Loading [MathJax]/jax/output/SVG/jax.js

An atomistic derivation of von-Kármán plate theory

  • Received: 01 July 2021 Revised: 01 February 2022 Published: 13 April 2022
  • Primary: 49J45, 70C20, 74K20

  • We derive von-Kármán plate theory from three dimensional, purely atomistic models with classical particle interaction. This derivation is established as a Γ-limit when considering the limit where the interatomic distance ε as well as the thickness of the plate h tend to zero. In particular, our analysis includes the ultrathin case where εh, leading to a new von-Kármán plate theory for finitely many layers.

    Citation: Julian Braun, Bernd Schmidt. An atomistic derivation of von-Kármán plate theory[J]. Networks and Heterogeneous Media, 2022, 17(4): 613-644. doi: 10.3934/nhm.2022019

    Related Papers:

    [1] Lifang Pei, Man Zhang, Meng Li . A novel error analysis of nonconforming finite element for the clamped Kirchhoff plate with elastic unilateral obstacle. Networks and Heterogeneous Media, 2023, 18(3): 1178-1189. doi: 10.3934/nhm.2023050
    [2] Andrea Braides, Margherita Solci, Enrico Vitali . A derivation of linear elastic energies from pair-interaction atomistic systems. Networks and Heterogeneous Media, 2007, 2(3): 551-567. doi: 10.3934/nhm.2007.2.551
    [3] Julian Braun, Bernd Schmidt . On the passage from atomistic systems to nonlinear elasticity theory for general multi-body potentials with p-growth. Networks and Heterogeneous Media, 2013, 8(4): 879-912. doi: 10.3934/nhm.2013.8.879
    [4] Manuel Friedrich, Bernd Schmidt . On a discrete-to-continuum convergence result for a two dimensional brittle material in the small displacement regime. Networks and Heterogeneous Media, 2015, 10(2): 321-342. doi: 10.3934/nhm.2015.10.321
    [5] Phoebus Rosakis . Continuum surface energy from a lattice model. Networks and Heterogeneous Media, 2014, 9(3): 453-476. doi: 10.3934/nhm.2014.9.453
    [6] Bernd Schmidt . On the derivation of linear elasticity from atomistic models. Networks and Heterogeneous Media, 2009, 4(4): 789-812. doi: 10.3934/nhm.2009.4.789
    [7] Mathias Schäffner, Anja Schlömerkemper . On Lennard-Jones systems with finite range interactions and their asymptotic analysis. Networks and Heterogeneous Media, 2018, 13(1): 95-118. doi: 10.3934/nhm.2018005
    [8] Leonid Berlyand, Volodymyr Rybalko . Homogenized description of multiple Ginzburg-Landau vortices pinned by small holes. Networks and Heterogeneous Media, 2013, 8(1): 115-130. doi: 10.3934/nhm.2013.8.115
    [9] Victor A. Eremeyev . Anti-plane interfacial waves in a square lattice. Networks and Heterogeneous Media, 2025, 20(1): 52-64. doi: 10.3934/nhm.2025004
    [10] Jose Manuel Torres Espino, Emilio Barchiesi . Computational study of a homogenized nonlinear generalization of Timoshenko beam proposed by Turco et al.. Networks and Heterogeneous Media, 2024, 19(3): 1133-1155. doi: 10.3934/nhm.2024050
  • We derive von-Kármán plate theory from three dimensional, purely atomistic models with classical particle interaction. This derivation is established as a Γ-limit when considering the limit where the interatomic distance ε as well as the thickness of the plate h tend to zero. In particular, our analysis includes the ultrathin case where εh, leading to a new von-Kármán plate theory for finitely many layers.



    The aim of this work is to derive von-Kármán plate theory from nonlinear, three-dimensional, atomistic models in a certain energy scaling as the interatomic distance ε and the thickness of the material h both tend to zero.

    The passage from atomistic interaction models to continuum mechanics (i.e., the limit ε0) has been an active area of research over the last years. In particular, this limit has been well studied for three-dimensional elasticity, cf., e.g., [3,1,19,6,10,17,5,4]. At the same time, there have emerged rigorous results deriving effective thin film theories from three-dimensional nonlinear (continuum) elasticity in the limit of vanishing aspect ratio (i.e., the limit h0), cf. [15,14,13,7,16]. First efforts to combine these passages and investigate the simultaneous limits ε0 and h0 were made in [11,20,21] for membranes (whose energy scales as the thickness h) and in [18] for Kirchhoff plates (whose energy scales like h3). In particular, this left open the derivation of the von-Kármán plate theory, which describes plates subject to small deflections with energy scale h5 and might even be the most widely used model for thin structures in engineering. Though we do want to mention [2] for a result regarding discrete von-Kármán plate theory that is motivated numerically and not physically.

    Our first aim is to close this gap. For thin films consisting of many atomic layers one expects the scales ε and h to separate so that the limit ε,h0 along hε is equivalent to first passing to the continuum limit ε0 and reducing the dimension from 3d to 2d in the limit h0. We will show in Theorem 2.1a) that this is indeed true.

    By way of contrast, for ultrathin films consisting of only a few atomic layers, more precisely, if ε,h0 such that the number of layers ν=hε+1 remains bounded, the classical von-Kármán theory turns out to capture the energy only to leading order in 1ν. The next aim is thus to derive a new finite layer version of the von-Kármán plate theory featuring additional explicit correction terms, see Theorem 2.1b). In view of the fabrication of extremely thin layers, such an analysis might be of some interest also in engineering applications. An interesting question related to such applications, which we do not address here, would be to extend our analysis to heterogeneous structures as in [9,8].

    Our third aim concerns a more fundamental modelling point of view which is based on the very low energy of the von-Kármán scaling: If the the plate is not too thick (more precisely, if h5ε30), we strengthen the previous results to allow for a much wider range of interaction models, that allow for much more physically realistic atomic interactions (compared to [14,13]) as they can now be invariant under reflections and no longer need to satisfy growth assumptions at infinity, see Theorems 2.2 and 2.3. In particular, this includes Lennard-Jones-type interaction models, see Example 3.

    Finally, on a technical note, the proof of the our main result set forth in Section 4 elucidates the appearance and structure of the correction terms in the ultrathin film regime. Both in [18] and the present contribution, at the core of the proof lies the identification of the limiting strain, which in the discrete setting can be seen as a 3×8 matrix rather than a 3×3 matrix. In [18] this has been accomplished with the help of ad hoc techniques that allowed to compare adjacent lattice unit cells. Now, for the proof of Proposition 4 we introduce a more general and flexible scheme to capture discreteness effects by splitting the deformation of a typical lattice unit cell into affine and non-affine contributions and passing to weak limits of tailor-made finite difference operators. While for hε these operators will tend to a differential operator in the limit, if hε, finite differences in the x3 direction will not become infinitesimal and lead to lower order corrections in 1ν.

    This work is organized as follows: In Section 2, we first describe the atomistic interaction model and then present our results. Our main theorem, Theorem 2.1, details the Γ-limits for both the thin (ν) and ultrathin (ν bounded) case. Theorems 2.2 and 2.3 then extend these results to more general and more physically realistic models. Section 3 contains a few technical tools to circumvent rigidity problems at the boundary and to compare continuous with discrete quantities. Using these tools we then prove our results in Section 4.

    Let SR2=R2×{0}R3 be an open, bounded, connected, nonempty set with Lipschitz boundary. To keep the notation simple we will only consider the cubic lattice. Let ε>0 be a small parameter describing the interatomic distance, then we consider the lattice εZ3. We denote the number of atom layers in the film by νN, ν2 and the thickness of the film by h=(ν1)ε. In the following let us consider sequences hn,εn,νn, nN, such that εn,hn0. The macroscopic reference region is Ωn=S×(0,hn) and so the (reference) atoms of the film are Λn=¯ΩnεnZ3. We will assume that the energy can be written as a sum of cell energies.

    More precisely, as in [18] we let z1,,z8 be the corners of the unit cube centered at 0 and write

    Z=(z1,,z8)=12(111111111111111111111111).

    Furthermore, by Λn=(xΛn(x+εn{z1,,z8}))(R2×(0,hn)) we denote the set of midpoints of lattice cells x+[εn/2,εn/2]3 contained in R2×[0,hn] for which at least one corner lies in Λn. Additionally, let w(x)=1εn(w(x+εnz1),,w(x+εnz8))R3×8. Then, we assume that the atomic interaction energy for a deformation map w:ΛnR3 can be written as

    Eatom(w)=xΛnW(x,w(x)), (1)

    where W(x,):R3×8[0,) only depends on those wi with x+εnziΛn, which makes (1) meaningful even though w is only defined on Λn.

    As a full interaction model with long-range interaction would be significantly more complicated in terms of notation and would result in a much more complicated limit for finitely many layers, we restrict ourselves to these cell energies.

    In the following we will sometimes discuss the upper and lower part of a cell separately. We write A=(A(1),A(2)) with A(1),A(2)R3×4 for a 3×8 matrix A.

    If the full cell is occupied by atoms, i.e., x+εnziΛn for all i, then we assume that W is is given by a homogeneous cell energy Wcell:R3×8[0,) with the addition of a homogeneous surface energy Wsurf:R3×4[0,) at the top and bottom. That means,

    W(x,w)={Wcell(w)ifx3(εn/2,hnεn/2),Wcell(w)+Wsurf(w(2))ifνn3andx3=hnεn/2,Wcell(w)+Wsurf(w(1))ifνn3andx3=εn/2,Wcell(w)+2i=1Wsurf(w(i))ifνn=2,andx3=hn/2.

    Example 1. A basic example is given by a mass-spring model with nearest and next to nearest neighbor interaction:

    Eatom(w)=α4x,xΛn|xx|=εn(|w(x)w(x)|εn1)2+β4x,xΛn|xx|=2εn(|w(x)w(x)|εn2)2.

    Eatom can be written in the form (1) by setting

    Wcell(w)=α161i,j8|zizj|=1(|wiwj|1)2+β81i,j8|zizj|=2(|wiwj|2)2

    and

    Wsurf(w1,w2,w3,w4)=α81i,j4|zizj|=1(|wiwj|1)2+β81i,j4|zizj|=2(|wiwj|2)2.

