
We study the effect of long-range interactions in non-convex one-dimensional lattice systems in the simplified yet meaningful assumption that the relevant long-range interactions are between M-neighbors for some M≥2 and are convex. If short-range interactions are non-convex, we then have a competition between short-range oscillations and long-range ordering. In the case of a double-well nearest-neighbor potential, thanks to a recent result by Braides, Causin, Solci, and Truskinovsky, we are able to show that such a competition generates M-periodic minimizers whose arrangements are driven by an interfacial energy. Given M, the shape of such minimizers is universal and independent of the details of the energies, but the number and shapes of such minimizers increase as M diverges.
Citation: Andrea Braides, Edoardo Voglino, Matteo Zanardini. Microstructures and anti-phase boundaries in long-range lattice systems[J]. Networks and Heterogeneous Media, 2024, 19(3): 992-1012. doi: 10.3934/nhm.2024044
[1] | 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 |
[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] | 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 |
[4] | 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 |
[5] | 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 |
[6] | Laura Sigalotti . Homogenization of pinning conditions on periodic networks. Networks and Heterogeneous Media, 2012, 7(3): 543-582. doi: 10.3934/nhm.2012.7.543 |
[7] | 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 |
[8] | 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 |
[9] | 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 |
[10] | Hervé Le Dret, Annie Raoult . Homogenization of hexagonal lattices. Networks and Heterogeneous Media, 2013, 8(2): 541-572. doi: 10.3934/nhm.2013.8.541 |
We study the effect of long-range interactions in non-convex one-dimensional lattice systems in the simplified yet meaningful assumption that the relevant long-range interactions are between M-neighbors for some M≥2 and are convex. If short-range interactions are non-convex, we then have a competition between short-range oscillations and long-range ordering. In the case of a double-well nearest-neighbor potential, thanks to a recent result by Braides, Causin, Solci, and Truskinovsky, we are able to show that such a competition generates M-periodic minimizers whose arrangements are driven by an interfacial energy. Given M, the shape of such minimizers is universal and independent of the details of the energies, but the number and shapes of such minimizers increase as M diverges.
In this paper we study the behavior of boundary-value minimum problems related to one-dimensional long-range lattice energies, which in the greatest generality can be stated as the asymptotic behavior as n→+∞ of solutions un={uni}i of the minimization of functionals of the form
En(u)=n∑k=1n∑i=kψk(ui−ui−k) | (1.1) |
on n+1-tuples u={ui}i of parameters with ui∈R, subjected to boundary conditions u0=0 and un=nℓ. In this generality, the problem is very complex and leads to a variety of different issues with competing short and long-range oscillations and concentration effect, except for the trivial case when all ψk are convex, for which the minimizer is essentially unique and coincides with the linear function ui=iℓ, except for boundary effects, which are asymptotically negligible upon some technical assumptions on ψk. Nevertheless, an averaged description of minimizers is possible in the spirit of Γ-convergence. To this end, we regard the energies En as defined on functions u:[0,1]→R, with domain the piecewise-affine functions defined, for some (n+1)-tuple {ui}i, as the interpolation of the points {(i/n,ui)}i=0,…,n. The Γ-convergence of such energies can then be studied with respect to the L1-convergence. Upon some growth hypotheses on ψk that ensure that limits of sequences un with energy of order n belong to some W1,p(0,1), the Γ-limit of 1nEn can be written as
F(u)=∫10ψhom(u′)dt, | (1.2) |
for a convex function ψhom resulting from a nonlinear homogenization process (see [5] and [2] for the higher-dimensional case).
The Γ-convergence above only ensures that the (interpolations of the) minimizers of boundary-value problems for En converge to the corresponding minimizer, or to one of the minimizers, of the analogous continuum boundary-value problem related to F, but brings no further information on their behavior, which may depend on ℓ. Note that if ψhom is strictly convex at ℓ then the unique minimizer is the linear function u(x)=ℓx, while at non-strictly convex points discrete solutions may converge to a particular choice among minimizers.
A particular class of energies for which an analysis of ψhom leading to a description of the convergence of discrete minimizers has been possible is the one studied in [7], where the non-convexity is confined to nearest-neighbour interactions governed by ψ=ψ1, and the long-range potentials are quadratic; that is, ψk(z)=akz2, with ak non negative, and the energies can be written as
En(u)=n∑i=1ψ(ui−ui−1)+n∑k=2n∑i=kak(ui−ui−k)2. | (1.3) |
In that case, the properties of minimizers can be linked to properties of the sequence ak. In particular, in [7] the case of double-well ψ is studied, for which a prototype is
ψ(z)=min{(z−1)2,(z+1)2}, | (1.4) |
in which case it is possible to describe the patterns of the minimizers by tracing whether the value zi=ui−ui−1 lies in one "well" (i.e., zi≤0) or the other one (i.e., zi≥0). As such patterns of minimizers are concerned, we recall the following interesting characterization of minimizers of energies (1.3) subjected to Dirichlet boundary conditions u0=0 and un=ℓn in the case when ak=0 for all k≥2 except for one value k=M: either
(a) minimizers ui are such that zi=ui−ui−1 tend (for n large) to be M-periodic with average ℓ, and take only two values: one, for m indices in the period, in one well and the second one, for M−m indices in the period, in the second well, or
(b) zi defined as above tends (for n large) to be a mixture of two periodic functions as in (a) with some ℓ′ and ℓ″ in the place of ℓ and for two consecutive values m and m+1 between 0 and M.
This characterization extends a formula known before when only a2≠0 (see [5]), in which case we have the only three possibilities that either the parameters zi tends to take a constant value in the first or in the second well, or that we have a 2-periodic pattern mixing values in both wells.
The appearance of microstructure is a recurring feature of non-convex variational systems. Such microstructures may be driven by a scale-free relaxation phenomenon described by convexification or quasi-convexification of the original energy densities (see, e.g., the books by Buttazzo [10] or Dacorogna [11]), or present more regular patterns at a specific scale due to competing long-range and short-range effects (as in the seminal paper by S. Müller [14]; see also [1]). Minimizers of En are similar to the latter, with oscillations both driven by short- and long-range microscopic interactions and by mesoscopic non-convexity.
Simple examples of variational problems exhibiting microscopic oscillations are lattice systems defined on "spin functions", i.e., functions taking only a finite number of values, the traditional choice being −1 and 1. If the energies are "frustrated", that is, the system presents interaction potentials that cannot be all separately minimized at the same time by a single function, then minimization may produce periodically modulated phases (see [3, Chapter 7]). Often, the determination of the period and shape of such minimizers is a nontrivial matter as in the case of infinite-range antiferromagnetic systems studied by Giuliani et al. [13], and has interesting continuum counterparts (see, e.g., [12]).
In the case of double-well problems, the location of the parameter in one or the other well relaxes the strict constraint that the parameter takes two values, that is, the constraint z∈{−1,1} is replaced by a potential ψ(z) where ψ is a strictly positive continuous function minimized in {−1,1}. To distinguish between them, we will call the first type of parameters "hard spins" and the second ones "soft spins". For the prototypical double-well potential ψ in Eq (1.4), it is clear that the two "wells" coincide with z negative and z positive. If also long-range interactions are taken into account, minimization for soft spins may produce patterns analogous to those for frustrated hard spins. Furthermore, the "soft" approach allows us to include more easily boundary-value problems as above.
In this paper we carry on a fine analysis of the energies En in (1.3) by examining not only minimizers, but also parameters ui whose energy in Eq (1.3) differs from the minimum by a finite quantity bounded as n tends to +∞. This is done by using a development by Γ-convergence ([4, Section 1.10] and [6]), and it is performed for double-well ψ and ak≠0 only for k=M, so that the descriptions (a) and (b) above provide the value of minima. The meaningful definition of convergence for functions u depends on whether we are in case (a) or (b) above. For simplicity of illustration, consider that the boundary datum ℓ is such that case (a) holds for some m. Then, given a sequence uni with bounded energy, the sequence is compact in the following sense: there are a finite number of indices inj, which we may suppose to converge after scaling, that is, inj/n→xj, such that in the intervals in the complement of such indices, each function coincides with an M-periodic minimizer ¯ui as in (a), up to an arbitrary small error. Hence, up to subsequences, each such sequence determines a continuum limit u whose derivative u′ takes values in the finite set Mm of M-periodic minimizers as in (a). A similar argument holds in case (b), for which we can conclude that the continuum limit u has derivative with values in Mm∪Mm+1. Once such a piecewise-affine limit is defined, we will prove that the Γ-limit has the form
F(u)=∑t∈S(u′)Φ(u′(t−),u′(t+)), | (1.5) |
where S(u′) denotes the discontinuity set of u′ in (0,1). This shows that minimization may give rise to microscopic patterns Mm, and microscopic incompatibility may give rise to interfaces between elements of Mm (antiphase boundaries) and/or interfaces between elements of Mm and Mm+1 (macroscopic interfaces). In order to avoid boundary effects, the analysis will be carried out under some periodicity assumptions.
It is interesting to note that even though the values of the slopes of microscopic minimizers in Mm depend on the average slope or boundary datum, ℓ, the set Mm has a 'universal' form, and its elements are in correspondence with M-tuples with m values equal to 1 and (m−1)-values equal to 0 (a prototypical version of this is in the case m=2; see [8]). A final remark is that the presence of M-th-neighbor interaction is often compared to that of a singular perturbation with a term containing the M-th derivative for continuum double-well problems. However, while in the continuum case the resulting phase-transition energy is essentially independent of M (see [9,15]), in the discrete case our result shows an increasing complexity of minimizers as M increases.
We will fix M∈N with M≥2 and functions ψ1,ψM:R→[0,+∞) satisfying the coerciveness condition
lim|z|→∞ψ1(z)|z|=+∞. | (2.1) |
We want to study the overall behavior of functionals with competing nearest-neighbor and M-th-neighbor interactions driven by the potential ψ1 and ψM, respectively, of the form
∑iψ1(ui+1−ui)+∑iψM(ui+M−uiM), | (2.2) |
defined on discrete functions indexed on Z. In our assumptions the potential ψ1 will be a double-well energy density which favors oscillations through non-convexity, while ψM is a convex potential favoring long-range ordering.
We preliminary analyze a scaled version, whose analysis will suggest a renormalization argument. We use a standard scaling procedure that allows us to use an analytic approach by Γ-convergence, introducing a reference interval [0,1] and the small parameter εn=1n with n∈N. The energies above, when we take into account the interaction involved on n+1 sites, now parameterized by εni with i∈{0,…,n}, take the form
n−1∑i=0ψ1(ui+1−uiεn)+n−1∑i=0ψM(ui+M−uiMεn). | (2.3) |
In this notation, ui=u(εni). Note that in the last sum we also take into account the values of ui for i∈{n,…,n+M−1}. In the sequel, in order not to have boundary effects, we will define ui for all values of i using some periodic conditions.
After this parameterization, we can identify such discrete functions with the piecewise-affine interpolation on [0,1] of the sites (iεn,ui). We define the space of such functions
An(0,1)={u:[0,1]→R continuous, and affine on (iεn,(i+1)εn),i∈{0,…,n−1}}, |
and the scaled functionals
En,M(u)=n−1∑i=0εnψ1(ui+1−uiεn)+n−1∑i=0εnψM(ui+M−uiMεn), | (2.4) |
if u∈An(0,1), while En,M(u)=+∞ if otherwise u∈L1(0,1)∖An(0,1). Since
n−1∑i=0εnψ1(ui+1−uiεn)=∫10ψ1(u′)dt, |
condition (2.1) ensures that functionals En,M are equicoercive in W1,1(0,1), namely, that if un is bounded in L1(0,1) and En,M(un)≤C<+∞, then, up to subsequences, un converges weakly in W1,1(0,1) and strongly in L1(0,1). The Γ-limit of En,M with respect to this convergence is described in the following result, where we also consider periodic conditions. To that end, we fix ℓ∈R and define
W1,1#,ℓ(0,1)={u∈W1,1loc(R):u(t)−ℓt is 1-periodic}, |
whose discrete counterpart is
A#n,ℓ(0,1)={u∈W1,1#,ℓ(0,1):u|[0,1]∈An(0,1)}. |
Theorem 2.1. The functionals En,M Γ-converge, with respect to the L1-topology, to the functional defined on W1,1(0,1) by
∫10ψ∗∗0(u′(t))dt, | (2.5) |
where ψ0 is given by:
ψ0(z)=ψM(z)+1Mmin{M∑k=1ψ1(zk):M∑k=1zk=Mz,z1,…zM∈R}. | (2.6) |
Furthermore, the convergence is compatible with the addition of periodic condition; that is, with fixed ℓ∈R, the functionals defined by
Eℓn,M(u)={En,M(u)u∈A#n,ℓ(0,1)+∞otherwise |
Γ-converge to
EℓM(u)={∫10ψ∗∗0(u′(t))dtu∈W1,1#,ℓ(0,1)+∞otherwiseinL1(0,1). |
The proof of this result can be found in [7] (see also [5] for the case M=2).
The main point of the analysis at the bulk scaling is the definition of ψ0, which will allow us to renormalize energies (2.2) by subtracting the affine term rℓ given by the tangent to ψ∗∗0 at ℓ and rewriting the sum as
n−1∑i=0(ψM(ui+M−uiM)+1MM−1∑k=0ψ1(ui+k+1−ui+k)−rℓ(ui+M−uiM)). | (2.7) |
This will be formalized as the computation of a higher-order Γ-limit starting from En,M. We will consider periodic boundary conditions. We also make the simplifying assumption that n is a multiple of M, the general case requiring a more complex notation being stated explicitly in Section 3.2, taking into account possible mismatch at the boundary due to incommensurability. In this case, after noting that minEℓM=ψ∗∗0(ℓ), and letting m=mℓ denote the slope of the straight line tangent to ψ∗∗0 at ℓ, we can consider the energies
E1n,M(u)=E1,ℓn,M(u):=Eℓn,M(u)−minEℓMεn=Eℓn,M(u)−ψ∗∗0(ℓ)εn=n−1∑i=0(ψM(ui+M−uiMεn)+1MM−1∑k=0ψ1(ui+k+1−ui+kεn)−ψ∗∗0(ℓ))=n−1∑i=0(ψM(ui+M−uiMεn)+1MM−1∑k=0ψ1(ui+k+1−ui+kεn)−ψ∗∗0(ℓ)−m(ui+M−uiMεn−ℓ))=n−1∑i=0(ψM(ui+M−uiMεn)+1MM−1∑k=0ψ1(ui+k+1−ui+kεn)−rℓ(ui+M−uiMεn)), | (2.8) |
where we have used that ∑n−1i=0(ui+M−ui)=Mℓ thanks to the n-periodicity of ui−ℓεni.
Until now we have made no assumptions on ψ1 and ψM. We study a particular case in which ψM is a strictly convex function and ψ1 is a double-well potential of the form
ψ1(z)=min{W1(z),W2(z)}, |
where W1 and W2 are two smooth convex functions. Note that this is not a very restrictive hypothesis since in the determination of the Γ-limit, only the values of W1 and W2 close to the bottom of the wells will be taken into account, so that more general ψ1 of the double-well type can be taken into account. We also make the assumption that W1 and W2 satisfy
Wi(z)≥cz2−1c, |
for some c>0. This structure provides a useful representation of ψ0. Indeed, we can distinguish two sets A1={x∈R|ψ1(x)=W1(x)} and A2={x∈R|ψ1(x)=W2(x)}. Then, we can rewrite the minimum problem in ψ0 as follows:
min{M∑k=1ψ1(zk):M∑k=1zk=Mz}=min0≤j≤Mmin{j∑k=1W1(zk)+M∑k=j+1W2(zk):M∑k=1zk=Mz,z1,…,zj∈A1,zj+1,…,zM∈A2}=min0≤j≤Mmin{jW1(z1)+(M−j)W2(z2):jz1+(M−j)z2=Mz,z1∈A1,z2∈A2}, |
where the last equality follows from the Jensen inequality. Then, we can define ¯ψj(z)=ψM(z)+1Mfj(z), where fj(z) is the value of the inner minimum problem of above. In this way, ψ0 can be represented as follows:
ψ0(z)=min0≤j≤M¯ψj(z). | (2.9) |
See Figure 1 for a typical form of ψ0.
Remark 2.2. A nontrivial result in [7] (see Theorem 4.1 therein) shows that when ψM is a parabola, ψ∗∗0 will alternate nondegenerate intervals [zlj,zrj] in which ψ∗∗0(z)=ψ0(z)=¯ψj(z) and intervals [zrj−1,zlj], for j≥1, in which ψ∗∗0(z)=rj(z) is a straight line. In this notation, zl0=−∞ and zrM=+∞. We will denote Kj=(zlj,zrj) and Jj=[zrj−1,zlj]. We note that the result holds for any strictly convex ψM with the same proof. Referring to (2.8), this description gives that if ℓ∈Kj, then
rℓ(z)=ψ0(ℓ)+ψ′0(ℓ)(z−ℓ)=¯ψj(ℓ)+¯ψ′j(ℓ)(z−ℓ), |
while if ℓ∈Jj, then rℓ is the affine function through (zrj−1,¯ψj−1(zrj−1)) and (zlj,¯ψj(zlj)).
Note that if we consider
Ein(u)=ψM(ui+M−uiMεn)+1MM−1∑k=0ψ1(ui+k+1−ui+kεn)−rℓ(ui+M−uiMεn), | (2.10) |
then these values are still positive since Ein(u)≥ψ∗∗0(ui+M−uiMεn)−rℓ(ui+M−uiMεn)≥0.
Before stating the convergence result, we need some definitions.
Definition 1. We define Mα⊂RM as the set of minimizers for the problem
min{M∑i=1ψ1(zi):M∑i=1zi=Mα}. | (P) |
We will then set
Mα={Mαifα∈int({z∈domψ0|ψ0(z)=ψ∗∗0(z)})⋃z∈Jj∩{ψ0(z)=ψ∗∗0(z)}Mzifα∈Jj. |
Remark 2.3. Mα and, consequently also Mα, are closed under permutation, that is, if z=(z1,…,zM)∈Mα, then for any permutation σ, (zσ(1),…,zσ(M)) also belongs to Mα.
Remark 2.4. When ψM is a parabola, we can use the structure of Kj and Jj from Remark 2.2 to simplify the Definition 1 as follows:
Mα={Mαifα∈M⋃j=0KjMzrj−1∪Mzljifα∈Jj, j∈{1,…,M}. |
Since ψ1(z)=min{W1(z),W2(z)} then, similarly to the way we found (2.9), we can define the minimum problems
min{jW1(z1)+(M−j)W2(z2):jz1+(M−j)z2=Mα,z1∈A1,z2∈A2}, | (Pj) |
such that minP=min{minPj,j=0,…,M}. By strict convexity, the problem (Pj) admits a unique minimizer (zj1,zj2), corresponding to (Mj) minimizers for (P), i.e., the number of possible M-tuple such that j entries are zj1 and M−j are zj2. Furthermore, for α∈[zlj,zrj], we have minP=minPj, and as a result the set Mα can be written as
Mα={σ((x1,…,xM)):x1=x2=⋯=xj=zj1,xj+1=⋯=xM=zj2,σpermutation}. |
In order to study the limit behavior of E1n,M, we need a notion of convergence of discrete un to a vectorial function.
Definition 2. Given u∈A#n,ℓ(0,1), we consider the subintervals
Ij=[(j−1)εn,jεn]forj∈Z |
and for i∈Z we define the set Ai=⋃Mk=1IiM+k, which is the union of M such consecutive intervals. For k∈{1,…,M}, we define uk the k-th M-interpolation of u as the piecewise-affine function obtained extending the slope zMi+k, where
zj:=uj−uj−1εn, |
to the whole Ai, that is, uk is defined by
{uk(0)=u0,u′k(t)=zMi+kift∈Ai |
Definition 3. We say that a sequence of discrete functions (un)n converges to ¯u=(¯u1,…,¯uM) in a functional topology (for instance, L1(0,1) or L∞(0,1)) if for each k the k-th M−interpolation unk converges to ¯uk in that topology.
Theorem 2.5 (Equi-coerciveness of the energies). If a sequence (un)n∈N satisfies supnE1n,M(un)<+∞, then, up to addition of a constant and extraction of a subsequence, the sequence converges uniformly to some M-tuple of piecewise-affine functions ¯u=(¯u1,…,¯uM) with ¯u′(t)∈Mℓ for almost every t. Moreover, there exists a finite set S⊂(0,1) such that un converges in W1,∞loc((0,1)∖S) to ¯u and the jump set S(¯u′) is contained in S.
This compactness theorem justifies the use of the convergence in Definition 3 for the Γ-limit of E1n,M. In order to describe it, we define the following interfacial energy.
Definition 4. Let ℓ∈R and z=(z1,z2,…zM),z′=(z′1,z′2,…z′M)∈Mℓ. The transition energy between z and z′ is defined by
Φ(ℓ)(z,z′):=infN∈Nminu{∑i∈Z(ψM(ui+M−uiM)+1MM−1∑k=0ψ1(ui+k+1−ui+k)−rℓ(ui+M−uiM)):u:Z→R,ui=uz(i)fori≤−N,ui=uz′(i)fori≥N}, | (2.11) |
where rℓ is the tangent line to ψ∗∗0 at ℓ computed at the point x and, for any z∈RM, uz is the piecewise affine function R⟶R defined as follows:
{u′z(t)=zkfort∈(k−1,k)+MZuz(0)=0 | (2.12) |
With this definition, we can state the Γ-convergence result as follows.
Theorem 2.6 (first-order Γ-limit with periodic boundary conditions). Assume that ψM is a nondegenerate convex parabola, ψ1 is of class C1(U), where U⊂R open is such that Mℓ⊂UM, and ψ∗∗0 is differentiable in ℓ. We define the domain
D#ℓ={u:R→RM|u(locally)piecewiseaffine,M∑k=1uk(t)−Mℓt1-periodic, u′1-periodic andu′(t)∈Mℓfor allt∈R∖S(u′)}. |
Then, E1n,M Γ-converges, with respect to the L∞ topology, to
E1M(u)=∑t∈S(u′)∩(0,1]Φ(ℓ)(u′(t−),u′(t+)), | (2.13) |
with domain D#ℓ.
In this section, for a greater generality, we consider L>0 and functions parameterized on [0,L] instead of [0,1]. In this case, εn=Ln, and we extend all the notation introduced in the case L=1. With fixed ℓ∈R, the energies we consider are directly written in the form
E1n,M(u):=n−1∑i=0(ψM(ui+M−uiMεn)+1MM−1∑k=0ψ1(ui+k+1−ui+kεn)−rℓ(ui+M−uiMεn)), | (3.1) |
defined on A#n,ℓ(0,L). Note that E1n,M(u)=∑n−1i=0Ein(u), with Ein be given by (2.10).
We can state the compactness result as in Theorem 2.5 independently of the boundary condition as follows.
Proposition 3.1. Let Ein be given by (2.10). If
supnn−1∑i=0Ein(un)=:C<+∞, |
then un satisfies the claim of Theorem 2.5 with the interval [0,L] in the place of the interval [0,1].
Proof. Let η>0, then we define the set
In(η):={i∈{0,1,…n−1}:Ein(un)>η}, |
and, as a consequence of the bound on the sum, we get that
supn#In(η)≤C(η)∼C/η<+∞. |
Therefore, if i∉In(η), by adding and subtracting ψ0(uni+M−uniMεn) to Ein(un), we obtain the two inequalities
0≤ψM(uni+M−uniMεn)+1M(M−1∑k=0ψ1(uni+k+1−uni+kεn))−ψ0(uni+M−uniMεn)≤η |
0≤ψ0(uni+M−uniMεn)−rℓ(uni+M−uniMεn)≤η. |
Note that, if z and (z1,…,zM) satisfy
a) ψM(z)+1M(∑M−1k=0ψ1(zk))−ψ0(z)≤η,for ∑M−1k=0zk=Mz
b) ψ0(z)−rℓ(z)≤η
then (z1,…,zM) is close to a minimizing M-tuple for the min in the definition of ψ0, while ψ0(z) is close to rℓ(z), the tangent line of ψ∗∗0 in ℓ. This means that if ψ0(ℓ)=ψ∗∗0(ℓ), then z is close to ℓ, while if ℓ is in some Jj, then z is close to either zrj−1 or zlj. Hence, for η small enough, we can find ε>0 so that if (a) and (b) are satisfied, then
dist((z1,…,zM),Mℓ)≤ε<12min{|z′−z"|:z′,z"∈Mℓ}. |
This entails that for each i∉In(η) there exists a unique zni=(zni,1,…,zni,M)∈Mℓ such that
|(uni+1−uniεn,…,uni+M−uni+M−1εn)−zni|≤ε,⟹|uni+k−uni+k−1εn−zni,k|≤ε, | (3.2) |
for all k∈{1,…,M}. We note that if as well i+1∉In(η), then the unique zni+1 must be a cyclic permutation of zni, specifically,
zni+1=(zni,2,…,zni,M,zni,1). |
Since In(η) is a finite set, we can identify Nn pairs of indices 0=j0≤i1<j1<i2<j2<⋯<iNn<jNn≤iNn+1=n such that In(η)={0,…n−1}∖∪Nnk=1{ik,…,jk}, and for all indices h between ik and jk, we have
|unh−unh−1εn−znik,[h−ik]+1|≤ε, |
where [h−ik] is the reminder of h−ik mod M.
Since C≥E1n,M(un)≥η(Nn−1), we must have that Nn is bounded with respect to n. Then, we can assume, up to extracting subsequences, that Nn is constantly equal to N. Similarly, since Mℓ is a finite set, we can assume znik=zk. With fixed k∈{2,…,N}, for each n, we chose an index ˜ik∈{jk−1+1,…,ik−1} and defined the sequence {˜xnk} such that ˜xnk=˜ikεn, then, up to subsequences, we can assume xnk→xk. However, |ik−jk−1| must be bounded independently of n and k, since
C≥E1n,M(un)≥N+1∑k=1η|ik−jk−1|. |
Then, for any i∈{jk−1+1,…,ik−1}, |˜xnk−iεn|≤Cηεn. Therefore, the whole {jk−1+1,…,ik−1}εn is converging to xk.
Finally, we can define the sets S=⋃Nk=1{xk} and, for a fixed small δ, Sδ=⋃Nk=1(xk−δ,xk+δ). By (3.2), we have that for n large enough, the s-th M-interpolation satisfies
supt∈(0,1)∖Sδ|(uns)′(t)−Zs(t)|≤ε, | (3.3) |
where Z(t)=(Z1(t),…,ZM(t)) is a piecewise-constant function such that Z(t)=zk for t∈(xk−1,xk).
Note that from the equi-coerciveness of En,M and hence also of E1n,M, we obtain that un,s is a pre-compact sequence in H1(0,L), so we can suppose that they converge uniformly. However, from (3.3), the subsequence is such that (uns)′ is converging to Zs. In conclusion, by applying a diagonal argument, we have proved that, up to subsequences, un converges in W1,∞loc((0,L)∖S) (in the sense of Definition 3) to a function u such that u′(t)=Z(t)∈Mℓ and the jump set S(u′)⊆S.
Note that in this proof we have not used the periodicity condition. If it is taken into account that we have to note that S(u′) can also contain the point 0.
We now compute the Γ-limit subjected to periodic boundary conditions, without the simplifying assumption that n is a multiple of M used for presentation purposes in the previous section. We will show that the limit exists and can be characterized if more in general n=q modulo M. The energy will have the same form as in the case q=0, but the characterization of the domain of the Γ-limit will depend on q, since M−interpolations do not necessarily inherit the periodicity condition of u. Note in particular that the limit as n→+∞ does not exist.
For n∈N, we let q∈{0,…,M−1} denote its class modulus M (n≡q mod M), then given u a piecewise-affine function in W1,1#,ℓ(0,L), for any index Mi+q with q∈{0,1,…,M−1}, its M-interpolations uk will satisfy:
uk,Mi+q−uk,Mi+q−1εn=uMi+k−uMi+k−1εn=un+Mi+k−un+Mi+k−1εn. |
Since n=Mr+q, we write
uk,Mi+k−uk,Mi+k−1εn=uM(i+r)+q+k−uM(i+r)+q+k−1εn, |
which is linked to the q+k interpolation (mod M) shifted of ⌊n+kM⌋. Thus, we have
u′k(t)=u′k+q(t+Mrεn)if q+k≤Mu′k(t)=u′k+q−M(t+M(r+1)εn)if q+k>M | (3.4) |
Proposition 3.2. Consider n(r)=Mr+q for a fixed q∈{0,…,M−1}. Consider a sequence (un(r))r∈N such that supn(r)E1n(r),M(un(r))≤C<+∞. Then, there exists a finite set S⊂(0,L] for which un(r) (up to subsequences) converges in W1,∞loc(R∖(S+LZ)) to a u=(u1,…,uM) as in Theorem 2.5, such that for any k=1,…,M we have
u′k(t)=u′q+k(t+L)ifq+k≤M,u′k(t)=u′k+q−M(t+L)ifq+k>M |
for almost every t∉S(u′).
Proof. The convergence is given by Proposition 3.1. What remains to prove is the periodicity property. Suppose at first q+k≤M. Fix t∉S(u′) and suppose t+L∉S(u′), then there exists a δ>0 such that Iδ(t)=[t−δ,t+δ]⊂R∖S(u′) and Iδ(t+L)⊂R∖S(u′). To facilitate the reading, in the following computations we take n=n(r). For every k fixed, (unk)′ converges uniformly on Iδ(t) and (unk+q)′ on Iδ(t+L). Since Mrεn=L(1−qn), we can rewrite (3.4) as
(unk)′(t)=(unk+q)′(t+L−Lqn(r)) |
Taking the limit in r, the left-hand side is converging to u′k(t), while for the righthand side we can notice that, since u′ is piecewise constant, u′k+q(t+L)=u′k+q(t+L−Lqn) for every n=n(r) large enough, so that
|(unk+q)′(t+L−Lqn)−u′k+q(t+L)|=|(unk+q)′(t+L−Lqn)−u′k+q(t+L−Lqn)|≤sups∈Iδ(t+L)|(unk+q)′(s)−u′k+q(s)|→r→+∞0. |
The case q+k>M follows similarly.
Remark 3.3. Note that at the limit we have the boundary conditions u′k(0+)=u′q+k(L+), where q+k is intended modulus M. This holds also if 0∈S(u′).
Proposition 3.4. Given (un)n∈N such that supnE1n,M(un)=:C<+∞. Consider 0=x0<x1<⋯<xN=L and α1,…,αN∈R such that Mαj⊆Mℓ/L and, by Theorem 2.5, un converges to u satisfying
u′(t)=zαj∈Mαj,fort∈(xj−1,xj). |
Then, ℓL=∑Nj=1αj(xj−xj−1).
Proof. Suppose L=1. By the boundary conditions, we have
ℓ=(unn−unM⌊n/M⌋)+⌊n/M⌋∑i=0unM(i+1)−unMi=(unn−unM⌊n/M⌋)+⌊n/M⌋∑i=0MεnunM(i+1)−unMiMεn. |
Now, with fixed δ>0, we define Ijδ=(xj−δ,xj+δ)∩[0,1], for j=0,…,N, and we call Sδ=∪Nj=0Ijδ. From Proposition 3.1, the interpolation derivative (unk)′(t) converges to zαjk on [xj−1+δ,xj−δ], then for an index i such that [Miεn,M(i+1)εn]⊂[xj−1+δ,xj−δ],
unM(i+1)−unMiMεn=1MM(i+1)−1∑m=Miunm+1−unmεn=1MM∑k=1unk,Mi+1−unk,Miεn→n→+∞1MM∑k=1zαjk=αj |
Let k−j,n be the smallest index multiple of M such that Mk+j,nεn≥xj+δ and similarly let k+j,n be the largest such that Mk−j,nεn≤xj−δ, so that limnεnMk−j,n=xj−δ and limnεnMk+j,n=xj+δ. We also define k−0,n=0 and k+N,n=⌊nM⌋. Then,
|k−j,n−1∑i=k+j−1,nMεnunM(i+1)−unMiMεn−αj(xj−xj−1−2δ)|≤|k−j,n−1∑i=k+j−1,nMεn(unM(i+1)−unMiMεn−αj)|+|αj(xj−δ−εnMk−j,n)|+|αj(xj−1+δ−εnMk+j−1,n)|, |
which tends to 0 as n tends to infinity. On the other hand, since supnE1n,M(un)=C, there exists a constant C1 independent from n such that for every index i we have
Ein(un)≤C⟹ψ1(uni+1−uniεn)−rℓ(uni+1−uniεn)≤C1, |
where we used the definition of Ein in Eq (2.10). Thus, from the superlinear growth of ψ1, we must have that there exists an R>0 such that |uni+1−uniεn|≤R. Then, we can conclude
|ℓ−N∑j=1k−j,n−1∑i=k+j−1,nMεnunM(i+1)−unMiMεn|=|unn−unMk+N,n+N∑j=0k+j,n−1∑i=k−j,nMεnunM(i+1)−unMiMεn|=|n−1∑i=Mk+N,nεnuni+1−uniεn+N∑j=0Mk+j,n−1∑i=Mk−j,nεnuni+1−uniεn|≤Rεn(MN∑j=0(k+j,n−k−j,n)+n−Mk+N,n). |
Sending n to +∞, we have
|ℓ−N∑j=1αj(xj−xj−1−2δ)|≤2RNδ |
Finally, for arbitrariness of δ>0, we must have ℓ=∑Nj=1αj(xj−xj−1).
Remark 3.5. Let Φ=Φ(ℓ) be as in Definition 4. We note that the minimum is well-defined because of Weierstrass' Theorem. Moreover, the infimum in N∈N in (2.11) can be replaced with the limit as N→+∞ because of the decreasing monotonicity. The terms inside the sums are 0 for i≥N or i≤−N−M when z,z′∈Mℓ. The function Φ:Mℓ×Mℓ→[0,∞] is a sub-additive function. In particular the functional defined by the righthand side of (2.13) is lower-semicontinuous.
Remark 3.6. Given ξ∈R, we can consider
Φ(z,z′,ξ):=infN∈Nminu{∑i∈Z(ψM(ui+M−uiM)+1MM−1∑k=0ψ1(ui+k+1−ui+k)−rℓ(ui+M−uiM)):u:Z→R,u=uz+ξ1fori≤−N,u=uz′+ξ2fori≥N,ξ=ξ2−ξ1} | (3.5) |
=infN∈Nminu{∑i∈Z(ψM(ui+M−uiM)+1MM−1∑k=0ψ1(ui+k+1−ui+k)−rℓ(ui+M−uiM))u:Z→R,u=uzfori≤−N,u=uz′+ξfori≥N}. | (3.6) |
The two formulas are equal because the sums which are involved are invariant under vertical translations of u.
Now note that Φ(z,z′,ξ)=Φ(z,z′) for every ξ∈R, under the assumption that ψM,ψ1 are C1(U) functions, where U⊂R open is such that Mℓ⊂UM. Indeed, if (uN)N is a minimizing sequence in the definition of Φ(z,z′,ξ), then we can consider the following new sequence:
˜u2N={uz′fori≥2Nuz′+2ξ−iξNforN≤i≤2NuNfori≤N. |
Observe that for varying N∈N, all ˜u2N's are competitors in the definition of Φ(z,z′). Let z∈Mα and z′∈Mα′ with α,α′∈R such that Mα⊆Mℓ and Mα′⊆Mℓ. We compute the difference between the involved sums for ˜u2N and the ones for uN by
2N∑i=N(ψM(α′−ξN)+1MM−1∑k=0ψ1(z′k−ξN)−rℓ(α′−ξN))+o(1N). | (3.7) |
Now observe that the C1 function f(η)=ψM(α′−η)+1M∑M−1k=0ψ1(z′k−η)−rℓ(α′−η) is always non-negative and it is equal to 0 if η=0, so that η=0 is a minimum point for f. Hence, by Fermat's Theorem f′(0)=0, and so
limN→+∞2N∑i=N(ψM(α′−ξN)+1MM−1∑k=0ψ1(z′k−ξN)−rℓ(α′−ξN))=0. |
In particular, we proved that Φ(z,z′)≤Φ(z,z′,ξ). Arguing in the same way, one can show the reverse inequality.
Remark 3.7. A last useful remark is that Φ(z,z′) is stable under cyclic permutations of the entries of z and z′. More precisely, if we define for q=1,…,M−1 the cyclic permutation σq:Mℓ→Mℓ given by
σq:z=(z1,…,zM)⟼σq(z):=(zM+1−q,…,zM,z1,…,zM−q), | (3.8) |
then Φ(σq(z),σq(z′))=Φ(z,z′). Indeed, if u is a competitor for the minimum problem in Φ(z,z′), then the translation (Tqu)i=ui−q preserves the value of the objective function in the definition of Φ and it satisfies
Tqu=uσq(z)fori≤−N−q,Tqu=uσq(z′)fori≥N−q. |
Thus, Tqu is a competitor for the minimum problem in Φ(σq(z),σq(z′)). Clearly we can argue in the same way with the inverse translation, proving the claim.
Theorem 3.8 (first-order Γ-limit with periodic boundary conditions). Assume that ψM is a nondegenerate strictly convex function, ψ1 is of class C1(U), where U⊂R open is such that Mℓ/L⊂UM, and ψ∗∗0 is differentiable in ℓL. Constraining n≡qmodM, we define the domain
D#q={u:R→RM|upiecewise affine,u′(t)∈Mℓ/Lforallt∈R∖S(u′),#(S(u′)∩(0,L])<+∞,M∑k=1uk(t)−MℓLtL-periodic, u′k(t)=u′[q+k]modM(t+L)fork=0,…,M−1}. |
Then E1n,M Γ-converges, with respect to L∞ topology, to
E1q(u)=∑t∈S(u′)∩(0,1]Φ(ℓ/L)(u′(t−),u′(t+)), |
with the domain dom(E1q)=D#q.
Proof. We first consider the case in which L=1 and n≡0 mod M.
Lower bound. Let un→u in L∞(0,1) be such that E1,ℓn,M(un)≤C<+∞ for every n∈N. Then, by Proposition 3.2, there exists a finite set S:={x1,…xN}⊂(0,1], with 0<x1<…xN−1<xN≤1, and there exist z1,…,zN∈Mℓ such that un (up to subsequences) converges in W1,∞loc(R∖(S+Z)) to a M-tuple u∈D#0 such that u′(t)=zj∈Mαj for t∈(xj−1+k,xj+k) and for all k∈Z for j∈{1,…,N}, with x0=xN−1. For j∈{1,...,N}, let (kjn)n be a sequence of natural numbers divisible by M such that
limn→+∞kjnεn−xj=0. | (3.9) |
Moreover, let (hjn)n be a sequence in MN such that
limn→+∞εnhjn=xj+xj−12. | (3.10) |
Now, we can write
E1n,M(un)=N∑j=1hj+1n−1∑i=hjn(ψM(uni+M−uniMεn)+1MM−1∑s=0ψ1(uni+s+1−uni+sεn)−r(uni+M−uniMεn)), | (3.11) |
where r=rℓ, and we have used the notation hN+1n=h1n+n. By periodicity we can choose the endpoints h1n and h1n+n without changing the sum. In order to recover Φ, we define
˜uni={uzj(i)−uzj+1(hjn−kjn)+unhjnεnfor i≤hjn−kjnuni+kjnεnfor hjn−kjn≤i≤hj+1n−kjnuzj+1(i)−uzj+1(hj+1n−kjn)+unhj+1nεnfor i≥hj+1n−kjn. | (3.12) |
Since zj∈Mαj, we note that for i≥hj+1n−kjn, we have
ψM(˜uni+M−˜uniM)+1MM−1∑s=0ψ1(˜uni+s+1−˜uni+s)=ψ0(αj+1)=r(αj+1)=r(˜uni+M−˜uniMεn) |
and similarly for i≤hjn−kjn−M. For hjn−kjn≤i≤hj+1n−kjn−M instead, we have
˜uni+1−˜uni=uni+kjn+1−uni+kjnεnand˜uni+M−˜uniM=uni+kjn+M−uni+kjnMεn. |
Then, defining ξnj=uzj+1(hj+1n−kjn)−uzj+1(hjn−kjn)−unhj+1n−unhjnεn, we can rewrite
hj+1n−1∑i=hjn(ψM(uni+M−uniMεn)+1MM−1∑s=0ψ1(uni+s+1−uni+sεn)−r(uni+M−uniMεn))= ∑i∈Z(ψM(˜uni+M−˜uniM)+1MM−1∑k=0ψ1(˜uni+k+1−˜uni+k)−r(˜uni+M−˜uniM))+ωn≥ Φ(zj,zj+1,ξnj)+ωn=Φ(zj,zj+1)+ωn. |
The error ωn comes from the difference in behavior between ˜uni and uni+kjn when hj+1n−kjn−M<i<hj+1n−kjn or hjn−kjn−M<i<hjn−kjn. More precisely, for hj+1n−kjn−M<i<hj+1n−kjn, ωn involves terms of the kind
ψM(uni+kjn+M−uni+kjnMεn)−ψM(˜uni+M−˜uniM), | (3.13) |
and similar terms where ψM is switched with ψ1 or −r. Instead, for the indices hjn−kjn−M<i<hjn−kjn, there are terms of the following type:
r(˜uni+M−˜uniM)−ψM(˜uni+M−˜uniM)−1MM−1∑k=0ψ1(˜uni+k+1−˜uni+k). | (3.14) |
To show that limnωn=0, it is sufficient to show that
{limn|uni+kjn+1−uni+kjnεn−(uzj+1(i+1)−uzj+1(i))|=0,for i∈(hj+1n−kjn−M,hj+1n−kjn+M),limn|uni+kjn+1−uni+kjnεn−(uzj(i+1)−uzj(i))|=0,for i∈(hjn−kjn−M,hjn−kjn+M). |
Indeed, by the definition of ˜un in (3.12), also ˜un satisfies analogous limits. Therefore, the differences of the kind (3.13) are negligible by the continuity of ψM, ψ1, and r. On the other hand, the terms of the form (3.14) will tend to
r(αj)−ψM(αj)−1MM−1∑k=0ψ1(uzj(i+k+1)−uzj(i+k))=0, |
as a consequence of the fact that zj∈Mαj. Let iM denote the residual class of i with respect to M. Then, by the definition of uzj+1 as in Eq (2.12), we have
uzj+1(i+1)−uzj+1(i)=(zj+1)iM+1. |
On the other hand, since there exists a compact around the midpoint of (xj,xj+1), which contains εn(hj+1n−M,hj+1n+M) for all n, we can use the locally uniform convergence of u′n to u′ to gain
uni+1+kjn−uni+1+kjnεn=(un)′iM+1(εn(i+kjn))⟶u′iM+1(xj+1+xj2)=(zj+1)iM+1for n→+∞. |
The case i∈(hjn−kjn−M,hjn−kjn+M) is analogous. This proves the claim and, consequently, that limnωn=0. In particular, putting everything together, we conclude
lim infnE1n,M(un)≥N∑j=1Φ(zj,zj+1). |
In the case q≠0, the only thing that changes is that we can not define hN+1n=h1n+n, because in this way it will not be divisible by M thus, we take hN+1n the highest index multiple of M below h1n+n. Consequently, in the decomposition (3.11), there will appear q residual terms of the kind:
ψM(uni+M−uniMεn)+1MM−1∑s=0ψ1(uni+s+1−uni+sεn)−r(uni+M−uniMεn), |
but since iεn is close to 1+x1+x02, that is, far from any critical point xj, by continuity we have that they tend to 0.
Upper bound. Let u be such that E10(u)<∞ and suppose, without loss of generality, that u(0)=0, then by definition there exist N∈N, z1,…zN∈Mℓ and we can write S(u′)∩(0,1]={x1,…xN}, with 0<x1<…xN−1<xN≤1, u′(t)=zj∈Mαj for t∈(xj−1+k,xj+k) and for all k∈Z, with x0=xN−1. Up to translations, which do not change the energy, we can always suppose that 1∉S(u′), that is, xN<1. Moreover, we have that
E10(u)=N∑j=1Φ(zj−1,zj). |
We first consider the case q=0. With fixed ε>0, there exists ˜N=˜N(ε)∈N multiple of M and N discrete functions vj such that
Φ(zj−1,zj)≤∑i∈ZψM(vji+M−vjiM)+1MM−1∑k=0ψ1(vji+k+1−vji+k)−rℓ(vji+M−vjiM)≤Φ(zj−1,zj)+ε | (3.15) |
and
vj={uzj−1 for i≤−˜Nuzj for i≥˜N |
for every j∈{1,…N}. Consider now the sequence of functions (un)n defined as follows
uni={εnv1i−knN+n−εnv1n−knN for 0≤i≤kn1−˜Nεnv2i−kn1+εnD2 for kn1−˜N≤i≤kn2−˜N…εnvNi−knN−1+εnDN for knN−1−˜N≤i≤knN−˜Nεnv1i−knN+εnD1 for knN−˜N≤i≤n, | (3.16) |
where knj:=min{k∈N:k≥xjn, and k is multiple of M} and
D2=v1kn1−knN+n−˜N−v2−˜N−v1n−knND3=D2+v2kn2−kn1−˜N−v3−˜N…DN=DN−1+vN−1knN−1−knN−2−˜N−vN−˜ND1=DN+vNknN−knN−1−˜N−v1−˜N. |
We note that unn−un0=∑Nj=1αj(knj−knj−1)εn, which in general is different from ℓ; therefore, in order to adjust the periodicity conditions, we apply a linear correction term ˜uni=uni+δnin, where δn=ℓ−∑Nj=1αj(knj−knj−1)εn. Since now the the boundary condition is satisfied, we can extend ˜un to R such that ˜un(t)−ℓt is 1-periodic. From Proposition 3.4 and the fact that |nxj−knj|≤M, we have
δn=N∑j=1αj(xj−xj−1−knjεn+knj−1εn)=O(εn)⟹∃C>0: (knj−knj−1)δn≤C ∀j,n. | (3.17) |
Since un is converging uniformly to u and δnx→0 uniformly in x∈[0,1], we have that ˜un→uin L∞([0,1]) (in the notion of 3). As for the convergence of the energy, if we let
Eni(u)=ψM(ui+M−uiMεn)+1MM−1∑k=0ψ1(ui+k+1−ui+kεn)−rℓ(ui+M−uiMεn), |
then, by Eq (3.15) we have
E10(u)+ε≥E1n,M(˜un)+N∑j=1(Rjn−knj−˜N−M∑i=knj−1+˜NEni(˜un)), |
where Rjn=∑˜N−1i=−˜N−M+1Eni(εnvj)−Eni+knj−1(˜un). However, this Rjn involves only a finite number of indices, independently from n; thus, since ˜uni+M−˜uniMεn=vji+M−vjiM+δn and ˜uni+1−˜uniεn=vji+1−vji+δn, by continuity of ψ1,ψM, and rℓ we have
limnN∑j=1Rjn=0. |
Instead, for i=knj−1−˜N,…,knj−˜N−M, we have vji−knj−1=uzj(i−knj−1) and then
Eni(˜un)=ψM(αj+δn)+1MM∑k=1ψ1((zj)k+δn)−rℓ(αj+δn)=:fj(δn). |
Note that we lost the dependence on i, so by (3.17) we have
knj−˜N−M∑i=knj−1+˜NEni(˜un)=(knj−knj−1−M)fj(δn)=(knj−knj−1−M)δn(fj(δn)δn)≤Cfj(δn)δn. |
By Fermat's Theorem we can conclude that limnfj(δn)δn=f′(0)=0; in conclusion
E10(u)+ε≥lim supnE1n,M(˜un), |
and the claim is proved by the arbitrariness of ε.
When q≠0, the proof proceeds similarly, but we have to pay attention to the boundary mismatch in the cycle of slopes in ˜u. This can be fixed by taking into account that, by Remark 3.7, we can write the energy as
E1q(u)=N∑j=1Φ(σq(zj−1),σq(zj)), |
where σq is defined as in Eq (3.8). Then, we can find some ˜N multiple of M and some vj as in Eq (3.15) with the condition
vj={uσq(zj−1)for i≤−˜Nuσq(zj)for i≥˜N. |
Hence, we define knj:=min{k∈N:k≥xjn and k≡q mod M} and we choose un as in Eq (3.16). In this way, the interpolations un are still converging uniformly to u because
uσq(zj)(i+knj+1)−uσq(zj)(i+knj)=uσq(zj)(i+q+1)−uσq(zj)(i+q)=uzj(i+1)−uzj(i). |
Since n−knN is divisible by M and unn−unn−1εn=σq(z1)M, we can extend un by periodicity on the whole R defining a recovery sequence.
The authors have equally contributed to the manuscript.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
Andrea Braides is an editorial board member for Networks and Heterogeneous Media and was not involved in the editorial review or the decision to publish this article. All authors declare that there are no competing interests.
The research reported in the present contribution was carried out as part of the projects "A Fluid-Structure Interaction tool for the protection of Clean Energy Production sites (FSI-CEP)" funded by the MUR Progetti di Ricerca di Rilevante Interesse Nazionale (PRIN) Bando 2022 PNRR–grant P20227CSJ5, and "Variational methods for stationary and evolution problems with singularities and interfaces" PRIN 2022J4FYNJ. AB is a member of GNAMPA of INdAM.
The authors declare there is no conflict of interest.
[1] |
G. Alberti, S. Müller, A new approach to variational problems with multiple scales, Comm. Pure Appl. Math., 54 (2001), 761–825. https://doi.org/10.1002/cpa.1013 doi: 10.1002/cpa.1013
![]() |
[2] |
R. Alicandro, M. Cicalese, A general integral representation result for continuum limits of discrete energies with superlinear growth, SIAM J. Math. Anal., 36 (2004), 1–37. https://doi.org/10.1137/S0036141003426471 doi: 10.1137/S0036141003426471
![]() |
[3] | R. Alicandro, A. Braides, M. Cicalese, M. Solci, Discrete Variational Problems with Interfaces, vol. 40 of Cambridge Monographs on Applied and Computational Mathematics, Cambridge University Press, Cambridge, 2024. |
[4] | A. Braides, Γ-convergence for Beginners, vol. 22 of Oxford Lecture Series in Mathematics and its Applications, Oxford University Press, Oxford, 2002. |
[5] | A. Braides, M. Gelli, M. Sigalotti, The passage from nonconvex discrete systems to variational problems in Sobolev spaces: the one-dimensional case, Trudy Mat. Inst. Steklova, 236 (2002), 408–427. |
[6] |
A. Braides, L. Truskinovsky, Asymptotic expansions by Γ-convergence, Continuum Mech. Thermodyn., 20 (2008), 21–62. https://doi.org/10.1007/s00161-008-0072-2 doi: 10.1007/s00161-008-0072-2
![]() |
[7] |
A. Braides, A. Causin, M. Solci, L. Truskinovsky, Beyond the classical Cauchy-Born rule, Arch Rational Mech Anal, 247 (2023), 107. https://doi.org/10.1007/s00205-023-01942-0 doi: 10.1007/s00205-023-01942-0
![]() |
[8] |
A. Braides, M. Cicalese, Surface energies in nonconvex discrete systems, Math Models Methods Appl Sci, 17 (2007), 985–1037. https://doi.org/10.1142/S0218202507002182 doi: 10.1142/S0218202507002182
![]() |
[9] | G. C. Brusca, D. Donati, M. Solci, Higher-order singular perturbation models for phase transitions, arXiv: 2402.13626, [Preprint], (2024) [cited 2024 September 26]. Available from: https://doi.org/10.48550/arXiv.2402.13626 |
[10] | G. Buttazzo, Semicontinuity, Relaxation and Integral Representation in the Calculus of Variations, vol. 207 of Pitman Research Notes in Mathematics Series, Longman Scientific & Technical, Harlow, 1989. |
[11] | B. Dacorogna, Direct Methods in the Calculus of Variations, vol. 78 of Applied Mathematical Sciences, Springer, New York, 2008. |
[12] |
S. Daneri, E. Runa, Exact periodic stripes for minimizers of a local/nonlocal interaction functional in general dimension, Arch Rational Mech Anal, 231 (2019), 519–589. https://doi.org/10.1007/s00205-018-1285-6 doi: 10.1007/s00205-018-1285-6
![]() |
[13] |
A. Giuliani, J. L. Lebowitz, E. H. Lieb, Checkerboards, stripes, and corner energies in spin models with competing interactions, Phys. Rev. B, 84 (2011), 064205. https://doi.org/10.1103/PhysRevB.84.064205 doi: 10.1103/PhysRevB.84.064205
![]() |
[14] |
S. Müller, Singular perturbations as a selection criterion for periodic minimizing sequences, Calc. Var, 1 (1993), 169–204. https://doi.org/10.1007/BF01191616 doi: 10.1007/BF01191616
![]() |
[15] | M. Solci, Free-discontinuity problems generated by higher-order singular perturbations, arXiv: 2402.10656, [Preprint], (2024) [cited 2024 September 26]. Available from: https://doi.org/10.48550/arXiv.2402.10656 |