. Zubov. Nonlinear Theory of Dislocations and Disclinations in Elastic Bodies. Lecture Notes in Physics Monographs. Springer, 1997.