    We will also allow for energy contributions from body forces fn:ΛnR3 given by

    Ebody(w)=xΛnw(x)fn(x).

    We will assume that the fn do not depend on x3, that fn(x)=0 for x in an atomistic neighborhood of the lateral boundary, see (17), and that there is no net force or first moment,

    xΛnfn(x)=0,xΛnfn(x)(x1,x2)T=0, (2)

    to not give a preference to any specific rigid motion. At last, we assume that after extension to functions ˉfn which are piecewise constant on each x+(εn2,εn2)2, xεnZ2, h3nˉfnf in L2(S).

    Overall, the energy is given as the sum

    En(w)=ε3nhn(Eatom(w)+Ebody(w)). (3)

    Due to the factor ε3nhn this behaves like an energy per unit (undeformed) surface area.

    Let us make some additional assumptions on the interaction energy. We assume that Wcell, Wsurf, and all W(x,) are invariant under translations and rotations, i.e., they satisfy

    W(A)=W(A+(c,,c)) and W(RA)=W(A)

    for any AR3×8 or AR3×4, respectively, and any cR3 and RSO(3). Furthermore, we assume that Wcell(Z)=W(x,Z)=0, which in particular implies Wsurf(Z(1))=Wsurf(Z(2))=0, where (Z(1),Z(2))=Z. At last we assume that W and Wcell are C2 in a neighborhood of Z, while Wsurf is C2 in neighborhood of Z(1).

    Since our model is translationally invariant, it is then equivalent to consider the discrete gradient

    ˉw(x)=1εn(w(x+εnz1)w,,w(x+εnz8)w)

    with

    w=188i=1w(x+εnzi)

    instead of w(x) for any x with x+εnziΛn for all i. In particular, the discrete gradient satisfies

    8i=1(ˉw(x))i=0.

    The bulk term is also assumed to satisfy the following single well growth condition.

    (G) Assume that there is a c0>0 such that

    Wcell(A)c0dist2(A,SO(3)Z)

    for all AR3×8 with 8i=1Ai=0.

    In the same way as in a pure continuum approach, it is convenient to rescale the reference sets to the fixed domain Ω=S×(0,1). For xR3 let us always write x=(x,x3)T with xR2. We define ˜Λn=H1nΛn and ˜Λn=H1nΛn with the rescaling matrix

    Hn=(10001000hn).

    A deformation w:ΛnR3 can be identified with the rescaled deformation y:˜ΛnR3 given by y(x)=w(Hnx). We then write En(y) for En(w). The rescaled discrete gradient is then given by

    (ˉny(x))i:=1εn(y(x+εn(zi),x3+εnhnzi3)y)=ˉw(Hnx)

    for x˜Λn, where now

    y=188i=1y(x+εn(zi),x3+εnhnzi3).

    For a differentiable v:ΩRk we analogously set nv:=vH1n=(v,1hn3v).

    In Section 3 we will discuss a suitable interpolation scheme with additional modifications at S to arrive at a ˜˜ynW1,2(Ω;R3) corresponding to yn. Furthermore, for sequences in the von-Kármán energy scaling we will expect yn and ˜˜yn to be close to a rigid motion xRn(x+cn) for some Rn,cn and will therefore be interested in the normalized deformation

    ˜yn:=RnT˜˜yncn, (4)

    which would then be close to the identity. The von-Kármán displacements in the limit will then be found as the limit objects of

    un(x):=1h2n10(˜yn)xdx3,and (5)
    vn(x):=1hn10(˜yn)3dx3. (6)

    To describe the limit energy, let Qcell(A)=D2Wcell(Z)[A,A] for AR3×8 and Qsurf(A)=D2Wsurf(Z(1))[A,A] for AR3×4. As the reference configuration is stress free, frame indifference implies

    D2Wcell(Z)[A,BZ]=0,D2Wsurf(Z(1))[A,BZ(1)]=0 (7)

    for all AR3×8, AR3×4 and all skew symmetric BR3×3. As in continuum elasticity theory this just follows from looking at 0=tAWcell(etBA)|t=0,A=Z.

    In particular,

    Qcell(BZ+c(1,,1))=Qsurf(BZ(1)+c(1,1,1,1))=0 (8)

    for all cR3 and all skew symmetric BR3×3.

    We introduce a relaxed quadratic form on R3×8 by

    Qrelcell(A)=minbR3Qcell(a1b2,,a4b2,a5+b2,,a8+b2)=minbR3Qcell(A+(be3)Z)=minbR3Qcell(A+sym(be3)Z).

    By Assumption (G) Qcell is positive definite on (R3e3)Z. Therefore, for each AR3×8 there exists a (unique) b=b(A) such that

    Qrelcell(A)=Qcell(A+(b(A)e3)Z)=Qcell(A+sym(b(A)e3)Z). (9)

    Here we used (7) to arrive at the symmetric version. Furthermore, the mapping Ab(A) is linear. (If ((vie3)Z)i=1,2,3 is a Qcell-orthonormal basis of (R3e3)Z, then b(A)=3i=1Qcell[(vie3)Z,A], where Qcell[,] denotes the symmetric bilinear form corresponding to the quadratic form Qcell().)

    At last, let us write

    Q2(A)=Qrelcell((A000)Z),Q2,surf(A)=Qsurf((A000)Z(1))

    for any AR2×2.

    We are now in place to state our main theorem in its first version.

    Theorem 2.1. (a) If νn, then 1h4nEnΓEvK with

    EvK(u,v,R):=S12Q2(G1(x))+124Q2(G2(x))+f(x)v(x)Re3dx,

    where G1(x)=symu(x)+12v(x)v(x) and G2(x)=()2v(x). More precisely, for every sequence yn with bounded energy 1h4nEn(yn)C, there exists a subsequence (not relabeled), a choice of RnSO(3),cnR3, and maps uW1,2(S;R2), vW2,2(S) such that (un,vn) given by (5), (6) and (4) satisfy unu in W1,2loc(S;R2), vnv in W1,2loc(S), RnR, and

    lim infn1h4nEn(yn)EvK(u,v,R).

    On the other hand, this lower bound is sharp, as for every uW1,2(S;R2), vW2,2(S), and RSO(3) there is a sequence yn such that unu in W1,2loc(S;R2), vnv in W1,2loc(S) (where we can take Rn=R, cn=0 without loss of generality) and

    limn1h4nEn(yn)=EvK(u,v,R).

    (b) If νnνN, then 1h4nEnΓE(ν)vK, to be understood in exactly the same way as in a), where

    E(ν)vK(u,v,R)=S12Qrelcell((G1(x)000)Z+12(ν1)G3(x))+ν(ν2)24(ν1)2Q2(G2(x))+1ν1Qsurf((G1(x)000)Z(1)+12v(x)4(ν1)M(1))+14(ν1)Q2,surf(G2(x))+νν1f(x)v(x)Re3dx.

    Here,

    G3(x)=(G2(x)000)Z+12v(x)M, (10)
    M=(M(1),M(2))=12e3(+1,1,+1,1,+1,1,+1,1), (11)
    Z=(Z(1),Z(2))=(z1,z2,z3,z4,+z5,+z6,+z7,+z8). (12)

    In the following we use the notation EvK(u,v), respectively, E(ν)vK(u,v), for the functionals without the force term.

    Example 2. Theorem 2.1 applies to the interaction energy of Example 1 if Wcell is augmented by an additional penalty term +χ(w) which vanishes in a neighborhood of SO(3)Z but is c>0 in a neighborhood of O(3)ZSO(3)Z, so as to guarantee orientation preservation.

    Remark 1. 1. The result in a) is precisely the functional one obtains by first applying the Cauchy-Born rule (in 3d) in order to pass from the discrete set-up to a continuum model and afterwards computing the (purely continuum) Γ-limit on the energy scale h5 as h0 as in [13]. Indeed, the Cauchy-Born rule associates the continuum energy density

    WCB(A)=Wcell(AZ)

    to the atomic interaction Wcell, and so Qcell(AZ)=D2WCB(Z)[A,A]=:QCB(A) for AR3×3, in particular,

    Q2(A)=minbR3QCB((A000)+be3).

    2. In contrast, for finite ν non-affine lattice cell deformations of the form AZ+aM, AR3×3, aR need to be taken into account. While AZ is non-affine in the out-of-plane direction, aM distorts a lattice unit cell in-plane in a non-affine way.

    3. Suppose that in addition Wcell and Wsurf satisfy the following antiplane symmetry condition:

    Wcell(w1,,w8)=Wcell(Pw5,,Pw8,Pw1,,Pw4),Wsurf(w1,,w4)=Wsurf(Pw1,,Pw4),

    where P is the reflection P(x,x3)=(x,x3). This holds true, e.g., in mass-spring models such as in Example 1. As both terms in G3 switch sign under this transformation, while the affine terms with G1 and G2 remain unchanged, one finds that the quadratic terms in E(ν)vK decouple in this case and we have

    E(ν)vK(u,v)=S12Q2(G1(x))+ν(ν2)24(ν1)2Q2(G2(x))+18(ν1)2Qrelcell(G3(x))+1ν1Q2,surf(G1(x))+(12v(x))216(ν1)3Qsurf(M(1))+14(ν1)Q2,surf(G2(x))dx=EvK(u,v)+S1ν1[Q2,surf(G1(x))+14Q2,surf(G2(x))]+18(ν1)2[Qrelcell(G3(x))13Q2(G2(x))]+116(ν1)3(12v(x))2Qsurf(M(1))dx.

    4. Standard arguments in the theory of Γ-convergence show that for a sequence (yn) of almost minimizers of En the in-plane displacement un, the out-of-plane displacement vn and the overall rotation Rn converge (up to subsequences) to a minimizer (u,v,R) of EvK, respectively, E(ν)vK.

    5. For the original sequence yn near the lateral boundary there can be lattice cells for which only a subset of their corners belong to Λn. As a consequence these deformation cannot be guaranteed to be rigid on such cells and the scaled in-plane and out-of-plane displacements may blow up. We thus chose to modify yn in an atomistic neighborhood of the lateral boundary so as to pass to the globally well behaved quantities ˜yn, see Section 3. For the original sequence yn, Theorem 2.1 implies a Γ-convergence result with respect to weak convergence in W1,2loc.

    One physically unsatisfying aspect of Theorem 2.1 is the strong growth assumption (G) which is in line with the corresponding continuum results [13]. The problem is actually two-fold. First, typical physical interaction potentials, like Lennard-Jones potentials, do not grow at infinity but converge to a constant with derivatives going to 0. And second, (G) also implies that Wcell(Z)>Wcell(Z). In particular, the atomistic interaction could not even be O(3)-invariant.

    Contrary to the continuum case, it is actually possible to remove these restrictions in our atomistic approach. Indeed, if one assumes ν5nε2n0 or equivalently h5n/ε3n0, then the von-Kármán energy scaling implies that the cell energy at every single cell must be small. In terms of the number of atom layers ν, this condition includes the case of fixed ν, as well as the case νn as long as this divergence is sufficiently slow, namely νnε2/5n.

    In this case, growth assumptions at infinity should no longer be relevant. In fact, we can replace (G) by the following much weaker assumption with no growth at infinity and full O(3)-invariance.

    (NG) Assume that Wcell(A)=Wcell(A) and that there is some neighborhood U of O(3)Z and a c0>0 such that

    Wcell(A)c0dist2(A,O(3)Z)

    for all AU with 8i=1Ai=0 and

    Wcell(A)c0

    for all AU with 8i=1Ai=0.

    One natural problem arising from this is that atoms that are further apart in the reference configuration can end up at the same position after deforming. In particular, due to the full O(3)-symmetry, neighboring cells can be flipped into each other without any cost to the cell energies, which completely destroys any rigidity that one expects in this problem.

    As a remedy, whenever we assume (NG), we will add a rather mild non-penetration term to the energy that can be thought of as a minimal term representing interactions between atoms that are further apart in the reference configuration. To make this precise, for small δ,γ>0 let V:R3×R3[0,] be any function with V(v,w)γ if |vw|<δ and V(v,w)=0 if |vw|2δ. Then define

    Enonpen(w)=x,ˉxΛnV(w(x)ε,w(ˉx)ε).

    Then, γ>0 ensures that there is a positive energy contribution whenever two atoms are closer than δε.

    The overall energy is then given by

    En(w)=ε3nhn(Eatom(w)+Ebody(w)+Enonpen(w)). (13)

    Theorem 2.2. Assume that ν5nε2n0, that fn=0, that En is given by (13), and that (G) is replaced by (NG). Then all the statements of Theorem 2.1 remain true, where now Rn,RO(3).

    Note that in this version, we assume fn=0. Indeed, if one were to include forces, one can typically reduce the energy by moving an atom infinitely far away in a suitable direction. Without any growth assumption in the interaction energy this can easily lead to infEn= and a loss of compactness. However, this is just a problem about global energy minimization. Not only should there still be well-behaved local minima of the energy, but the energy barrier in between should become infinite in the von-Kármán energy scaling.

    In the spirit of local Γ-convergence, we can thus consider the set of admissible functions

    Sδ={w:ΛnR3suchthatdist(ˉw(x),SO(3)Z)<δforallxΛn},

    where Λn labels 'interior cells' away from the lateral boundary, cf. Section 3. This leads us to the total energy

    En(w)={ε3nhn(Eatom(w)+Ebody(w))ifwSδ,else. (14)

    We then have a version of the Γ-limit that does allow for forces.

    Theorem 2.3. Assume that ν5nε2n0, that En is given by (14) with δ>0 sufficiently small, and that (G) is replaced by (NG). Then all the statements of Theorem 2.1 remain true. Furthermore, there is an infinite energy barrier in the sense that

    limninf{1h4nEn(w):wSδSδ/2}=.

    Remark 2. 1. For n large enough, the energy barrier implies that minimizers of the restricted energy (13) correspond to local minimizers of the unrestricted energy (3). The results thus implies convergence of local minimizers of (3) in Sδ.

    2. To formulate it differently, if a sequence (wn) is not separated by a diverging (unrestricted) energy barrier from the reference state id, i.e. each wn can be connected by a continuous path of deformations (wtn)t[0,1] with equibounded energy Eatom(wtn)+Ebody(wtn), then wnSδ for large n. This implies convergence of minimizers of the unrestricted energy under the assumption that a diverging energy barrier cannot be overcome.

    3. As the energy only has to be prescribed in Sδ, Theorem 2.3 also describes local minimizers of energy functionals which are invariant under particle relabeling for point configurations which after labeling with their nearest lattice site by {w(x):xΛn} belong to Sδ, where their energy can be written in the form (14).

    Example 3. In the setting of Theorems 2.2 and 2.3, Example 2 can be generalized to energies of the form

    Eatom(w)=α4x,xΛn|xx|=εnV1(|w(x)w(x)|εn1)+β4x,xΛn|xx|=2εnV2(|w(x)w(x)|εn2),

    where V1,V2 are pair interaction potentials with Vi(0)=0, Vi C2 in a neighborhood of 0 and Vi(r)c0min{r2,1} for some c0>0. (This is satisfied, e.g., for the Lennard-Jones potential r(1+r)122(1+r)6+1.) Due to the non-penetration term in (13) no additional penalty terms for orientation preservation are necessary. Most notably, it is not assumed that Vi(r) as r.

    We first extend a lattice deformation slightly beyond Λn and in doing so possibly modify it near the lateral boundary S×[0,hn] where lattice cells might not be completely contained in ˉΩn. Then we interpolate so as to obtain continuum deformations to which the continuum theory set forth in [12,13] applies.

    For xΛn, with Λn as defined at the beginning of Section 2, we set

    Qn(x)=x+(εn2,εn2)3.

    and also write Qn(ξ)=Qn(x) whenever ξQn(x).

    On a cell that has a corner outside of Λn there is no analogue to (G) (or (NG)) and hence no control of w(x) in terms of W(x,w(x)). For this reason we modify our discrete deformations w:ΛnR3 near the lateral boundary of Ωn.

    Let Sn={xS:dist(x,S)>2εn} and note that, for εn>0 sufficiently small, Sn is connected with a Lipschitz boundary. (This follows from the fact that S can be parameterized with finitely many Lipschitz charts.) If xΛn is such that ¯Qn(x)(Sn×R), we call Qn(x) an inner cell and write xΛn. The corners of these cells are the interior atom positions Λn=Λn+εn{z1,,z8} and the part of the specimen made of such inner cells is denoted

    Ωinn=(xΛn¯Qn(x)).

    Recall the definition of Λn from Section 2 and set

    ˉΛn=Λn+{z1,,z8},Ωoutn=(xΛn¯Qn(x)).

    The (lateral) boundary cells Qn(x) are those for which

    xΛn:=ΛnΛn.

    Later we will also use the rescaled versions of these sets which are denoted ˜Λn=H1nΛn, ˜ˉΛn=H1nˉΛn, ˜Λn=H1nΛn, ˜Λn=H1nΛn, (˜Λn)=Λn. The rescaled lattice cells are ˜Qn(x)=H1nQn(Hnx).

    If w:ΛnR3 is a lattice deformation, following [19] we define a modification and extension w:ˉΛnR3 as follows. First we set w(x)=w(x) if xΛn. Now partition Λn into the 8 sublattices Λn,i=Λnεn(zi+2Z3). We apply the following extension procedure consecutively for i=1,,8:

    For every cell Q=Qn(x) with xΛn,i such that there exists a neighboring cell Q=Qn(x), i.e. sharing a face with Q, on the corners of which w has been defined already, we extend w to all corners of Q by choosing an extension w such that dist2(ˉw(x),SO(3)Z) is minimal.

    As a result of this procedure, w will be defined on every corner of each cell neighboring an inner cell. Now we repeat this procedure until w is extended to ˉΛn, i.e., to every corner of all inner and boundary cells. Since S is assumed to have a Lipschitz boundary, the number of iterations needed to define w on all boundary cells is bounded independently of ε.

    Our modification scheme guarantees that the rigidity and displacements of boundary cells can be controlled in terms of the displacements, respectively, rigidity of inner cells, see [19,Lemmas 3.2 and 3.4]1:

    1We apply these lemmas without a Dirichlet part of the boundary, i.e., Lε(Ω)= in the notation of [19]. Note also that there is a typo in the statement of these lemmas. The set Bε should read {ˉxLε(Ω)Lε(Ω):ˉxVε}, which in our notation (and without Dirichlet part of the boundary) is a subset of Λn.

    Lemma 3.1. There exist constants c,C>0 (independent of n) such that for any w:ΛnR3 and RSO(3)

    xΛn|ˉw(x)RZ|2CxΛn|ˉw(x)RZ|2

    as well as

    xΛndist2(ˉw(x),SO(3)Z)CxΛndist2(ˉw(x),SO(3)Z).

    For the sake of notational simplicity, we will sometimes write w instead of w.

    Let w:ˉΛnR3 be a (modified and extended) lattice deformation. We introduce two different interpolations: ˜w and ˉw. ˜wW1,2(Ωoutn;R3) is obtained by a specific piecewise affine interpolation scheme as in [18,19] which in particular associates the exact average of atomic positions to the center and to the faces of lattice cells. This will allow for a direct application of the results in [13] on continuum plates. By way of contrast, ˉw is a piecewise constant interpolation on the lattice Voronoi cells of ˉΛn. The advantage of this interpolation will be that a discrete gradient of w translates into a continuum finite difference operator acting on ˉw.

    Let xΛn. In order to define ˜w on the cube ¯Q(x) we first set ˜w(x)=188i=1w(x+εnzi). Next, for the six centers v1,,v6 of the faces F1,,F6 of [12,12]3 we set ˜w(x+εnvi)=14jw(x+εnzj), where the sum runs over those j such that zj is a corner of the face with center vi. Finally, we interpolate linearly on each of the 24 simplices

    co(x,x+εnvk,x+εnzi,x+εnzj)

    with |zizj|=1, |zivk|=|zjvk|=12, i.e., whose corners are given by the cube center and the center and two neighboring vertices of one face. Note that for this interpolation

    ˜w(x)=Q(x)˜w(ξ)dξ, (15)
    ˜w(x+εnvk)=x+εnFk˜w(ζ)dζ, (16)

    for every face x+εnFk of Q(x).

    For the second interpolation we first let Voutn:=(xˉΛn(x+[εn2,εn2]3)) and then define ˉwL2(Voutn;R3) by ˉw(ξ)=w(x) for all ξx+(εn2,εn2)3, xˉΛn. Note that

    ˉˉw(x)=1εn(ˉw(x+εnz1)ˉw,,ˉw(x+εnz8)ˉw)

    with ˉw=188i=1ˉw(x+εnzi) defines a piecewise constant mapping on Ωoutn such that

    ˉˉw(ξ)=ˉw(x)wheneverξQn(x),xΛn.

    It is not hard to see that the original function controls the interpolation and vice versa.

    Lemma 3.2. There exist constants c,C>0 such that for any (modified, extended and interpolated) lattice deformation ˜w:ΩoutnR3 and any cell Q=Qn(x), xΛn,

    c|ˉw(x)|2ε3nQ|˜w(ξ)|2dξC|ˉw(x)|2.

    Proof. After translation and rescaling we may without loss assume that εn=1 and Q=(0,1)3, hence x=(12,12,12)T. The claim then is an immediate consequence of the fact that both

    ˜w|ˉ˜w(x)|and˜w˜wL2(Q;R3×3)

    are norms on the finite dimensional space of continuous mappings ˜w which are affine on each co(x,x+vk,x+zi,εnzj) with |zizj|=1, |zivk|=|zjvk|=12, and which have Q˜w(ξ)dξ=0.

    Lemma 3.3. There exist constants c,C>0 such that for any (modified, extended and interpolated) lattice deformation ˜w:ΩoutnR3 and any cell Q=Qn(x), xΛn,

    cdist2(ˉw(x),SO(3)Z)ε3nQdist2(˜w(ξ),SO(3))dξCdist2(ˉw(x),SO(3)Z).

    This is in fact [19,Lemma 3.6]. We include a simplified proof.

    Proof. After translation and rescaling we may without loss assume that εn=1 and Q=(0,1)3. The geometric rigidity result [14,Theorem 3.1] (indeed, an elementary version thereof) yields

    cminRSO(3)˜wR2L2(Q)Qdist2(˜w(ξ),SO(3))dξCminRSO(3)˜wR2L2(Q).

    By definition also

    dist2(ˉw(x),SO(3)Z)=minRSO(3)|ˉw(x)RZ|.

    The claim then follows from applying Lemma 3.2 to ξ˜w(ξ)Rξ for each RSO(3).

    For a sequence wn of (modified and extended) lattice deformations wn:ˉΛnR3 with interpolations ˜wn:ΩoutnR3 and ˉwn:VoutnR3 we consider the rescaled deformations ˜˜yn:˜ΩoutnR3 defined by

    ˜˜yn(x):=˜wn(Hnx)with˜Ωoutn:=H1nΩoutn

    and ˉˉyn:˜VoutnR3 defined by

    ˉˉyn(x):=ˉwn(Hnx)with˜Voutn:=H1nVoutn.

    (Later we will normalize by a rigid change of coordinates to obtain ˜yn and ˉyn.) Their rescaled (discrete) gradients are

    n˜˜yn(x):=˜wn(Hnx)andˉnˉˉyn(x):=ˉˉwn(Hnx)

    for all x˜Ωoutn. Finally, the force fn after extension to ˉΛn is assumed to satisfy

    fn(x)=0forxˉΛnΛn (17)

    and its the piecewise constant interpolation is ˉfn:˜VoutnR3.

    Remark 3. Suppose νn=ν constant. We note that for a sequence of mappings yn:ΛnR3, if ˜˜yny in L2(Ω;R3) then y is continuous in x3 and affine in x3 on the intervals (i1ν1,iν1), i=1,,ν. Similarly, if ˉˉyny in L2(S×(12(ν1),2ν12(ν1));R3), then y is constant in x3 on the intervals (2i12(ν1),2i+12(ν1)), i=0,,ν1.

    Suppose y,yL2(Ω;R3) are piecewise affine, respectively, constant in x3 as detailed above with y(x,x3)=y(x,iν1) if x3(2i12(ν1),2i+12(ν1)), i=0,,ν1. It is not hard to see that the following are equivalent.

    ˜˜yy in L2(Ω;R3).

    ˉˉyy in L2(S×(12(ν1),2ν12(ν1));R3).

    ε3nhnx˜Λn|yn(x)x+(εn2,εn2)2×(εn2hn,εn2hn)y(ξ)dξ|20.

    The same is true in case νn for y=y if in the second statement S×(12(ν1),2ν12(ν1)) is replaced by Ω.

    In particular, limiting deformations do not depend on the interpolation scheme.

    For the compactness we will heavily use the corresponding continuum rigidity theorem from [12,Theorem 3] and [13,Theorem 6]:

    Theorem 4.1. Let yW1,2(Ω;R3) and set I=I(y)=Ωdist2(ny,SO(3))dx. Then there exists maps R:SSO(3) and ˜RW1,2(S;R3×3) with |˜R|C, and a constant RSO(3) such that

    nyR2L2(Ω)CI, (18)
    R˜R2L2(S)CI, (19)
    ˜R2L2(S)CIh2n, (20)
    nyR2L2(Ω)CIh2n, (21)
    RR2Lp(S)CpIh2n, p<. (22)

    Crucially, none of the constants depend on n, y, or I.

    Furthermore, we will also use the continuum compactness result [12,Lemmas 4 and 5] and [13,Lemma 1,Eq. (96),and Lemma 2] based on the previous rigidity result applied to some sequence (ˆyn).

    Theorem 4.2. Let ˆynW1,2(Ω;R3) with I(ˆyn)Ch4n. Then there are RnSO(3), cnR3 as well as a uW1,2(S;R2) and a vW2,2(S) such that yn=RnTˆyncn satisfies

    nynRn2L2(Ω)Ch4n (23)
    Rn˜Rn2L2(S)Ch4n (24)
    ˜R2L2(S)Ch2n (25)
    nynId2L2(Ω)Ch2n (26)
    Ω(nyn)12(nyn)21dx=0. (27)

    And, up to extracting subsequences,

    1h2n10ynxdx3=:unuinW1,2(S;R2), i=1,2, (28)
    1hn10(yn)3dx3=:vnvinW1,2(S;R), (29)
    nynIdhn=:AnA=e3vve3inL2(Ω;R3×3), (30)
    2sym(RnId)h2nA2inLp(S;R3×3), p<, (31)
    RTnnynIdh2nGinL2(Ω;R3×3), (32)

    where the upper left 2×2 submatrix G of G is given by

    G(x)=G1(x)+(x312)G2(x), (33)

    with

    symG1=12(u+(u)T)+vv,G2=()2v. (34)

    The following proposition allows us to apply these continuum results.

    Proposition 1. In the setting of Theorem 2.1, consider a sequence wn with

    En(wn)Ch4n (35)

    Then,

    0I(˜˜yn)=Ωdist2(n˜˜yn,SO(3))dxCh4n. (36)

    Here, ˜˜ynW1,2(Ω;R3) is the rescaled, modified, and interpolated version of wn according to Section 3.

    In the setting of Theorem 2.3 the statement remains is true as well, while in the setting of Theorem 2.2 (36) is still true but now ˜˜yn is the rescaled, modified, and interpolated version of either wn or wn where the correct sign does depend on wn.

    Proof. Rescaling the wn and applying the modification and interpolation steps from Section 3, we have sequences ˜˜ynW1,2(Ω;R3) and ˉˉynL2(Ω;R3). In particular, we can use Theorem 4.1 for this sequence.

    Take Rn according to Theorem 4.2. Then by Lemmas 3.1 and 3.3,

    ε3nhnx˜Λn|ˉnˉˉyn(x)RnZ|2CΩ|n˜˜y(x)Rn|2dxCInh2n.

    A standard discrete Poincaré-inequality then shows

    ε3nhnx˜Λn|ˉˉyn(x)Rn(xhx3)ˉcn|2ε3nhnx˜Λn|ˉnˉˉyn(x)RnZ|2CInh2n

    for a suitable ˉcnR3. Now fn does not depend on x3, vanishes close to S where the modification takes place, and satisfies xΛnfn=0, as well as xΛnfnx=0. Hence, we see that

    ε3nhnEbody(wn)=ε3nhnx˜Λnfn(x)yn(x)=ε3nhnx˜Λnfn(x)(ˉˉyn(x)Rn(xhx3)ˉcn).

    Using ˉfnL2(S)Ch3n and abbreviating I(˜˜yn)=In, we thus find

    |ε3nhnEbody(wn)|CInh2n.

    On the other hand, due to (G) and Lemmas 3.1 and 3.3 we have

    ε3nhnEatom(wn)c0ε3nhnx(˜Λn)dist2(ˉnyn(x),SO(3)Z)cε3nhnx˜Λndist2(ˉnˉˉyn(x),SO(3)Z)cIn.

    Hence,

    0InCε3nhnEatom(wn)Ch4n+Cε3nhn|Ebody(wn)|Ch4n+CInh2n.

    We thus have

    0InCh4n.

    All these statements remain true in the setting of Theorem 2.3 as the Assumptions (G) and (NG) are equivalent on Sδ.

    Now, consider the setting of Theorem 2.2 with Assumption (NG) instead of (G), as well as fn=0 and ν5nε2n0 with the energy given by (13). Using (35), we find

    0Wcell(ˉw(x))Ch5nε3n

    for every xΛn and

    0V(wn(ˉx)εn,wn(ˉˉx)εn)Ch5nε3n

    for all ˉx,ˉˉxΛn. As h5nε3n0, for n large enough, the right hand side is strictly smaller then c0 or γ, respectively. Therefore, for all n large enough we have

    ˉwn(x)UforallxΛn

    and

    |wn(ˉx)wn(ˉˉx)|>εnδ (37)

    for all ˉx,ˉˉxΛn.

    ˉwn(x)U implies Wcell(ˉwn(x))c0dist2(ˉwn(x),O(3)Z). In particular, we thus find

    dist2(ˉwn(x),O(3)Z)Ch5nε3n.

    Again, for n large enough, this means that every xΛn the discrete gradient ˉwn(x) is arbitrarily close to O(3)Z and thus very close to σn(x)SO(3)Z with a unique σn(x){±1}. We now want to show that the sign σn(x) is the same for all x in the interior cells. As the interior of the union of all these cells is connected, it suffices to show that σn is the same on any two cells that share a (d1)-face. Indeed, if that were false, we would have some x,x in cells that share a (d1)-face such that

    dist2(ˉwn(x),O(3)Z)=|ˉwn(x)QZ|2Ch5nε3n,

    and

    dist2(ˉwn(x),O(3)Z)=|ˉwn(x)+QZ|2Ch5nε3n,

    with Q,QSO(3). Without loss of generality assume x=x+εne3. Then

    ˉwn(x)(0,b)T=ˉwn(x)(b,0)T

    for all bR4 with ibi=0. In particular choosing b=(1,+1,+1,1) and b=(1,1,+1,+1), we get |(Q+Q)ei|Ch5nε3n for i=1,2. As Q,QSO(3), we find |(QQ)e3|Ch5nε3n. Overall, we see that both deformed cells are almost on top of each other. More specifically,

    |w(x+εnz1)w(x+εnz5)|=|w(x+εnz5)w(x+εnz1)+w(x+εnz5)w(x+εnz1)|εn(|Qz5Qz1Qz1+Qz5|+Ch5nε3n)=εn(|(QQ)e3|+Ch5nε3n)εnCh5nε3nδεn

    for n large enough. This is a contradiction to the non-penetration condition (37).

    That means, we have

    ε3nhnxΛndist2(σnˉwn(x),SO(3)Z)Ch4n

    for an x-independent σn{±1}. Applying the modification and interpolation procedure from Section 3 to σnwn as in the case (G) above, we find

    Ωdist2(n˜˜y(x),SO(3)Z)dxCh4n.

    Now we can directly apply Theorems 4.1 and 4.2 for the continuum objects ˜˜yn. In particular, for ˜yn=RnT˜˜yncn as defined in (4) and corresponding un and vn as in (5), respectively, (6), after extracting a subsequence from (28) and (29) we get that

    unuinW1,2(S;R2),vnvinW1,2(S;R). (38)

    For later we also introduce ˉyn=RnTˉˉyncn.

    We will also use the following finer statement.

    Proposition 2. In the setting of Theorem 4.2, applied to ˜˜yn and with ˜yn=RnT˜˜yncn, we have

    1h2n((˜yn)x)=:ˆunˆuinW1,2(Ω;R2), (39)
    1hn(˜yn)3=:ˆvnˆvinW1,2(Ω), (40)

    where

    ˆu(x)=u(x)(x312)v(x), (41)
    ˆv(x)=v(x)+(x312). (42)

    Proof. According to Korn's inequality

    ˆunW1,2(Ω;R2)C(symˆunL2(Ω;R2×2)+ˆunx3L2(Ω;R2)+|Ωskewˆundx|+|Ωˆundx|).

    According to Theorem 4.2, symˆun is bounded in L2 by (23) and (31). Furthermore, skewˆundx=0 by (27), and ˆundx is bounded due to (28). As

    (ˆun)ix3=1hn(n˜ynId)i3,

    i=1,2, this term is bounded in L2 as well. This shows compactness. To identify the limit and thus show convergence of the entire sequence, note that

    10ˆundx3uinW1,2(S;R2),

    by (28) and

    (ˆun)ix3=1hn(n˜ynId)i3vxiinL2(Ω),

    for i=1,2 by (30).

    (26) and (29) in Theorem 4.2 also show that ˆvn is bounded in W1,2(Ω) with ˆvnx31 and

    10ˆvndx3v.

    As a first consequence, we will now describe the limiting behavior of the force term Ebody(wn)=Ebody(yn)=x˜ˉΛnfn(x)yn(x), where fn(x)=fn(x) satisfies (2), (17) and h3nˉfnf in L2(S).

    Note that the forces considered are a bit more general than in [13].

    Proposition 3. Let yn be a sequence with En(yn)Ch4n and suppose that (38) holds true for ˜yn, un, vn as defined in (4), (5), (6). Assume that RnR. Then

    ε3nh5nEbody(yn){Sf(x)v(x)Re3dx,ifνn,νν1Sf(x)v(x)Re3dx,ifνn=νconstant,

    as n.

    Proof. In terms of the extended and interpolated force density we have

    ε3nh5nEbody(yn)=1h4n˜Voutnˉfn(x)ˉˉyn(x)dx=1h4n˜Voutnˉfn(x)(ˉˉyn(x)Rn(x0)Rncn)dx=˜Voutnh3nRnTˉfn(x)h1n(ˉyn(x0))dx.

    By Proposition 2, h1n(˜yn(x0))ˆve3 in L2(Ω;R3) with ˆv as in (40) and so Remark 3 shows that

    ε3nh5nEbody(yn)ΩRTf(x)ˆv(x)e3dx=Ωf(x)v(x)Re3dx

    if νn, where in the last step we have used that (2) together with fn(x)=fn(x) also implies that xΛnx3fn(x)=0. If νn=ν constant, then Remark 3 gives

    ε3nh5nEbody(yn)1ν1ν1j=0SRTf(x)ˆv(x,jν1)e3dx=νν1Sf(x)v(x)Re3dx

    with an analogous argument for the last step.

    To show the lower bounds in our Γ-convergence results, we have to understand the limit of the discrete strain. Let (yn) satisfy En(yn)Ch4n and set

    ˉGn:=1h2n(RTnˉnˉynZ).

    By Proposition 1 (˜˜yn) satisfies the assumptions of Theorem 4.2 and Proposition 2 so that, after a rigid change of coordinates, ˜yn satisfies (23)–(34) and (39)–(42). In particular, by (32) we know that for a subsequence the continuum strain converges as

    1h2n(RTnn˜ynId)GinL2(Ω;R3×3),

    where G satisfies (33) and (34).

    For the discussion of discrete strains, recall that we defined

    Z=(z1,z2,z3,z4,+z5,+z6,+z7,+z8),M=12e3(+1,1,+1,1,+1,1,+1,1).

    We define a projection P acting on maps via

    Pf(x)=k/(ν1)(k1)/(ν1)f(x,t)dtifk1ν1x3<kν1

    in case νnν< and P=id in case νn.

    Proposition 4. Let (yn)n satisfy En(yn)Ch4n with 1h2n(RTnn˜ynId)G in L2(Ω;R3×3). Then,

    ˉGnˉG:={GZ,ifνn,PGZ+12(ν1)G3,ifνnνN,

    in L2(Ω;R3×8), where G3 is as in Theorem 2.1.

    Proof. The compactness follows from Theorem 4.2. On a subsequence (not relabeled) we thus find ˉGnˉG. As RnId in L2 while being uniformly bounded, we also find

    RnˉGn=1h2n(ˉnˉynRnZ)ˉG.

    We have

    limn1h2n(RTnn˜ynId)=limn1h2n(n˜ynRn)=G,

    weakly in L2(Ω;R3×3) where G satisfies (33) and (34).

    In order to discuss the discrete strains in more detail, we separate affine and non-affine contributions. We say that a bR8 is affine if it is an element of the linear span of b0,b1,b2,b3, where b0=(1,,1) and bi=ZTei, i=1,2,3. Any bR8 which is perpendicular to all affine vectors is called non-affine. I.e., a non-affine b is characterized by 8i=1bi=0 and Zb=0.

    We begin by identifying the easier to handle affine part of the limiting strain. By construction we have RnˉGnb00 and so ˉGb0=0=GZb0. For i{1,2,3} we use that on any ˜Qn(x), x˜Λn,

    ˉnˉyn(x)b1=12εn((y2+y3+y6+y7)(y1+y4+y5+y8)),

    where yi=˜yn(x+εn(zi),x3+εnhnzi3). So, using (16) for ˜yn,

    ˉnˉyn(x)b1=2εnx+{εn2}×(εn2,εn2)×(εn2hn,εn2hn)˜yn(ξ+εne1)˜yn(ξ)dξ=2˜Qn(x)1˜yn(ξ)dξ.

    Analogous arguments yield

    ˉnˉyn(x)b2=2˜Qn(x)2˜yn(ξ)dξandˉnˉyn(x)b3=2hn˜Qn(x)3˜yn(ξ)dξ.

    By Pn we denote the projection which maps functions to piecewise constant functions via Pnf(x)=˜Qn(x)f(ξ)dξ on ˜Qn(x). Then Pn[RnˉGn]ˉG. On the other hand, observing that ZZT=2Id3×3, we find

    Pn[RnˉGn]bi=2h2nPn[i˜ynRnei]2PGei=PGZbi,i=1,2

    and

    Pn[RnˉGn]b3=2h2nPn[h1n3˜ynRne3]2PGe3=PGZb3.

    In summary we get that for every affine bR8

    ˉGb=PGZb. (43)

    For the discussion of the non-affine part of the strain we fix a non-affine bR8, i.e., a b satisfying 8i=1bi=0, Zb=0, and write bT=((b(1))T,(b(2))T), where b(1),b(2)R4. Let Z2dim:=((z1),(z2),(z3),(z4))R2×4 be the matrix of two-dimensional directions. Then Z2dim(b(1)+b(2))=0 and 4i=1b(1)i=4i=1b(2)i=0. We introduce the difference operator

    ˉ2dimf(x):=1εn(f(x+εn(zi),x3)144j=1f(x+εn(zj),x3))i=1,2,3,4.

    The idea is now to separate differences into in-plane and out-of-plane differences, as all in-plane differences are infinitesimal, while out-of-plane differences stay non-trivial if νnν and have to be treated more carefully.

    Using

    ˉnˉyn(x)=(ˉ2dimnˉyn(xεn2hne3),ˉ2dimnˉyn(x+εn2hne3))+12hn˜Qn(x)3˜yn(ξ)dξ(1,1,1,1,+1,+1,+1,+1)

    we find

    RnˉGn(x)b=1h2nˉnˉyn(x)b=1h2n(ˉ2dimnˉyn(x+εn2hne3)ˉ2dimnˉyn(xεn2hne3))b(2) (44)
    +1h2nˉ2dimnˉyn(xεn2hne3)(b(1)+b(2)), (45)

    where we have used that 4i=1b(1)i=4i=1b(2)i=0.

    First consider the term (45). Since ˉ2dimn¯id(xεn2hne3)=Z2dim and Z2dim(b(1)+b(2))=0, for any φCc(Ω) and i=1,2 by (39) and Remark 3 we have

    1h2neTiΩˉ2dimn(ˉyn¯id)(xεn2hne3)(b(1)+b(2))φ(x)dx=1h2neTiΩ(ˉyn¯id)(xεn2hne3)(ˉ2dimn)φ(x)(b(1)+b(2))dxΩˆui(˜x)φ(x)Z2dim(b(1)+b(2))dx=0, (46)

    where, either ˜x=x (if νn), or ˜x=(x,(ν1)x3ν1) (if νn=ν is constant).

    For the third component, we instead have

    1h2neT3Ωˉ2dimnˉyn(xεn2hne3)(b(1)+b(2))φ(x)dx=1hnεn(νn1)eT3Ω(ˉ2dimnˉyn(xεn2hne3)nˉyn(xεn2hne3)Z2dim)(b(1)+b(2))φ(x)dx=1(νn1)εnΩ(ˉyn)3(xεn2hne3)hn((ˉ2dimn)φ(x)+nφ(x)Z2dim)(b(1)+b(2))dx.

    Now,

    1εn((ˉ2dimn)φ(x)+nφ(x)Z2dim)(122φ(x)[(zi),(zi)]184j=12φ(x)[(zj),(zj)])i=1,...,4

    uniformly. Therefore, (40) gives

    1h2neT3Ωˉ2dimnˉyn(xεn2hne3)(b(1)+b(2))φ(x)dx0, (47)

    if νn. For νn=ν constant however, using (40) and (42) we find

    1h2neT3Ωˉ2dimnˉyn(xεn2hne3)(b(1)+b(2))φ(x)dx1(ν1)Ωˆv(x,(ν1)x3ν1)(122φ(x)[(zi),(zi)])i=1,...,4(b(1)+b(2))dx=1(ν1)Ω(122v(x)[(zi),(zi)])i=1,...,4(b(1)+b(2))φ(x)dx, (48)

    where we have used that 8i=1bi=0.

    We still need to find the limit of (44). For any test function φCc(Ω;R3) we find

    Ω1h2n(ˉ2dimnˉyn(x+εn2hne3)ˉ2dimnˉyn(xεn2hne3))b(2)φ(x)dx=εnhnΩ1εnhn(ˉyn(x+εn2hne3)ˉyn(xεn2hne3))(ˉ2dimn)Pnφ(x)b(2)dx=1h2nΩ˜Qn(x)(ˉyn(ξ+εn2hne3)ˉyn(ξεn2hne3))dξ(ˉ2dimn)Pnφ(x)b(2)dx=εnh3nΩ3˜yn(x)(ˉ2dimn)Pnφ(x)b(2)dx=εnhnΩPnAn(x)e3(ˉ2dimn)φ(x)b(2)dx.

    Here the penultimate step is true by our specific choice of interpolation to define ˜yn, whereas the last step follows from (30) and ˉ2dimn1hne3=0. If νn this converges to 0. In case νn=ν constant we obtain from (30)

    limnΩ1h2n(ˉ2dimnˉyn(x+εn2hne3)ˉ2dimnˉyn(xεn2hne3))b(2)φ(x)dx=1ν1ΩPA(x)e3φ(x)Z2dimb(2)dx=1ν1Ω(1v(x),2v(x),0)φ(x)Z2dimb(2)dx=1ν1Ω(2v(x)Z2dimb(2)0)φ(x)dx. (49)

    Summarizing (46), (47), (48), and (49), we see that for non-affine b we have ˉGb=0 in case νn and

    ˉGb=(1ν12v(x)Z2dimb(2)1ν14i=1122v(x)[(zi),(zi)](b(1)+b(2))i)=(12(ν1)2v(x)Z2dim(b(2)b(1))12(ν1)8i=12v(x)[(zi),(zi)]bi)18(ν1)Δv(x)8j=1bje3

    as 8j=1bj=0, if νnν.

    Elementary computations show that for the affine basis vectors bk, k{0,1,2,3},

    Z2dim((bk)2(bk)1)=0

    and also

    8i=12v(x)[(zi),(zi)]bki14Δv(x)8j=1bkj=0.

    Thus combining with (43), for every bR8 we get

    ˉGb=GZb

    if νn and

    ˉGb=PGZb+(12(ν1)2v(x)Z2dim(b(2)b(1))12(ν1)8i=12v(x)[(zi),(zi)]bi)18(ν1)Δv(x)8j=1bje3.

    if νn=ν is constant. So ˉG=GZ if νn and

    ˉG=PGZ12(ν1)(2v(x)000)Z+12(ν1)e3(2v(x)[(zi),(zi)])i=1,,818(ν1)Δv(x))e3(1,,1).

    with Z as in (12) if νn=ν is constant. Noting that

    2v(x)[(zi),(zi)]={14(11v(x)+212v(x)+22v(x))ifi{1,3,5,7},14(11v(x)212v(x)+22v(x))ifi{2,4,6,8},

    with M as in (11) this can be written as

    ˉG=PGZ12(ν1)(2v(x)000)Z+12(ν1)12v(x)M.

    Last, we note that subsequences were indeed not necessary, as the limit is characterized uniquely.

    Having established convergence of the strain, the lim inf inequality in Theorems 2.1, 2.2 and 2.3 can now be shown by a careful Taylor expansion of W(x,), cf. [14,13,18].

    Proof of the lim inf inequality in Theorems 2.1, 2.2 and 2.3. The lim inf inequality in Theorem 2.3 is an immediate consequence of the lim inf inequality in Theorem 2.1 applied to a cell energy Wcell of the form

    Wcell(A)={Wcell(A),ifdist(A,SO(3)Z)<δ,dist2(A,SO(3)Z),ifdist(A,SO(3)Z)δ.

    Furthermore, in view of Proposition 3 it suffices to establish the lower bound for fn=0.

    Assume that (yn) is a sequence of atomistic deformations such that

    supnEn(yn)<

    so that by Proposition 1 its modification and interpolation (˜yn) verifies the assertions of Theorem 4.2. Set

    ˉGn:=1h2n(RTnˉnˉynZ).

    By frame indifference and nonnegativity of the cell energy we have

    ε3nh5nEn(yn)ε3nh5nx(˜Λn)W((x,hnx3),ˉnˉyn(x))=1h4nΩinnW(εn(x1εn+12,x2εn+12,hnx3εn+12),Z+h2nˉGn(x))dx.

    First assume that νn as n. Due to nonnegativity of Wsurf we can estimate

    ε3nh5nEn(yn)1h4nΩχn(x)Wcell(Z+h2nˉGn(x))dx=Ω12Qcell(χn(x)ˉGn(x))h4nχn(x)ω(|h2nˉGn(x)|)dx,

    where χn is the characteristic function of {xΩinn:ˉGh1n}Ω and

    ω(t):=sup{|12Qcell(F)Wcell(Z+F)|:FR3×8with|F|t}

    so that t2ω(t)0 as t0. Since ˉG2n is bounded in L1(Ω;R3×8) and

    χn(h2nˉGn)2ω(h2nˉGn)0

    uniformly,

    h4nχnω(h2nˉGn)=ˉG2nχn(h2nˉGn)2ω(h2nˉGn)0inL1(Ω;R3×8).

    Moreover, χn1 boundedly in measure and so by Proposition 4 χnˉGnˉG=GZ, where G satisfies (33) and (34). By lower semicontinuity it follows that

    lim infnε3nh5nEn(yn)12ΩQcell(ˉG(x))dx12ΩQrelcell(ˉG(x))dx=12ΩQrelcell((G1(x)+(x312)G2(x)000)Z)dx.

    Integrating the last expression over x3(0,1) and noting that the integral of the cross terms vanish we obtain

    lim infnε3nh5nEn(yn)EvK(u,v).

    Now suppose that νnνN. We let χn as above but now define

    ω(t):=sup{|12Qcell(F)Wcell(Z+F)|:FR3×8with|F|t}+2sup{|12Qsurf(F)Wsurf(Z(1)+F)|:FR3×4with|F|t}

    so that still t2ω(t)0 as t0. With ˉG(x)=(ˉG(1)(x),ˉG(2)(x)) we have

    lim infnε3nh5nEn(yn)12ΩQcell(ˉG(x))dx+12(ν1)SQsurf(ˉG(1)(x,12(ν1)))+Qsurf(ˉG(2)(x,2ν32ν2))dx,

    where we have used that ˉG is constant on S×(0,1ν1) and on S×(ν2ν1,1). Here (see Eq. (10) for G3),

    ˉG(1)(x,12ν2)=1ν10G(x,x3)dx3Z(1)+12(ν1)G(1)3(x),ˉG(2)(x,2ν32ν2)=1ν2ν1G(x,x3)dx3Z(2)+12(ν1)G(2)3(x).

    The bulk part is estimated as

    121212Qcell(ˉG(x))dx312(ν1)ν1k=1Qrelcell((sym(PG)(x,2k12ν2)000)Z+12(ν1)G3(x))=12(ν1)ν1k=1Qrelcell((symG1(x)+2kν2ν2G2(x)000)Z+12(ν1)G3(x))=12(ν1)ν1k=1[Qrelcell((symG1(x)000)Z+12(ν1)G3(x))+(2kν)2(2ν2)2Qrelcell((G2(x)000)Z)]=12Qrelcell((symG1(x)000)Z+12(ν1)G3(x))+ν(ν2)24(ν1)2Qrelcell((G2(x)000)Z),

    where we have used that ν1k=1(2kν)2(2ν2)3=ν(ν2)24(ν1)2.

    For the surface part first note that by (8), for any A=(aij)R3×3 and BR3×4 we have

    Qsurf(AZ(1)+B)=Qsurf(AZ(1)+B+(a3e3e3a3)Z(1)+(a3+a3)(1,1,1,1))=Qsurf((A000)Z(1)+B)=Qsurf((symA000)Z(1)+B),

    where a3 denotes the third column, a3 the third row and A=(aij)1i,j2 the upper left 2×2 part of A. Thus also

    Qsurf(AZ(2)+B)=Qsurf(AZ(1)+a3(1,1,1,1)+B)=Qsurf((symA000)Z(1)+B).

    It follows that

    Qsurf(ˉG1(x,12ν2))=Qsurf((symG1(x)ν22ν2G2(x)000)Z(1)+12(ν1)G(1)3(x))=Qsurf((symG1(x)12G2(x)000)Z(1)+12v(x)4(ν1)M(1)),Qsurf(ˉG2(x,2ν32ν2))=Qsurf((symG1(x)+ν22ν2G2(x)000)Z(1)+12(ν1)G(2)3(x))=Qsurf((symG1(x)+12G2(x)000)Z(1)+12v(x)4(ν1)M(1))),

    and so

    Qsurf(ˉG1(x,12ν2))+Qsurf(ˉG2(x,2ν32ν2))=2Qsurf((symG1(x)000)Z(1)+12v(x)4(ν1)M(1))+12Qsurf((G2(x)000)Z(1)),

    Adding bulk and surface contributions and integrating over x we arrive at

    lim infnε3nh5nEn(yn)S12Qrelcell((symG1(x)000)Z+12(ν1)G3(x))+ν(ν2)24(ν1)2Qrelcell((G2(x)000)Z)+1ν1Qsurf((symG1(x)000)Z(1)+12v(x)4(ν1)M(1))+14(ν1)Qsurf((G2(x)000)Z(1))dx=E(ν)vK(u,v).

    Note that in the Theorem 2.1 the skew symmetric part of G1 is then just set to be zero as it does not impact the energy.

    Without loss of generality we assume that R=Id. (For general R one just considers the sequence Ryn with yn as in (50) and Rn=R below.

    If u:SR2 and v:SR are smooth up to the boundary, we choose a smooth extension to a neighborhood of S and define the lattice deformations yn:˜ˉΛnR3 by restricting to ˜ˉΛn the mapping yn:¯˜ΩoutnR3, defined by

    yn(x)=(xhnx3)+(h2nu(x)hnv(x))h2n(x312)((v(x))T0)+h3nd(x,x3) (50)

    for all x¯˜Ωoutn. Here d:¯˜ΩoutnR3 will be determined later, see (55) and (56) for films with many, respectively, a bounded number of layers. In both cases, d is smooth and bounded in W1,(¯˜Ωoutn;R3) uniformly in n.

    We let Rn=Id and cn=0 for all n and define ˜ynW1,2(˜Ωoutn;R3) as in (4) by interpolating as in Section 3 (more precisely, descaling to wn and then interpolating and rescaling) to obtain ˜yn=˜˜yn. Analogously we let ˉy=ˉˉy. We define un and vn as in (5) and (6), respectively. It is straightforward to check that indeed unu in W1,2(S;R2) and vnv in W1,2(S).

    In order to estimate the energy of yn we need to compute its discrete gradient. Instead of directly calculating ˉˉyn=(ˉ1ˉyn,,ˉ8ˉyn) it is more convenient to first determine ˉDyn=(ˉD1yn,,ˉD8yn) which for each x˜Λn(εn2,εn2,εn2hn) is defined by

    ˉDiyn(x)=1εn[yn(ˆx+εn((ai),h1nai3))yn(ˆx)],

    where for x˜Ωoutn we have set

    ˆx=(εnx1εn,εnx2εn,(νn1)x3νn1),

    so that ˜Qn(x)=ˆx+(0,εn)2×(0,(νn1)). We set ai=12(1,1,1)T+zi{0,1}3 and write A:=(a1,,a8)=Z+12(1,1,1)T(1,1,1,1,1,1,1,1)T. Note that

    ˉDiyn(x)=ˉiˉyn(x)ˉ1ˉyn(x)andˉiˉyn(x)=ˉDiyn(x)188j=1ˉDjyn(x). (51)

    In particular, if ˉDyn(x) is affine, i.e., ˉDyn(x)=FA for some FR3×3, then

    ˉiˉyn(x)=Fai188j=1Faj=F(ai12(1,1,1)T)=Fzi (52)

    and so ˉˉyn(x)=FZ.

    For x in a fixed cell ˜Qn(x)=ˆx+(0,εn)2×(0,(νn1)), Taylor expansion of yn (restricted to ¯˜Qn(x)) yields

    ˉDiyn(x)=yn(ˆx)(ai)+h1n3yn(ˆx)ai3+εn2()2yn(ˆx)[(ai),(ai)]+εnh1n2j=1j3yn(ˆx)aijai3+εnh2n233yn(ˆx)(ai3)2+ε2n63((yn)1(ζ1εn),(yn)2(ζ2εn),(yn)2(ζ2εn))T[((ai),h1nai3),((ai),h1nai3),((ai),h1nai3)]

    for some ζεnˆx+[0,εn]2×[0,εnh1n]. Plugging in (50) we get

    ˉDiyn(x)=((Id2×20)+(h2nu(ˆx)hnv(ˆx))h2n(ˆx312)((v(ˆx))T0)+h3nd(ˆx))(ai)+h1n((0hn)+0h2n((v(ˆx))T0)+h3n3d(ˆx))ai3+εnhn2(0()2v(ˆx)[(ai),(ai)])+O(εnh2n)εnhn((v(ˆx))T0)(ai)ai3+O(εnh2n)+εnhn233d(ˆx)(ai3)2+ε2n6333(d1(ζ1εn),d2(ζ2εn),d3(ζ3εn))T(ai3)3+O(ε2nhn).

    It follows that

    ˉDiyn(x)=(Id3×3+hn(hnu(ˆx)(v(ˆx))Tv(ˆx)0)h2n(ˆx312)(()2v(ˆx)000)+h2n(03×23d(ˆx)))ai+εnhn((()2v(ˆx)(ai)ai312()2v(ˆx)[(ai),(ai)])+1233d(ˆx)(ai3)2)+ε2n6333(d1(ζ1εn),d2(ζ2εn),d3(ζ3εn))T(ai3)3+O(εnh2n+ε2nhn).

    We define the skew symmetric matrix B(ˆx)=Bn(ˆx) by

    B(ˆx)=(h2n2(u(ˆx)(u(ˆx))T)hn(v(ˆx))Thnv(ˆx)0)+h2n2(02×23d(ˆx)(3d(ˆx))T0),

    where we have written d=(d1,d2)T for d=(d1,d2,d3)T, and consider the special orthogonal matrix

    eB(ˆx)=Id3×3B(ˆx)+12B2(ˆx)+O(|B(ˆx)|3)=Id3×3hn(02×2(v(ˆx))Tv(ˆx)0)h2n2(u(ˆx)(u(ˆx))T+v(ˆx)v(ˆx)3d(ˆx)(3d(ˆx))T|v(ˆx)|2)+O(|hn|3).

    Now compute

    eB(ˆx)ˉDiyn(x)=ˉDiyn(x)hn(02×2(v(ˆx))Tv(ˆx)0)(Id3×3+hn(02×2(v(ˆx))Tv(ˆx)0))aih2n2(u(ˆx)(u(ˆx))T+v(ˆx)v(ˆx)3d(ˆx)(3d(ˆx))T|v(ˆx)|2)ai+O(h3n+εnh2n+ε2nhn)=(Id3×3+h2n(symu(ˆx)+12v(ˆx)v(ˆx)0012|v(ˆx)|2)h2n(ˆx312)(()2v(ˆx)000)+h2n(02×2123d(ˆx)12(3d(ˆx))T3d3(ˆx)))ai+εnhn((()2v(ˆx)(ai)ai312()2v(ˆx)[(ai),(ai)])+1233d(ˆx)(ai3)2)+ε2n6333(d1(ζ1εn),d2(ζ2εn),d3(ζ3εn))T(ai3)3+O(h3n+εnh2n+ε2nhn). (53)

    Here, the error term is uniform in ˆx.

    We can now conclude the proof of Theorems 2.1, 2.2 and 2.3.

    Proof of the lim sup inequality in Theorems 2.1, 2.2 and 2.3. As the discrete gradient ˉnˉyn is uniformly close to SO(3)Z, the following arguments apply to show that yn defined by (50) serves as a recovery sequence in all three theorems. Moreover, in view of Proposition 3 it suffices to construct recovery sequences for fn=0.

    We first specialize now to the case νn. For

    G(x)=G1(x)+(x312)G2(x)=symu(x)+12v(x)v(x)(x312)()2v(x). (54)

    choosing d(x)=x3d0(x)+x23x32d1(x) with

    d0(x)=argminbR3Qcell[(G1(x)0012|v(x)|2)Z+(be3)Z],d1(x)=argminbR3Qcell[(G2(x)000)Z+(be3)Z] (55)

    according to (9), from (52) and (53) we obtain

    eB(ˆx)ˉˉyn(x)=(Id3×3+h2n(G(ˆx)0012|v(ˆx)|2)+h2nsym((d0(ˆx)+(ˆx312)d1(ˆx))e3))Z+O(h3n+εnhn)

    and, Taylor expanding Wcell, we see that due to the smoothness of u and v the piecewise constant mappings xh4nWcell(ˉˉyn(x))=h4nWcell(eB(ˆx)ˉˉyn(x)) converge uniformly to

    12Qrelcell((G000)Z)=12Q2(G).

    This shows that

    limnh4nEn(yn)=12SQ2(G(x))dx=S12Q2(G1(x))+124Q2(G2(x))dx=EvK(u,v)

    and thus finishes the proof in case νn.

    Now suppose that εnhn1ν1. Abbreviating ()2v(ˆx)=G2(ˆx)=G2=(fij)R2×2, we observe that

    (2G2(ai)ai3(ai)TG2(ai))i=1,,8=(000002f112f112f122f12000002f212f212f222f220f11μ,νfμνf220f11μ,νfμνf22),

    and hence, with b=b(ˆx)=((11+12)v(ˆx),(21+22)v(ˆx),0)T=(f11+f12,f21+f22,0)T,

    (2G2(ai)ai3(ai)TG2(ai))i=1,,8(e3bbe3)A=(0000f11+f12f11+f12f11f12+f11f120000f21+f22f21+f22f21f22f21f220f120f210f120f21),=(G2000)(Z+Z)+12f12(2Me3(1,,1))=(G2000)A12b(1,,1)+(G2000)Z+f122(2Me3(1,,1)).

    This shows that

    (()2v(ˆx)(ai)ai312()2v(ˆx)[(ai),(ai)])i=1,,8=12(e3bbe3+(G2000))A14(b+e3)(1,,1)+12(G2000)Z+12f12M.

    We define the affine part of the strain G(x)=G1(x)+(x312)G2(x) as in (54). The non-affine part is abbreviated by 12(ν1)G3(x) as in (10). Then using (53) we can write

    eB(ˆx)ˉˉyn(x)=[Id3×3+h2n(G(ˆx,ˆx3+12(ν1))0012|v(ˆx)|2)+h2nsym(3d(ˆx))e3)+h2n2(ν1)(e3b(ˆx)b(ˆx)e3)]Z+h2n2(ν1)G3(ˆx)+O(h3n)+[εnhn233d(ˆx)+ε2n6333(d1(ζ1εn),d2(ζ2εn),d3(ζ3εn))T](z13,,z83),

    where we have used (52) and (51).

    We set

    d0(x)=argmindR3Qcell[(G1(x)0012|v(x)|2)Z+sym(de3)Z+12(ν1)G3(x)],d1(x)=argmindR3Qcell[(G2(x)000)Z+sym(de3)Z]

    according to (9) and define d:S×[0,1]R, S a neighborhood of S, inductively by d(x,0)=0 and

    d(x,j1ν1+t)=d(x,j1ν1)+td0(x)+t2jν2(ν1)d1(x)ift[j1ν1,jν1], (56)

    for j=1,,ν1. Then d is smooth in x and piecewise linear in x3, more precisely, affine in x3 in between two atomic layers: On S×[j1ν1,jν1], j{1,,ν1}, it satisfies

    3d(x)=d0(x)+2jν2(ν1)d1(x)=d0(x)+(ˆx312+12(ν1))d1(x)

    since ˆx3=ˆx3(x)=j1ν1. Taylor expanding Wcell, we see that the piecewise constant mappings xh4nWcell(ˉˉyn(x))=h4nWcell(eB(ˆx)ˉˉyn(x)) converge uniformly on S×[j1ν1,jν1] to

    12Qrelcell((G1(x)+2jν2(ν1)G2(x)000)Z+12(ν1)G3(x))

    for each j{1,,ν1}. Since 1ν1ν1j=12jν2(ν1)=0 and 1ν1ν1j=1(2jν2(ν1))2=ν(ν2)12(ν1)2, this shows

    1h4n˜ΩoutnWcell(ˉˉyn(x))dxS12Qrelcell((G1(x)000)Z+12(ν1)G3(x))+ν(ν2)24(ν1)2Qrelcell((G2(x)000)Z)dx. (57)

    For the surface part we write ˉˉyn=([ˉˉyn](1),[ˉˉyn](2)) and use that the piecewise constant mappings S×[0,1ν1]R,

    xh4nWsurf([ˉˉyn(x)](1))=h4nWsurf([eB(ˆx)ˉˉyn(x)](1)),

    converge uniformly to

    12Qsurf((G1(x)ν22(ν1)G2(x)000)Z+12(ν1)G3(x))=12Qsurf((symG1(x)12G2(x)000)Z(1)+12v(x)4(ν1)M(1)).

    Similarly, the mappings S×[ν2ν1,1]R,

    xh4nWsurf([ˉˉyn(x)](2))=h4nWsurf([eB(ˆx)ˉˉyn(x)](2)),

    converge uniformly to

    12Qsurf((symG1(x)+12G2(x)000)Z(1)+12v(x)4(ν1)M(1)).

    So with Soutn such that ˜Ωoutn=Soutn×(0,1),

    1h4n(ν1)SoutnWsurf([ˉˉyn(x,12(ν1))](1))+Wsurf([ˉˉyn(x,2ν32(ν1))](2))dxS1ν1Qsurf((symG1(x)000)Z(1)+12v(x)4(ν1)M(1))+14(ν1)Qsurf((G2(x)000)Z(1))dx. (58)

    Summarizing (58) and (57), we have shown that

    limnh4nEn(yn)=limnε3nh5nx˜ΛnW(x,ˉyn(x))=E(ν)vK(u,v)

    as n, where we have also used that the contribution of the lateral boundary cells ε3nh5nx˜ΛnW(x,ˉyn(x)) is negligible in the limit n.

    Proof of the energy barrier in Theorem 2.3. If a sequence of wnSδ satisfies the energy bound En(wn)Ch4n, then the proof of Proposition 1 shows ε3nhnEatom(wn)Ch4n. Hence,

    dist2(ˉwn(x),SO(3)Z)CEatom(wn)Ch5nε3n=C(νn1)5ε2n,

    which tends to 0 by assumption. This implies that wnSδ/2 for n large enough.

    This work was supported by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under project number 285722765, as well as the Engineering and Physical Sciences Research Council (EPSRC) under the grant EP/R043612/1.



    [1] A general integral representation result for continuum limits of discrete energies with superlinear growth. SIAM J. Math. Anal. (2004) 36: 1-37.
    [2] Numerical solution of a Föppl–von Kármán model. SIAM J. Numer. Anal. (2017) 55: 1505-1524.
    [3] From molecular models to continuum mechanics. Arch. Ration. Mech. Anal. (2002) 164: 341-381.
    [4] Connecting atomistic and continuous models of elastodynamics. Arch. Ration. Mech. Anal. (2017) 224: 907-953.
    [5]

    J. Braun and B. Schmidt, Existence and convergence of solutions of the boundary value problem in atomistic and continuum nonlinear elasticity theory, Calc. Var. Partial Differential Equations, 55 (2016), 36pp.

    [6] On the passage from atomistic systems to nonlinear elasticity theory for general multi-body potentials with p-growth. Netw. Heterog. Media (2013) 8: 879-912.
    [7] Confining thin elastic sheets and folding paper. Arch. Ration. Mech. Anal. (2008) 187: 1-48.
    [8] Energy minimizing configurations of pre-strained multilayers. J. Elasticity (2020) 140: 303-335.
    [9]

    M. de Benito Delgado and B. Schmidt, A hierarchy of multilayered plate models, ESAIM Control Optim. Calc. Var., 27 (2021), 35pp.

    [10] Cauchy-Born rule and the stability of crystalline solids: Static problems. Arch. Ration. Mech. Anal. (2007) 183: 241-297.
    [11] A scheme for the passage from atomic to continuum theory for thin films, nanotubes and nanorods. J. Mech. Phys. Solids (2000) 48: 1519-1540.
    [12] The Föppl-von Kármán plate theory as a low energy Γ-limit of nonlinear elasticity. C. R. Math. Acad. Sci. Paris (2002) 335: 201-206.
    [13] A hierarchy of plate models derived from nonlinear elasticity by gamma-convergence. Arch. Ration. Mech. Anal. (2006) 180: 183-236.
    [14] A theorem on geometric rigidity and the derivation of nonlinear plate theory from three-dimensional elasticity. Comm. Pure Appl. Math. (2002) 55: 1461-1506.
    [15] The nonlinear membrane model as variational limit of nonlinear three-dimensional elasticity. J. Math. Pures Appl. (9) (1995) 74: 549-578.
    [16] Interpenetration of matter in plate theories obtained as Γ-limits. ESAIM Control Optim. Calc. Var. (2017) 23: 119-136.
    [17] Justification of the Cauchy-Born approximation of elastodynamics. Arch. Ration. Mech. Anal. (2013) 207: 1025-1073.
    [18] A derivation of continuum nonlinear plate theory from atomistic models. Multiscale Model. Simul. (2006) 5: 664-694.
    [19] On the derivation of linear elasticity from atomistic models. Netw. Heterog. Media (2009) 4: 789-812.
    [20] On the passage from atomic to continuum theory for thin films. Arch. Ration. Mech. Anal. (2008) 190: 1-55.
    [21] Qualitative properties of a continuum theory for thin films. Ann. Inst. H. Poincaré C Anal. Non Linéaire (2008) 25: 43-75.
  • This article has been cited by:

    1. Mario Santilli, Bernd Schmidt, A Blake-Zisserman-Kirchhoff theory for plates with soft inclusions, 2023, 175, 00217824, 143, 10.1016/j.matpur.2023.05.005
    2. Bernd Schmidt, Jiří Zeman, A continuum model for brittle nanowires derived from an atomistic description by $$\Gamma $$-convergence, 2023, 62, 0944-2669, 10.1007/s00526-023-02562-y
    3. Manuel Friedrich, Leonard Kreutz, Konstantinos Zemas, Derivation of effective theories for thin 3D nonlinearly elastic rods with voids, 2024, 34, 0218-2025, 723, 10.1142/S0218202524500131
    4. Bernd Schmidt, Jiří Zeman, A Bending-Torsion Theory for Thin and Ultrathin Rods as a \(\boldsymbol{\Gamma}\)-Limit of Atomistic Models, 2023, 21, 1540-3459, 1717, 10.1137/22M1517640
  • Reader Comments
  • © 2022 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(807) PDF downloads(179) Cited by(4)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog