Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

Some computable quasiconvex multiwell models in linear subspaces without rank-one matrices

  • In this paper we apply a smoothing technique for the maximum function, based on the compensated convex transforms, originally proposed by Zhang in [1] to construct some computable multiwell non-negative quasiconvex functions in the calculus of variations. Let KEMm×n with K a finite set in a linear subspace E without rank-one matrices of the space Mm×n of real m×n matrices. Our main aim is to construct computable quasiconvex lower bounds for the following two multiwell models with possibly uneven wells:

    i) Let f:KEE be an L-Lipschitz mapping with 0L1/α and H2(X)=min{|PEXAi|2+α|PEXf(Ai)|2+βi:i=1,2,,k}, where α>0 is a control parameter, and

    ii) H1(X)=α|PEX|2+min{|Ui(PEXAi)|2+γi:i=1,2,,k}, where AiE with Ui:EE invertible linear transforms for i=1,2,,k. If the control paramenter α>0 is sufficiently large, our quasiconvex lower bounds are 'tight' in the sense that near each 'well' the lower bound agrees with the original function, and our lower bound are of C1,1. We also consider generalisations of our constructions to other simple geometrical multiwell models and discuss the implications of our constructions to the corresponding variational problems.

    Citation: Ke Yin, Kewei Zhang. Some computable quasiconvex multiwell models in linear subspaces without rank-one matrices[J]. Electronic Research Archive, 2022, 30(5): 1632-1652. doi: 10.3934/era.2022082

    Related Papers:

    [1] Haijun Wang, Gege Kang, Ruifang Zhang . On optimality conditions and duality for multiobjective fractional optimization problem with vanishing constraints. Electronic Research Archive, 2024, 32(8): 5109-5126. doi: 10.3934/era.2024235
    [2] Zhensheng Zhou, Lin Wang, Xue Zou, Fei Wang, Zaijun Zhang, Xiaobo Yan . The first hitting time analysis of evolutionary algorithms based on renewal process. Electronic Research Archive, 2024, 32(5): 2994-3015. doi: 10.3934/era.2024137
    [3] Yiyuan Qian, Haiming Song, Xiaoshen Wang, Kai Zhang . Primal-dual active-set method for solving the unilateral pricing problem of American better-of options on two assets. Electronic Research Archive, 2022, 30(1): 90-115. doi: 10.3934/era.2022005
    [4] Yi Liu, Lei Gao, Tianxu Zhao . Partially doubly strictly diagonally dominant matrices with applications. Electronic Research Archive, 2023, 31(5): 2994-3013. doi: 10.3934/era.2023151
    [5] Qiuning Du, Fang Li . Some elementary properties of Laurent phenomenon algebras. Electronic Research Archive, 2022, 30(8): 3019-3041. doi: 10.3934/era.2022153
    [6] Waleed Mohamed Abd-Elhameed, Omar Mazen Alqubori, Abdulrahman Khalid Al-Harbi, Mohammed H. Alharbi, Ahmed Gamal Atta . Generalized third-kind Chebyshev tau approach for treating the time fractional cable problem. Electronic Research Archive, 2024, 32(11): 6200-6224. doi: 10.3934/era.2024288
    [7] Sida Lin, Dongyao Yang, Jinlong Yuan, Changzhi Wu, Tao Zhou, An Li, Chuanye Gu, Jun Xie, Kuikui Gao . A new computational method for sparse optimal control of cyber-physical systems with varying delay. Electronic Research Archive, 2024, 32(12): 6553-6577. doi: 10.3934/era.2024306
    [8] Xin Liu, Guang-Xin Huang . New error bounds for the tensor complementarity problem. Electronic Research Archive, 2022, 30(6): 2196-2204. doi: 10.3934/era.2022111
    [9] Abdelkader Lamamri, Mohammed Hachama . Approximate solution of the shortest path problem with resource constraints and applications to vehicle routing problems. Electronic Research Archive, 2023, 31(2): 615-632. doi: 10.3934/era.2023030
    [10] Mehmet Ali Özarslan, Ceren Ustaoğlu . Extended incomplete Riemann-Liouville fractional integral operators and related special functions. Electronic Research Archive, 2022, 30(5): 1723-1747. doi: 10.3934/era.2022087
  • In this paper we apply a smoothing technique for the maximum function, based on the compensated convex transforms, originally proposed by Zhang in [1] to construct some computable multiwell non-negative quasiconvex functions in the calculus of variations. Let KEMm×n with K a finite set in a linear subspace E without rank-one matrices of the space Mm×n of real m×n matrices. Our main aim is to construct computable quasiconvex lower bounds for the following two multiwell models with possibly uneven wells:

    i) Let f:KEE be an L-Lipschitz mapping with 0L1/α and H2(X)=min{|PEXAi|2+α|PEXf(Ai)|2+βi:i=1,2,,k}, where α>0 is a control parameter, and

    ii) H1(X)=α|PEX|2+min{|Ui(PEXAi)|2+γi:i=1,2,,k}, where AiE with Ui:EE invertible linear transforms for i=1,2,,k. If the control paramenter α>0 is sufficiently large, our quasiconvex lower bounds are 'tight' in the sense that near each 'well' the lower bound agrees with the original function, and our lower bound are of C1,1. We also consider generalisations of our constructions to other simple geometrical multiwell models and discuss the implications of our constructions to the corresponding variational problems.



    On the occasion of the 75th birthday of Professor E. N. Dancer

    In this paper we use the formula for the smooth approximations of the maximum function in Rn based on compensated convex transforms [1] (see Definition 2 below) to construct computable quasiconvex lower bounds with multiwell structure in the calculus of variations. By 'computable' we mean that the evaluation of the quasiconvex lower bounds only involves projection onto a linear subspace E without rank-one matrices in Mm×n, and a projection onto the simplex Δk where k2 is the number of 'wells'. The complexity of the former is O(mn), while the complexity of the latter is theoretically O(klogk) for the algorithm proposed in [2,3] or observed O(k) in the algorithm proposed in [4].

    The maximum function

    fm(x)=max{x1,,xm},x=(x1,x2,,xm)Rm (1.1)

    plays an important role in many optimization problems, in the sense that many problems can be proposed as minimization of the maximum function composed with other functions [5]. In [1] a smooth approximation for the maximum function was introduced by the upper compensated convex transform Cuλ(fm) for λ>0, where the formula for the transform is given by

    Cuλ(fm)(x)=λ|x|2λdist2(x,co(Δm2λ))+14λ,xRm, (1.2)

    where co(Δm/(2λ)) is the convex hull of the set of the scaled basis vectors Δm/(2λ)={e1/(2λ),e2/(2λ),,em/(2λ)} with ej=(δj,1,δj,1,,δj,m)Rm where δj,k=1 if j=k and δj,k=0 if jk for j=1,2,,m are Kronecker delta. Let Pco(Δm/(2λ))x be the projection of xRm to the simplex co(Δm/(2λ)), we can write

    Cuλ(fm)(x)=λ|x|2λ|xPco(Δm2λ)x|2+14λ (1.3)

    and the complexity for evaluatiing Cuλ(fm)(x) is the same as finding the convex projection Pco(Δm2λ)x. There are numerical schemes with complexity O(mlogm) to compute the convex projection above [2]. Our computable quasiconvex functions rely on this fact. Some generalisations of approximations by compensated convex transforms to maximum-like functions can be found in [6].

    It is known [1] that Cuλ(fm)(x) is a C1,1 approximation of fm(x) from above, that is, fm(x)Cuλ(fm)(x) and limλ+Cuλ(fm)(x)=fm(x) [1], and in general compensated convex trnasforms are 'tight' approximations. For the special case of the maximum function, we will show in Proposition 2 below that Cuλ(fm)(x) is a tight approximation in the sense that if the vector x=(x1,x2,,xm) is sorted with the order x1x2xm, then Cuλ(fm)(x)=fm(x) if x1x21/(2λ).

    For the minimum function gm(x)=min{x1,,xm}=max{x1,,xm} for xRm, it is easy to verify that the lower compensated convex transform of gm is given by

    Clλ(gm)(x)=Cu(fm)(x)=λdist2(x,co(Δm2λ))λ|x|214λ. (1.4)

    In this paper we apply the lower compensated convex transforms to two multiwell models in the calculus of variations. Compensated convex transforms have been applied to problems involving geometric singular extraction, shape analysis and approximation and interpolations of sampled data [7,8,9,10,11].

    The first type of multiwell functions is defined as follows. Let α>0 be a control parameter. Let EMm×n be a linear subspace without rank-one matrices and let f:KEE be an L-Lipschitz mapping with 0L1/α, where E is the orthogonal complement of E in Mm×n and K={Ai:i=1,2,,k}E is a finite set. We define the multiwell function H2(X) with quadratic growth by

    H2(X)=min{|PEXAi|2+α|PEXf(Ai)|2+βi:i=1,2,,k},XMm×n (1.5)

    for k2. If we define the quadratic function qi(X)=|PEXAi|2+α|PEXf(Ai)|2+βi for i=1,2,,k, then we may write

    H2(X)=min{qi(X):i=1,2,,k}.

    We see that the point Xi=Ai+f(Ai) is a local minimum point of H2() if for ji,

    qi(Ai+f(Ai))=βi<qj(Ai+f(Ai)=|AiAj|2+α|f(Ai)f(Aj)|2+βj.

    A sufficient condition for every Xi=Ai+f(Ai) to be a local minimum point is |AiAj|2>|βiβj| for 1ijk. So, under this additional restriction, H2(X) has a multiwell structure with each Xi=Ai+f(Ai) a local minimum point.

    The second type of multiwell functions is in the form

    H1(X)=α|PEX|2+min{|Ui(PEXAi)|2+γi:i=1,2,,k}, (1.6)

    where AiE and Ui:EE are invertible linear transforms and γi0 for i=1,2,,k. Again we can show that every AiE is a local minimum point of H1 if |Uj(AiAj)|2>|γiγj|.

    Our aim is to use the lower compensated convex transform Cuλ(fk) to construct composite functions that are quasiconvex lower bounds of H1 and H2. We will also consider simple generalisations of these constructions.

    Next we discuss backgrounds and motivations. The Quasiconvex function in the sense of Morrey [12] is an important tool in the vectorial calculus of variations [13], with applications in nonlinear elasticity [14] and material microstructure. For example, under the growth condition 0F(X)C|X|p+C1 for XMm×n, the variational integral I(u):=ΩF(Du)dx is weakly lower-semicontinuous in the Sobolev space W1,p(Ω,Rm) if and only if F is quasiconvex [15], where Mm×n is the linear space of real m×n matrices.

    A continuous function F:Mm×nR is quasiconvex [12] if for every XMm×n, for every open set ΩRn and for every ϕC0(Ω,Rm), we have

    ΩF(X+Dϕ(x))dxΩF(X). (1.7)

    Convex functions are quasiconvex but the converse is not true. A large class of functions, called polyconvex functions [14] are quasiconvex. Quasiconvex functions satisfying certain geometric conditions have been constructed by studying the quasiconvex envelope [13] of a given function with the multiwell property which describes the macroscopic properties of material microstructure for non-quasiconvex multiwell models [16,17]. A simplified multiwell model is in the form of an (Euclidean) p-distance function to a compact set K in Mm×n [18,19,20]. A typical example is the p-distance function distp(X,{I,I}) to the set {I,I} for XM2×2,1p<2, whose quasiconvex envelope Qdistp(X,{I,I}) vanishes exactly at I and IM2×2. So far a more explicit geometric lower bound of Qdistp(X,{I,I}) above is still not known. Some more explicit quasiconvex functions can be obtained by establishing lower bounds for the quasiconvex envelope via the so-called 'translation method' (see e.g., [21]). In particular, the explicit formula of the quasiconvex envelope for a double-well model [20] and the systematic study of restrictions of microstructure [22] lead to the study of linear subspaces of matrices without rank-one matrices, which are simple linear 'elliptic' objects in Mm×n. Here a linear subspace E is called simple linear elliptic if divPE(grad) is a linear elliptic operator satisfying the strong Legendre-Hadamard ellipticity condition for some constant coefficient.

    The following is a motivating example for our construction of computable quasiconvex lower bounds for a class of multiwell functions [1].

    Let EMm×n be a linear subspace without rank-one matrices, where m,n2. It is known [22] that there exists ε>0 such that

    |PEab|2ε|a|2|b|2,aRm,bRn,

    where PE:Mm×nE is the Euclidean orthogonal projection of Mm×n to the orthogonal complement E of E, aRm and bRn are treated as column vectors and ab=abTMm×n with bT the transpose of b.

    Now we optimise the above construction [1] by defining

    λ0=inf{|PEab|2:aRm,bRn,|a|=|b|=1},

    where |a| and |b| are the Euclidean norms of a and b in Rm and Rn respectively, so that we have 0<λ0<1 and

    |PEab|2λ0|ab|2=λ0|PEab|2+λ0|PEab|2

    hence if we define λE=λ0/(1λ0), we have

    |PEab|2λE|PEab|2 (1.8)

    for all aRm and bRn. Thus the quadratic form qE(X)=|PEX|2λE|PEX|2 is a rank-one convex quadratic form as qE(ab)0 for all aRm and bRn, hence is a quasiconvex function (see, e.g., [23]).

    Now we consider a simple multiwell model. Let EMm×n be a linear subspace without rank-one matrices and let KE be a closed set. Consider the (Euclidean) squared-distance function dist2(X,K)=dist2(PE(X),K)+|PEX|2, then (see [1])

    FλE(X)=co(dist2(PE(),K)+λE|PE()|2)(X)+qE(X)

    is a quasiconvex lower bound of the squared-distance function as co(dist2(PE(),K)+λE|PE()|2)(X) is a convex function and qE(X)=|PEX|2λE|PEX|2 is a rank-one convex quadratic form.

    However, in the above construction, even when KE is a finite set, in general, the computation of the convex envelope co(dist2(PE(),K)+λE|PE()|2)(X) is not a simple task whose complexity is generally not known. As a general reference, it is known that to determine the convexity of a quartic polynomial is an NP-hard question [24]. The computation of the values of quasiconvex functions whose existence are known is one of the motivations for us to find computable quasiconvex functions with multiwell structure. For numerical computation of the rank-one convex envelope for general functions, we refer to [25]. On the other hand, for numerical computation of compensated convex transforms in the discrete setting, there are efficient methods [26].

    Before we state our main results, let us first introduce some preliminary notions and results.

    Definition 1. (Quasiconvex functions [12,13,14]) Suppose f:Mm×nR is continuous. Then f is quasiconvex if

    Gf(X+Dϕ(x))dxGf(X)dx

    XMm×n, GRn open and ϕC0(G,Rm).

    For a continuous function f:Mm×nR bounded below, the quasiconvex envelope Qf:Mm×nR is the largest quesiconvex function satisfying Q(f)f. For the precise definition, we refer to [13]. In this paper we only consider quasiconvex lower bound g of a given function f.

    The translation method (see, e.g., [21]) is a simple and effective method for finding quasiconvex lower bounds. Suppose f:Mm×nR is continuous and bounded below and assume that g:Mm×nR is quasiconvex, then the function h defined by

    h(X):=co[fg](X)+g(X) (1.9)

    is quasiconvex and satisfies h(X)f(X) for all XMm×n. We call h a translation lower bound which is quasiconvex as the sum of a convex function and a quasiconvex function remains quasiconvex.

    The following example, which is a generalisation of a double-well function in [20] using a different method, can be used to show how the translation method applies to multiwell models when the wells are contained in subspaces without rank-one matrices [1]. This example is one of the motivations for the definition of compensated convex transforms [1].

    Example 1. Let EMm×n be a linear subspace without rank-one matrices where m,n2. Let KE be a non-empty closed set. $

    Consider the squared Euclidean distance function to the set K

    H(X)=dist2(X,K),XMm×n,

    which is not quasiconvex. We see that

    dist2(X,K)=|PEX|2+dist2(PEX,K).

    Let λE>0 be defined by

    λE=λ01λ0withλ0=inf{|PEab|2:aRm,bRn,|a|=|b|=1}. (1.10)

    Then the translation bound by the rank-one convex quadratic form qE(X):=|PEX|2λE|PEX|2 given by

    G(X):=co[dist2(PE,K)+λE|PE|2](X)+[|PEX|2λE|PEX|2] (1.11)

    is the quasiconvex envelope of dist2(X,K). Note that rank-one convex quadratic functions are quasiconvex [23]. For detailed calculations, we refer to [1].

    Next we briefly describe the main tool used in this paper: the compensated convex transforms.

    Motivated from the translation method, in particular, formula (1.9), compensated convex transforms were introduced in [1].

    Definition 2. Let f:RnR be a continuous function with at most quadratic growth and let λ>0 be large if needed, we define λ-parametrised convexity-based transforms.

    The Lower compensated convex transform is defined by

    Clλ(f)(x):=co[λ||2+f()](x)λ|x|2,xRn.

    The Upper compensated convex transform is defined by

    Cuλ(f)(x):=λ|x|2co[λ||2f()](x),xRn

    where co[g]= convex envelope of g, and || denotes the Euclidean norm.

    We also have Cuλ(f)(x)=Clλ(f)(x), and both Clλ(f) and Cuλ(f) are 'tight' approximations of f from below and above respectively as λ+ in the sense that if f is of C1,1 in a neiighbourhood of x0Rn, then when λ>0 is sufficiently large, we have f(x0)=Clλ(f)(x0)=Cuλ(f)(x).

    Next we briefly describe the properties of the smooth approximation of the finite maximum function fm(x)=max{x1,x2,,xm} for x=(x1,x2,,xm)Rm by the upper compensated convex transform given by [1]

    Cuλ(fm)(x)=λ|x|2λdist2(x,co(Δm2λ))+14λ.

    If we consider the minimum function gm(x)=min{xi:1i}=fm(x), then the lower compensated convex transform Clλ(gm)(x) satisfies

    Clλ(gm)(x)=Cuλ(fm)(x)=λdist2(x,Δm2λ)λ|x|214λ.

    It is known that there are numerical schemes to compute the convex projection PΔm/(2λ) with complexity O(mlogm), e.g., [2].

    The following are some properties of Cuλ(fm) and Clλ(gm). We will discuss their proofs in Section 5.

    Proposition 2. i) The function Cuλ(fm) is a C1,1 approximation of fm with linear growth and

    |DCuλ(fm)(x)DCuλ(fm)(y)|8λ|xy|.

    ii) Tight approximation: Assume x=(x1,x2,,xm) is 'sorted' in the increasing order: x1x2xm, then fm(x)=Cuλ(fm)(x) if and only ifx1x212λ.

    iii) Similar to ii), if x=(x1,x2,,xm) is 'sorted' in the decreasing order: x1x2xm, Then gm(x)=Clλ(gm)(x) iff x2x112λ.

    iv) Error estimate:

    fm(x)Cuλ(fm)(x)fm(x)+12λ.

    vi) Cuλ(fm) is a 'monotone' convex function [1]:

    if xy in the sense that xiyi for i=1,2,,m, then Cuλ(fm)(x)Cuλ(fm)(y).

    The plan for the rest of the paper is as follows. In Section 2 we consider the multiwell function H2 with quadratic growth. We give conditions on parameters involved to construct quasiconvex lower bounds H2 by Clλ(gm) so that the resulting quasiconvex functions are computable. In Section 3 we consider the multiwell function H1 with mixed linear-quadratic growth and take a slightly different approach from the method used for H2. We present a generalisation of H1 in Section 4 and conclude the paper with the proof of tight approximation of Cuλ(fm).

    In this section we construct quasiconvex lower bounds for the multiwell function H2 by the lower compensated convex transform Clλ(gk)() where gk(x)=min{x1,x2,,xk} is the minimum function defined on Rk.

    Let EMm×n be a linear subspace without rank-one matrices. Let K={A1,A2,,Ak}E and let f:KE be a Liptschitz function with a small Lipschitz constant L0.

    Recall the definition of H2:Mm×nR:

    H2(X)=min{|PE(XAi)|2+α|PEXf(Ai)|2+βi:i=1,2,,k},XMm×n. (2.1)

    In this section we give conditions on K, α>0, βi and L>0 so that the quasiconvex lower bound G2 defined below preserves the shape of H2 near the graph of f defined by Γf={(Ai+f(Ai):i=1,2,,k}, i.e., H2(X)=G2(X) if X is near Γf.

    We define

    qi(X):=|PE(XAi)|2+α|PEXf(Ai)|2+βi, (2.2)

    so that

    H2(X)=q(X)+min{i(X):i=1,2,,k},XMm×n, (2.3)

    where

    q(X):=|PE(X)|2+α|PE(X)|22A1PE(X)2αf(A1)PE(X),i(X)=2(AiA1)PE(X)2α(f(Ai)f(A1))PE(X)+ci, (2.4)

    with ci=|Ai|2+α|f(Ai))|2+βi. Note that 1(X)=c1 is a constant.

    We also define

    Lk(X)=(1(X),,k(X)). (2.5)

    The following is the main result of this section.

    Theorem 3. Suppose EMm×n is a linear subspace without rank-one matrices with the ellipticity constant λE>0 defined by (1.10). Let

    D=maxij|AiAj|,d=minij|AiAj|,β=maxij|βiβj|

    Suppose f:KE is a Lipschitz mapping with Lipschitz constant 0L1/α. If d2>2β, α>1,

    1+λEα>32(1+λE)(k1)D2d22βandλ=1+λEα8(1+λE)(k1)D2. (2.6)

    Then

    G2(X):=q(X)Cuλ(fk(Lk(X))),XMm×n (2.7)

    is a quasiconvex lower bound of H2(X), where fk is the maximum function in Rk.

    AlsoG2(X)=H2(X) if

    |PEXAi|2+|PEXf(Ai)|214λ,i=1,2,,k. (2.8)

    Remark 4. The assumption d2>2β implies that every 'energy well' Xi:=Ai+f(Ai) for i=1,2,,k is a local minimum point of H2.

    Also if X is close to the 'well' Xi, the quasiconvex lower bound G2(X) agrees with H2(X).

    The computation of G2(X) requires the computation of Cuλ(fk(Lk(X)) which has the complexity of O(klogk).

    Proof of Theorem 3 By the the formula for the upper transform in the definition of G2(X), we have

    G2(X):=[q(X)λ|Lk(X)|2]+(λdist2(Lk(X),co(Δk2λ))14λ).

    Since the function λdist2(Lk(X),co(Δk2λ))14λ is convex as Lk(X) is an affine mapping, hence this function is convex. So, we need to show:

    i) the quadratic function q(X)λ|Lk(X)|2 is a rank one convex quadratic function, which implies that G2(X) is quasiconvex.

    ii) near Ai+f(Ai) we have j(X)i(X)+1/(2λ) so that Cuλ(fk)(Lk(X))=i(X) for all ji hence G2(X)=H2(X).

    Proof of i) Let τk=8(k1)D2. We may write

    q(X)λ|Lk(X)|2=q(X)λ|Lk(X)|2=|PEX|2+α|PEX|24λki=2[((AiA1)PEX)2+α2((f(Ai)f(A1))PEX)2]8λαki=2((AiA1)PEX)((f(Ai)f(A1))PEX)+affine terms(1τkλ)|PEX|2+(ατkλ)|PEX|2+affine terms,whereτk=8(k1)D2.

    We observe that

    ατkλ=α11+λE>0

    and

    1τkλ=λE(1α)1+λE<0.

    Now let XMm×n be a rank-one matrix, we have

    (1τkλ)|PEX|2+(ατkλ)|PEX|2(λEα+1(1+λE)τkλ)|PEX|2=0.

    Thus q(X)λ|Lk(X)|2 is a rank-one convex quadratic function, hence is quasiconvex. Therefore G2(X) is quasiconvex on Mm×n.

    Proof of ii) We need to show that for each i, if (2.8) holds, i.e.,

    |PEXAi|2+|PEXf(Ai)|214λ,

    then for each ji, we have j(X)i(X)+12λ. Equivalently we need to prove that

    q(X)+i(X)+12λq(X)+j(X)qi(X)+12λqj(X).

    We have

    qj(X)qi(X)12λ=|AjAi|2+α|f(Aj)f(Ai)|22(PEXAi)(AjAi)2α(PEXf(Ai))(f(Aj)f(Ai))+βjβi12λ|AjAi|22|AjAi||PEXAi|2α|f(Aj)f(Ai)||PEXf(Ai)|β12λ|AjAi|22λ|AjAi|β12λ.

    Next we show that

    |AjAi|22λ|AjAi|β12λ>0 (2.9)

    if (2.6) and (2.8) hold.

    Let x1<0<x2 be the two roots of the quadratic polynomial x22λxβ12λ. It remains to show that |AjAi|>x2. From (2.6) we see that

    λ>4d22β,

    so

    x2=12λ+β+1λ<12d22β4+d2+2β4<12d22β+12d2+2β<d<|AjAi|.

    Note that by our assumption on f, we have |f(Aj)f(Ai)||AiAj|/α. Thus by Property iii) in Proposition 2, we have Cuλ(fk(Lk(X))=Clλ(gk(Lk(X))=gk(Lk(X))=i(X) as j(X)i(X)+12λ, hence H2(X)=G2(X).

    Remark 5. Due to Property iv) of the upper compensated convex transform in Propostion 2, we have the error estimate

    G2(X)H2(X)G2(X)+12λ,XMm×n.

    Therefore if α>0 is large, then G2(X) is a very good quasiconvex lapproximation of H2(X).

    The structure of the quasiconvex lower bound G2() of the multiwell function H2() suggests that G2() is of C1,1(Mm×n) and is of quadratic growth for XMm×n. Therefore when we consider the variational integral

    I2(u)=ΩG2(Du(x))dx

    for minimisers or more general stationary points [27] the natural space would be W1,2(Ω,Rm). This is in contrast with quasiconvex lower bounds G1 for H1 which will be discussed in the next section, where we will see that G1 has a mixed growth which might lead to some more challenges for us to choose a proper function space to accommodate such energy density (integrands).

    The following figure (Figure 1) gives an illustration of an example of a three well model H2 restricted to ER2, where f0 :

    H2(x,y)=min{(x+2)2+y2+1,x2+9(y+2)21,(x2)2+(y2)2}
    Figure 1.  Views of H2(x,y) from two angles and a quasiconvex lower bound (λ=0.25).

    with the three wells at (2,0),(0,2),(2,2), with different heights and a quasiconvex lower bound (λ=0.25) restricted on E.

    In this section we construct quasiconvex lower bounds for the multiwell function H1 defined in (1.6) by taking the lower compensated convex transform Clλ(gk)()=Cuλ(fk)(), where gk(x) and fk(x) are the minimum and maximum functions defined on Rk respectively.

    Similar to what we have defined in Section 2, let EMm×n be a linear subspace without rank-one matrices and let K={A1,A2,,Ak}E be a finite set with k distinct elements.

    We recall the definition of H1:Mm×nR:

    H1(X)=α|PEX|2+min{hi(PEX):i=1,2,,k},XMm×n, (3.1)

    where

    hi(Y)=|Ui(YAi)|2+γi,,YE

    with Ui:EE an invertible linear transform and γi0 for i=1,2,,k. We also define

    qi(X)=α|PEX|2+hi(PEX),XMm×n

    for i=1,2,,k.

    Let Fk(Y)=(h1(Y),h2(Y),,hk(Y)) for YE. We define for λ>0.

    G1(X)=α|PEX|2Cuλ(fk)(Fk(PEX))=λ|dist2(Fk(PEX),Δk2λ)+[α|PEX|2λ|Fk(PEX)|2].

    Note that α|PEX|2λ|Fk(PEX)|2 is a quadratic function.

    Let d=min1ijk|AiAj|, γ=max1ikγi, and umax=max1ikUiop, where Uiop is the operator norm of Ui. Let uj=inf{|UjY|:YE,|Y|=1} for j=1,2,,k. Since Uj:EE is an invertible linear transform for j=1,2,,k, we see that uj>0 for all j=1,2,,k. We define umin=min{uj:j=1,2,,k}. Clearly, umin>0. If γ is small enough and α>0 is sufficiently large, we can establish a similar result for H1 as we have done for H2.

    Theorem 6. Suppose umind>2γ. Let λE>0 be defined by (1.10). If

    λ:=λEαku2max>(1+2)u2max+1/2umind2γ,

    then

    G1(X)=α|PEX|2Cuλ(fk)(Fk(PEX))

    is a quasiconvex lower bound of H1(X) for XMm×n and G1(X)=H1(X) if |PEXAi|1/λ for i=1,2,,k.

    The condition for α is a rough sufficient condition as we have just used the fact

    a+b2a+ba+b,fora,b0

    in our estimates for gj(PEX)gi(PEX)+1/(2λ) for ji.

    The condition umind>2γ is essentially a sufficient geometric assumption which means the minimum distance among the wells |AjAi| is larger that the maximum height γ so that any given well Ai is a genuine 'well' with Ai a local minimum point of the energy density G1(X) not swallowed by other wells.

    Proof of Theorem 6 Observe that by the formula for the upper transform of the maximum function we have

    G1(X)=[α|PEX|2λ|Fk(PEX)|2]+λdist2(Fk(PEX),co(Δk2λ))14λ.

    We need to show that α|PEX|2λ|Fk(PEX)|2 is a rank-one convex quadratic function and

    Xdist2(Fk(PEX),co(Δk2λ))

    is convex, that is, for YE,

    Ydist2(Fk(Y),Δk2λ)

    is convex.

    Note that hj(Y)=|Uj(YAj)|2+γ2j is a non-negative convex function for YE and for j=1,2,,k. We define, for uRk, the function

    f(u)=dist2(u,Δk2λ).

    Clearly f(u) is convex. The key observation is that f(u) is also positively increasing in the sense that if uv0, i.e., uivi0 for all of the corresponding components, then f(u)f(v). This requires some more detailed structural properties of the convex projection PΔk/(2λ). We have

    Lemma 7. Let

    f(u)=dist2(u,Δm2λ),uRm.

    If u=(u1,,um),h=(h1,,hm)Rm satisfy that u0 and h0 componentwise in the sense that ui0 and hi0 for i=1,2,,m, then Df(u)h0. Consequently, f(u)f(v) if uv0 componentwise.

    Proof We have

    f(u)=dist2(u,Δm2λ)=|uPΔm2λ(u)|2=|u+PΔm2λ(u)|2,

    where the convex projection is in the form

    PΔm2λ(u)=mj=1λj2λej

    with λj0 for j=1,2,,m and mj=1λj=1. If u0 and h0 componentwise, then

    Df(u)h=Ddist2(u,Δm2λ)h=2(u+PΔm2λ(u))h=2mj=1ujhj+2mi=1λi2λeih=2mj=1ujhj+2mi=1λi2λhi0

    as uihi0, λi0 and eih=hi0 for i=1,2,,m with e1,e2,,em the standard Euclidean basis vectors.

    The last claim follows from the fundamental theorem of calculus that if uv0, i.e., uivi0 for i=1,2,,m, then f(u)f(v)=10Df(tu+(1t)v)(uv)dt0 as tu+(1t)v0 and uv0 componentwise.

    Proof of Theorem 6 (continued) Now we can show that

    V1(X):=dist2(Fk(PEX),co(Δk2λ))=f(Fk(PEX)),XMm×n

    is convex, where f:RkR is defined as in Lemma 7 with m=k. Let 0<t<1 and X,YMm×n, since Fk(PEX)=(h1(PE(X)),h2(PE(X)),,hk(PE(X))) and every component function hj(PE(X))0 is convex, we first have

    hj(PE(tX+(1t)Y))=hj(tPEX+(1t)PEY))thj(PEX)+(1t)hj(PEY),j=1,2,,k

    hence Fk(PE(tX+(1t)Y))tFk(PEX)+(1t)Fk(PEY) componentwise.

    Since f is positively increasing, we have

    V1(tX+(1t)Y)=f(Fk(PE(tX+(1t)Y)))f(tFk(PEX)+(1t)Fk(PEY)).

    Also f is convex, which implies

    f(tFk(PEX)+(1t)Fk(PEY))tf(Fk(PEX))+(1t)f(Fk(PEY))=tV1(X)+(1t)V1(Y).

    Therefore V1(X) is a convex function in Mm×n.

    Next we show that the quadratic function Q1(X)=α|PEX|2λ|Fk(PEX)|2 is a rank-one convex quadratic function in Mm×n. We have for every XMm×n, we have

    Q1(X)=α|PEX|2λkj=1h2j(PEX)=α|PEX|2λkj=1(|Ui(PEXAi)|2+γi)α|PEX|2|(λkj=1(|Ui(PEX)|2)+L(PEX),

    where

    L(PEX)=λkj=1(2Ui(PEX)Ui(Ai)+|Ui(Ai)|2+γj)

    is an affine function of PEX hence is a convex function. So, we only need to show that the quadratic form

    α|PEX|2λkj=1|Ui(PEX)|2

    is rank-one convex. We have, for every rank-one matrix XMm×n,

    α|PEX|2|(λkj=1|Ui(PEX)|2αλE|PEX|2|λu2maxk|PEX|2=(αλEλu2maxk)|PEX|2=0,

    as αλE=λu2k.

    Next we prove that G1(X)=H1(X) if |PEXAi|1/λ for each i=1,2,,k. By definition of the upper transform, G1(X)=H1(X) if and only if Cuλ(fk)(Fk(PEX))=min{h1(PEX),h2(PEX),,hk(PEX)}. By Property iii) in Proposition 2, if |PEXAi|1/λ, we show that hj(PEX)hi(PEX)+1/(2λ) for all j=1,2,,k with ji so that Cuλ(fk)(Fk(PEX))=hi(PEX) hence G1(X)=H1(X). If we write PEX=Ai+YE, the assumption |PEXAi|1/λ implies |Y|1/λ. We have

    hj(PEX)hi(PEX)+12λ|Uj(PEXAj)|2+γj|Ui(PEXAi)|2+γi+12λ|Uj(PEXAj)|+γj2|Ui(PEXAi)|+γi+12λ|Uj(AiAj+Y)|+γj2|Ui(Y)|+γi+12λ|Uj(AiAj)||Uj(Y)|2|Ui(Y)|+γi+12λumin|AiAj|2|Uj(Y)|2+|Ui(Y)|+γ+12λumind2(umax|Y|2+umax|Y|+γ+12λ)umind2γ(1+2)umaxλ+12λas|Y|1/λλ(1+2)umax+1/2umind2γ.

    The last inequality holds as we have assumed that umind>2γ and λ=λEαku2max is large if α>0 is large.

    Remark 8. In Theorem 6, if we consider the special case that Ui=I - the identity transform, then we have umax=umin=1 and in this special case the assumptions will be much simpler as we can assume that

    λ=λEαk>1+2+1/2d2γ.

    Remark 9. We may generalise Theorem 6 to deal with more complicated multiwell models. Even for a single non-elliptic well model in the form

    H1(X)=α|PEX|2+min{|UiPEX|:i=1,2,,k},XMm×n, (3.2)

    where Ui:EE is an invertible linear transform, we see that under the assumption that

    λ=λEαk, (3.3)

    we can show that the corresponding lower bound G1(X) is still a quasiconvex lower bound. However, at X=0 we have H1(0)=0 but G1(0)<0. This is due to the fact that

    Cuλ(fk(Fk(0))=λdist2(0,Δk/(2λ))14λ=λ|PΔk/(2λ)(0)|214λ=λ|kj=1ej2kλ|214λ=k14kλ<0.

    Here we have used the fact that the distance between 0 and the simplex Δk/(2λ) is attained at the center of the simplex kj=1ej2kλ.

    At any point where |UjPEX||UiPEX|+1/(2λ) for all ji, we still have H1(X)=G1(X). Thus if we just wish to construct a quasiconvex function with the 'desired' geometric feature which is differentiable except at 0, then we can make a simple lift by considering G1(X)+(k1)/(4kλ).

    Remark 10. From Theorem 6 we see that both the multiwell function H1 and its quasiconvex lower bound G1 are of mixed growth. In the subspace EMm×n both H1 and G1 are of quadratic growth. In the subspace E, both H1 and G1 are of linear growth.

    If the height γj>0 for all j=1,2,,k, then we see that G1 is at least of C1,1loc(Mm×n). However, if for some j, γj=0, then G1 is not differentiable at Aj.

    For the variational integral I(u)=ΩG1(Du)dx, if we consider the Dirichlet problem, say u=0 on Ω, the natural space is W1,20(Ω,Rm) where ΩRn is, say, a bounded Lipchitz domain. The main reason is that I(u) is coercive in W1,20(Ω,Rm) because we have

    Ω|PEDϕ(x)|2dxλEΩ|PEDϕ(x)|2dx

    for ϕW1,20(Ω,Rm). As G1(PEX)c0 for some c0R we have both ΩG1(Du)dxα|PEDu|2dx+c0meas(Ω) and ΩG1(Du)dxαλE|PEDu|2dx+c0meas(Ω).

    If we consider the natural boundary value problem for I(u)=ΩG1(Du)dx under the constraint, say Ωu(x)dx=0, then W1,2(Ω,Rm) does not seem to be the right space to study such a variational integral as I(u) is not coercive in this space. We have α|PEX|2Cuλ(fk)(Fk(PEX)) and it can be verified that there are c0>0, c1>0, C0>0 and C1>0 such that c0|PEX|c1Cuλ(fk)(Fk(PEX))C0|PEX|+C1. However, the integral Ω|PEDu|2dx does not contribute to the coercivity of I(u) in the subspace E. An example is in [28] that if we consider the two dimensional conformal subspace EM2×2 which does not have rank-one matrices, then E=Eˉ is the two-dimensional anti-conformal subspace. If u=u1+iu2 is a holomophic function in Ω and we define v:=(u1,u2):ΩR2, then we have |PEˉDv|2=|ˉu|2=0 by the Cauchy-Riemann equations. Therefore Ω|PEˉDv|2dx does not contribute to the coercivity of I(v) under the condition Ωvdx=0.

    It seems that some spaces with mixed growth condition might be the correct spaces to accommodate such variational integrals. Furthermore the study of (partial) regularity of minimisers and more general critical points for I(u) under the natural boundary condition seems to be a challenging question.

    The following figure (Figure 2) gives and illustration of an example of a three well model H1 restricted to ER2 in the form

    min{9(x+2)2+y2+1,x2+9(y+2)2+0.52,(x2)2+(y2)2}
    Figure 2.  H1(x,y), lower bound G1(x,y) with λ=1/2, and their sublevel set at 1.8211.

    with the three 'anisotropic' wells and with different heights at (2,0),(0,2),(2,2) (λ=1/2).

    The following figure (Figure 3) gives and illustration of an example of a single non-elliptic well model H1 restricted to ER2 in the form

    min{x2/100+y2,(xcos(π/3)+sin(π/3)2/100+(xsin(π/3)ycos(pπ/3))2,(xcos(2π/3)+sin(2π/3)2/100+(xsin(2π/3)ycos(2π/3))2}
    Figure 3.  H1(x,y), lower bound G1(x,y) with λ=0.25, and their difference H1(x,y)G1(x,y).

    with the three 'anisotropic' wells and with different heights at (2,0),(0,2),(2,2) (λ=1/2).

    Suppose dim(E)=s. Using an orthonormal basis of E given by B1,B2,,Bs and let xi=PEXBi and x=(x1,x2,,xs)Rs, we define Li(x)=Aixbi for yRs with AiMmi×s and biRmi with 1mi<s and i=1,2,k.

    We may consider the following more general model

    H1(X)=α|PEX|2+V0(PEX),

    where

    V0(PE(X)=min{|Aixbi|:i=1,2,,k},xRd,

    or more generally,

    Vγ(x)=min{|Aixbi|2+γ2i:i=1,2,,k},xRd,

    where |Aixbi| is the Euclidean norm in Rmi and γi0. For the function V0, the zero set of can be the union of finitely many planes.

    Let Fk(x)=(|A1xb1|,,|Akxbk|). We can approximate H1(x) from below by G1(x)=α|PEX|2Cuλ(fk)(Fk(x)), with x=(PEXB1,.(PEXB1,,(PEXBk).

    We can use G1 to define quasiconvex functions as before. However due to the special feature of Cuλ(fk)(x), at intersections of the planes defined by the zero set of V0(x), we see that Cuλ(fk)(Fk(x))<0 at points of intersections. As we commented earlier, this is due to the fact that if x1=x2x3xm, then by Lemma 12 in Section 5 and the fact that Clλ(g2(x1,x2))=Cuλ(f2(x1,x2)) in R2, we have

    Clλ(gm)(x)=Cuλ(fm(x))Cuλ(f2(x1,x2))=x118λ=gm(x)18λ.

    If x1=x2=0, we see that

    14λCuλ(fm(x))18λ<0.

    Example 11. Let H1(x,y)=2min{|x|,|y|,(x1)2+(y1)2} and let G1(x,y)=Cuλ(f3)(F3(x,y)) $

    with F3(x,y)=(|x|,|y|,(x1)2+(y1)2)R3. The graphs of H1(x,y) and G1(x,y) with λ=1/2 are shown in Figure 4.

    Figure 4.  Views of H1, G1 and the difference H1G1 in the 2d subspace E (λ=0.5).

    In this section we prove the tight approximation of the upper compensated convex transform of the maximum function (Proposition 2 (ii)) . Recall from [1] that

    Cuλ(fm)(x)=λ|x|2λdist2(x,co(Δm2λ))+14λ

    for xRm and λ>0. Items i) and iv) of Proposition 2 were established in [1] with the maximum function as a special case. The fact that The function Cuλ(fm) is a C1,1 approximation of fm and satisfies

    |DCuλ(fm)(x)DCuλ(fm)(y)|8λ|xy|,

    for x,yRm is a consequence of a general result in [1] (Theorem 4.1). The error estimate in iv) that

    fm(x)Cuλ(fm)(x)fm(x)+12λ

    for xRm was established in [1] (Theorem 5.1) which also contains the estimate |DCuλ(fm)(x)|1 for xRm. Thus the statement in i) that Cuλ(fm) is of linear growth is a direct consequence of this gradient estimate.

    Before we prove Proposition 2 (ii) we state the following simple lemma which can be verified through a direct calculation using the definition of the upper compensated convex transform.

    Lemma 12. Let f2(x)=max{x1,x2} for x=(x1,x2)R2. We have

    Cuλ(f2)(x)=λ|x|2λdist2(x,Δ2/(2λ))+14λ={x1+x22+λ2(x1x2)2+18λ,|x1x2|12λ,f2(x),|x1x2|12λ. (4.1)

    Next we prove ii), the 'Tight approximation' property: Assume x=(x1,x2,,xm) is 'sorted' in the increasing order: x1x2xm, then fm(x)=Cuλ(fm)(x)=x1 if and only if x1x212λ. Note that Item iii) for the lower transform of the minimum function can be proved by using the identity Clλ(gm)(x)=Cuλ(fm(x)).

    Proof of Proposition 2 Item ii) If x(0)1x(0)21/(2λ), we see that fm is differentiable at x(0). By the translation invariance property of compensated convex transforms [7], we have,

    Cuλ(fm)(x(0))=co(λ|x(0)|2fm())(x(0)).

    The conclusion in this case follows if we can show that co(λ|x(0)|2fm())(x(0))=fm(x(0)). To prove this we only need to show that the tangent plane for fm defined by (x)=fm(x(0))fm(x(0))(xx(0)) lies below the graph of the function λ|xx(0)|2fm(x) for all xRm, that is,

    fm(x(0))fm(x(0))(xx(0))λ|xx(0)|2fm(x),xRm. (4.2)

    Since fm(x)=x1 for x near x(0), we have fm(x(0))=x(0)1 and fm(x(0))=e1, hence (4.2) is equivalent to

    x(0)1(x1x(0)1)λ|xx(0)|2xk, (4.3)

    where xk=fm(x) for some k{1,2,,m} by definition. If k=1, then inequality (4.3) clearly holds. If k1, we have, as (4.3) is equivalent to

    A:=λ|xx(0)|2(xkx(0)k)+(x1x(0)1)+(x(0)1x(0)k)0. (4.4)

    If we complete squares in A defined in (4.4) above, we obtain

    A=mjk,1λ(xjx(0)j)2+λ(xkx(0)k12λ)2+λ(x1x(0)1+12λ)2+(x(0)1x(0)k)12λ.

    Therefore A0 if and only if

    x(0)1x(0)k12λ0.

    Since k1 we have

    x(0)1x(0)kx(0)1x(0)212λ.

    The conclusion follows.

    The work of the first author is partially supported by NSFC grants 11801200 and 12141107. We wish to thank the anonymous referees for their helpful comments and suggestions which have greatly improved the presentation of this paper.

    The authors declare there is no conflicts of interest.



    [1] K. Zhang, Compensated convexity and its applications, Ann. Inst. H. Poincaré Anal. Non Linéaire, 25 (2008), 743–771. https://doi.org/10.1016/j.anihpc.2007.08.001 doi: 10.1016/j.anihpc.2007.08.001
    [2] Y. Chen, X. Ye, Projection onto a simplex, arXiv preprint, (2011), arXiv: 1101.6081.
    [3] C. W. Combettes, S. Pokutta, Complexity of linear minimization and projection on some sets, Oper. Res. Lett., 49 (2021), 565–571. https://doi.org/10.1016/j.orl.2021.06.005 doi: 10.1016/j.orl.2021.06.005
    [4] L. Condat, Fast projection onto the simplex and the l1 ball, Math. Program., 158 (2016), 575–585. https://doi.org/10.1007/s10107-015-0946-6 doi: 10.1007/s10107-015-0946-6
    [5] E. Y. Pee, J. O. Royset, On solving large-scale finite minimax problems using exponential smoothing, J. Optim. Theory Appl., 148 (2011), 390–421. https://doi.org/10.1023/B:JOTA.0000006685.60019.3e doi: 10.1023/B:JOTA.0000006685.60019.3e
    [6] K. Zhang, Convex analysis based smooth approximations of maximum functions and squared distance functions, J. Nonlinear Convex Anal., 9 (2008), 379–406.
    [7] K. Zhang, A. Orlando, E. Crooks, Compensated convexity and Hausdorff stable geometric singularity extractions, Math. Models Methods Appl. Sci., 25 (2015), 747–801. https://doi.org/10.1142/S0218202515500189 doi: 10.1142/S0218202515500189
    [8] K. Zhang, A. Orlando, E. Crooks, Compensated convexity and Hausdorff stable extraction of intersections for smooth manifolds, Math. Models Methods Appl. Sci., 25 (2015), 839–873. https://doi.org/10.1142/S0218202515500207 doi: 10.1142/S0218202515500207
    [9] K. Zhang, E. Crooks, A. Orlando, Compensated convexity, multiscale medial axis maps and sharp regularity of the squared distance function, SIAM J. Math. Anal., 47 (2015), 4289–4331. https//doi.org/10.1137/140993223 doi: 10.1137/140993223
    [10] K. Zhang, E. Crooks, A. Orlando, Compensated convexity methods for approximations and interpolations of sampled functions in Euclidean spaces: theoretical foundations, SIAM J. Math. Anal., 48 (2016), 4126–4154. https://doi.org/10.1137/15M1045673 doi: 10.1137/15M1045673
    [11] K. Zhang, E. Crooks, A. Orlando, Compensated convexity methods for approximations and interpolations of sample functions in Euclidean spaces: Applications to sparse data, contour lines and inpainting, SIAM J. Imaging Sci., 11 (2018), 2368–2428. https://doi.org/10.1137/17M116152X doi: 10.1137/17M116152X
    [12] C. B. Morrey Jr, Multiple Integrals in the Calculus of Variations, Springer-Verlag, New York, 1966. https://doi.org/10.1007/978-3-540-69952-1
    [13] B. Dacorogna, Direct Methods in the Calculus of Variations, 2nd edition, Springer-Verlag, New York, 1989. https://doi.org/10.1007/978-0-387-55249-1
    [14] J. M. Ball, Convexity conditions and existence theorems in nonlinear elasticity, Arch. Rational Mech. Anal., 63 (1977), 337–403. https://doi.org/10.1007/BF00279992 doi: 10.1007/BF00279992
    [15] E. Acerbi, N. Fusco, Semicontinuity problems in the calculus of variations, Arch. Rational Mech. Anal., 86 (1984), 125–145. https://doi.org/10.1007/BF00275731 doi: 10.1007/BF00275731
    [16] J. M. Ball, R. D. James, Fine phase mixtures as minimizers of energy, Arch. Rational Mech. Anal., 100 (1987), 13–52. https://doi.org/10.1007/BF00281246 doi: 10.1007/BF00281246
    [17] J. M. Ball, R. D. James, Proposed experimental tests of a theory of fine microstructures and the two-well problem, Philos. Trans. R. Soc. A, 338 (1992), 389–450. https://doi.org/10.1098/rsta.1992.0013 doi: 10.1098/rsta.1992.0013
    [18] V. Šverák, Quasiconvex functions with subquadratic growth, Proc. Royal Soc. London Ser. A, 433 (1991), 723–F725. https://doi.org/10.1098/rspa.1991.0073 doi: 10.1098/rspa.1991.0073
    [19] K. Zhang, A construction of quasiconvex functions with linear growth at infinity, Ann. Scuola Norm. Sup. Pisa Cl. Sci., 19 (1992), 313–326.
    [20] R. V. Kohn, The relaxation of a double-well energy, Cont. Mech. Therm., 3 (1991), 981–1000. https://doi.org/10.1007/BF01135336 doi: 10.1007/BF01135336
    [21] N. B. Firoozye, Optimal use of the translation method and relaxations of variational problems, Comm. Pure Appl. Math., 44 (1991), 643–678. https://doi.org/10.1002/cpa.3160440603 doi: 10.1002/cpa.3160440603
    [22] K. Bhattacharya, N. B. Firoozye, R. D. James, R.V. Kohn, Restrictions on microstructures, Proc. Royal Soc. Edinb. - A, 124 (1994), 843–878. https://doi.org/10.1017/S0308210500022381 doi: 10.1017/S0308210500022381
    [23] M. Giaquinta, Introduction to regularity theory for nonlinear elliptic systems, Lectures in Mathematics ETH Zürich, Birkhüser Verlag, Basel, 1993. https://doi.org/10.1007/978-88-7642-443-4
    [24] A. A. Ahmadi, A. Olshevsky, P. Parrilo, J. N. Tsitsiklis, NP-hardness of deciding convexity of quartic polynomials and related problems, Math. Program., 137 (2013), 453–476. https://doi.org/10.1007/s10107-011-0499-2 doi: 10.1007/s10107-011-0499-2
    [25] G. Dolzmann, Numerical computation of rank-one convex envelopes, SIAM J. Numer. Anal., 36 (1999), 1621–1635. https://doi.org/10.1137/S0036142997325581 doi: 10.1137/S0036142997325581
    [26] K. Zhang, E. Crooks, A. Orlando, Compensated convexity on bounded domains, mixed Moreau envelopes and computational methods, Appl. Math. Model., 94 (2021), 688–720. https://doi.org/10.1016/j.apm.2021.01.040 doi: 10.1016/j.apm.2021.01.040
    [27] K. Zhang, Mountain pass solutions for a double-well energy, J. Differ. Equ., 182 (2002), 490–510. https://doi.org/10.1006/jdeq.2001.4113 doi: 10.1006/jdeq.2001.4113
    [28] K. Zhang, An elementary derivation of the generalized Kohn-Strang relaxation formulae, J. Convex Anal., 9 (2002), 269–285.
  • 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(1822) PDF downloads(66) Cited by(0)

Figures and Tables

Figures(4)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog