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

Oscillation criteria for sublinear and superlinear first-order difference equations of neutral type with several delays

  • Received: 02 February 2022 Revised: 17 June 2022 Accepted: 18 June 2022 Published: 01 August 2022
  • MSC : 34C10, 39A10, 39A12, 39A21

  • The purpose of this paper is to investigate the oscillatory behaviour of a class of first-order sublinear and superlinear neutral difference equations. Some conditions are established by applying Banach's Contraction mapping principle, Knaster-Tarski fixed point theorem and using several inequalities. We provide some examples to illustrate the outreach of the main results.

    Citation: Mohamed Altanji, Gokula Nanda Chhatria, Shyam Sundar Santra, Andrea Scapellato. Oscillation criteria for sublinear and superlinear first-order difference equations of neutral type with several delays[J]. AIMS Mathematics, 2022, 7(10): 17670-17684. doi: 10.3934/math.2022973

    Related Papers:

    [1] Roberto Alicandro, Giuliano Lazzaroni, Mariapia Palombaro . Derivation of a rod theory from lattice systems with interactions beyond nearest neighbours. Networks and Heterogeneous Media, 2018, 13(1): 1-26. doi: 10.3934/nhm.2018001
    [2] Roberto Alicandro, Andrea Braides, Marco Cicalese . Phase and anti-phase boundaries in binary discrete systems: a variational viewpoint. Networks and Heterogeneous Media, 2006, 1(1): 85-107. doi: 10.3934/nhm.2006.1.85
    [3] 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
    [4] Thomas Hudson . Gamma-expansion for a 1D confined Lennard-Jones model with point defect. Networks and Heterogeneous Media, 2013, 8(2): 501-527. doi: 10.3934/nhm.2013.8.501
    [5] 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
    [6] Claudio Canuto, Anna Cattani . The derivation of continuum limits of neuronal networks with gap-junction couplings. Networks and Heterogeneous Media, 2014, 9(1): 111-133. doi: 10.3934/nhm.2014.9.111
    [7] 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
    [8] Michele Gianfelice, Marco Isopi . On the location of the 1-particle branch of the spectrum of the disordered stochastic Ising model. Networks and Heterogeneous Media, 2011, 6(1): 127-144. doi: 10.3934/nhm.2011.6.127
    [9] 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
    [10] 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
  • The purpose of this paper is to investigate the oscillatory behaviour of a class of first-order sublinear and superlinear neutral difference equations. Some conditions are established by applying Banach's Contraction mapping principle, Knaster-Tarski fixed point theorem and using several inequalities. We provide some examples to illustrate the outreach of the main results.



    In this paper we study an atomistic model for (possibly heterogeneous) nanowires. We consider a scaling of the energy that corresponds to a reduction of the system from N dimensions to one dimension and, in addition, accounts for transitions between different equilibria.

    Specifically, in the homogeneous case, we study the asymptotic behaviour of the energy defined by

    εε(u):=i,jZN|ij|R(|u(εi)u(εj)|ε|ij|)p, (0.1)

    where p>1 and u is a deformation of the portion of the lattice εZN modelling the nanowire; the small parameter ε>0 represents the atomic distance and R>0 is sufficiently large to include a certain number of interactions beyond nearest neighbours. The above sum is taken over a "thin" domain, i.e., a domain consisting of a few lines of atoms (for the precise formula see (1.4)); as the lattice distance converges to zero, we perform a discrete to continuum limit and a dimension reduction simultaneously.

    This model was first studied in [14,15] under the assumption that the admissible deformations satisfy the non-interpenetration condition, namely, that the Jacobian determinant of a suitably defined piecewise affine interpolation of u is positive. Here we remove such assumption and we show that, by incorporating into the energy the effect of interactions in a certain finite range, one can recover the results of [14,15] and get even further insight into the problem. More precisely, we obtain an effective energy that accounts for the effects of changes of orientation in the lattice. The latter are thus allowed, but energetically penalised. We remark that, in dimension two, our analysis corresponds to the first-order Γ-limit of a functional of the kind studied in [1,18] without non-interpenetration assumptions. We also point out that the effects of long range interactions in non-convex lattice systems have already been analysed in [7,5,6] in the one-dimensional case.

    For the scaling of (0.1), we obtain a complete description of the Γ-limit with respect to two different topologies (Theorems 5.1 and 5.4). It turns out that the Γ-limit with respect to the topology used in [14,15] is trivial (see Remark 4), that is, one can exhibit recovery sequences for which the gradient always lies in the same energy well up to an asymptotically vanishing correction. In order to see the effects of changes of orientation in the nanowire, we introduce a stronger topology which is sensitive to them. In this case, for each change of orientation, the Γ-limit gives a finite positive contribution which is characterised by a discrete optimal transition problem. Moreover, one can prove that if we prescribe affine boundary conditions of the type xBx with dist(B;SO(N)) sufficiently small, then recovery sequences for minimisers will always preserve orientation (Remark 6). In this respect our model is consistent with the non-interpenetration condition. On the other hand, we also show that minimisers may violate such condition if we add to the functionals pathological loading terms which force the deformations to overcome the energetic barrier between equilibria with opposite orientation (see Section 6 and Remark 7).

    The Γ-limit is nontrivial, also in the weaker topology, when one considers heterogeneous nanowires, which consist of components with different equilibria, arranged longitudinally; i.e., the interface between the components is a cross-section of the rod. In this case, we prove an estimate on the minimal energy spent to match the equilibria. Precisely, denoting by kN the number of atomic layers of the nanowire, we show that the minimal cost grows faster than kN1. The proof of such result (Theorem 2.2) follows as an application of [2,Theorem 3.1]. Such lower bound is to be compared with the estimate that one can prove in the case of a two-or three-dimensional model accounting for dislocations. This is discussed in Section 7 where we compare the minimal energy of heterogeneous defect-free systems and the minimal energy of heterogeneous systems containing dislocations. It turns out that for sufficiently large values of k, the latter are energetically preferred since their energy may grow exactly like kN1 (see Remark 8). In this respect our result is consistent with the one proven in [14,15] under the non-interpenetration assumption. We recall that the first variational justification of dislocation nucleation in nanowire heterostructures was obtained in [17] in the context of non-linear elasticity. This result was later generalised to a discrete to continuum setting in [14,15] under the non-interpenetration condition, and is here validated without the latter assumption. More recently, variational models for misfit dislocations at semi-coherent interfaces and in elastic thin films have been proposed in [10] and [11] respectively.

    The paper is organised as follows. In Section 1 we introduce the model. In Section 1 we introduce the minimal costs to bridge different equilibria and study their dependence on the thickness of the nanowire. In Sections 3-5, performing a discrete to continuum limit and a dimension reduction simultaneously, we characterise the Γ-limit of the energy functional for different choices of the topology (Theorems 5.1 and 5.4). All the results are stated in the general case of heterogeneous nanowires. In Section 6 we discuss the effect of boundary conditions on the Γ-limit and briefly study a model including external forces (only in the homogeneous case, for simplicity). In the final part of the paper, Section 7, we compare the model for defect-free nanowires with models including dislocations at the interface, showing that the latter are energetically favoured.

    Notation. We recall some basic notions of geometric measure theory for which we refer to [3]. Given a bounded open set ΩRN, N2, and M1, BV(Ω;RM) denotes the space of functions of bounded variation; i.e., of functions uL1(Ω;RM) whose distributional gradient Du is a Radon measure on Ω with |Du|(Ω)<+, where |Du| is the total variation of Du. If uBV(Ω;RM), the symbol u stands for the density of the absolutely continuous part of Du with respect to the N-dimensional Lebesgue measure LN. We denote by Ju the jump set of u, by u+ and u the traces of u on Ju, and by νu(x) the measure theoretic inner normal to Ju at x, which is defined for HN1-a.e. xJu, where HN1 is the (N1)-dimensional Hausdorff measure. A function uBV(Ω;RM) is said to be a special function of bounded variation if DuuLN is concentrated on Ju; in this case one writes uSBV(Ω;RM). Given a set EΩ, we denote by P(E,Ω) its relative perimeter in Ω and by E its reduced boundary. We recall that a partition {Ei}iN of Ω is called a Caccioppoli partition if iNP(Ei,Ω)<+. Given a rectifiable set KΩ, we say that a Caccioppoli partition {Ei}iN of Ω is subordinated to K if for every iN the reduced boundary Ei of Ei is contained in K, up to a HN1-negligible set.

    For N2, MN×N is the set of real N×N matrices, GL+(N) is the set of matrices with positive determinant, O(N) is the set of orthogonal matrices, and SO(N) is the set of rotations. We denote by I the identity matrix and J the reflection matrix such that Je1=e1 and Jei=ei for i=2,,N, where {ei:i=2,,N} is the canonical basis in RN. The symbol co(X) stands for the convex hull of a set X in MN×N. Moreover, given N+1 points x0,x1,,xNRN, we denote by [x0,x1,,xN] the simplex determined by all convex combinations of those points.

    Finally, U is the class of subsets of (L,L) that are disjoint union of a finite number of open intervals.

    In the paper, the same letter C denotes various positive constants whose precise value may change from place to place.

    We study the dimension reduction of a discrete model for heterogeneous nanowires. Let L>0, kN, Ωkε:=(L,L)×(kε,kε)N1. Up to an affine deformation HGL+(N), we can reduce to the case where the lattice is ZN. Thus we consider the discrete thin domain Lε(k)RN defined as

    Lε(k):=εZN¯Ωkε, (1.1)

    where ¯Ωkε is the union of all hypercubes with vertices in εZN that have non-empty intersection with Ωkε. In the physically relevant case of N=3, the set Lε(k) models the crystal structure of a nanowire of length 2L and thickness 2kε, where k is the number of parallel atomic planes. We will nonetheless state all the results for a general N, since their proof does not depend on the dimension. Notice that in definition (1.1) the dependence on k is explicit; this parameter will indeed play a major role in the subsequent analysis.

    The bonds between the atoms are defined by means of the so-called Kuhn decomposition, which is relevant for modelling some specific Bravais lattices. (See [2,Remark 2.6] for details on the treatment of some lattices in dimension two and three, such as the hexagonal or equilateral triangular, the face-centred cubic, and the body-centred cubic.) First we define a partition T0 of the unit cube (0,1)N into N-simplices: we say that TT0 if the (N+1)-tuple of its vertices belongs to the set

    {{0,ei1,ei1+ei2,,ei1+ei2++eiN}:(12Ni1i2iN)SN},

    where SN is the set of permutations of N elements; see Figure 1. Next, we define T as the periodic extension of T0 to all of RN. We say that two nodes x,yZN are contiguous if there exists a simplex TT that has both x and y as its vertices. We set

    Figure 1. 

    The six tetrahedra in the Kuhn decomposition of a three-dimensional cube

    .
    B1:={ξRN:x and x+ξ are contiguous}. (1.2)

    If both simplices [x0,x1,,xN] and [y0,x1,,xN] belong to T, then we say that [x0,x1,,xN] and [y0,x1,,xN] are neighbouring simplices (i.e., they share a facet) and x0 and y0 are opposite vertices. We set

    B2:={ξRN:x and x+ξ are opposite vertices}, (1.3)

    and remark that, by periodicity, B1 and B2 do not depend on x.

    We assume that Lε(k) is composed of two species of atoms, occupying the points contained in the subsets

    Lε(k):={xLε(k):x1<0},L+ε(k):={xLε(k):x10},

    respectively, where x=(x1,,xN). The two species of atoms are characterised by equilibrium distances given by ε and λε, respectively, where λ(0,1] is fixed; the case λ(0,1) models a heterogeneous nanowire, while the case λ=1 refers to a homogeneous nanowire. Specifically, the total interaction energy relative to a deformation u:Lε(k)RN is defined as

    ε1,λε(u,k):=xLε(k)ξB1B2x+εξLε(k)c(ξ)||u(x+εξ)u(x)|ε|Hξ||p+xL+ε(k)ξB1B2x+εξLε(k)c(ξ)||u(x+εξ)u(x)|ελ|Hξ||p, (1.4)

    where p>1, HGL+(N), and the coefficient c(ξ) is equal to some c1>0 for ξB1 and to c2>0 for ξB2.

    To simplify the presentation, we restrict our attention to the case of p-harmonic potentials, though our analysis applies, without any significant change, to more general potentials satisfying polynomial growth conditions. More precisely, we may replace ε1,λε(u,k) with

    xLε(k)ξB1B2x+εξLε(k)ϕ1(ξ,|u(x+εξ)u(x)|e|Hξ|)+xL+ε(k)ξB1B2x+εξLε(k)ϕλ(ξ,|u(x+εξ)u(x)|eλ|Hξ|),

    where ϕ:ZN×R[0,+) is a positive potential such that

    C1|z|pϕμ(ξ,z)C2|z|pfor μ=λ,1,

    for some positive constants C1,C2. One could consider the case of potentials depending also on x and satisfying suitable periodicity assumptions: this would require a more delicate analysis and would lead to a more complex formula for the Γ-limit.

    In principle, all the results that we present in the sequel extend to the case when the two components of the nanowire have equilibria of the form H and H+, where H,H+GL+(N). We have chosen to analyse the case when H+=λH, since this is particularly meaningful in applications where one has misfit between two crystalline materials with the same lattice structure but different lattice distance at equilibrium (see e.g. [9,13]).

    We study the limit behaviour of ε1,λε(,k) as ε0+, thus performing simultaneously a discrete to continuum limit and a dimension reduction to a one-dimensional system. The limit functional was derived in [14,15] by means of Γ-convergence, under the assumption that the admissible deformations fulfil the non-interpenetration condition, namely, that the Jacobian determinant of (the piecewise affine interpolation of) any deformation is strictly positive almost everywhere. The non-interpenetration assumption was used in several parts of the analysis; in particular, it was needed to prove that the limit functional (dependent on k) scales like kN as k.

    The main novelty of the present paper is that we remove the non-interpenetration assumption made in [14,15], allowing for changes of orientations. Furthermore, in the study of the Γ-limit we define a stronger topology that accounts for such changes. In the proof of the new results, only those parts that differ from [14,15] will be shown in details.

    In the sequel of the paper we will often consider the rescaled domain 1εΩkε, which converges, as ε0+, to the unbounded strip

    Ωk,:=R×(k,k)N1.

    We define the associated lattice and subsets

    L(k):=ZN¯Ωk,,L(k):={xL(k):x1<0},L+(k):={xL(k):x10},

    where ¯Ωk, is the union of all hypercubes with vertices in ZN that have non-empty intersection with Ωk,. For u:L(k)RN we define

    E1,λ(u,k):=xL(k)ξB1B2x+ξL(k)c(ξ)||u(x+ξ)u(x)||Hξ||p+xL+(k)ξB1B2x+ξL(k)c(ξ)||u(x+ξ)u(x)|λ|Hξ||p. (1.5)

    We identify every deformation u of the lattice Lε(k) by its piecewise affine interpolation with respect to the triangulation εT. By a slight abuse of notation, such extension is still denoted by u. We can then define the domain of the functional (1.4) as

    Aε(Ωkε):={uC0(¯Ωkε;RN):u piecewise affine, u constant on ΩkεεT TT}.

    Similarly, for (1.5) we define

    A(Ωk,):={uC0(¯Ωk,;RN):u piecewise affine, u constant on Ωk,T TT}.

    As customary in dimension reduction problems, we rescale the domain Ωkε to a fixed domain Ωk, independent of ε, by introducing the change of variables z(x):=(x1,εx2,,εxN). Accordingly, for each uAε(Ωkε) we define ˜u(x):=u(z(x)). Moreover we set Ωk:=A1ε(Ωkε)=(L,L)×(k,k)N1, where AεMN×N is the diagonal matrix

    Aε:=diag(1,ε,,ε); (1.6)

    i.e., z(x)=Aεx. In this way we can recast the functionals (1.4) defined over varying domains into functionals defined on deformations of the fixed domain Ωk. Precisely we set

    I1,λε(˜u,k):=E1,λε(u,k)for ˜u˜Aε(Ωk), (1.7)

    with

    ˜Aε(Ωk):={˜uC0(A1ε(¯Ωkε);RN):˜u piecewise affine, ˜u constant on Ωk(A1εεT) TT}.

    For later use it will be convenient to set the following notation:

    Ωk:=(L,0)×(k,k)N1,Ω+k:=(0,L)×(k,k)N1.

    We recall that, throughout the paper, I is the identity matrix and J is the reflection matrix such that Je1=e1 and Jei=ei for i=2,,N.

    We will study the Γ-limit of the sequence I1,λε(,k) as ε0+ for every fixed k. For this purpose we introduce the quantity γ(P1,P2;k) for P1,P2O(N)λO(N), which represents the minimum cost of a transition from a well to another. Specifically, for each P1O(N) and P2λO(N) we define

    γ(P1,P2;k):=inf{E1,λ(v,k):M>0, vA(Ωk,),v=P1H for x1(,M),v=P2H for x1(M,+)}; (2.1a)

    for P1,P2O(N)

    γ(P1,P2;k):=inf{E1,1(v,k):M>0, vA(Ωk,),v=P1H for x1(,M),v=P2H for x1(M,+)}, (2.1b)

    where

    E1,1(v,k):=xL(k)ξB1B2x+ξL(k)c(ξ)||v(x+ξ)v(x)||Hξ||p;

    for P1,P2λO(N)

    γ(P1,P2;k):=inf{Eλ,λ(v,k):M>0,vA(Ωk,),v=P1Hforx1(,M),v=P2Hforx1(M,+)}, (2.1c)

    where

    Eλ,λ(v,k):=xL(k)ξB1B2x+ξL(k)c(ξ)||v(x+ξ)v(x)|λ|Hξ||p.

    The next proposition shows that the relevant quantities defined through (2.1) are in fact four: the minimal costs of the transition at the interface between the energy wells O(N) and λO(N) are provided in (2.2b) and (2.2c); the minimal cost of the transition between SO(N) and O(N)SO(N) is provided in (2.2d); the one of the transition between λSO(N) and λO(N)SO(N) is given in (2.2e). Moreover, the constants in (2.2d) and in (2.2e) are related by the proportionality rule (2.3).

    Proposition 1. For each kN, the function γ satisfies for every R,RSO(N) and Q,QO(N)SO(N)

    γ(R,R;k)=γ(Q,Q;k)=γ(λR,λR;k)=γ(λQ,λQ;k)=0, (2.2a)
    γ(R,λR;k)=γ(Q,λQ;k)=γ(I,λI;k), (2.2b)
    γ(R,λQ;k)=γ(Q,λR;k)=γ(I,λJ;k), (2.2c)
    γ(R,Q;k)=γ(Q,R;k)=γ(I,J;k),and (2.2d)
    γ(λR,λQ;k)=γ(λQ,λR;k)=γ(λI,λJ;k). (2.2e)

    Moreover,

    γ(λP1,λP2;k)=λpγ(P1,P2;k)forevery P1,P2O(N). (2.3)

    Proof. First one notices that γ(P1,P2;k)=γ(JP1,JP2;k). Hence, the proof of (2.2) relies on the construction of low energy transitions between two given rotations or two given rotoreflections, see [14,Proposition 2.4]. Finally, standard comparison arguments yield (2.3).

    We now prove estimates on the asymptotic behaviour of γ(I,λI) and γ(I,λJ) as k, which have interesting consequences towards the comparison of this model with those accounting for dislocations in nanowires, see Section 7 below. Indeed, in Theorem 2.2 below we show that for λ1 (heterogeneous nanowire) these constants grow faster than kN1, while it is known that the corresponding minimum cost for nanowires with dislocations scales like kN1 (see discussion at the end of Section 7). In contrast, we remark that for λ=1 one has γ(I,I)=0 and γ(I,J)CkN1. An essential tool in the proof of Theorem 2.2 is the following result.

    Theorem 2.1.[2,Theorem 3.1] Let uεAε((0,1)N) be a sequence such that

    εN1ξB1B2x,x+εξεZN(0,1)N||uε(x+εξ)uε(x)|ε|Hξ||p<C. (2.4)

    Then there are a subsequence (not relabelled) and a function uW1,((0,1)N;RN) such that uεu in Lp((0,1)N;MN×N) and

    uSBV((0,1)N;O(N)H). (2.5)

    Specifically, u is a collection of an at most countable family of rigid deformations, i.e., there exists a Caccioppoli partition {Ei}iN subordinated to the reduced boundary {uSO(N)H}, such that

    u(x)=iN(RiHx+bi)χEi(x), (2.6)

    where RiO(N) and biRN. Moreover, if EiEj, then detRidetRj=1 and EiEj is flat, i.e., the measure theoretic normal vector to EiEj is constant (up to the sign).

    We now prove the main result of this section.

    Theorem 2.2. Let λ(0,1) and (P1,P2){(I,λI),(I,λJ)}. There exists C>0 such that

    γ(P1,P2;k)CkN. (2.7)

    Moreover,

    limkγ(P1,P2;k)kN1=+. (2.8)

    Proof. The upper bound (2.7) is proven by comparing test functions for γ(P1,P2;k) with those for γ(P1,P2;1). Namely, let vA(Ω1,) be such that v(x)=P1Hx for every xL(k) and v(x)=P2Hx for every xL+(k); in particular, v=P1H for x1(,1) and v=P2H for x1(0,+). Then one defines uA(Ωk,) by u(x):=kv(x/k), which yields γ(P1,P2;k)E1,λ(u,k)CE1,λ(v,1)kN, and thus γ(P1,P2;k)Cγ(P1,P2;1)kN. Note that in the previous inequalities one uses the fact that vL and that the energy of the interactions in B2 can be bounded, using the Mean Value Theorem, by the energy of the interactions in B1.

    For the proof of the lower bound (2.8) we will use Theorem 2.1 in each of the subsets (1,0)×(1,1)N1 and (0,1)×(1,1)N1. By contradiction, suppose that there exist a sequence kj and a sequence {uj}A(Ωkj,) such that

    1kN1jE1,λ(uj,kj)<C, (2.9)

    for some positive C. Define vj:Ω1,RN as vj(x):=1kjuj(kjx). Accordingly, we consider the rescaled lattices

    Lj:=1kjZN¯Ω1,,L+j:=Lj{x1>0},Lj:=Lj{x1<0}.

    Expressing E1,λ(uj,kj) in terms of vj, one finds

    E1,λ(uj,k)=xLjξ/kjB1B2x+ξ/kjLjc(ξ)||vj(x+ξkj)vj(x)|1kj|Hξ||p+xL+jξ/kjB1B2x+ξ/kjLjc(ξ)||vj(x+ξkj)vj(x)|1kjλ|Hξ||p. (2.10)

    The above term controls the (piecewise constant) gradient of vj. From (2.9), (2.10), and Theorem 2.1 we deduce that, up to subsequences, vjv in Lp((1,1)N;MN×N), for some vW1,((1,1)N;RN), where vO(N)H for a.e. x(1,0)×(1,1)N1 and vλO(N)H for a.e. x(0,1)×(1,1)N1. Precisely,

    v(x)=iN(RiHx+ai)χEi(x)+jN(λQjHx+bj)χE+j(x),

    where Ri,QjO(N), ai,bjRN, and {Ei} (respectively {E+j}) is a Caccioppoli partition of (1,0)×(1,1)N1 (respectively of (0,1)×(1,1)N1). Then, since {Ei}{E+j} is a Caccioppoli partition of (1,1)N, by the local structure of Caccioppoli partitions (see e.g. [3,Theorem 4.17]), we find that, for HN1-a.e. x{0}×(k,k)N1, xEiE+j for some i,j (where E denotes the reduced boundary of E). Therefore, using a blow-up argument and the fact that vW1,((1,1)N;RN), we deduce that there exist rank-1 connections between O(N)H and λO(N)H; see [2,Lemma 3.2]. This implies in particular that λ=1, which is a contradiction to λ(0,1). Hence (2.8) follows.

    Remark 1. An estimate similar to (2.8) was proven in [14,15] (for a hexagonal lattice in dimension two and a class of three-dimensional lattices) via a different argument, based on the non-interpenetration condition. In fact, in [14,15] a stronger result is proven, namely, that γ(I,λI;k) scales like kN.

    The non-interpenetration assumption turns out to be necessary if the energy involves only nearest neighbour interactions; indeed, in such a case, one can exhibit deformations that violate the non-interpenetration condition and for which (2.8) does not hold, see [14,Section 4.2]. Such deformations, which consist of suitable foldings of the lattice, would be energetically expensive (and, in particular, would not provide a counterexample to (2.8)) in the present setting, exactly because of the effect of the interactions across neighbouring cells. It is the latter ones that prevent folding phenomena and allow one to prove (2.8), via Theorem 2.1.

    Before characterising the Γ-convergence for the rescaled functionals (1.7), we show a compactness theorem for sequences with equibounded energy, as well as bounds from above and from below on those functionals in terms of the changes of orientation in the wire. Such bounds will be used in the proof of the Γ-convergence results, Theorems 5.1 and 5.4.

    Essential tools for the compactness and the lower bound are provided by the following rigidity estimates.

    Theorem 3.1. [12,Theorem 3.1] Let N2, and let 1<p<. Suppose that URN is a bounded Lipschitz domain. Then there exists a constant C=C(U) such that for each uW1,p(U;RN) there exists a constant matrix RSO(N) such that

    uRLp(U;MN×N)C(U)dist(u,SO(N))Lp(U). (3.1)

    The constant C(U) is invariant under dilation and translation of the domain.

    It is convenient to define the energy of a single simplex T with vertices x0,,xN,

    Ecell(uF;T):=Nij=0||F(xixj)||H(xixj)||pfor every FMN×N,

    where uF is the affine map uF(x):=Fx. The following lemma provides a lower bound on Ecell(uF;T) in terms of the distance of F from O(N). It will be instrumental in using Theorem 3.1.

    Lemma 3.2. [2,Lemma 2.2] There exists a constant C>0 such that

    distp(F,SO(N)H)CEcell(uF;T)FMN×N:detF0, (3.2a)
    distp(F,(O(N)SO(N))H)CEcell(uF;T)FMN×N:detF0. (3.2b)

    The next lemma asserts that if in two neighbouring simplices the sign of detu has different sign, then the energy of those two simplices is larger than a positive constant. It will be convenient to define the energetic contribution of the interactions within two neighbouring simplices T=[x0,x1,,xN], S=[y0,x1,,xN] as

    Ecell(u;ST):=Nij=0||u(xi)u(xj)||H(xixj)||p
    +Nj=1||u(y0)u(xj)||H(y0xj)||p+||u(y0)u(x0)||H(y0x0)||p.

    Lemma 3.3. [2,Lemma 2.3] There exists a positive constant C0 (depending on H) with the following property: if two neighbouring N-simplices S, T have different orientations in the deformed configuration, i.e.,

    det(u|S)det(u|T)0,

    then Ecell(u;ST)C0.

    Lemma 3.2 will allow us to apply Theorem 3.1. More precisely, in the part of the wire with x1(L,0) we use (3.1) or its "symmetric" version for O(N)SO(N) in subdomains that scale in such a way that the constant of the rigidity estimate does not change; for x1(0,L) we use corresponding estimates for λSO(N) or λ(O(N)SO(N)). Thus we approximate the deformation gradient with piecewise constant matrices in O(N), respectively λO(N).

    Due to the fact that a minimum energy has to be paid for each change of orientation, see Lemma 3.3, the parts with positive determinant do not mix with those with negative determinant. Hence, passing to the weak* limit we obtain functions taking values in co(SO(N))co(O(N)SO(N)), respectively λco(SO(N))λco(O(N)SO(N)). Here, co(X) denotes the convex hull of a set X in MN×N.

    Remark 2. It is well known that co(SO(N))co(O(N)SO(N)): indeed, the intersection always contains the zero matrix, here denoted by 0. In dimension N=2, one can see that

    co(SO(2))={(αββα):α2+β21},co(O(2)SO(2))={(αββα):α2+β21}.

    In particular, co(SO(2))co(O(2)SO(2))={0}. For N>2, the intersection is nontrivial. For example, co(SO(3))co(O(3)SO(3)) contains the matrix 13I. Moreover, one can see that

    co(SO(N))co(O(N)SO(N))

    for N\ge2.

    Henceforth, the symbol \mathcal{U} stands for the class of subsets of (-L, L) that are disjoint union of a finite number of open intervals.

    Proposition 2. Let {\tilde{u}}_\varepsilon\in {\tilde {\mathcal{A}}}_{\varepsilon}(\Omega _{k}) be a sequence such that

    \label{succ-limitata} \limsup\limits_{\varepsilon\to 0^{+}} \mathcal{I}_{\varepsilon }^{1, \lambda }({\tilde{u}}_\varepsilon, k) \leq C \, . (3.3)

    Then there exist functions {\tilde{u}} \in W^{1, \infty}(\Omega _k;\mathbb{R}^N), d_1, \dots, d_N\in L^{\infty}(\Omega _k;\mathbb{R}^N), and a subsequence (not relabelled) such that

    (3.4)

    and {\tilde{u}}, d_1, \dots, d_N, are independent of x_2, \dots, x_N, i.e., \partial_{j}{\tilde{u}} = \partial_{j} d_i = 0, for each i = 2, \dots, N and j = 2, \dots, N. Moreover, there exists U\in\mathcal{U} such that

    \label{charact} (\partial_1 {\tilde{u}} \, | \, d_2 \, | \cdots \, | d_N)\in \begin{cases} {\rm{co}}(SO(N))H &\;\;\;\;\; a.e.\;\;in\ (-L, 0)\cap U \, , \\ {\rm{co}}(O(N) {\setminus} SO(N))H &\;\;\;\;\; a.e.\;\;in\ (-L, 0)\setminus U\, , \\ \lambda\, {\rm{co}}(SO(N))H & \;\;\;\;\;a.e.\;\;in\ (0, L) \cap U \, , \\ \lambda\, {\rm{co}}(O(N) {\setminus} SO(N))H & \;\;\;\;\;a.e.\;\;in\ (0, L)\setminus U \, , \end{cases} (3.5)

    and

    \label{gammaliminf} \begin{split} \liminf\limits_{\varepsilon\to 0^{+}} \mathcal{I}_{\varepsilon }^{1, \lambda }(\tilde u_\varepsilon, k)\geq&\ \gamma(I, J;k) \, \mathcal{H}^0(\partial U\cap(-L, 0)) + \gamma(\lambda I, \lambda J;k) \, \mathcal{H}^0(\partial U\cap(0, L)) \\ & + \gamma(I, \lambda I;k) \left[1-\chi_{\partial U}(0) \right] + \gamma(I, \lambda J;k)\, \chi_{\partial U}(0) \, . \end{split} (3.6)

    Remark 3. The right-hand side of (3.6) contains different contributions. The first term corresponds to the minimal energy needed to bridge a rotation with a rotoreflection, or viceversa, in the left part of the nanowire; the energy spent depends on the number of changes of orientation, i.e., on the cardinality of \partial U. The second term plays an analogous role for the right part of the nanowire. The remaining terms describe the interfacial energy spent to bridge the two energy wells O(N)H and \lambda\, O(N)H: that contribution also depends on whether or not the orientation is preserved across the interface, i.e., on whether 0 is an inner or external, or boundary point for U.

    Proof. (Compactness) The assumption (3.3) implies that \{\nabla u_\varepsilon\}, resp. \{\nabla {\tilde{u}}_\varepsilon A_\varepsilon^{-1}\}, is uniformly bounded in L^{\infty}(\Omega _k\varepsilon ;{{\mathbb{M}}^{N\times N}}), respectively L^{\infty}(\Omega _k;{{\mathbb{M}}^{N\times N}}). (Recall that u_\varepsilon(x) = \tilde{u}_\varepsilon(A_\varepsilon^{-1} x).) Therefore there exist a subsequence of \{{\tilde{u}}_\varepsilon\} (not relabelled) and functions {\tilde{u}}\in W^{1, \infty}(\Omega _k;\mathbb{R}^N) and d_i \in L^{\infty}(\Omega _k;\mathbb{R}^N) for i = 2, \dots, N, such that \partial_{1}{\tilde{u}}_\varepsilon \mathop * \limits_ \rightharpoonup \partial_{1} {\tilde{u}} weakly* in L^{\infty}(\Omega _k;{{\mathbb{M}}^{N\times N}}), where {\tilde{u}} is independent of x_i for all i = 2, \dots, N, and \frac{1}\varepsilon\partial_{i}{\tilde{u}}_\varepsilon \mathop * \limits_ \rightharpoonup d_i for each i = 2, \dots, N.

    In order to show \partial_{j}d_i = 0 and (3.5), we apply the rigidity estimate (3.1) to the sequence u_\varepsilon. To this aim, we divide the domain \overline{\Omega}_{k\varepsilon} into subdomains that are the Cartesian product of intervals (a_i, a_i +\varepsilon), a_i\in\varepsilon\mathbb{Z}, and the cross-section (k\varepsilon, k\varepsilon)^{N-1}. We first observe that, by Lemma 3.3 and assumption (3.3), the number of changes of orientation of u_\varepsilon is uniformly bounded in \varepsilon . More precisely, we can find a uniformly bounded number of subdomains (a_i, a_i +\varepsilon)\times(k\varepsilon, k\varepsilon)^{N-1}, i\in I_\varepsilon, \# I_\varepsilon \le C, such that if i\notin I_\varepsilon then \det\nabla u_\varepsilon has constant sign in (a_i, a_i +\varepsilon)\times(k\varepsilon, k\varepsilon)^{N-1}. In each of these subdomains, we use (3.2) to apply the rigidity estimate (3.1), or its "symmetric" version for O(N){\setminus}SO(N).

    Specifically, for each a_i with a_i < 0 and i\notin I_\varepsilon, there exists P_\varepsilon (a_i)\in O(N)H such that

    \int_{({a_i},{a_i} + \varepsilon ) \times {{(k\varepsilon ,k\varepsilon )}^{N - 1}}} |\nabla {u_\varepsilon } - {P_\varepsilon }({a_i}){|^p}{\mkern 1mu} {\rm{d}}x \le C\int_{({a_i},{a_i} + \varepsilon ) \times {{(k\varepsilon ,k\varepsilon )}^{N - 1}}} \;{\rm{dis}}{{\rm{t}}^p}(\nabla {u_\varepsilon },O(N)H){\mkern 1mu} {\rm{d}}x{\mkern 1mu} ,

    and for every a_i >0 with i\notin I_\varepsilon there exists P_\varepsilon (a_i)\in \lambda \, O(N) H such that

    \int_{({a_i},{a_i} + \varepsilon ) \times {{(k\varepsilon ,k\varepsilon )}^{N - 1}}} |\nabla {u_\varepsilon } - {P_\varepsilon }({a_i}){|^p}{\mkern 1mu} {\rm{d}}x \le C\int_{({a_i},{a_i} + \varepsilon ) \times {{(k\varepsilon ,k\varepsilon )}^{N - 1}}} \;\;{\rm{dis}}{{\rm{t}}^p}(\nabla {u_\varepsilon },\lambda {\mkern 1mu} O(N)H){\mkern 1mu} {\rm{d}}x{\mkern 1mu} .

    Moreover for i\in I_\varepsilon we set P_\varepsilon(a_i) = I if a_i <0 and P_\varepsilon(a_i) = \lambda I if a_i\ge0. By interpolation one defines a piecewise constant matrix field P_\varepsilon:(-L, L)\to O(N)H \cup \lambda \, O(N) H such that P_\varepsilon(x_{1}) = P_\varepsilon(a_i) if x_{1}\in (a_i, a_i+\varepsilon). Summing up over i and rescaling the variables, one gets

    \begin{align} &\int_{\Omega _k^-} |\nabla {\tilde{u}}_\varepsilon A_\varepsilon^{-1} - P_\varepsilon(x_1)|^{p} \, \text{d}x \leq C\!\! \int_{A_\varepsilon^{-1}(\overline{\Omega }_{k\varepsilon}) \cap\{x_1 < 0\}} \!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\! \text{dist}^{p}(\nabla {\tilde{u}}_\varepsilon A_\varepsilon^{-1}, O(N)H ) \, \text{d}x \leq C\varepsilon\, , \end{align} (3.7a)
    \begin{align}&\int_{\Omega _k^+} |\nabla {\tilde{u}}_\varepsilon A_\varepsilon^{-1} - P_\varepsilon(x_1)|^{p} \, \text{d}x \leq C\!\! \int_{A_\varepsilon^{-1}(\overline{\Omega }_{k\varepsilon}) \cap\{x_1 > 0\}} \!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\! \text{dist}^{p}(\nabla {\tilde{u}}_\varepsilon A_\varepsilon^{-1}, \lambda\, O(N) H ) \, \text{d}x \leq C\varepsilon\, , \end{align} (3.7b)

    where the last inequality of each line follows by applying Lemma 3.2 to each subdomain with i\notin I_\varepsilon and by recalling that each subdomain has volume proportional to \varepsilon after rescaling.

    We now define the sets

    \begin{align*} K_\varepsilon &: = \{ a_i^\varepsilon \in (-L, L) \colon P_\varepsilon(x_1) \in SO(N)H \cup \lambda\, SO(N) H \;\;\text{ for }\;\; x_1\in [a_i^\varepsilon, a_i^\varepsilon + \varepsilon)\} \, , \\ U_\varepsilon &: = \bigcup\limits_{a_i^\varepsilon\in K_\varepsilon} [a_i^\varepsilon, a_i^\varepsilon + \varepsilon)\, , \end{align*}

    and remark that Lemma 3.2, Lemma 3.3, and assumption (3.3) imply that the cardinality of \partial U_\varepsilon is uniformly bounded. Therefore the sequence \{\chi_{U_\varepsilon}\} converges, up to subsequences, to \chi_{U} strongly in L^1(-L, L), where

    \label{U} U = \bigcup\limits_{i = 1}^n (\alpha_i, \beta_i) \, , \;\;\;\;\;-L\le\alpha_1 < \beta_1 < \alpha_2 < \beta_2 < \dots < \alpha_n < \beta_n\le L \, . (3.8)

    Since we can write

    \begin{split} P_\varepsilon(x_1) = & R_\varepsilon(x_1)\big(\chi_{U_\varepsilon\cap (-L, 0)} H + \chi_{U_\varepsilon\cap (0, L)}\lambda H \big) \\ &+ JR_\varepsilon(x_1)\big((1-\chi_{U_\varepsilon\cap (-L, 0)}) H + (1-\chi_{U_\varepsilon\cap (0, L)})\lambda H \big)\, , \end{split}

    where R_\varepsilon : (-L, L) \to SO(N) is piecewise constant, we deduce that P_\varepsilon converges, up to subsequences, to some P\in L^\infty((-L, L);{{\mathbb{M}}^{N\times N}}) in the weak* topology of L^\infty((-L, L);{{\mathbb{M}}^{N\times N}}). From (3.7) it follows that the weak* limit of \nabla {\tilde{u}}_\varepsilon A_\varepsilon^{-1} coincides with P and therefore does not depend on x_j for each j = 2, \dots, N. Moreover, inclusion (3.5) follows from the fact that \chi_{U_\varepsilon}P_\varepsilon converges weakly* to \chi_{U}P.

    (Lower bound) Inequality (3.6) is proven by a standard argument which can be found, for example, in [14,16,17]. We will briefly sketch the main ideas and refer the reader to [14,16,17] for full details. First recall that \partial U consists of a finite number of points, cf. (3.8). Since \chi_{U_\varepsilon} \to \chi_U and since the number of points of \partial U_\varepsilon is uniformly bounded, one can find \sigma >0, \alpha_i^\varepsilon \to \alpha_i, \beta_i^\varepsilon \to \beta_i such that

    \begin{align}&(\alpha_i^\varepsilon -2\sigma, \alpha_i^\varepsilon - \sigma )\subset (-L, L)\setminus U_\varepsilon \, , \;\;\;\;\;(\alpha_i^\varepsilon +\sigma, \alpha_i^\varepsilon +2 \sigma ) \subset U_\varepsilon \, , \end{align} (3.9a)
    (\beta_i^\varepsilon -2\sigma, \beta_i^\varepsilon - \sigma ) \subset U_\varepsilon \, ,\;\;\;\;\; (\beta_i^\varepsilon +\sigma, \beta_i^\varepsilon +2 \sigma ) \subset (-L, L)\setminus U_\varepsilon \, . (3.9b)

    Moreover, if \sigma is sufficiently small, all the intervals (\alpha_i^\varepsilon -2\sigma, \alpha_i^\varepsilon + 2 \sigma ) and (\beta_i^\varepsilon -2\sigma, \beta_i^\varepsilon + 2 \sigma ) are mutually disjoint and therefore it suffices to prove the lower bound for one of such intervals. Suppose that \alpha_i^\varepsilon \in (0, L) and define

    v_\varepsilon(x_1, x_2, \dots, x_N) : = \tfrac{1}\varepsilon{\tilde{u}}_\varepsilon(\varepsilon x_1 + \alpha_i^\varepsilon, x_2, \dots, x_N) = \tfrac{1}\varepsilon u_\varepsilon(\varepsilon x_1 + \alpha_i^\varepsilon, \varepsilon x_2, \dots, \varepsilon x_N).

    Then, \nabla v_\varepsilon (x) = \nabla {\tilde{u}}_\varepsilon (\varepsilon x_1 + \alpha_i^\varepsilon, x_2, \dots, x_N) A_\varepsilon^{-1} = \nabla u_\varepsilon(\varepsilon x_1 + \alpha_i^\varepsilon, \varepsilon x_2, \dots, \varepsilon x_N), and, by (3.7), we have

    \label{ci-siamo-quasi} \begin{split}&\int_{(-\frac{2\sigma}\varepsilon, -\frac{\sigma}\varepsilon)\times (-k, k)^{N-1} } \text{dist}^{p}(\nabla v_\varepsilon, \lambda(O(N){\setminus} SO(N)) H ) \, \text{d}x \\&+ \int_{(\frac{\sigma}\varepsilon, \frac{2\sigma}\varepsilon)\times (-k, k)^{N-1}} \text{dist}^{p}((\nabla v_\varepsilon, \lambda\, SO(N)H) \, \text{d}x \leq C \, . \end{split} (3.10)

    From (3.10), Theorem 3.1 and the Poincaré inequality, we deduce that there exists a unit interval contained in (-\frac{2\sigma}\varepsilon, -\frac{\sigma}\varepsilon) such that in the Cartesian product of such interval with the cross-section (-k, k)^{N-1}, the W^{1, p}-norm of the difference between v_\varepsilon and an affine map of the form \lambda Q H x + a, with Q\in O(N){\setminus}SO(N) and a\in \mathbb{R}^N, is bounded by C \varepsilon/\sigma. By the same argument one can find a unit interval contained in (\frac{\sigma}\varepsilon, \frac{2\sigma}\varepsilon) such that in the Cartesian product of such interval with the cross-section (-k, k)^{N-1}, the W^{1, p}-norm of the difference between v_\varepsilon and an affine map of the form \lambda RH x + b, with R\in SO(N) and b\in \mathbb{R}^N, is bounded by C \varepsilon/\sigma. By gluing the function v_\varepsilon with these maps on such intervals, one can define a function \hat{v}_\varepsilon\in \mathcal{A}_{\infty}(\Omega _{k, \infty}) that is a competitor for \gamma (\lambda J, \lambda I;k) and such that (cf.(2.2))

    \mathcal{I}_\varepsilon ^{1,\lambda }({\tilde u_\varepsilon },k){|_{(\alpha _i^\varepsilon - 2\sigma ,\alpha _i^\varepsilon + 2\sigma ) \times {{( - k,k)}^{N - 1}}}} \ge \mathcal{E} _\infty ^{\lambda ,\lambda }({\hat v_\varepsilon },k) - C\frac{\varepsilon }{\sigma }{\mkern 1mu} ,

    where \mathcal{I}_{\varepsilon }^{1,\lambda } (\tilde u_\varepsilon, k)|_{(\alpha_i^\varepsilon -2\sigma, \alpha_i^\varepsilon + 2\sigma )\times (-k, k)^{N-1}} only takes into account the interactions between atoms lying in the subset (\alpha_i^\varepsilon -2\sigma, \alpha_i^\varepsilon + 2\sigma )\times (-k, k)^{N-1}. Arguing in a similar way for the other intervals in (3.9) yields (3.6).

    We prove that the bound (3.6) is in fact optimal.

    Proposition 3. Let F\in L^{\infty}((-L, L);{{\mathbb{M}}^{N\times N}}) and U\in\mathcal{U} satisfy

    \label{cohull} F\in \begin{cases} {\rm{co}}(SO(N))H\;\;\;\;\; a.e.{\rm{ }}\;in\ (-L, 0)\cap U \, , \\ {\rm{co}}(O(N){\setminus} SO(N))H\;\;\;\;\; a.e.{\rm{ }}\;in\ (-L, 0)\setminus U \, , \\ \lambda\, {\rm{co}}(SO(N))H\;\;\;\;\; a.e.{\rm{ }}\;in\ (0, L)\cap U \, , \\ \lambda\, {\rm{co}}(O(N){\setminus} SO(N))H\;\;\;\;\; a.e.{\rm{ }}\;in\ (0, L)\setminus U \, . \end{cases} (4.1)

    Then there exists a sequence \{{\tilde{u}}_\varepsilon\}\subset {\tilde {\mathcal{A}}}_{\varepsilon}(\Omega _{k}) such that

    \label{1328061} {\tilde{u}}_\varepsilon A_\varepsilon^{-1}\mathop * \limits_ \rightharpoonup F \;\;\;\;\; weakly*{\rm{ }}in\;\;\;\;\;\ L^{\infty}(\Omega _k;{{\mathbb{M}}^{N\times N}})\, , (4.2)

    and

    \limsup\limits_{\varepsilon\to 0^{+}} \mathcal{I}_{\varepsilon }^{1, \lambda }(\tilde u_\varepsilon, k)\leq\ \gamma(I, J;k) \, \mathcal{H}^0(\partial U\cap(-L, 0)) + \gamma(\lambda I, \lambda J;k) \, \mathcal{H}^0(\partial U\cap(0, L)) \\+ \gamma(I, \lambda I;k) \left[1-\chi_{\partial U}(0) \right] + \gamma(I, \lambda J;k)\, \chi_{\partial U}(0) (4.3)

    Proof. Using a standard approximation argument we may assume that x_1\mapsto F(x_1) is piecewise constant, with values in O(N)H for a.e. x_1\in(-L, 0) and values in \lambda\, O(N)H for a.e. x_1\in(0, L). We may also assume that this approximation process does not modify the set U of (4.1). More precisely, there exist m, n\in\mathbb{Z}, m < 0, n\ge0, -L = a_m < a_{m+1} < \dots < a_{-1} < a_0 = 0 < a_1 < \dots < a_n < a_{n+1} = L, and R_{i}\in O(N) for i = m, \dots, -1, 0, \dots, n such that

    F = \sum\limits_{i = m}^{-1}\chi_{(a_i, a_{i+1})}R_{i}H+ \sum\limits_{i = 0}^{n}\chi_{(a_i, a_{i+1})}\lambda R_{i}H

    and

    U = {\mathop{\rm int}} \bigcup \big\{ [a_i, a_{i+1}] \colon R_i\in SO(N) \, , \ m\le i\le n-1 \big\} \, .

    The following construction is similar to that in [14,Proposition 3.2], so we will show the details only for what concerns the changes of orientation. We introduce a mesoscale \{\sigma_\varepsilon\} such that \varepsilon \ll\sigma_\varepsilon\ll 1 as \varepsilon\to0^+. Next we define \tilde u_\varepsilon in the sets of the type (a_i{+}\sigma_\varepsilon, a_{i+1}{-}\sigma_\varepsilon)\times(-k, k)^{N-1} in such a way that its gradient equals R_i H A_\varepsilon if a_{i+1}\le0 and equals \lambda R_i H A_\varepsilon if a_i\ge 0. This determines \tilde u_\varepsilon in those regions, up to some additive constants that will have to be fixed at the end of the construction in order to make \tilde u_\varepsilon continuous.

    We now complete the definition of \tilde u_\varepsilon in the sets of the type (a_i{-}\sigma_\varepsilon, a_{i}{+}\sigma_\varepsilon)\times(-k, k)^{N-1}. Let us first assume i < 0, i.e., a_i < 0. Since R_{i-1} and R_{i} may be in SO(N) or in O(N){\setminus} SO(N), one can have four cases. If both R_{i-1} and R_i are in SO(N), it is possible to define \tilde u_\varepsilon by interpolating R_{i-1} and R_i so that the cost of the transition has order O(\varepsilon/\sigma_\varepsilon), so it gives no contribution to (4.3); we refer to [14] for details. The case R_{i-1}, R_i\in O(N){\setminus} SO(N) is completely analogous.

    If R_{i-1}\in SO(N) and R_i\in O(N){\setminus} SO(N) or viceversa, we define \tilde u_\varepsilon in the set (a_i{-}\sigma_\varepsilon, a_{i}{+}\sigma_\varepsilon)\times(-k, k)^{N-1} as a rescaling of a quasiminimiser of (2.1b). More precisely, we fix \eta>0 and apply the definition of \gamma (R_{i-1}, R_i;k), thus finding M>0 and v\in \mathcal{A}_{\infty}(\Omega _{k, \infty}) such that

    \nabla v = R_{i-1}H \ \text{for} \ x_1\in(-\infty, -M)\, , \,\;\;\;\;\; \nabla v = R_{i}H \ \text{for} \ x_1\in(M, +\infty)

    and

    \mathcal{E} _{\infty }^{1, 1}(v, k) \leq \gamma (I, J;k) + \eta \, ,

    where we used also Proposition 1. With this at hand, we define \tilde u_\varepsilon in the set (a_i{-}\sigma_\varepsilon, a_{i}{+}\sigma_\varepsilon)\times(-k, k)^{N-1} as

    \tilde u_\varepsilon(x): = \varepsilon v(\tfrac{1}{\varepsilon}A_\varepsilon x) + b \, .

    The constant vector b in the last equation is chosen in such a way that \tilde u_\varepsilon is continuous. Since each point of \partial U gives the same contribution \gamma (I, J;k) to the upper bound, we obtain the first term of (4.3).

    The case i>0, i.e., a_i>0, is treated similarly to i < 0 and gives rise to the second term of (4.3). Finally, for i = 0, i.e., a_i = a_0 = 0, we argue as above and define \tilde u_\varepsilon by using a rescaling of a quasiminimiser of (2.1a) and applying the definition of \gamma (R_{-1}, \lambda R_0;k). We then get an interfacial contribution in (4.3) that differs in the two cases 0\in \partial U and 0\notin\partial U.

    In the next theorem we characterise the \Gamma-limit of the sequence \{ \mathcal{I}_{\varepsilon }^{1,\lambda }(\cdot, k) \} with respect to the weak* convergence in W^{1, \infty}(\Omega _k;\mathbb{R}^N); see [4,8] for an introduction to \Gamma-convergence. As it can be inferred from the compactness result in Proposition 2, the domain of the \Gamma-limit turns out to be

    \label{gammadomain} \begin{split} \mathcal{A}^{1,\lambda }(k): = \big\{ u\in W^{1, \infty}(\Omega _k;\mathbb{R}^N)\colon & \partial_{2} u = \dots = \partial_{N} u = 0 \ \text{a.e. in}\ \Omega _k\, , \\ & |\partial_1 u |\leq 1 \ \text{a.e. in}\ \Omega _k^- \, , \ |\partial_1 u |\leq \lambda \ \text{a.e. in}\ \Omega _k^+ \big\} \, . \end{split} (5.1)

    We show that on such domain the \Gamma-limit is constant. Hence, the macroscopic description of the model is similar to that of [14,15]; in particular, it does not have memory of the changes of orientation in minimising sequences. In order to keep track of the orientation changes, we need to introduce a stronger topology for the \Gamma-convergence, as we see in Theorem 5.4.

    Theorem 5.1. The sequence of functionals \{ \mathcal{I}_{\varepsilon }^{1, \lambda }(\cdot, k)\} \Gamma-converges, as \varepsilon\to 0^{+}, to the functional

    \mathcal{I}^{1,\lambda }(u, k) = \begin{cases} \gamma (k) & \;\;\;\;\;if\ u\in\mathcal{A}^{1,\lambda }(k)\, , \\ +\infty &\;\;\;\;\; otherwise, \end{cases} (5.2)

    with respect to the weak* convergence in W^{1, \infty}(\Omega _k;\mathbb{R}^N), where

    \label{gammamin} \gamma(k): = \min\big\{ \gamma(I, \lambda I;k), \gamma(I, \lambda J;k) \big\} \, . (5.3)

    Proof. (Liminf inequality) Let {\tilde{u}}_\varepsilon \in {\tilde {\mathcal{A}}}_{\varepsilon}(\Omega _{k}) be a sequence of functions converging to a function u weakly* in W^{1, \infty}(\Omega _k;\mathbb{R}^N). We have to show that

    \mathcal{I}^{1,\lambda }(u, k)\leq \liminf\limits_{\varepsilon \to 0^+} \mathcal{I}_{\varepsilon }^{1,\lambda }({\tilde{u}}_\varepsilon, k) \, .

    We assume that \liminf_{\varepsilon \to 0^+} \mathcal{I}_{\varepsilon }^{1,\lambda }({\tilde{u}}_\varepsilon, k) \le C, the other case being trivial. By applying Proposition 2 we find a set U\in\mathcal{U} and functions {\tilde{u}} \in W^{1, \infty}(\Omega _k;\mathbb{R}^N), d_2, \dots, d_N\in L^{\infty}(\Omega _k;\mathbb{R}^N) independent of x_2, \dots, x_N, such that (3.4), (3.5), and (3.6) hold. This implies that \partial_1u = \partial_1{\tilde{u}} a.e., the function u is independent of x_2, \dots, x_N, and u\in \mathcal{A}^{1,\lambda }(k). Notice that the right-hand side of (3.6) is greater than or equal to \gamma(k), since \gamma(\cdot, \cdot;k) is positive.

    (Limsup inequality) Given a function u\in W^{1, \infty}(\Omega _k;\mathbb{R}^N) we have to find a sequence \{{\tilde{u}}_\varepsilon\}\subset {\tilde {\mathcal{A}}}_{\varepsilon}(\Omega _{k}) such that {\tilde{u}}_\varepsilon \mathop * \limits_ \rightharpoonup u weakly* in W^{1, \infty}(\Omega _k;\mathbb{R}^N) and

    \limsup\limits_{\varepsilon\to 0^+} \mathcal{I}_{\varepsilon }^{1,\lambda } ({\tilde{u}}_\varepsilon, k)\leq \mathcal{I}^{1,\lambda }(u, k) \, . (5.4)

    We assume that u\in\mathcal{A}^{1,\lambda }(k), the other case being trivial.

    The construction of the recovery sequence depends on the precise value of the minimum in (5.3). Since we do not know such value, we explain how to proceed in the case when \gamma(k) is any of the two quantities therein.

    \bullet If \gamma(k) = \gamma(I, \lambda I;k), we set U: = (-L, L) and, following e.g. [16,Theorem 4.1], we construct measurable functions d_{2}, \dots, d_N\in L^{\infty}(\Omega _k;\mathbb{R}^{N}), independent of x_2, \dots, x_N, such that

    (\partial_1 u \, | \, d_2 \, | \, \cdots \, | \, d_N)\in \begin{cases} {\rm{co}}(SO(N))H & \text{a.e. in}\ \Omega _k^-\, , \\ \lambda\, {\rm{co}}(SO(N))H & \text{a.e. in}\ \Omega _k^+\, . \end{cases}

    \bullet If \gamma(k) = \gamma(I, \lambda J;k) we set U: = (-L, 0) and construct d_{2}, \dots, d_N in such a way that

    (\partial_1 u \, | \, d_2 \, | \, \cdots \, | \, d_N)\in \begin{cases} {\rm{co}}(SO(N))H & \text{a.e. in}\ \Omega _k^-\, , \\ \lambda\, {\rm{co}}(O(N){\setminus}SO(N))H & \text{a.e. in}\ \Omega _k^+\, . \end{cases}

    Proposition 3 can be now applied to F: = (\partial_1 u \, | \, d_2 \, | \, \cdots \, | \, d_N), hence providing us with a sequence \{{\tilde{u}}_\varepsilon\}\subset {\tilde {\mathcal{A}}}_{\varepsilon}(\Omega _{k}) satisfying (4.2)-(4.3). In particular we have \nabla{\tilde{u}}_\varepsilon \mathop * \limits_ \rightharpoonup \nabla u weakly* in L^{\infty}(\Omega _k;{{\mathbb{M}}^{N\times N}}) and (5.4) holds because of the choice of U and the definition of \gamma(k).

    Remark 4. As long as the \Gamma-convergence is taken with respect to the weak* topology of W^{1, \infty}(\Omega _k;\mathbb{R}^N), (5.2) only accounts for the cost of transitions at the interface between the two species of atoms. Indeed, away from the interface it is always possible to construct recovery sequences without mixing rotations and rotoreflections, as done in the proof of the limsup inequality; such transitions have low interaction energy, since \gamma(I, I) = \gamma(J, J) = 0, see also Proposition 1. In particular, for \lambda = 1 the limit functional is trivial, since \mathcal{I}^{1,1}(u, k) = 0 if u\in\mathcal{A}^{1,1}(k).

    Below we show that, if a stronger topology is chosen, the value of the \Gamma-limit changes. The resulting limit functional depends on an internal variable, D in (5.7), that keeps track of the changes of orientation throughout the thin wire. In fact, different transitions between the energy wells must now be employed according to the value of D; two examples are provided in Figure 2.

    Figure 2. 

    Two possible recovery sequences for the profile at the centre of the figure. Here we picture only a part of the wire containing just one species of atoms, therefore the transition at the interface is not represented. A kink in the profile may be reconstructed by folding the strip, i.e., mixing rotations and rotoreflections (left); or by a gradual transition involving only rotations or only rotoreflections (right). In the limit, the former recovery sequence gives a positive cost, while the latter gives no contribution. If the stronger topology is chosen, the appropriate recovery sequence will depend on the value of the internal variable, which defines the orientation of the wire

    .

    We introduce the sequence of functionals defined for u\in W^{1, \infty}(\Omega _k;\mathbb{R}^N) and D\in L^\infty(\Omega _k;{{\mathbb{M}}^{N\times N}}) by

    \widehat {\cal I}_\varepsilon ^{1,\lambda }(\tilde u,D,k): = \left\{ {\begin{array}{*{20}{l}} {{\cal I}_\varepsilon ^{1,\lambda }(\tilde u,k)\;\;\;\;\;{\rm{if}}\;\tilde u \in {{\widetilde {\cal A}}_\varepsilon }({\Omega _k})\;{\rm{and}}\;D = ({\partial _1}\tilde u{\mkern 1mu} |{\mkern 1mu} {\varepsilon ^{ - 1}}{\partial _2}\tilde u{\mkern 1mu} |{\mkern 1mu} \cdots {\mkern 1mu} |{\mkern 1mu} {\varepsilon ^{ - 1}}{\partial _N}\tilde u){\mkern 1mu} ,}\\ { + \infty \;\;\;\;\;{\rm{otherwise}}{\rm{.}}} \end{array}} \right.

    In the next theorem we study the \Gamma-limit of the sequence \left\{ {\widehat {\cal I}_\varepsilon ^{1,\lambda }( \cdot , \cdot ,k)} \right\} as \varepsilon\to0^+ with respect to the weak* convergence in W^{1, \infty}(\Omega _k;\mathbb{R}^N)\times L^\infty(\Omega _k;{{\mathbb{M}}^{N\times N}}). As a consequence of Proposition 2, the domain of the \Gamma-limit turns out to be

    \begin{align} & {{\widehat{\mathcal{A}}}^{1,\lambda }}(k):=\{(u,D):u\in {{\mathcal{A}}^{1,\lambda }}(k),\ D\in {{L}^{\infty }}({{\Omega }_{k}};{{\mathbb{M}}^{N\times N}}),\ \ \\ & \ \ \ D{{e}_{1}}={{\partial }_{1}}u,\ {{\partial }_{2}}D=\ldots ={{\partial }_{N}}D=0\ \text{a}\text{.e}\text{. in}\ {{\Omega }_{k}},\ \ \ \ \\ & \ D\in \text{co}(SO(N))H\cup \text{co}(O(N)\setminus SO(N))H\ \text{a}\text{.e}\text{. in}\ \Omega _{k}^{-}, \\ & D\in \lambda \text{co}(SO(N))H\cup \lambda \text{co}(O(N)\setminus SO(N))H\ \text{a}\text{.e}\text{. in}\ \Omega _{k}^{+}\}, \\ \end{align}

    where \mathcal{A}^{1,\lambda }(k) is defined by (5.1). It is convenient to introduce the following definition, where the functional \mathcal{J} coincides with the right-hand sides of (3.6) and (4.3).

    Definition 5.2. Given (u,D) \in {\widehat {\cal A}^{1,\lambda }}(k), let \mathcal{U}(u, D) be the collection of all subsets U\in\mathcal{U} such that

    \label{coso} D \in \begin{cases} {\rm{co}}(SO(N))H& \text{for a.e.}\ x_1\in(-L, 0)\cap U \, , \\ {\rm{co}}(O(N){\setminus} SO(N))H& \text{for a.e.}\ x_1\in(-L, 0)\setminus U \, , \\ \lambda\, {\rm{co}}(SO(N))H& \text{for a.e.}\ x_1\in(0, L)\cap U \, , \\ \lambda\, {\rm{co}}(O(N){\setminus} SO(N))H& \text{for a.e.}\ x_1\in(0, L)\setminus U \, . \end{cases} (5.5)

    For U\in\mathcal{U}(u, D) we set

    \begin{split} \mathcal{J}(U) : = &\ \gamma(I, J;k) \, \mathcal{H}^0(\partial U\cap(-L, 0)) + \gamma(\lambda I, \lambda J;k) \, \mathcal{H}^0(\partial U\cap(0, L)) \\ & + \gamma(I, \lambda I;k) \left[1-\chi_{\partial U}(0) \right] + \gamma(I, \lambda J;k)\, \chi_{\partial U}(0) \end{split}

    and

    {\mathcal{J}_{\min }}(u, D) : = \min\limits_{U\in \mathcal{U}(u, D)} \mathcal{J}(U) \, . (5.6)

    The last definition will be used to apply Propositions 2 and 3 towards the characterisation of the \Gamma-limit with respect to the stronger topology. To this end, each pair (u,D)\in {{\widehat{\mathcal{A}}}^{1,\lambda }}(k) is associated with a set U realising (5.5). Such U is in general not unique, since {\rm{co}}(SO(N)) \cap {\rm{co}}(O(N){\setminus} SO(N)) \neq \emptyset. Therefore, we choose it to be "optimal", i.e., minimising (5.6). Notice that the minimum in (5.6) is attained since

    \begin{split} &\{ \mathcal{J}(U) \colon U \in \mathcal{U}(u, D) \} \\ \subset& \{ m_1 \gamma(I, J;k) + m_2 \gamma(I, \lambda I;k)+ m_3 \gamma(I, \lambda J;k)+ m_4 \gamma(\lambda I, \lambda J;k) \colon m_i\in \mathbb{N} \} \, . \end{split}

    A minimiser needs not be unique as shown in the following example.

    Example 5.3. Fix a_1 < a_2 < 0 and assume that D(x_1)\in (O(N){\setminus}SO(N))H for x_1 < a_1, D(x_1) = 0 for a_1 < x_1 < a_2, D(x_1)\in SO(N)H for a_2 < x_1 < 0, and D(x_1)\in \lambda\, SO(N)H for x_1>0. Then any interval of the type U = (a, +\infty), with a_1\le a\le a_2, is a minimiser of (5.6).

    Theorem 5.4. The sequence of functionals \{\widehat{\mathcal{I}}_{\varepsilon }^{1,\lambda }(\cdot ,\cdot ,k)\} \Gamma-converges, as \varepsilon\to 0^{+}, to the functional

    {{\widehat{\mathcal{I}}}^{1,\lambda }}(u,D,k):=\left\{ \begin{array}{*{35}{l}} {{\mathcal{J}}_{\min }}(u,D) & if\ (u,D)\in {{\widehat{\mathcal{A}}}^{1,\lambda }}(k), \\ +\infty & otherwise\text{, } \\ \end{array} \right. (5.7)

    with respect to the weak* convergence in W^{1, \infty}(\Omega _k;\mathbb{R}^N)\times L^\infty(\Omega _k;{{\mathbb{M}}^{N\times N}}), where {\mathcal{J}_{\min }}(u, D) is defined by (5.6).

    Proof. The liminf inequality is obtained by applying Proposition 2 and arguing as in Theorem 5.1. Also the derivation of the limsup inequality is similar to the one performed in Theorem 5.1; let us simply point out that, while in the proof of Theorem 5.1 the matrix field F needed to be reconstructed, here we set F: = D and choose U as a minimiser of (5.6). The conclusion follows by applying Proposition 3.

    Remark 5. We underline that Theorem 5.4 provides a nontrivial \Gamma-limit also in the case when \lambda = 1. Indeed, one has {{\widehat{\mathcal{I}}}^{1,1}}(u,D,k)=\gamma (I,J;k){{\mathcal{H}}^{0}}(\partial U\cap (-L,L)) if (u,D)\in {{\widehat{\mathcal{A}}}^{1,1}}(k) and U miminises (5.6), where \gamma (I, J;k)>0.

    In the present section we discuss how the previous results extend to the case when the functional (1.4) is complemented by boundary conditions or external forces. Although our considerations apply to the case of general H\in GL^+(N) and \lambda \in (0, 1], for simplicity we will focus on the case H = I and \lambda = 1. We will also test the consistency of the present model with the non-interpenetration condition by looking at minimisers of the \Gamma-limit when boundary conditions or forces are prescribed. We will see that the continuum limit that keeps track of such constraints is the one provided by the stronger topology (5.7).

    Boundary conditions. Let B^-, B^+\in GL^+(N) and suppose that the functional (1.4) is now defined on deformations u\in\mathcal{A}_{\varepsilon}(\Omega _k\varepsilon ) that satisfy

    \begin{cases} \nabla u(x) = B^-x&\text{ if }-L < x_1 < -L +\varepsilon\, , \\ \nabla u(x) = B^+x&\text{ if }L -\varepsilon < x_1 < L \, . \end{cases} (6.1)

    It is easy to see that while the compactness result of Proposition 2 remains valid, the \Gamma-limit (5.2) will now contain additional terms corresponding to the minimal energy spent to fix the atoms in the vicinity of the lateral boundaries. However, such extra terms do not depend on the limiting deformations, therefore they do not encode any information about the behaviour of minimising sequences. As far as the stronger topology is concerned, one can see that the limit functional (5.7) will contain the additional quantities \gamma (B^-, P;k) and \gamma (P, B^+;k) defined, for P\in\{I, J\}, by

    \begin{split} \gamma(B^-, P;k): = \inf\big\{ \mathcal{E} _M^{1,1}(v, k) \colon&M > 0\, , \ v\in \mathcal{A}_{\infty}(\Omega _{k, \infty})\, , \\ & \nabla v = B^- \ \text{for} \ x_1\in(-\infty, -M)\, , \, \\& \nabla v = P \ \text{for} \ x_1\in(M, +\infty) \big\}\, , \end{split} (6.2)
    \begin{split} \gamma(P, B^+;k): = \inf\big\{ \mathcal{E} _M^{1,1}(v, k) \colon&M > 0\, , \ v\in \mathcal{A}_{\infty}(\Omega _{k, \infty})\, , \\ & \nabla v = P \ \text{for} \ x_1\in(-\infty, -M)\, , \, \\& \nabla v = B^+ \ \text{for} \ x_1\in(M, +\infty) \big\}\, , \end{split} (6.3)

    where \mathcal{E} _M^{1,1} is as in (2.1b), except that the sum is taken over all atoms contained in the bounded strip (-M, M)\times (-k, k)^{N-1}. The choice of P = I or P = J depends on whether or not \pm L \in \partial \bar U, where \bar U is a minimiser of (5.6). Precisely, if -L \in \partial \bar U (resp. L \in \partial \bar U), then in (6.2) (resp. (6.3)) we take P = I, otherwise we take P = J.

    Remark 6. By Proposition 2 and the properties of \Gamma-convergence, minimisers of (1.4) subjected to (6.1) converge, up to subsequences, to minimisers of (5.7) complemented with the above extra terms. Moreover, if \text{dist}(B^{\pm};SO(N)) is sufficiently small, then such minimisers will not have transitions between \text{co}\big(SO(N)\big) and \text{co}\big(O(N){\setminus} SO(N)\big). This follows from the fact that

    \gamma(I, B^{\pm};k)\to 0 \;\;\;\;\; \text{as} \;\;\;\;\; \text{dist}\;\;\;\;\;(B^{\pm};SO(N))\to 0

    and therefore, as long as \gamma(I, B^{+};k)+\gamma(B^{-}, I;k) < \gamma(I, J;k), the optimal transitions will fulfil the non-interpenetration condition. In this respect the quantity \gamma(I, J;k) can be regarded as an energetic barrier that must be overcome in order to have folding effects.

    External forces. We study a class of tangential/radial forces acting along the rod. Let F_1\in \mathbb{R}^N, F_2, \dots, F_{N}\in C^0(\mathbb{R}^N;\mathbb{R}^N) be a collection of vector fields such that F_i = F_i(x_1) for every 2 = 1, \dots, N. We denote by F the matrix field whose columns are F_1, \dots, F_N. For each u \colon\mathcal{L}_\varepsilon(k)\to\mathbb{R}^N, consider the functional

    \begin{align} & {{\mathcal{F}}_{\varepsilon }}(u,k):= \\ & \sum\limits_{(\pm {{L}_{\varepsilon }},{{x}_{2}},\ldots ,{{x}_{N}})\in {{\mathcal{L}}_{\varepsilon }}(k)}{}{{F}_{1}}\cdot (u({{L}_{\varepsilon }},{{x}_{2}},\ldots ,{{x}_{N}})-u(-{{L}_{\varepsilon }},{{x}_{2}},\ldots ,{{x}_{N}})) \\ & +\sum\limits_{({{x}_{1}},\pm \varepsilon k,\ldots ,{{x}_{N}})\in {{\mathcal{L}}_{\varepsilon }}(k)}{}{{F}_{2}}({{x}_{1}})\cdot (u({{x}_{1}},\varepsilon k,{{x}_{3}},\ldots ,{{x}_{N}})-u({{x}_{1}},-\varepsilon k,{{x}_{3}},\ldots ,{{x}_{N}}))+\cdots \\ & \ldots +\sum\limits_{({{x}_{1}},\ldots ,{{x}_{N-1}},\pm \varepsilon k)\in {{\mathcal{L}}_{\varepsilon }}(k)}{}{{F}_{N}}({{x}_{1}})\cdot (u({{x}_{1}},\ldots ,{{x}_{N-1}},\varepsilon k)-u({{x}_{1}},\ldots ,{{x}_{N-1}},-\varepsilon k)), \\ \end{align} (6.4)

    where L_\varepsilon: = L if L is an integer multiple of \varepsilon , and L_\varepsilon: = ([L/\varepsilon] +1)\varepsilon otherwise. The functional \mathcal{F}_\varepsilon consists of several terms: the first sum represents a tangential force, while the other terms define a radial force acting on the external atoms of the lattice and enforcing the average displacements along the coordinate directions e_2, \dots, e_N to be aligned with the given vector fields F_2, \dots, F_{N}. Note that \mathcal{F}_\varepsilon(u, k) can be written as

    \begin{split} \mathcal{F}_\varepsilon(u, k) = \\& \sum\limits_{x_1 = -L_\varepsilon}^{L_\varepsilon-\varepsilon} \ \sum\limits_{(x_1, \dots, x_N)\in \mathcal{L}_\varepsilon(k)}\!\!\!\! \!\!\!\!\!\!\! F_1 {\cdot} \big( u(x_1+\varepsilon, \dots, x_N) - u(x_1, \dots, x_N) \big) \\ +& \sum\limits_{x_2 = -\varepsilon k}^{\varepsilon(k-1)} \ \sum\limits_{(x_1, \dots, x_N)\in \mathcal{L}_\varepsilon(k)}\!\!\!\!\!\!\!\!\!\!\! F_2(x_1) \cdot \big( u(x_1, x_2+\varepsilon, \dots, x_N) - u(x_1, x_2, \dots, x_N) \big)+ \cdots \\ \dots+& \sum\limits_{x_N = -\varepsilon k}^{\varepsilon(k-1)} \ \sum\limits_{(x_1, \dots, x_N)\in \mathcal{L}_\varepsilon(k)}\!\!\!\!\!\!\!\!\!\!\! F_N(x_1) \cdot \big( u(x_1, \dots, x_N+\varepsilon) - u(x_1, \dots, x_N) \big) \, , \end{split}

    hence we have that \mathcal{F}_\varepsilon(u, k)\simeq\frac{1}{\varepsilon^{N-1}}\int {\Omega _{k\varepsilon}}{F:\nabla u}{\rm{d}}x.

    Introducing the new variables z(x): = A_\varepsilon x defined by (1.6), and adopting the notation used in Section 1, (6.4) can be equivalently expressed in terms of {\tilde{u}}(x) : = u(z(x)), namely

    \begin{split} &\ {\tilde{\mathcal{F}}}_\varepsilon({\tilde{u}}, k): = \\ &\!\!\!\!\!\!\! \sum\limits_{(\pm L_\varepsilon, x_2, \dots, x_N)\in A_\varepsilon^{-1}\mathcal{L}_\varepsilon(k)} \!\!\!\!\!\!\!\!\!\!\! F_1 {\cdot} \big( {\tilde{u}}(L_\varepsilon, x_2, \dots, x_N) - {\tilde{u}}(- L_\varepsilon, x_2, \dots, x_N) \big) \\ +& \!\!\!\!\!\!\! \sum\limits_{(x_1, \pm k, \dots, x_N)\in A_\varepsilon^{-1}\mathcal{L}_\varepsilon(k)} \!\!\!\!\!\!\!\!\!\!\! F_2(x_1) {\cdot} \big( {\tilde{u}}(x_1, k, x_3, \dots, x_N) - {\tilde{u}}(x_1, - k, x_3, \dots, x_N) \big)+ \dots \\ \dots+& \!\!\!\!\!\!\! \sum\limits_{(x_1, \dots, x_{N-1}, \pm k)\in A_\varepsilon^{-1}\mathcal{L}_\varepsilon(k)}\!\!\!\!\! \!\!\!\!\!\! F_N(x_1) {\cdot} \big( {\tilde{u}}(x_1, \dots, x_{N{-}1}, k) - {\tilde{u}}(x_1, \dots, x_{N-1}, - k) \big) \\ = &\ \mathcal{F}_\varepsilon(u, k)\, , \end{split}

    so that {\tilde{\mathcal{F}}}_\varepsilon({\tilde{u}}, k)\simeq \int {\Omega _{k}}{F:(\nabla {\tilde{u}}\, A_\varepsilon^{-1})}{\rm{d}}{x}. We can then address the study of the asymptotic behaviour of the sequence

    {{\mathcal{G}}_{\varepsilon }}(\tilde{u},D,k):=\widehat{\mathcal{I}}_{\varepsilon }^{1,1}(\tilde{u},D,k)-{{\widetilde{\mathcal{F}}}_{\varepsilon }}(\tilde{u},k),\ \ \ \ \tilde{u}\in {{\widetilde{\mathcal{A}}}_{\varepsilon }}({{\Omega }_{k}}),D\in {{L}^{\infty }}({{\Omega }_{k}};{{\mathbb{M}}^{N\times N}}). (6.5)

    Note that in this context we cannot use the weak* convergence in W^{1, \infty}(\Omega _k;\mathbb{R}^N), since this does not control {\tilde{\mathcal{F}}}_\varepsilon({\tilde{u}}), which is in fact a term depending on D = \nabla{\tilde{u}}\, A_\varepsilon^{-1}. This justifies the choice of \widehat{\mathcal{I}}_{\varepsilon }^{1,1} rather than \mathcal{I}^{1,1}_\varepsilon in the definition of \mathcal{G}_\varepsilon: in order to control both terms in the right hand side of (6.5), we use the stronger topology provided by the weak* convergence in W^{1, \infty}(\Omega _k;\mathbb{R}^N)\times L^\infty(\Omega _k;{{\mathbb{M}}^{N\times N}}). The force term is indeed a continuous perturbation of \widehat{\mathcal{I}}_{\varepsilon }^{1,1} with respect to such topology. We observe that

    \frac{C}\varepsilon \int_{\Omega _k} ( |\nabla {\tilde{u}}\, A_\varepsilon^{-1}|^p -1 ) \, \text{d}x \leq \frac{C}\varepsilon\int_{{\Omega }_{k}}\text{dist}^{p}(\nabla {\tilde{u}}\, A_\varepsilon^{-1}, O(N)) \, \text{d}x \leq \widehat{\mathcal{I}}_{\varepsilon }^{1,1}({\tilde{u}}, k)

    and

    {{\tilde{\mathcal{F}}}_{\varepsilon }}(\tilde{u},k)\le C(\int_{{{\Omega }_{k}}}{|}F{{|}^{{{p}'}}}\text{d}x+\int_{{{\Omega }_{k}}}{|}\nabla \tilde{u}A_{\varepsilon }^{-1}{{|}^{p}}\text{d}x).

    Let now \{({\tilde{u}}_\varepsilon, D_\varepsilon)\} \in {\tilde {\mathcal{A}}}_\varepsilon(\Omega _{k}) \times L^\infty(\Omega _k;{{\mathbb{M}}^{N\times N}}) be a sequence such that

    \limsup\limits_{\varepsilon\to 0^+} \mathcal{G}_\varepsilon(\tilde u_\varepsilon, D_\varepsilon, k) \leq C\, .

    The previous inequalities imply that || \nabla {\tilde{u}}\, A_\varepsilon^{-1} ||_{L^p(\Omega _k;{{\mathbb{M}}^{N\times N}})} is equibounded, which in turn implies that \limsup_{\varepsilon\to 0^{+}} \widehat{\mathcal{I}}_{\varepsilon }^{1,1}({\tilde{u}}_\varepsilon, k) \leq C and thus the conclusions of Proposition 2 are still valid. (See also [16,Remark 4.2] for similar results.) Taking also into account Theorem 5.4, we derive the following result.

    Theorem 6.1. The following results hold:

    (Compactness) Let \{({\tilde{u}}_\varepsilon, D_\varepsilon)\} \in {\tilde {\mathcal{A}}}_\varepsilon(\Omega _{k}) \times L^\infty(\Omega _k;{{\mathbb{M}}^{N\times N}}) be a sequence such that

    \limsup\limits_{\varepsilon\to 0^+} \mathcal{G}_\varepsilon(\tilde u_\varepsilon, D_\varepsilon, k) \leq C\, .

    Then there exists (\tilde{u},D)\in {{\widehat{\mathcal{A}}}^{1,1}}(k) and a subsequence (not relabelled) such that

    (\Gamma-limit) The sequence of functionals \{ \mathcal{G}_\varepsilon \} \Gamma-converges, as \varepsilon\to 0^{+}, to the functional

    \mathcal{G}(u,D,k):={{\widehat{\mathcal{I}}}^{1,1}}(u,D,k)-\tilde{\mathcal{F}}(D,k), (6.6)

    with respect to the weak* convergence in W^{1, \infty}(\Omega _k;\mathbb{R}^N)\times L^\infty(\Omega _k;{{\mathbb{M}}^{N\times N}}), where

    \tilde{\mathcal{F}}(D,k):={{(2k)}^{N-1}}\int_{-L}^{L}{(}{{F}_{1}}\cdot {{d}_{1}}+\cdots +{{F}_{N}}\cdot {{d}_{N}})\text{d}{{x}_{1}}

    for D{ = }(d_1| \cdots | d_N).

    As a consequence of the previous theorem and the standard properties of \Gamma-convergence we infer the following result about convergence of minima and minimisers.

    Corollary 1. We have that

    \begin{align} & \underset{\varepsilon \to 0}{\mathop{\lim }}\,\min \{{{\mathcal{G}}_{\varepsilon }}(u,D):(u,D)\in {{\widetilde{\mathcal{A}}}_{\varepsilon }}({{\Omega }_{k}})\times {{L}^{\infty }}({{\Omega }_{k}};{{\mathbb{M}}^{N\times N}})\} \\ & =\min \{\mathcal{G}(u,D,k):(u,D)\in {{\widehat{\mathcal{A}}}^{1,1}}(k)\}. \\ \end{align}

    Moreover if (u_\varepsilon, D_\varepsilon)\in {\tilde {\mathcal{A}}}_\varepsilon(\Omega _{k}) \times L^\infty(\Omega _k;{{\mathbb{M}}^{N\times N}}) is such that

    \lim\limits_{\varepsilon\to 0} \mathcal{G}_\varepsilon (u_\varepsilon, D_\varepsilon) = \lim\limits_{\varepsilon\to 0} \min\{\mathcal{G}_\varepsilon (u, D) \colon (u, D)\in {\tilde {\mathcal{A}}}_\varepsilon(\Omega _{k}) \times L^\infty(\Omega _k;{{\mathbb{M}}^{N\times N}}) \} \, ,

    then any cluster point (\overline u, \overline D) of (u_\varepsilon, D_\varepsilon) with respect to the weak* convergence in W^{1, \infty}(\Omega _k;\mathbb{R}^N)\times L^\infty(\Omega _k;{{\mathbb{M}}^{N\times N}}) is a minimiser for \min \{\mathcal{G}(u,D,k):(u,D)\in {{\widehat{\mathcal{A}}}^{1,1}}(k)\}.

    We now come back to the question of the consistency of the model with the non-interpenetration condition. In this context we cannot expect that minimisers of (6.5) preserve orientation for the whole class of loads defined above. This is clarified in the following remark.

    Remark 7. Minimisers of the functional defined by (6.6) may have transition points between the two wells SO(N) and O(N){\setminus}SO(N). Suppose for instance that F_1, \dots, F_N satisfy the following properties: there exist n_1, \dots, n_N \in S^{N-1}, a\in (-L, L), such that (n_1| \cdots | n_N)\in SO(N), F_i(x_1) = f_{i}(x_1)n_i for each i = 1, \dots, N, f_1\in\mathbb{R}, f_{i} > 0 in (-L, L) for each i = 1, \dots, N-1, f_{N} > 0 in (-L, a), f_{N} < 0 in (a, L).

    Define \overline D: = (n_1| \cdots | n_N) if x_1\in(-L, a), and \overline D: = (n_1| \cdots| n_{N-1} | -n_N) if x_1\in(a, L). Note that ({{x}_{1}}{{n}_{1}},\bar{D})\in {{\widehat{\mathcal{A}}}^{1,1}}(k), and \overline D has a transition point at x_1 = a. Denote by \widehat{\mathcal{A}}_{0}^{1,1}(k) the subset of {{\widehat{\mathcal{A}}}^{1,1}}(k) of deformations with no transitions; i.e., \widehat{\mathcal{A}}_{0}^{1,1}(k):=\{(u,D)\in {{\widehat{\mathcal{A}}}^{1,1}}(k):{{\widehat{\mathcal{I}}}^{1,1}}(u,D,k)=0\}. It is easy to see that

    \begin{align} & C:=\underset{(u,D)\in \widehat{\mathcal{A}}_{0}^{1,1}(k)}{\mathop{\min }}\,-\tilde{F}(D,k) > -\tilde{F}(\bar{D},k) \\ & =-{{(2k)}^{N-1}}(\sum\limits_{i=1}^{N-1}{\int_{-L}^{L}{{{f}_{i}}}}\text{d}{{x}_{1}}+\int_{-L}^{a}{{{f}_{N}}}\text{d}{{x}_{1}}-\int_{a}^{L}{{{f}_{N}}}\text{d}{{x}_{1}}). \\ \end{align}

    Therefore, if f_1, \dots, f_N are such that

    -\widetilde F(\overline D, k) + \gamma(I, J;k) < C,

    then it is energetically preferred to have a transition at a, namely, all minimisers of \mathcal{G} are given by (x_1 n_1 + b, \overline D), with b any vector in \mathbb{R}^N. In contrast, if f_N is always positive, then minimisers will not display any transition.

    The lattice mismatch in heterostructured materials, corresponding to \lambda\neq 1 in the model described in this section, can be relieved by creation of dislocations; i.e., line defects of the crystal structure. We refer to [9,13,19] for an account of the literature on dislocations in nanowires. A model for discrete heterostructured nanowires accounting for dislocations was studied in [14,15] under the assumption that deformations fulfil the non-interpenetration condition. In this paper we have chosen to consider only defect-free configurations in order to both simplify the exposition and to pose emphasis on the difficulties to overcome when the non-interpenetration assumption is removed. In the final part of the paper, we outline the results that can be obtained when dislocations are accounted for.

    Following the ideas of [14], in dimension N = 2 we introduce other possible models where the reference configuration represents a lattice with dislocations. More precisely, we fix \rho\in[\lambda, 1] and set

    \mathcal{L}_\varepsilon(\rho, k): = \mathcal{L}_\varepsilon^-(1, k)\cup\mathcal{L}_\varepsilon^+(\rho, k) \, ,

    where

    \begin{align*} \mathcal{L}_\varepsilon^-(1, k) &: = \phantom{\rho} \varepsilon \mathbb{Z}^2 \cap \overline\Omega _{k\varepsilon} \cap \{x_1 < 0\} \, , \\ \mathcal{L}_\varepsilon^+(\rho, k) &: = \rho \varepsilon \mathbb{Z}^2 \cap \overline\Omega _{k\varepsilon} \cap \{x_1\ge0\} \, , \end{align*}

    and \overline\Omega _\varepsilon is as in (1.1). For \rho\neq1, the number of atomic layers parallel to e_1 is different in the two sublattices (for sufficiently large k); this can be regarded as a system containing dislocations at the interface.

    In presence of dislocations, the choice of the interactions and of the equilibria strongly depends on the lattice that one intends to model. Therefore, in this section we focus on the simplest situation of hexagonal (or equilateral triangular) Bravais lattices in dimension two and we fix

    H: = \begin{pmatrix} 1&-\tfrac12 \\ 0&\tfrac{\sqrt3}2 \end{pmatrix} \, .

    The lattice H \mathcal{L}_\varepsilon(\rho, k) consists of two Bravais hexagonal sublattices with different lattice constants \varepsilon and \rho\varepsilon, respectively; see Figure 3.

    Figure 3. 

    Lattices with dislocations: choice of the interfacial nearest neighbours in \mathcal{L}_\varepsilon(\rho, k) and H \mathcal{L}_\varepsilon(\rho, k) for a Delaunay triangulation

    .

    The bonds between nearest and next-to-nearest neighbours are defined first in the lattice H \mathcal{L}_\varepsilon(\rho, k). To this end, one chooses a Delaunay triangulation of H \mathcal{L}_\varepsilon(\rho, k) as defined in [14,Section 1]. Two points x, y of the lattice are said to be nearest neighbours if there is a lattice point z such that the triangle [x, y, z] is an element of the triangulation. Two points x, y are next-to-nearest neighbours if there are z_1, z_2 such that [x, z_1, z_2] and [y, z_1, z_2] are elements of the triangulation. These definitions coincide with the usual notions of nearest and next-to-nearest neighbours away from the interface. We underline that other choices of interfacial bonds are possible to derive our main results. Indeed, one may start from any triangulation of the lattice satisfying the following properties: the number of nearest neighbours of each point has to be uniformly bounded by a constant independent of \varepsilon , while the length of the bonds in H \mathcal{L}_\varepsilon(\rho, k) has to be uniformly bounded by a constant C_\varepsilon = C\varepsilon.

    Once the bonds in the lattice H \mathcal{L}_\varepsilon(\rho, k) are defined, we define the bonds of a point x\in\mathcal{L}_\varepsilon(\rho, k) as follows:

    \begin{align*} B_1(x) &: = \{ \xi\in\mathbb{R}^N\colon Hx, \ H(x{+}\xi) \in H \mathcal{L}_\varepsilon(\rho, k) \ \text{are nearest neighbours} \} \, , \\ B_2(x) &: = \{ \xi\in\mathbb{R}^N\colon Hx, \ H(x{+}\xi) \in H \mathcal{L}_\varepsilon(\rho, k) \ \text{are next-to-nearest neighbours} \} \, . \end{align*}

    We remark that if x_1\le-2\varepsilon, then B_1(x) = B_1 and B_2(x) = B_2, while if x_1\ge\rho\varepsilon, then B_1(x) = \rho B_1 and B_2(x) = \rho B_2, where B_1, B_2 are as in (1.2)-(1.3). The total interaction energy is

    \begin{split} &\mathcal{E} _{\varepsilon }^{1, \lambda }(u, \rho, k) : = \\ &\!\!\! \sum\limits_{\substack{ x\in \mathcal{L}^-_\varepsilon (\rho, k)\\ \xi\in B_1(x) }} \!\!\!\ c_1 \left|\frac{|u(x+\xi)-u(x)|}e-1\right|^p + \!\!\! \sum\limits_{\substack{ x\in \mathcal{L}^+_\varepsilon(\rho, k)\\ \xi\in B_1(x) }} \!\!\! c_1 \left|\frac{|u(x+ \xi)-u(x)|}e-\lambda \right|^p \\ +&\!\!\! \sum\limits_{\substack{ x\in \mathcal{L}^-_\varepsilon (\rho, k)\\ \xi\in B_2(x) }} \!\!\!\ c_2 \left|\frac{|u(x+\xi)-u(x)|}e-\sqrt3\right|^p + \!\!\! \sum\limits_{\substack{ x\in \mathcal{L}^+_\varepsilon(\rho, k)\\ \xi\in B_2(x) }} \!\!\! c_2 \left|\frac{|u(x+ \xi)-u(x)|}e-\sqrt3\lambda \right|^p \, . \end{split}

    Notice that away from the interface all bonds in the reference configuration are in equilibrium if \rho = \lambda; instead, interfacial bonds are never in equilibrium. The equilibrium distance of two atoms at the interface is in fact an average of the equilibrium distances of the two sublattices. Generalisations of this energy can be considered as described in [14,Section 4].

    The results shown in detail in this paper for the defect-free case (corresponding to \rho = 1) can be extended to models with dislocations (\rho\neq1) without significant changes in the proof. Thus we obtain a \Gamma-convergence result for the rescaled functionals \mathcal{I}_{\varepsilon }^{1, \lambda }(\cdot, \rho, k) defined as in (1.7). (Notice that the definition of the admissible functions is given as in the dislocation-free case, with the only variant that the gradients are constant on the elements of the triangulation introduced to define the interfacial bonds.) Before stating the theorem we introduce the lattices

    \begin{align*} \mathcal{L}_{\infty}(\rho, k)&: = \mathcal{L}_{\infty}^-(1, k)\cup\mathcal{L}_{\infty}^+(\rho, k) \, , \\ \mathcal{L}_{\infty}^-(1, k) &: = \phantom{\rho} \mathbb{Z}^2 \cap \overline\Omega _{k, \infty} \cap \{x_1 < 0\} \, , \\ \mathcal{L}_{\infty}^+(\rho, k) &: = \rho \mathbb{Z}^2 \cap \overline\Omega _{k, \infty} \cap \{x_1\ge0\} \, , \end{align*}

    where the triangulation is chosen in analogy with the one for \mathcal{L}_\varepsilon(\rho, k). We also set

    \begin{split} \gamma(P_1, \lambda P_2;\rho, k): = \inf\big\{ &\mathcal{E} _{\infty }^{1, \lambda }(v, \rho, k) \colon M > 0\, , \ v\in \mathcal{A}_{\infty}(\Omega _{k, \infty})\, , \\ & \nabla v = P_1H \ \text{for} \ x_1\in(-\infty, -M)\, , \, \\&\nabla v = \tfrac\lambda\rho P_2 H \ \text{for} \ x_1\in(M, +\infty) \big\}\, , \end{split}

    with

    \begin{split} &\mathcal{E} _{\infty }^{1, \lambda }(u, \rho, k) : = \\ &\!\!\! \sum\limits_{\substack{ x\in \mathcal{L}^-_{\infty}(\rho, k)\\ \xi\in B_1(x) }} \!\!\!\ c_1 \Big||u(x+\xi)-u(x)|-1\Big|^p + \!\!\! \sum\limits_{\substack{ x\in \mathcal{L}^+_{\infty}(\rho, k)\\ \xi\in B_1(x) }} \!\!\! c_1 \Big||u(x+ \xi)-u(x)|-\lambda \Big|^p \\ +&\!\!\! \sum\limits_{\substack{ x\in \mathcal{L}^-_{\infty}(\rho, k)\\ \xi\in B_2(x) }} \!\!\!\ c_2 \Big||u(x+\xi)-u(x)|-\sqrt3\Big|^p + \!\!\! \sum\limits_{\substack{ x\in \mathcal{L}^+_{\infty}(\rho, k)\\ \xi\in B_2(x) }} \!\!\! c_2 \Big||u(x+ \xi)-u(x)|-\sqrt3\lambda \Big|^p \, . \end{split}

    Theorem 7.1. The sequence of functionals \{ \mathcal{I}_{\varepsilon }^{1, \lambda }(\cdot, \rho, k)\} \Gamma-converges, as \varepsilon\to 0^{+}, to the functional

    {{\mathcal{I}}^{1,\lambda }}(u,\rho ,k)=\left\{ \begin{array}{*{35}{l}} \gamma (\rho ,k) & if\ u\in {{\mathcal{A}}^{1,\lambda }}(\rho ,k), \\ +\infty & otherwise\text{, } \\ \end{array} \right.

    with respect to the weak* convergence in W^{1, \infty}(\Omega _k;\mathbb{R}^N), where

    \begin{split} \mathcal{A}^{1,\lambda }(\rho, k): = \big\{ u\in W^{1, \infty}(\Omega _k;\mathbb{R}^N)\colon & \partial_{2} u = 0 \ a.e.\text{ }in\ \Omega _k\, , \\ & |\partial_1 u |\leq 1 \ a.e.\text{ }in\ \Omega _k^- \, , \ |\partial_1 u |\leq \tfrac\lambda\rho \ a.e.\text{ }in\ \Omega _k^+ \big\} \end{split}

    and

    \gamma(\rho, k): = \min\big\{ \gamma(I, \lambda I;\rho, k), \gamma(I, \lambda J;\rho, k) \big\} \, .

    The stronger topology introduced in Theorem 5.4 allows us to take into account the cost of "folding" the lattice using rotoreflections, giving deeper insight into deformations that bridge different equilibria. Indeed, it is possible to combine Theorems 5.4 and 7.1 giving the \Gamma-convergence in the stronger topology for models with dislocations; we omit the full statement for brevity.

    Remark 8. It is easy to see that for \rho = \lambda

    C_1 k\le \gamma (\lambda, k)\le C_2 k

    for some constants C_1, C_2>0. To obtain the estimate from above it is sufficient to consider the identical deformation and recall that the maximal length of a bond and the maximal number of bonds per atom in the lattice \mathcal{L}_{\infty}(\rho, k) are uniformly bounded. This configuration corresponds to the case when dislocations are uniformly distributed along the interface between the two sublattices. (Recall that here N = 2 and that the length of the interface is 2k.) In contrast, the cost of a defect-free configuration (\rho = 1) is superlinear as already shown in Theorem 2.2. In fact, following the same proof it is possible to conclude that whenever \rho\neq\lambda one has

    \gamma (\rho, k)\le C_\rho \, k^2 \;\;\;\;\;\text{and} \;\;\;\;\;\lim\limits_{k\to\infty}\frac{\gamma (\rho, k)}{k} = +\infty \, .

    This gives a mathematical proof of the experimentally observed fact that dislocations are preferred in order to relieve the lattice mismatch when the thickness of the specimen is sufficiently large. We recall that a similar result was proven in [14,15] (under the non-interpenetration assumption), see also Remark 1.

    The results sketched here for hexagonal lattices can be obtained also for other lattices by adapting the technique to each specific case. In particular, we refer to [15] for details on the rigidity of face-centred and body-centred cubic lattices in dimension three.



    [1] R. P. Agarwal, Difference equations and inequalities: Theory, methods, and applications, Marcel Dekker, New York, 2000.
    [2] R. P. Agarwal, M. Bohner, S. R. Grace, D. O'Regan, Discrete oscillation theory, Hindawi Publishing Corporation, New York, 2005.
    [3] T. Candan, Existence of nonoscillatory solutions to first order neutral differential equations, Electon. J. Differ. Equ., 2016 (2016), 1–11.
    [4] G. E. Chatzarakis, G. N. Miliaras, Asymptotic behaviour in neutral difference equations with variable coefficients and more than one day arguments, J. Math. Comput. Sci., 1 (2012), 32–52.
    [5] L. H. Erbe, Q. Kong, B. G. Zhang, Oscillation theory for functional differential equations, Marcel Dekker Inc., New York, 1995.
    [6] Y. Gao, G. Zhang, Oscillation of nonlinear first order neutral difference equations, Appl. Math. E-Notes, 1 (2001), 5–10.
    [7] D. A. Georgiou, E. A Grove, G. Ladas, Oscillations of neutral difference equations, Appl. Anal., 33 (1989), 243–253. https://doi.org/10.1080/00036818908839876 doi: 10.1080/00036818908839876
    [8] G. Gomathi Jawahar, Oscillation properties of certain types of first order neutral delay difference equations, Int. J. Sci. Eng. Res., 4 (2013), 1197–1201.
    [9] I. Györi, G. Ladas, Oscillation theory of delay differential equations with applications, Clarendon Press, Oxford, 1991.
    [10] J. R. Graef, E. Thandapani, S. Elizabeth, Oscillation of first order nonlinear neutral difference equations, Indian J. Pure Appl. Math., 36 (2005), 503–512.
    [11] B. Karpuz, Sharp conditions for oscillation and nonoscillation of neutral difference equations, In: M. Bohner, S. Siegmund, R. Simon Hilscher, P. Stehlik, Difference equations and discrete dynamical systems with applications, ICDEA 2018, Springer Proceedings in Mathematics & Statistics, Vol. 312, Springer, Cham, 2020. https://doi.org/10.1007/978-3-030-35502-9_11
    [12] V. Kolmanovski, A. Myshkis, Introduction to the theory and applications of functional differential equations, Kluwer, Dordrecht, 1999.
    [13] F. Kong, Existence of nonoscillatory solutions of a kind of first order neutral differential equation, Math. Commun., 22 (2017), 151–164.
    [14] X. Lin, Oscillation of solutions of neutral difference equations with a nonlinear neutral term, Comput. Math. Appl., 52 (2006), 439–448. https://doi.org/10.1016/j.camwa.2006.02.009 doi: 10.1016/j.camwa.2006.02.009
    [15] Q. Li, C. Wang, F. Li, H. Liang, Z. Zhang, Oscillation of sublinear difference equations with positive neutral term, J. Appl. Math. Comput., 20 (2006), 305–314. https://doi.org/10.1007/BF02831940 doi: 10.1007/BF02831940
    [16] A. Murugesan, R. Suganthi, Oscillation behaviour of first order nonlinear functional neutral delay difference equations, Int. J. Differ. Equ., 12 (2017), 1–13.
    [17] A. D. Myschkis, Lineare Differentialgleichungen mit nacheilendem Argument, VEB Deutscher Verlag der Wissenschaften, Berlin, 1955.
    [18] S. S. Santra, D. Baleanu, K. M. Khedher, O. Moaaz, First-order impulsive differential systems: Sufficient and necessary conditions for oscillatory or asymptotic behavior, Adv. Differ. Equ., 2021 (2021), 283 https://doi.org/10.1186/s13662-021-03446-1 doi: 10.1186/s13662-021-03446-1
    [19] O. Öcalan, Oscillation criteria for systems of difference equations with variable coefficients, Appl. Math. E-Notes, 6 (2006), 119–125.
    [20] O. Öcalan, O. Duman, Oscillation analysis of neutral difference equations with delays, Chaos, Solitons, Fract., 39 (2009), 261–270. https://doi.org/10.1016/j.chaos.2007.01.094 doi: 10.1016/j.chaos.2007.01.094
    [21] O. Öcalan, Linearized oscillation of nonlinear difference equation with advanced arguments, Arch. Math. (Brno), 45 (2009), 203–212.
    [22] N. Parhi, A. K. Tripathy, Oscillation of forced nonlinear neutral delay difference equations of first order, Czech. Math. J., 53 (2003), 83–101.
    [23] N. Parhi, A. K. Tripathy, Oscillation of a class neutral difference equations of higher order, J. Math. Anal. Appl., 28 (2003), 756–774. https://doi.org/10.1016/S0022-247X(03)00298-1 doi: 10.1016/S0022-247X(03)00298-1
    [24] H. Péics, Positive solutions of neutral delay difference equations, Novi. Sad. J. Math., 35 (2005), 111–122.
    [25] S. H. Saker, M. A. Arahet, New oscillation criteria of first order neutral delay difference equations of Emden-Fowler type, J. Comput. Anal. Appl., 29 (2021), 252–265.
    [26] S. S. Santra, D. Majumder, R. Bhattacharjee, T. Ghosh, Asymptotic behaviour for first-order difference equations I, In: Applications of networks, sensors and autonomous systems analytics, Studies in Autonomic, Data-driven and Industrial Computing, Singapore: Springer, 2022. https://doi.org/10.1007/978-981-16-7305-4_4
    [27] X. H. Tang, X. Lin, Necessary and sufficient conditions for oscillation of first-order nonlinear neutral difference equations, Comput. Math. Appl., 55 (2008), 1279–1292. https://doi.org/10.1016/j.camwa.2007.06.009 doi: 10.1016/j.camwa.2007.06.009
    [28] M. K. Yildiz, H. Öğünmez, Oscillatory results of higher order nonlinear neutral delay difference equations with a nonlinear neutral term, Hacet. J. Math. Stat., 43 (2014), 809–814.
    [29] G. Zhang, Oscillation for nonlinear neutral difference equations, Appl. Math. E-Notes, 2 (2002), 22–24.
  • This article has been cited by:

    1. Silvio Fanzon, Marcello Ponsiglione, Riccardo Scala, Uniform distribution of dislocations in Peierls–Nabarro models for semi-coherent interfaces, 2020, 59, 0944-2669, 10.1007/s00526-020-01787-5
    2. Prashant K. Jha, Timothy Breitzman, Kaushik Dayal, Discrete-to-Continuum Limits of Long-Range Electrical Interactions in Nanostructures, 2023, 247, 0003-9527, 10.1007/s00205-023-01869-6
  • 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(1763) PDF downloads(129) Cited by(3)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog