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

On the extreme rays of the cone of 3×33\times 3 quasiconvex quadratic forms: Extremal determinants vs extremal and polyconvex forms

Davit Harutyunyan and Narek Hovsepyan University of California Santa Barbara, harutyunyan@math.ucsb.eduTemple University, narek.hovsepyan@temple.edu
Abstrakt

This work is concerned with the study of the extreme rays of the convex cone of 3×33\times 3 quasiconvex quadratic forms (denoted by 𝒞3{\cal C}_{3}). We characterize quadratic forms f𝒞3,f\in{\cal C}_{3}, the determinant of the acoustic tensor of which is an extremal polynomial, and conjecture/discuss about other cases. We prove that in the case when the determinant of the acoustic tensor of a form f𝒞3f\in{\cal C}_{3} is an extremal polynomial other than a perfect square, then the form must itself be an extreme ray of 𝒞3;{\cal C}_{3}; when the determinant is a perfect square, then the form is either an extreme ray of 𝒞3{\cal C}_{3} or polyconvex; and finally, when the determinant is identically zero, then the form ff must be polyconvex. The zero determinant case plays an important role in the proofs of the other two cases. We also make a conjecture on the extreme rays of 𝒞3,{\cal C}_{3}, and discuss about weak and strong extremals of 𝒞d{\cal C}_{d} for d3,d\geq 3, where it turns out that several properties of 𝒞3{\cal C}_{3} do not hold for 𝒞d{\cal C}_{d} for d>3,d>3, and thus case d=3d=3 is special. These results recover all previously known results (to our best knowledge) on examples of extreme points of 𝒞3{\cal C}_{3} that were proved to be such. Our results also improve the ones proven by the first author and Milton [Comm. Pure Appl. Math., Vol. 70, Iss. 11, Nov. 2017, pp. 2164-2190] on weak extremals in 𝒞3{\cal C}_{3} (or extremals in the sense of Milton) introduced in [Comm. Pure Appl. Math., Vol. XLIII, 63-125 (1990)].

In the language of positive biquadratic forms, quasiconvex quadratic forms correspond to nonnegative biquadratic forms and the results read as follows: If the determinant of the 𝒚\bm{y} (or 𝒙\bm{x}) matrix of a 3×33\times 3 nonnegative biquadratic form in 𝒙,𝒚3\bm{x},\bm{y}\in\mathbb{R}^{3} is an extremal polynomial that is not a perfect square, then the form must be an extreme ray of the convex cone of 3×33\times 3 nonnegative biquadratic forms (𝒞3);({\cal C}_{3}); if the determinant is identically zero, then the form must be a sum of squares; if the determinant is a nonzero perfect square, then the form is either an extreme ray of 𝒞3,{\cal C}_{3}, or is a sum of squares.

The proofs are all established by means of several classical results from linear algebra, convex analysis (geometry), real algebraic geometry, and the calculus of variations.

Keywords:   Quasiconvex quadratic forms, positive biquadratic forms, sums of squares, polyconvexity, rank-one convexity.

Mathematics Subject Classification:  12D15, 12E10, 15A63, 49J40, 70G75, 74B05, 74B20,

1 Introduction

Let us point out from the onset that as we are applied mathematicians, the paper is written in the applied mathematics/calculus of variations language. However, the subject is in the intersection of the fields of applied mathematics/calculus of variations and real algebraic geometry/convex geometry, thus we have drawn some appropriate links between those two fields of mathematics in terms of language and results that we can understand.
Quasicovex quadratic forms and sums of squares: From applied mathematics to real algebraic geometry. Quasiconvexity is a central subject in the calculus of variations and in applied mathematics. It was introduced by Morrey in 1952 [Reference,Reference] and has several equivalent definitions, among which the simplest looking one is as follows [Reference]: Let n,N,n,N\in\mathbb{N}, and let the function f:N×nf\colon\mathbb{R}^{N\times n}\to\mathbb{R} be Borel measurable and locally bounded. Then ff is said to be quasiconvex, if

f(𝝃)[0,1]nf(𝝃+φ(x))𝑑x,f(\bm{\xi})\leq\int_{[0,1]^{n}}f(\bm{\xi}+\nabla\varphi(x))dx, (1.1)

for all matrices 𝛏N×n\bm{\xi}\in\mathbb{R}^{N\times n} and all functions φW01,([0,1]n,N).\varphi\in W_{0}^{1,\infty}([0,1]^{n},\mathbb{R}^{N}). Under some appropriate growth conditions and some continuity conditions on the Lagrangian f,f, it is known that quasiconvexity of ff in the gradient variable is equivalent to the fact that the energy functional

E(𝒚)=Ωf(x,𝒚(x),𝒚)𝑑xE(\bm{y})=\int_{\Omega}f(x,\bm{y}(x),\nabla\bm{y})dx

is weakly lower semicontinuous in an appropriate Sobolev space [Reference,Reference,Reference,Reference,Reference]; the weak lower semicontinuity of the energy EE in turn implies the existence of global minimizers for EE in the Sobolev space under consideration. The rank-one convexity condition, known to be a weaker than the quasiconvexity condition [Reference,Reference], occurs when considering the second variation of the energy functional E(𝒚).E(\bm{y}). It reads as follows: Let n,Nn,N\in\mathbb{N} and let f:N×n.f\colon\mathbb{R}^{N\times n}\to\mathbb{R}. Then ff is said to be rank-one-convex, if

f(λ𝑨+(1λ)𝑩)λf(𝑨)+(1λ)f(𝑩),f(\lambda\bm{A}+(1-\lambda)\bm{B})\leq\lambda f(\bm{A})+(1-\lambda)f(\bm{B}), (1.2)

for all λ[0,1]\lambda\in[0,1] and 𝐀,𝐁N×n\bm{A},\bm{B}\in\mathbb{R}^{N\times n} such that rank(𝐀𝐁)1.\mathrm{rank}(\bm{A}-\bm{B})\leq 1. In linear elasticity a necessary condition for a body containing a linearly elastic homogeneous material with elasticity tensor 𝑪\bm{C} to be stable, when the displacement is fixed at the boundary, is the rank-one convexity condition. In elasticity, when the material phase separates the displacement field (with no cracking), the displacement must still be continuous across the phase boundaries. Such phase separation is most easily seen in shape memory materials such as Nitinol. A simple geometry for the phase separated material is a laminate of the phases, and the continuity of the displacement field forces the difference of the displacement gradient in one phase minus the displacement field in the second phase to be a rank-one tensor. Thus to avoid this layering transformation the energy ff as a function of the displacement gradient must be rank one convex. More generally, to avoid separation at the microscale into other geometries of possibly lower energy (with affine boundary conditions on the displacement 𝒖\bm{u} at the boundary Ω\partial\Omega of the body) the energy f(𝒖(𝒙))f(\nabla{\bm{u}}(\bm{x})) has to be a quasiconvex function of 𝒖(𝒙)\nabla{\bm{u}(\bm{x})} [Reference,Reference].

It is known that in the case when ff is a quadratic form, it is quasiconvex if and only if it is rank-one convex [Reference,Reference,Reference], which reduces to the so-called Legendre-Hadamard condition:

f(𝒙𝒚)0,for all𝒙N,𝒚n,f(\bm{x}\otimes\bm{y})\geq 0,\quad\text{for all}\quad\bm{x}\in\mathbb{R}^{N},\bm{y}\in\mathbb{R}^{n}, (1.3)

where 𝒙𝒚\bm{x}\otimes\bm{y} is the tensor product of the vectors 𝒙\bm{x} and 𝒚\bm{y} with (𝒙𝒚)ij=xiyj,(\bm{x}\otimes\bm{y})_{ij}=x_{i}y_{j}, for 1iN,1jn.1\leq i\leq N,1\leq j\leq n. It is then clear that quasiconvex quadratic forms in applied mathematics correspond to nonnegative biquadratic forms in real algebraic geometry. Let 𝒞N,n{\cal C}_{N,n} denote the convex cone of N×nN\times n quasiconvex quadratic forms, where we set 𝒞n=𝒞n,n.{\cal C}_{n}={\cal C}_{n,n}. Another convexity condition in the calculus of variations is the polyconvexity condition introduced by Ball [Reference], which is known to be an intermediate condition between the standard convexity and quasiconvexity. A function f:N×nf\colon\mathbb{R}^{N\times n}\to\mathbb{R} is called polyconvex, if there exists a convex function g:Kg\colon\mathbb{R}^{K}\to\mathbb{R} such that f(𝛏)=g(M1,,MK)f(\bm{\xi})=g(M_{1},\dots,M_{K}) where MiM_{i} are all the minors (including the first order ones) of the matrix 𝛏N×n\bm{\xi}\in\mathbb{R}^{N\times n}. Terpstra [Reference] proved that in the special case when ff is a quadratic form, then ff is polyconvex if and only if it can be written as a convex quadratic form plus a linear combination of the second order minors of 𝝃,\bm{\xi}, see also [Reference]. This means that polyconvex quadratic forms in applied mathematics correspond to biquadratic forms that are sums of squares in real algebraic geometry. A characterization of symmetric polyconvexity has been recently given in [Reference]. Also, a characterization of rank-one (quasiconvex) quadratic forms depending only on the strain is given by Zhang [Reference] using Morse index. Ball showed that in the case N=n,N=n, the determinant of the gradient function 𝒚\nabla\bm{y} is a Null-Lagrangian, and one has weak convergence of determinants det(𝒚m)\det(\nabla\bm{y}_{m}) under the weak convergence of the fields {𝒚m}\{\bm{y}_{m}\} in a Sobolev space W1,p(n<p<),W^{1,p}(n<p<\infty), thus the same classical theory of existence of global minimizers for convex Lagrangians goes through for polyconvex Lagrangians ff too [Reference]. There is no known algorithm that checks (analytically or even numerically) if the given function is quasiconvex or not, and it is surprisingly very complex even for simple functions f,f, while checking the polyconvexity of a function ff can be straightforward in many cases. This makes polyconvexity much easier to deal with. The present work continues the line of studying extreme rays and the so-called Milton extremals (or simply weak extremals) of 𝒞3,{\cal C}_{3}, initiated in [Reference] and further developed in [Reference,Reference,Reference]. Namely, we study the elements of 𝒞3{\cal C}_{3} that have an extremal acoustic tensor determinant as a polynomial, and characterize them. For the convenience of the reader, we next present definitions of weak and strong extremals (extreme rays) of 𝒞3,{\cal C}_{3}, (for the definition of the acoustic tensor see the paragraph right before Thorem LABEL:th:1.1).

Definition 1.1.

A quasiconvex quadratic form f(𝛏):N×nf(\bm{\xi})\colon\mathbb{R}^{N\times n}\to\mathbb{R} (f𝒞N,nf\in{\cal C}_{N,n}) is called

  • (i)

    A weak (or Milton) extremal, if one can not subtract a convex form from it, other then a multiple of itself, preserving the quasiconvexity of f.f.

  • (i)

    An extreme ray of 𝒞N,n{\cal C}_{N,n} (or a strong extremal), if one can not subtract a quasiconvex form from it, other then a multiple of itself, preserving the quasiconvexity of f.f.

It is not difficult to prove that even the notion of weak extremality in 𝒞N,n{\cal C}_{N,n} has the Krein-Milman property [Reference]. Extremals (weak or strong) are known to play an important role in the theory of composites as suggested by the work [Reference], especially when bounding effective properties of composites (such as shear or bulk moduli in elasticity for instance), in particular, the simplest forms of extremals that are the 2×22\times 2 minors of 𝝃\bm{\xi} (which are also Null-Lagrangians), are the basis of the so-called translation method of Murat and Tartar [Reference,Reference] or Cherkaev and Gibiansky [Reference], see also the works [Reference,Reference,Reference,Reference,Reference,Reference,Reference] and the books [Reference,Reference]. Special forms of extremals have been used by Kang and Milton in [Reference] to prove bounds on the volume fractions of two materials in a three dimensional body from boundary measurements. Extremal quasiconvex forms are also the best choice of quasiconvex functions for obtaining series expansions for effective tensors that have an extended domain of convergence, and thus analyticity properties as a function of the component moduli on this domain (see section 14.8, and page 373 of section 18.2 of [Reference]).

It is easy to see that any nontrivial extremal quadratic form (different from the square of a linear form or linear combination of 2×22\times 2 minors) is automatically an example of a quasiconvex quadratic form that is not polyconvex. Note, as proven by Terpstra [Reference], that a quadratic form is polyconvex if and only if it is the sum of a convex form and a linear combination of second order minors of the matrix 𝝃.\bm{\xi}. It was an open question in the applied mathematics community to find an explicit example of a quadratic form that is not polyconvex, until Serre provided one [Reference] in 1981. Surprisingly such an example was already provided in linear algebra/real algebraic geometry community by Choi [Reference] six years earlier in 1975, which had not been known to the applied mathematics communities until very recent times (we believe until the year 2019). Two years later Choi and Lam provided another, even more beautiful explicit example of such a form in [Reference]:

f(𝝃)=ξ112+ξ222+ξ332+ξ122+ξ232+ξ3122(ξ11ξ22+ξ22ξ33+ξ33ξ11),f(\bm{\xi})=\xi_{11}^{2}+\xi_{22}^{2}+\xi_{33}^{2}+\xi_{12}^{2}+\xi_{23}^{2}+\xi_{31}^{2}-2(\xi_{11}\xi_{22}+\xi_{22}\xi_{33}+\xi_{33}\xi_{11}), (1.4)

where they prove that the new example is in fact an extreme ray (the first such explicit example) of 𝒞3;{\cal C}_{3}; see also [Reference]. In fact it is an open question whether weak and strong extremals of 𝒞3{\cal C}_{3} are the same, while for 𝒞d,{\cal C}_{d}, d4d\geq 4 they are different, see next section. The first author and Milton came up with the Choi-Lam example later in [Reference] being unaware of it (as the applied mathematics community was unaware of it) due to the lack of communication between the two communities/fields. Nonnegative biquadratic forms have been a central subject of interest in the real algebraic geometry community, such as extreme points of the convex cone 𝒞d{\cal C}_{d} [Reference,Reference], separability and inseparability of positive linear maps [Reference,Reference,Reference], maximal possible number of their zeros and connections with extremality [Reference,Reference]. In particular the problem of expressing a nonnegative homogeneous polynomial as a sum of squares is very famous in real algebraic geometry [Reference,Reference,Reference,Reference,Reference,Reference,Reference,Reference]. In 1888 Hilbert raised the question of whether any nonnegative polynomial over reals can be expressed as a sum of squares of rational functions, which was solved in the affirmative by Artin [Reference]. For the problem of sums of squares of polynomials we refer to the recent surveys by Blekherman and coauthors [Reference,Reference]. Another very important related problem in applied mathematics, concerning sixth order homogeneous polynomials in three variables and determinants of 𝒚\bm{y}-matrices of quadratic forms is, whether or not any such polynomial, in particular the well known Robinson’s polynomial, is a determinant of a 𝒚\bm{y}-matrix, and it is open as well [Reference,Reference]. In [Reference] the authors construct the first examples of nonnegative biquadratic forms with a tensor in (3)4,(\mathbb{R}^{3})^{4}, that have maximal number of nontrivial zeros, namely ten of them. Note that by our result in Theorem 2.1, the latter are extreme rays of 𝒞3,{\cal C}_{3}, as their 𝒚\bm{y}-matrix determinants are scalar multiples of the generalized Robinson’s polynomial [Reference,Reference], which is an extremal polynomial.

Recall that an 2n2n-homogeneous polynomial P(𝐱)P(\bm{x}) in the variable 𝐱=(x1,x2,,xm)\bm{x}=(x_{1},x_{2},\dots,x_{m}) is said to be an extremal polynomial, if deg(P)=2n,\mathrm{deg}(P)=2n, P(𝐱)0P(\bm{x})\geq 0 for all 𝐱m,\bm{x}\in\mathbb{R}^{m}, and P(𝐱)P(\bm{x}) can not be split into the sum of two linearly independent polynomials P1P_{1} and P2P_{2} having the same properties.

Some of the above results were used in [Reference] to come up with a sufficient condition for a form f(𝝃)=𝝃𝑪𝝃T𝒞df(\bm{\xi})=\bm{\xi}\bm{C}\bm{\xi}^{T}\in{\cal C}_{d} to be a weak extremal, where 𝝃d×d\bm{\xi}\in\mathbb{R}^{d\times d} and 𝑪(d)4.\bm{C}\in(\mathbb{R}^{d})^{4}. Namely, let a rank-one matrix 𝝃d×d\bm{\xi}\in\mathbb{R}^{d\times d} be given as 𝝃=𝒙𝒚,\bm{\xi}=\bm{x}\otimes\bm{y}, where 𝒙,𝒚d,\bm{x},\bm{y}\in\mathbb{R}^{d}, d3.d\geq 3. Then one can write f(𝝃)=f(𝒙𝒚)=𝒙T(𝒚)𝒙T,f(\bm{\xi})=f(\bm{x}\otimes\bm{y})=\bm{x}T(\bm{y})\bm{x}^{T}, where T(𝒚)T(\bm{y}) is a d×dd\times d matrix, called the acoustic tensor (or just 𝒚\bm{y}-matrix) of f,f, with entries being quadratic forms in 𝒚.\bm{y}. The following results have been proven in [Reference] (we combine Theorems 3.4-3.7 in one).

Theorem 1.2.

Let the quadratic form f(𝛏)=𝛏𝐂𝛏T,f(\bm{\xi})=\bm{\xi}\bm{C}\bm{\xi}^{T}, where 𝛏d×d\bm{\xi}\in\mathbb{R}^{d\times d} and 𝐂(d)4,\bm{C}\in(\mathbb{R}^{d})^{4}, d3d\geq 3 be quasiconvex. Then

  • (i)

    If the determinant det(T(𝒚))\det(T(\bm{y})) is an irreducible (over the reals) extremal polynomial, then the form ff is a weak extremal.

  • (ii)

    Assume d=3.d=3. If the determinant det(T(𝒚))\det(T(\bm{y})) is an extremal that is not a perfect square, then ff is a weak extremal.

  • (iii)

    Assume d=3.d=3. If det(T(𝒚))0\det(T(\bm{y}))\equiv 0 then the form ff is either a weak extremal or polyconvex.

  • (iv)

    Assume d=3.d=3. If the determinant det(T(𝒚))\det(T(\bm{y})) is a perfect square (note that this automatically implies that it is an extremal polynomial as can be seen easily), then ff is either a weak extremal, polyconex, or the sum of a polyconvex and a weak extremal forms, where the extremal form has identically zero acoustic tensor determinant.

We improved the result in (ii) for forms having linear elastic orthotropic symmetry in [Reference], showing that in fact under the extremality and non-square condition on the determinant det(T(𝒚)),\det(T(\bm{y})), the form ff must in fact be an extreme ray of 𝒞3.{\cal C}_{3}. In the present manuscript we study forms f𝒞3f\in{\cal C}_{3} in questions (ii)-(iv) for strong extremality, see Theorems 2.1-2.2 in the next section. We also conjecture about weak versus strong extremals of 𝒞3,{\cal C}_{3}, about extremality or the vanishing property of the acoustic tensor determinant versus weak or strong extremality or polyconvexity of f(𝝃)f(\bm{\xi}) for d=3d=3 and d4,d\geq 4, see next section.

2 Main Results

Let dd\in\mathbb{N}, (d3d\geq 3) and 𝑪=(Cijkl)(d)4\bm{C}=(C_{ijkl})\in(\mathbb{R}^{d})^{4} be a fourth order tensor with usual symmetries:

Cijkl=Ckjil=Cilkj,1i,j,k,ld.C_{ijkl}=C_{kjil}=C_{ilkj},\qquad 1\leq i,j,k,l\leq d. (2.1)

In what follows we will regard the matrix 𝝃=(ξij)d×d\bm{\xi}=(\xi_{ij})\in\mathbb{R}^{d\times d} as a d2d^{2}-vector so that the quadratic form f(𝝃)f(\bm{\xi}) will be given by

f(𝝃)=𝝃𝑪𝝃T=1i,j,k,ldCijklξijξkl,f(\bm{\xi})=\bm{\xi}\bm{C}\bm{\xi}^{T}=\sum_{1\leq i,j,k,l\leq d}C_{ijkl}\xi_{ij}\xi_{kl}, (2.2)

which will be applicable in the context of elasticity. As already noted, in the special case when 𝝃=𝒙𝒚\bm{\xi}=\bm{x}\otimes\bm{y} is a rank-one matrix, where 𝒙,𝒚d,\bm{x},\bm{y}\in\mathbb{R}^{d}, the quadratic form ff reduces to

f(𝒙𝒚)=𝒙(𝒚𝑪𝒚T)𝒙T=𝒙𝑻(𝒚)𝒙T,f(\bm{x}\otimes\bm{y})=\bm{x}(\bm{y}\bm{C}\bm{y}^{T})\bm{x}^{T}=\bm{x}\bm{T}(\bm{y})\bm{x}^{T}, (2.3)

where 𝑻(𝒚)=𝒚𝑪𝒚Td×d\bm{T}(\bm{y})=\bm{y}\bm{C}\bm{y}^{T}\in\mathbb{R}^{d\times d} is the acoustic tensor (or simply the 𝒚\bm{y}-matrix) of f.f. Also, as mentioned above, it turns out that the determinant of 𝑻(𝒚)\bm{T}(\bm{y}) tells quite a lot about the form f,f, which is quite unexpected [Reference]. We will focus on the case when det(𝑻(𝒚))\det(\bm{T}(\bm{y})) is an extremal polynomial. The following are the main results of the paper. The first theorem refers to the cases (ii) and (iv) in Thereom 1.2.

Theorem 2.1.

Let f(𝛏)=𝛏𝐂𝛏T𝒞3,f(\bm{\xi})=\bm{\xi}\bm{C}\bm{\xi}^{T}\in{\cal C}_{3}, where 𝛏3×3\bm{\xi}\in\mathbb{R}^{3\times 3} and 𝐂(3)4\bm{C}\in(\mathbb{R}^{3})^{4} is a fourth order tensor with usual symmetries as in (2.1). Assume that the determinant of the 𝐲\bm{y}-matrix of f(𝐱𝐲)f(\bm{x}\otimes\bm{y}) is an extremal polynomial. Then one has the following:

  • 1.

    If det(𝑻(𝒚))\det(\bm{T}(\bm{y})) is not a perfect square, then ff must be an extreme ray of 𝒞3.{\cal C}_{3}.

  • 2.

    If det(𝑻(𝒚))\det(\bm{T}(\bm{y})) is a perfect square, then ff is either an extreme ray of 𝒞3{\cal C}_{3} or polyconvex.

The next theorem refers to the case (iii) in Thereom 1.2. It will also be a major factor in the proof of the main Theorem 2.1.

Theorem 2.2.

Let f(𝛏)=𝛏𝐂𝛏T𝒞3,f(\bm{\xi})=\bm{\xi}\bm{C}\bm{\xi}^{T}\in{\cal C}_{3}, where 𝛏3×3\bm{\xi}\in\mathbb{R}^{3\times 3} and 𝐂(3)4\bm{C}\in(\mathbb{R}^{3})^{4} is a fourth order tensor with usual symmetries as in (2.1). Assume that the determinant of the 𝐲\bm{y}-matrix of f(𝐱𝐲)f(\bm{x}\otimes\bm{y}) is identically zero. Then ff must be a polyconvex form.

Several remarks are in order.

Remark 2.3 (The case d=3d=3).

Let d=3.d=3. Note first that the Choi-Lam example in (1.4) gives

det𝑻(𝒚)=y14y22+y24y32+y34y123y12y22y32,\det{\bm{T}(\bm{y})}=y_{1}^{4}y_{2}^{2}+y_{2}^{4}y_{3}^{2}+y_{3}^{4}y_{1}^{2}-3y_{1}^{2}y_{2}^{2}y_{3}^{2},

which is known to be an extremal polynomial; this falls into Theorem 2.1. An example of a polyconvex ff that has a perfect square or zero determinant would be f(𝛏)=i=13ξii2f(\bm{\xi})=\sum_{i=1}^{3}\xi_{ii}^{2} or f(𝛏)=ξ112.f(\bm{\xi})=\xi_{11}^{2}. However, we are not aware of an example of an ff that is non-polyconvex, is an extreme ray of 𝒞3{\cal C}_{3} such that det(𝐓(𝐲))\det(\bm{T}(\bm{y})) is a perfect square. We believe that if f𝒞3f\in{\cal C}_{3} with det𝐓(𝐲)\det{\bm{T}(\bm{y})} being a perfect square, then ff must in fact be polyconvex. However, at the moment we have no proof for the statement.

Remark 2.4 (The case d4d\geq 4).

Note that if one only assumes that det𝐓(𝐲)\det{\bm{T}(\bm{y})} is en extremal polynomial (not necessarily irreducible), then f(𝛏)f(\bm{\xi}) has to be neither a weak extremal of 𝒞d{\cal C}_{d} nor polyconvex. A counterexample would be

f(𝝃)=ξ112+ξ222+ξ332+ξ122+ξ232+ξ3122(ξ11ξ22+ξ22ξ33+ξ33ξ11)+k=3dξkk2,f(\bm{\xi})=\xi_{11}^{2}+\xi_{22}^{2}+\xi_{33}^{2}+\xi_{12}^{2}+\xi_{23}^{2}+\xi_{31}^{2}-2(\xi_{11}\xi_{22}+\xi_{22}\xi_{33}+\xi_{33}\xi_{11})\ +\sum_{k=3}^{d}\xi_{kk}^{2},

which has the acoustic tensor determinant

det𝑻(𝒚)=(y14y22+y24y32+y34y123y12y22y32)k=3dyk2,\det{\bm{T}(\bm{y})}=(y_{1}^{4}y_{2}^{2}+y_{2}^{4}y_{3}^{2}+y_{3}^{4}y_{1}^{2}-3y_{1}^{2}y_{2}^{2}y_{3}^{2})\prod_{k=3}^{d}y_{k}^{2},

which is clearly an extremal polynomial. However, obviously ff is neither a weak extremal nor polyconvex.

Remark 2.5 (The case d4d\geq 4).

Another thing to note is that in the case d4,d\geq 4, one can put together two copies of the Choi-Lam form to achieve an example f𝒞df\in{\cal C}_{d} that is a weak but not a strong extremal of 𝒞d.{\cal C}_{d}. Namely, it is easy to see that the form

f(𝝃)\displaystyle f(\bm{\xi}) =ξ112+ξ222+ξ332+ξ122+ξ232+ξ3122(ξ11ξ22+ξ22ξ33+ξ33ξ11)\displaystyle=\xi_{11}^{2}+\xi_{22}^{2}+\xi_{33}^{2}+\xi_{12}^{2}+\xi_{23}^{2}+\xi_{31}^{2}-2(\xi_{11}\xi_{22}+\xi_{22}\xi_{33}+\xi_{33}\xi_{11})
+ξ222+ξ332+ξ442+ξ232+ξ342+ξ4122(ξ22ξ33+ξ33ξ44+ξ44ξ22)\displaystyle+\xi_{22}^{2}+\xi_{33}^{2}+\xi_{44}^{2}+\xi_{23}^{2}+\xi_{34}^{2}+\xi_{41}^{2}-2(\xi_{22}\xi_{33}+\xi_{33}\xi_{44}+\xi_{44}\xi_{22})

is a weak extremal of 𝒞d.{\cal C}_{d}. This implies that weak and strong extremals of 𝒞d{\cal C}_{d} are in general different for d4.d\geq 4.

Remark 2.6 (The case d4d\geq 4).

Taking again the Choi-Lam example f(𝛏)f(\bm{\xi}) we have det𝐓(𝐲)0\det{\bm{T}(\bm{y})}\equiv 0 for d4.d\geq 4. This shows that Theorem 2.2 fails for d4.d\geq 4.

Finally, we make following conjecture.

Conjecture 2.7 (The case d=3d=3).

Any non-polyconvex weak extremal f𝒞3f\in{\cal C}_{3} is an extreme ray of 𝒞3.{\cal C}_{3}. Moreover, if f𝒞3f\in{\cal C}_{3} is a non-polyconvex extreme ray of 𝒞3,{\cal C}_{3}, then det𝐓(𝐲)\det{\bm{T}(\bm{y})} is en extremal polynomial different from a perfect square.

The motivation behind this conjecture is the yet unproven fact that any nonnegative sixth degree homogeneous polynomial P(𝒚)P(\bm{y}) in three variables (𝒚3\bm{y}\in\mathbb{R}^{3}) is necessarily the determinant of the acoustic tensor T(𝒚)T(\bm{y}) of an element f𝒞3,f\in{\cal C}_{3}, e.g., [Reference,Reference,Reference]. A weaker statement, that every real multivariate polynomial has a symmetric determinantal representation is known to be true, and was recently proven by Helton, McCullough, and Vinnikov [Reference], see also [Reference,Reference].

3 Proof of Theorem 2.1

Proof of Theorem 2.1.

We will be utilizing Theorem 2.2 in the proof here; the proof of which is postponed until Section 4. We will be carrying out some steps applicable to both cases in Theorem 2.1, and at the same time considering each case separately if necessary. Assume in contradiction that ff is not an extreme ray of 𝒞3,{\cal C}_{3}, thus there exists a form f1𝒞3f_{1}\in{\cal C}_{3} such that f1f_{1} and ff are linearly independent satisfying the inequalities

0f1(𝒙𝒚)f(𝒙𝒚),for all𝒙,𝒚3.0\leq f_{1}(\bm{x}\otimes\bm{y})\leq f(\bm{x}\otimes\bm{y}),\quad\text{for all}\quad\bm{x},\bm{y}\in\mathbb{R}^{3}. (3.1)

We will prove that in the first case this is not possible, while in the second case this leads to the conclusion that ff is polyconvex. Denote f(𝒙𝒚)=𝒙𝑻(𝒚)𝒙T,f(\bm{x}\otimes\bm{y})=\bm{x}\bm{T}(\bm{y})\bm{x}^{T}, f1(𝒙𝒚)=𝒙𝑻1(𝒚)𝒙T,f_{1}(\bm{x}\otimes\bm{y})=\bm{x}\bm{T}^{1}(\bm{y})\bm{x}^{T}, 𝑻(𝒚)=(tij(𝒚))i,j=13,\bm{T}(\bm{y})=(t_{ij}(\bm{y}))_{i,j=1}^{3}, and 𝑻1(𝒚)=(tij1(𝒚))i,j=13.\bm{T}^{1}(\bm{y})=(t_{ij}^{1}(\bm{y}))_{i,j=1}^{3}. Consider the determinant det(𝑻(𝒚)λ𝑻1(𝒚))\det(\bm{T}(\bm{y})-\lambda\bm{T}^{1}(\bm{y})) as a polynomial in λ:\lambda\in\mathbb{R}:

P(λ)\displaystyle P(\lambda) =det(𝑻(𝒚)λ𝑻1(𝒚))\displaystyle=\det(\bm{T}(\bm{y})-\lambda\bm{T}^{1}(\bm{y})) (3.2)
=det(𝑻(𝒚))λi,j=13tij1(𝒚)cof(𝑻(𝒚))ij+λ2i,j=13tij(𝒚)cof(𝑻1(𝒚))ijλ3det(𝑻1(𝒚)),\displaystyle=\det(\bm{T}(\bm{y}))-\lambda\sum_{i,j=1}^{3}t^{1}_{ij}(\bm{y})\mathrm{cof}(\bm{T}(\bm{y}))_{ij}+\lambda^{2}\sum_{i,j=1}^{3}t_{ij}(\bm{y})\mathrm{cof}(\bm{T}^{1}(\bm{y}))_{ij}-\lambda^{3}\det(\bm{T}^{1}(\bm{y})),

which will be a key factor in the analysis. The determinant above gives rise to the coefficients of λk,\lambda^{k}, for k=0,1,2,3k=0,1,2,3 that are homogeneous polynomials of 𝒚\bm{y} of degree six, which turn out to have to satisfy certain monotonicity properties proven in [Lemma 4.1, Reference] and given in the lemma below.

Lemma 3.1.

Let nn\in\mathbb{N} satisfy n2n\geq 2 and let 𝐀,𝐁𝕄symn×n\bm{A},\bm{B}\in\mathbb{M}_{sym}^{n\times n} be symmetric positive semi-definite matrices such that 𝐀𝐁\bm{A}\geq\bm{B} in the sense of quadratic forms. Then for any integers 1k<mn1\leq k<m\leq n one has the inequality

1(nm)Mm(𝑩)Mm(𝑩)cof𝑨(Mm(𝑩))1(nk)Mk(𝑩)Mk(𝑩)cof𝑨(Mk(𝑩)),\frac{1}{{n\choose m}}\sum_{M_{m}(\bm{B})}M_{m}(\bm{B})\mathrm{cof}_{\bm{A}}(M_{m}(\bm{B}))\leq\frac{1}{{n\choose k}}\sum_{M_{k}(\bm{B})}M_{k}(\bm{B})\mathrm{cof}_{\bm{A}}(M_{k}(\bm{B})), (3.3)

where the number (nm){n\choose m} is the binomial coefficient, and the sum Mm(𝐁)\sum_{M_{m}(\bm{B})} is taken over all mm-th order minors Mm(𝐁)M_{m}(\bm{B}) of 𝐁,\bm{B}, and cof𝐀(Mm(𝐁))\mathrm{cof}_{\bm{A}}(M_{m}(\bm{B})) denotes the cofactor of the minor in the matrix 𝐀,\bm{A}, obtained by choosing the same rows and columns as to get the minor Mm(𝐁)M_{m}(\bm{B}) in 𝐁.\bm{B}.

Due to (3.1), we have 𝑻(𝒚)𝑻1(𝒚)\bm{T}(\bm{y})\geq\bm{T}^{1}(\bm{y}) for all 𝒚3\bm{y}\in\mathbb{R}^{3} in the sense of quadratic forms, thus Lemma 3.1 implies the inequalities

03det(𝑻1(𝒚))i,j=13tij(𝒚)cof(𝑻1(𝒚))iji,j=13tij1(𝒚)cof(𝑻(𝒚))ij3det(𝑻(𝒚)),𝒚3.0\leq 3\det(\bm{T}^{1}(\bm{y}))\leq\sum_{i,j=1}^{3}t_{ij}(\bm{y})\mathrm{cof}(\bm{T}^{1}(\bm{y}))_{ij}\leq\sum_{i,j=1}^{3}t^{1}_{ij}(\bm{y})\mathrm{cof}(\bm{T}(\bm{y}))_{ij}\leq 3\det(\bm{T}(\bm{y})),\ \ \ \bm{y}\in\mathbb{R}^{3}. (3.4)

Hence the polynomials 3det(𝑻1(𝒚))3\det(\bm{T}^{1}(\bm{y})), i,j=13tij(𝒚)cof(𝑻1(𝒚))ij,\sum_{i,j=1}^{3}t_{ij}(\bm{y})\mathrm{cof}(\bm{T}^{1}(\bm{y}))_{ij}, i,j=13tij1(𝒚)cof(𝑻(𝒚))ij,\sum_{i,j=1}^{3}t^{1}_{ij}(\bm{y})\mathrm{cof}(\bm{T}(\bm{y}))_{ij}, being in between zero and the extremal polynomial 3det(𝑻(𝒚))3\det(\bm{T}(\bm{y})) must be scalar multiples of det(𝑻(𝒚)),\det(\bm{T}(\bm{y})), i.e., we have

det(𝑻1(𝒚))\displaystyle\det(\bm{T}^{1}(\bm{y})) =αdet(𝑻(𝒚)),\displaystyle=\alpha\det(\bm{T}(\bm{y})), (3.5)
i,j=13tij(𝒚)cof(𝑻1(𝒚))ij\displaystyle\sum_{i,j=1}^{3}t_{ij}(\bm{y})\mathrm{cof}(\bm{T}^{1}(\bm{y}))_{ij} =βdet(𝑻(𝒚)),\displaystyle=\beta\det(\bm{T}(\bm{y})),
i,j=13tij1(𝒚)cof(𝑻(𝒚))ij\displaystyle\sum_{i,j=1}^{3}t^{1}_{ij}(\bm{y})\mathrm{cof}(\bm{T}(\bm{y}))_{ij} =γdet(𝑻(𝒚)),\displaystyle=\gamma\det(\bm{T}(\bm{y})),
for someα,β,γ\displaystyle\text{for some}\quad\alpha,\beta,\gamma 0.\displaystyle\geq 0.

Consequently we get from (3.2) and (3.5) the key identity

det(𝑻(𝒚)λ𝑻1(𝒚))=(1γλ+βλ2αλ3)det(𝑻(𝒚))=φ(λ)det(𝑻(𝒚)),𝒚3,λ.\det(\bm{T}(\bm{y})-\lambda\bm{T}^{1}(\bm{y}))=(1-\gamma\lambda+\beta\lambda^{2}-\alpha\lambda^{3})\det(\bm{T}(\bm{y}))=\varphi(\lambda)\det(\bm{T}(\bm{y})),\ \ \ \bm{y}\in\mathbb{R}^{3},\lambda\in\mathbb{R}. (3.6)

In the next step we note that the polynomial φ\varphi does not have roots in (,1)(-\infty,1), more precisely

φ(λ)>0, forλ(,1).\varphi(\lambda)>0,\quad\text{ for}\quad\lambda\in(-\infty,1). (3.7)

Indeed, for λ0\lambda\leq 0 we have by the conditions α,β,γ0\alpha,\beta,\gamma\geq 0 that φ(λ)1.\varphi(\lambda)\geq 1. Choosing a point 𝒚03\bm{y}^{0}\in\mathbb{R}^{3} such that det(𝑻(𝒚0))>0\det(\bm{T}(\bm{y}^{0}))>0, we have for any λ(0,1)\lambda\in(0,1) by Lemma 3.1 that

φ(λ)\displaystyle\varphi(\lambda) =1det(𝑻(𝒚0))det(𝑻(𝒚0)λ𝑻1(𝒚0))\displaystyle=\frac{1}{\det(\bm{T}(\bm{y}^{0}))}\det(\bm{T}(\bm{y}^{0})-\lambda\bm{T}^{1}(\bm{y}^{0}))
=1det(𝑻(𝒚0))det[(1λ)𝑻(𝒚0)+λ(𝑻(𝒚0)𝑻1(𝒚0))]\displaystyle=\frac{1}{\det(\bm{T}(\bm{y}^{0}))}\det[(1-\lambda)\bm{T}(\bm{y}^{0})+\lambda(\bm{T}(\bm{y}^{0})-\bm{T}^{1}(\bm{y}^{0}))]
(1λ)3,\displaystyle\geq(1-\lambda)^{3},

as 𝑻(𝒚0)𝑻1(𝒚0)\bm{T}(\bm{y}^{0})\geq\bm{T}^{1}(\bm{y}^{0}) in the sense of quadratic forms. Note also that the equality α=0\alpha=0 is impossible as it would mean by (3.5) that

det(𝑻1(𝒚))=αdet(𝑻(𝒚))=0,𝒚3,\det(\bm{T}^{1}(\bm{y}))=\alpha\det(\bm{T}(\bm{y}))=0,\qquad\bm{y}\in\mathbb{R}^{3},

i.e., the quasiconvex form f1f^{1} has an identically zero acoustic tensor determinant, thus by Theorem 2.2 it must be polyconvex. Invoking again the characterization theorem for polyconvex quadratic forms by Terpstra [Reference], we infer that f1f^{1} is a sum of squares (at least one), which means by (3.1) that in fact one can subtract a perfect square form ff still preserving the quasiconvexity of f,f, i.e., ff is not a weak extremal, which contradicts part (ii) of Theorem 1.2. Consequently we must have α>0\alpha>0 and det(𝑻1(𝒚))=αdet(𝑻(𝒚))>0\det(\bm{T}^{1}(\bm{y}))=\alpha\det(\bm{T}(\bm{y}))>0 whenever det(𝑻(𝒚))>0.\det(\bm{T}(\bm{y}))>0. Also it is important to note that φ\varphi is necessarily a third degree polynomial. Choose again 𝒚03\bm{y}_{0}\in\mathbb{R}^{3} (as above) such that det(𝑻(𝒚0))>0.\det(\bm{T}(\bm{y}_{0}))>0. Hence setting 𝑨=𝑻(𝒚0)\bm{A}=\bm{T}(\bm{y}_{0}) and 𝑩=𝑻1(𝒚0)\bm{B}=\bm{T}^{1}(\bm{y}_{0}) we have det(𝑨)det(𝑩)>0,\det(\bm{A})\geq\det(\bm{B})>0, where 𝑨,𝑩3×3\bm{A},\bm{B}\in\mathbb{R}^{3\times 3} are symmetric positive definite matrices, thus the square root 𝑩1/2\bm{B}^{1/2} and the inverse 𝑩1/2\bm{B}^{-1/2} exist and are symmetric. Next we have from (3.2) and (3.6),

P(λ)\displaystyle P(\lambda) =det(𝑨λ𝑩)\displaystyle=\det(\bm{A}-\lambda\bm{B})
=det(𝑩1/2(𝑩1/2𝑨𝑩1/2λ𝑰)𝑩1/2)\displaystyle=\det(\bm{B}^{1/2}(\bm{B}^{-1/2}\bm{A}\bm{B}^{-1/2}-\lambda\bm{I})\bm{B}^{1/2})
=det(𝑩)det(𝑩1/2𝑨𝑩1/2λ𝑰)\displaystyle=\det(\bm{B})\det(\bm{B}^{-1/2}\bm{A}\bm{B}^{-1/2}-\lambda\bm{I})
=det(𝑨)φ(λ),\displaystyle=\det(\bm{A})\varphi(\lambda),

which gives

φ(λ)=αdet(𝑩1/2𝑨𝑩1/2λ𝑰),\varphi(\lambda)=\alpha\cdot\det(\bm{B}^{-1/2}\bm{A}\bm{B}^{-1/2}-\lambda\bm{I}),

thus the roots of φ\varphi are real as φ\varphi is a scalar multiple of the characteristic polynomial of the symmetric matrix 𝑩1/2𝑨𝑩1/2.\bm{B}^{-1/2}\bm{A}\bm{B}^{-1/2}. On the other hand (3.7) implies that all three roots of φ\varphi belong to the interval [1,).[1,\infty). Denoting them by 1λ1λ2λ31\leq\lambda_{1}\leq\lambda_{2}\leq\lambda_{3} we have

φ(t)=(1γλ+βλ2αλ3)=α(λλ1)(λλ2)(λλ3),\varphi(t)=(1-\gamma\lambda+\beta\lambda^{2}-\alpha\lambda^{3})=-\alpha(\lambda-\lambda_{1})(\lambda-\lambda_{2})(\lambda-\lambda_{3}),

and we have by Vieta’s theorem the formulae

α=1λ1λ2λ3,β=λ1+λ2+λ3λ1λ2λ3,γ=λ1λ2+λ2λ3+λ3λ1λ1λ2λ3,\alpha=\frac{1}{\lambda_{1}\lambda_{2}\lambda_{3}},\quad\beta=\frac{\lambda_{1}+\lambda_{2}+\lambda_{3}}{\lambda_{1}\lambda_{2}\lambda_{3}},\quad\gamma=\frac{\lambda_{1}\lambda_{2}+\lambda_{2}\lambda_{3}+\lambda_{3}\lambda_{1}}{\lambda_{1}\lambda_{2}\lambda_{3}}, (3.8)

which will be utilized in the next steps. Next introduce the biquadratic form

g(𝒙𝒚)=f(𝒙𝒚)λ1f1(𝒙𝒚).g(\bm{x}\otimes\bm{y})=f(\bm{x}\otimes\bm{y})-\lambda_{1}f^{1}(\bm{x}\otimes\bm{y}).

The strategy from here on is to either prove that gg is identically zero, which will imply that f1f_{1} is a multiple of ff, or otherwise to arrive at a contradiction in the case when det(𝑻(𝒚))\det(\bm{T}(\bm{y})) is not a perfect square, or prove that ff is polyconvex in the case when det(𝑻(𝒚))\det(\bm{T}(\bm{y})) is a perfect square. We have from (3.6) that det(𝑺(𝒚))0\det(\bm{S}(\bm{y}))\equiv 0 for 𝒚3,\bm{y}\in\mathbb{R}^{3}, where 𝑺(𝒚)=(sij(𝒚))1i,j3\bm{S}(\bm{y})=(s_{ij}(\bm{y}))_{1\leq i,j\leq 3} is the acoustic tensor of g,g, i.e. g(𝒙𝒚)=𝒙𝑺(𝒚)𝒙Tg(\bm{x}\otimes\bm{y})=\bm{x}\bm{S}(\bm{y})\bm{x}^{T}. Note that as ff and f1f_{1} are linearly independent, then the form gg is not identically zero. Next we aim to prove that the diagonal entries of the cofactor matrix cof(𝑺(𝒚))\mathrm{cof}(\bm{S}(\bm{y})) are nonnegative. If they all vanish identically, then there is nothing to prove. Assume for instance cof(𝑺(𝒚))33\mathrm{cof}(\bm{S}(\bm{y}))_{33} does not vanish identically, then the set {𝒚3:cof(𝑺(𝒚))33=0}\{\bm{y}\in\mathbb{R}^{3}:\mathrm{cof}(\bm{S}(\bm{y}))_{33}=0\} is a null set, thus because det(𝑺(𝒚))0,\det(\bm{S}(\bm{y}))\equiv 0, the last row of 𝑺\bm{S} must be a linear combination of the first two for a.e. 𝒚3,\bm{y}\in\mathbb{R}^{3}, thus we obtain the form

𝑺(𝒚)=[[1.5]s11s12rs11+qs12s12s22rs12+qs22rs11+qs12rs12+qs22r2s11+q2s22+2rqs12],\bm{S}(\bm{y})=\begin{bmatrix}[1.5]s_{11}&s_{12}&rs_{11}+qs_{12}\\ s_{12}&s_{22}&rs_{12}+qs_{22}\\ rs_{11}+qs_{12}&rs_{12}+qs_{22}&r^{2}s_{11}+q^{2}s_{22}+2rqs_{12}\end{bmatrix}, (3.9)

where the linear combination coefficients rr and qq are rational functions given by

r(𝒚)=cof(𝑺(𝒚))13cof(𝑺(𝒚))33,q(𝒚)=cof(𝑺(𝒚))23cof(𝑺(𝒚))33.r(\bm{y})=\frac{\mathrm{cof}(\bm{S}(\bm{y}))_{13}}{\mathrm{cof}(\bm{S}(\bm{y}))_{33}},\qquad q(\bm{y})=-\frac{\mathrm{cof}(\bm{S}(\bm{y}))_{23}}{\mathrm{cof}(\bm{S}(\bm{y}))_{33}}. (3.10)

Note that (3.9) also yields the form of the cofactor matrix cof(𝑺):\mathrm{cof}(\bm{S}):

cof(𝑺)=[[1.5]r2cof(𝑺)33rqcof(𝑺)33rcof(𝑺)33rqcof(𝑺)33q2cof(𝑺)33qcof(𝑺)33rcof(𝑺)33qcof(𝑺)33cof(𝑺)33].\mathrm{cof}(\bm{S})=\begin{bmatrix}[1.5]r^{2}\cdot\mathrm{cof}(\bm{S})_{33}&rq\cdot\mathrm{cof}(\bm{S})_{33}&-r\cdot\mathrm{cof}(\bm{S})_{33}\\ rq\cdot\mathrm{cof}(\bm{S})_{33}&q^{2}\cdot\mathrm{cof}(\bm{S})_{33}&-q\cdot\mathrm{cof}(\bm{S})_{33}\\ -r\cdot\mathrm{cof}(\bm{S})_{33}&-q\cdot\mathrm{cof}(\bm{S})_{33}&\mathrm{cof}(\bm{S})_{33}\end{bmatrix}. (3.11)

Now using the equality f=g+λ1f1f=g+\lambda_{1}f_{1} and formula (3.2) we get

det(𝑻(𝒚))\displaystyle\det(\bm{T}(\bm{y})) =det(𝑺(𝒚)+λ1𝑻1(𝒚))\displaystyle=\det(\bm{S}(\bm{y})+\lambda_{1}\bm{T}^{1}(\bm{y}))
=det(𝑺(𝒚))+λ1i,j=13tij1(𝒚)cof(𝑺(𝒚))ij+λ12i,j=13sij(𝒚)cof(𝑻1(𝒚))ij+λ13det(𝑻1(𝒚))\displaystyle=\det(\bm{S}(\bm{y}))+\lambda_{1}\sum_{i,j=1}^{3}t^{1}_{ij}(\bm{y})\mathrm{cof}(\bm{S}(\bm{y}))_{ij}+\lambda_{1}^{2}\sum_{i,j=1}^{3}s_{ij}(\bm{y})\mathrm{cof}(\bm{T}^{1}(\bm{y}))_{ij}+\lambda_{1}^{3}\det(\bm{T}^{1}(\bm{y}))
=λ1i,j=13tij1(𝒚)cof(𝑺(𝒚))ij+λ12i,j=13sij(𝒚)cof(𝑻1(𝒚))ij+λ13det(𝑻1(𝒚)),\displaystyle=\lambda_{1}\sum_{i,j=1}^{3}t^{1}_{ij}(\bm{y})\mathrm{cof}(\bm{S}(\bm{y}))_{ij}+\lambda_{1}^{2}\sum_{i,j=1}^{3}s_{ij}(\bm{y})\mathrm{cof}(\bm{T}^{1}(\bm{y}))_{ij}+\lambda_{1}^{3}\det(\bm{T}^{1}(\bm{y})),

hence owing to the first equality in (3.5) we obtain

i,j=13tij1(𝒚)cof(𝑺(𝒚))ij=(1λ1λ12α)det(𝑻(𝒚))λ1i,j=13sij(𝒚)cof(𝑻1(𝒚))ij.\sum_{i,j=1}^{3}t^{1}_{ij}(\bm{y})\mathrm{cof}(\bm{S}(\bm{y}))_{ij}=\left(\frac{1}{\lambda_{1}}-\lambda_{1}^{2}\alpha\right)\det(\bm{T}(\bm{y}))-\lambda_{1}\sum_{i,j=1}^{3}s_{ij}(\bm{y})\mathrm{cof}(\bm{T}^{1}(\bm{y}))_{ij}. (3.12)

We have further utilizing the first two identities in (3.5), that

i,j=13sij(𝒚)cof(𝑻1(𝒚))ij\displaystyle\sum_{i,j=1}^{3}s_{ij}(\bm{y})\mathrm{cof}(\bm{T}^{1}(\bm{y}))_{ij} =i,j=13(tijλ1tij1)(𝒚)cof(𝑻1(𝒚))ij\displaystyle=\sum_{i,j=1}^{3}(t_{ij}-\lambda_{1}t_{ij}^{1})(\bm{y})\mathrm{cof}(\bm{T}^{1}(\bm{y}))_{ij} (3.13)
=(β3λ1α)det(𝑻(𝒚)),\displaystyle=(\beta-3\lambda_{1}\alpha)\det(\bm{T}(\bm{y})),

thus owing back to (3.12) we obtain

i,j=13tij1(𝒚)cof(𝑺(𝒚))ij=(1λ1+2λ12αλ1β)det(𝑻(𝒚)).\sum_{i,j=1}^{3}t^{1}_{ij}(\bm{y})\mathrm{cof}(\bm{S}(\bm{y}))_{ij}=\left(\frac{1}{\lambda_{1}}+2\lambda_{1}^{2}\alpha-\lambda_{1}\beta\right)\det(\bm{T}(\bm{y})). (3.14)

Consequently recalling (3.8) and (3.11) we get from (3.14) after some simple algebra,

cof(𝑺(𝒚))33[(r,q,1)𝑻1(𝒚)(r,q,1)T]=(λ2λ1)(λ3λ1)λ1λ2λ3det(𝑻(𝒚)).\mathrm{cof}(\bm{S}(\bm{y}))_{33}\cdot[(-r,-q,1)\bm{T}^{1}(\bm{y})(-r,-q,1)^{T}]=\frac{(\lambda_{2}-\lambda_{1})(\lambda_{3}-\lambda_{1})}{\lambda_{1}\lambda_{2}\lambda_{3}}\det(\bm{T}(\bm{y})). (3.15)

The last equality suggests considering the following cases separately.
Case 1: λ1<λ2.\lambda_{1}<\lambda_{2}.
Case 2: λ1=λ2<λ3.\lambda_{1}=\lambda_{2}<\lambda_{3}.
Case 3: λ1=λ2=λ3.\lambda_{1}=\lambda_{2}=\lambda_{3}.
Case 1.
In this case by the fact that det(𝑻(𝒚))>0\det(\bm{T}(\bm{y}))>0 a.e. in 3\mathbb{R}^{3} and by the positive semi-definiteness of 𝑻1(𝒚),\bm{T}^{1}(\bm{y}), the equality (3.15) immediately implies that cof(𝑺(𝒚))330\mathrm{cof}(\bm{S}(\bm{y}))_{33}\geq 0 for all 𝒚3.\bm{y}\in\mathbb{R}^{3}. Therefore we have cof(𝑺(𝒚))ii0\mathrm{cof}(\bm{S}(\bm{y}))_{ii}\geq 0 for all 𝒚3,\bm{y}\in\mathbb{R}^{3}, i=1,2,3.i=1,2,3. Consider next the following two cases.
Case 1a: One of the diagonal entries of S(y)\bm{S}(\bm{y}) is definite.
Case 1b: All of the diagonal entries of S(y)\bm{S}(\bm{y}) are indefinite.
Case 1a.
In this case if say s11(𝒚)s_{11}(\bm{y}) is positive semidefinite, then we get by Silvester’s criterion that 𝑺(𝒚)\bm{S}(\bm{y}) is positive semidefinite, thus the form gg will become a quasiconvex quadratic form that has zero acoustic tensor determinant, thus by Theorem 2.2 it must be polyconvex. As gg is not identically zero, it must be a sum of squares, containing at least one square, thus the condition f=λ1f1+gf=\lambda_{1}f^{1}+g will imply that ff is not a weak extremal, which contradicts Theorem 1.2 in the case when det(𝑻(𝒚))\det(\bm{T}(\bm{y})) is not a perfect square. Considering the case when det(𝑻(𝒚))\det(\bm{T}(\bm{y})) is a perfect square, note that part (iv) of Theorem 1.2 together with Theorem 2.2 imply that ff has to be either a weak extremal or polyconvex. The case of a weak extremal is again ruled out by the equality f=λ1f1+gf=\lambda_{1}f^{1}+g as f1f_{1} is quasiconvex and gg is nonzero and polyconvex. Thus we conclude that ff is polyconvex. Now, in the case when all of sii(𝒚)s_{ii}(\bm{y}) are negative semidefinite, then again by Silvester’s criterion we have that 𝑺(𝒚)\bm{S}(\bm{y}) must be negative semidefinite. Recall next the following classical linear algebra (convex analysis) theorem. It has to do with the fact that the convex cone of all n×nn\times n positive semidefinite symmetric matrices is self-dual.

Theorem 3.2.

Let nn\in\mathbb{N} and let 𝐀=(aij),𝐁=(bij)n×n\bm{A}=(a_{ij}),\bm{B}=(b_{ij})\in\mathbb{R}^{n\times n} be symmetric positive semidefinite matrices. Then the inner product of 𝐀\bm{A} and 𝐁\bm{B} is nonnegative:

𝑨:𝑩=i,j=1naijbij0.\bm{A}\colon\bm{B}=\sum_{i,j=1}^{n}a_{ij}b_{ij}\geq 0.

Note next that (3.13) implies the equality

i,j=13sij(𝒚)cof(𝑻1(𝒚))ij=λ2+λ32λ1λ1λ2λ3det(𝑻(𝒚)),\sum_{i,j=1}^{3}s_{ij}(\bm{y})\mathrm{cof}(\bm{T}^{1}(\bm{y}))_{ij}=\frac{\lambda_{2}+\lambda_{3}-2\lambda_{1}}{\lambda_{1}\lambda_{2}\lambda_{3}}\det(\bm{T}(\bm{y})), (3.16)

where the right hand side is strictly positive a.e. in 3,\mathbb{R}^{3}, while the left hand side is nonpositive due to Thereom 3.2 and the fact that 𝑺\bm{S} is negative semidefinite and cof(𝑻1)\mathrm{cof}(\bm{T}^{1}) is positive semidefinite. This gives a contradiction.
Case 1b. We start by recalling the following theorem by Marcellini [Corollary 1, Reference].

Theorem 3.3 (Marcellini).

Let Q1Q_{1} and Q2Q_{2} be two quadratic forms in n,\mathbb{R}^{n}, with Q2Q_{2} indefinite. If Q1(𝛏)=0Q_{1}(\bm{\xi})=0 for every 𝛏\bm{\xi} such that Q2(𝛏)=0,Q_{2}(\bm{\xi})=0, then there exists λ\lambda\in\mathbb{R} such that Q1=λQ2.Q_{1}=\lambda Q_{2}.

From the fact that

cof(𝑺)33=s11s22s1220,\mathrm{cof}(\bm{S})_{33}=s_{11}s_{22}-s_{12}^{2}\geq 0,

we have s12(𝒚)=0s_{12}(\bm{y})=0 whenever s11(𝒚)=0.s_{11}(\bm{y})=0. As s11s_{11} is indefinite, we have by Marcellini’s theorem that

s12=a12s11for somea12.s_{12}=a_{12}s_{11}\quad\text{for some}\quad a_{12}\in\mathbb{R}. (3.17)

We aim to prove next that s22s_{22} is a multiple of s11s_{11} as well. To that end we recall another lemma proven in [Reference].

Lemma 3.4.

Assume Q(𝛏)Q(\bm{\xi}) is an indefinite quadratic form in nn variables that vanishes at a point 𝛏𝟎=(ξ10,ξ20,,ξn0).\bm{\xi^{0}}=(\xi_{1}^{0},\xi_{2}^{0},\dots,\xi_{n}^{0}). Then given any open neighbourhood UU of the point 𝛏𝟎\bm{\xi^{0}} there exist two open subsets U1,U2UU_{1},U_{2}\subset U such that

Q(𝝃)<0,𝝃U1andQ(𝝃)>0,𝝃U2.Q(\bm{\xi})<0,\quad\bm{\xi}\in U_{1}\quad\text{and}\quad Q(\bm{\xi})>0,\quad\bm{\xi}\in U_{2}.

Let us prove that s11(𝒚)=0s_{11}(\bm{y})=0 implies s22(𝒚)=0.s_{22}(\bm{y})=0. Assume in contradiction s11(𝒚0)=0s_{11}(\bm{y}^{0})=0 and s22(𝒚0)0s_{22}(\bm{y}^{0})\neq 0 for some 𝒚03.\bm{y}^{0}\in\mathbb{R}^{3}. Let the open neighborhood UU of 𝒚0\bm{y}^{0} be such that s22s_{22} does not vanish and does not change sign in U.U. Then by Lemma 3.4, the form s11s_{11} admits both positive and negative values within U,U, thus we can find a point 𝒚1U\bm{y}^{1}\in U such that s11(𝒚1)s22(𝒚1)<0,s_{11}(\bm{y}^{1})s_{22}(\bm{y}^{1})<0, which contradicts the condition cof(𝑺)33=s11s22s1220\mathrm{cof}(\bm{S})_{33}=s_{11}s_{22}-s_{12}^{2}\geq 0. Consequently s11(𝒚)=0s_{11}(\bm{y})=0 implies s22(𝒚)=0,s_{22}(\bm{y})=0, thus again by Marcellini’s theorem above we get

s22=a22s11 for somea22.s_{22}=a_{22}s_{11}\quad\text{ for some}\quad a_{22}\in\mathbb{R}. (3.18)

The above analysis carried out for all other diagonal elements of the cofactor matrix cof(𝑺)\mathrm{cof}(\bm{S}) yields the form of the matrix 𝑺:\bm{S}:

𝑺(𝒚)=s11(𝒚)[[1.5]a11a12a13a12a22a23a13a23a33],\bm{S}(\bm{y})=s_{11}(\bm{y})\cdot\begin{bmatrix}[1.5]a_{11}&a_{12}&a_{13}\\ a_{12}&a_{22}&a_{23}\\ a_{13}&a_{23}&a_{33}\end{bmatrix}, (3.19)

where the matrix 𝑨=(aij)\bm{A}=(a_{ij}) has zero determinant and nonnegative second order principal minors, thus by Silvester’s criterion it is either positive or negative semidefnite. In both cases we have by (3.16) that

s11(𝒚)[𝑨:cof(𝑻1(𝒚))]=λ2+λ32λ1λ1λ2λ3det(𝑻(𝒚)),s_{11}(\bm{y})\cdot[\bm{A}\colon\mathrm{cof}(\bm{T}^{1}(\bm{y}))]=\frac{\lambda_{2}+\lambda_{3}-2\lambda_{1}}{\lambda_{1}\lambda_{2}\lambda_{3}}\det(\bm{T}(\bm{y})), (3.20)

where the right hand side of (3.20) takes strictly positive values a.e. in 3,\mathbb{R}^{3}, while the left hand side does not, due to the constant sign of the inner product 𝑨:cof(𝑻1(𝒚)),\bm{A}\colon\mathrm{cof}(\bm{T}^{1}(\bm{y})), the semi-definiteness of s11,s_{11}, and Lemma 3.4. This does finish Case 1b.
Case 2. Like in Case 1 we first prove that all diagonal entries of cof(𝑺)\mathrm{cof}(\bm{S}) are nonnegative. For the entry cof(𝑺)33\mathrm{cof}(\bm{S})_{33} we have that it is either identically zero, or if not then the zero set {𝒚3:cof(𝑺)33(𝒚)=0}\{\bm{y}\in\mathbb{R}^{3}\ :\ \mathrm{cof}(\bm{S})_{33}(\bm{y})=0\} has zero Lebesgue measure, thus the steps leading to (3.15) go through and we obtain by (3.15) that

cof(𝑺(𝒚))33[(r,q,1)𝑻1(𝒚)(r,q,1)T]0.\mathrm{cof}(\bm{S}(\bm{y}))_{33}\cdot[(-r,-q,1)\bm{T}^{1}(\bm{y})(-r,-q,1)^{T}]\equiv 0. (3.21)

For the points 𝒚3\bm{y}\in\mathbb{R}^{3} with cof(S(𝒚))330\mathrm{cof}(S(\bm{y}))_{33}\neq 0 we will get that (r,q,1)𝑻1(𝒚)(r,q,1)T=0,(-r,-q,1)\bm{T}^{1}(\bm{y})(-r,-q,1)^{T}=0, hence, as the vector (r,q,1)0(-r,-q,1)\neq 0 we obtain det(𝑻1(𝒚))=0\det(\bm{T}^{1}(\bm{y}))=0 by the positive semi-definiteness of 𝑻1.\bm{T}^{1}. Consequently as the zero determinant set {𝒚:det(𝑻1(𝒚))=0}\{\bm{y}\ :\ \det(\bm{T}^{1}(\bm{y}))=0\} is the same as the set {𝒚:det(𝑻(𝒚))=0},\{\bm{y}\ :\ \det(\bm{T}(\bm{y}))=0\}, which is a null set, then cof(𝑺(𝒚))330\mathrm{cof}(\bm{S}(\bm{y}))_{33}\equiv 0. This argument yields the conditions:

cof(𝑺(𝒚))11=cof(𝑺(𝒚))22=cof(𝑺(𝒚))330.\mathrm{cof}(\bm{S}(\bm{y}))_{11}=\mathrm{cof}(\bm{S}(\bm{y}))_{22}=\mathrm{cof}(\bm{S}(\bm{y}))_{33}\equiv 0. (3.22)

Once (3.22) is established we will get the desired results following the steps in Case 1a and Case 1b, where (3.16) and the fact that the coefficient λ2+λ32λ1λ1λ2λ3\frac{\lambda_{2}+\lambda_{3}-2\lambda_{1}}{\lambda_{1}\lambda_{2}\lambda_{3}} on the right hand side is positive were used.
Case 3. The proof of this case immediately follows from (3.5), (3.6), and (3.8). Indeed bearing in mind that α>0\alpha>0 and det(𝑻1)=αdet(𝑻)>0\det(\bm{T}^{1})=\alpha\det(\bm{T})>0 a.e. in 3\mathbb{R}^{3}, putting together (3.5), (3.6), and (3.8) and making the change of variables t=λλ1t=\frac{\lambda}{\lambda_{1}} we obtain

det((λ1𝑻1)1/2𝑻(λ1𝑻1)1/2t𝑰)=(1t)3,t,𝒚3.\det((\lambda_{1}\bm{T}^{1})^{-1/2}\bm{T}(\lambda_{1}\bm{T}^{1})^{-1/2}-t\bm{I})=(1-t)^{3},\qquad t\in\mathbb{R},\ \bm{y}\in\mathbb{R}^{3}.

The last identity implies that the diagonal form of the symmetric matrix (λ1𝑻1)1/2𝑻(λ1𝑻1)1/2(\lambda_{1}\bm{T}^{1})^{-1/2}\bm{T}(\lambda_{1}\bm{T}^{1})^{-1/2} must coincide with the identity matrix for all 𝒚3,\bm{y}\in\mathbb{R}^{3}, thus we get 𝑻=λ1𝑻1,\bm{T}=\lambda_{1}\bm{T}^{1}, i.e., 𝑻1\bm{T}^{1} is a multiple of 𝑻.\bm{T}.

4 Proof of Theorem 2.2

Proof of Theorem 2.2.

Let us mention that some parts of the proof are borrowed from [Reference] with minor changes, but we choose to repeat them here for the convenience of the reader. Assume the quadratic form f(𝝃)=𝝃𝑪𝝃T=𝒙T𝑻(𝒚)𝒙f(\bm{\xi})=\bm{\xi}\bm{C}\bm{\xi}^{T}=\bm{x}^{T}\bm{T}(\bm{y})\bm{x} is quasiconvex such that det(𝑻(𝒚))0\det(\bm{T}(\bm{y}))\equiv 0 for 𝒚3.\bm{y}\in\mathbb{R}^{3}. We will basically prove here that if the entries of a symmetric matrix 𝑻(𝒚)3×3\bm{T}(\bm{y})\in\mathbb{R}^{3\times 3} are quadratic forms in 𝒚3,\bm{y}\in\mathbb{R}^{3}, such that 𝑻(𝒚)\bm{T}(\bm{y}) is positive semidefinite for all 𝒚3\bm{y}\in\mathbb{R}^{3} and det(𝑻(𝒚))0,\det(\bm{T}(\bm{y}))\equiv 0, then the biquadratic form f(𝒙𝒚)=𝒙T𝑻(𝒚)𝒙f(\bm{x}\otimes\bm{y})=\bm{x}^{T}\bm{T}(\bm{y})\bm{x} is a sum of squares. We can without loss of generality assume that the third row of 𝑻\bm{T} is a linear combination of the first two for a.e. 𝒚3\bm{y}\in\mathbb{R}^{3} with rational coefficients a(𝒚)a(\bm{y}) and b(𝒚),b(\bm{y}), thus due to the symmetry, the matrix 𝑻(𝒚)\bm{T}(\bm{y}) must have the form

𝑻(𝒚)=[[1.5]t11t12at11+bt12t12t22at12+bt22at11+bt12at12+bt22a2t11+b2t22+2abt12],\bm{T}(\bm{y})=\begin{bmatrix}[1.5]t_{11}&t_{12}&at_{11}+bt_{12}\\ t_{12}&t_{22}&at_{12}+bt_{22}\\ at_{11}+bt_{12}&at_{12}+bt_{22}&a^{2}t_{11}+b^{2}t_{22}+2abt_{12}\end{bmatrix}, (4.1)

where the rational functions aa and bb are given by

a=cof(𝑻)13cof(𝑻)33,b=cof(𝑻)23cof(𝑻)33.a=\frac{\mathrm{cof}(\bm{T})_{13}}{\mathrm{cof}(\bm{T})_{33}},\qquad b=-\frac{\mathrm{cof}(\bm{T})_{23}}{\mathrm{cof}(\bm{T})_{33}}. (4.2)

For the form ff we get

f(𝒙𝒚)\displaystyle f(\bm{x}\otimes\bm{y}) =x12t11+2x1x2t12+x22t22+2x1x3(at11+bt12)\displaystyle=x_{1}^{2}t_{11}+2x_{1}x_{2}t_{12}+x_{2}^{2}t_{22}+2x_{1}x_{3}(at_{11}+bt_{12}) (4.3)
+2x2x3(at12+bt22)+x32(a2t11+b2t22+2abt12).\displaystyle+2x_{2}x_{3}(at_{12}+bt_{22})+x_{3}^{2}(a^{2}t_{11}+b^{2}t_{22}+2abt_{12}).

Next we have from the fact det𝑻(𝒚)=0\det{\bm{T}(\bm{y})}=0 for all y3y\in\mathbb{R}^{3} that

rank(cof(𝑻(𝒚))1,for all𝒚3.\mathrm{rank}(\mathrm{cof}(\bm{T}(\bm{y}))\leq 1,\quad\text{for all}\quad\bm{y}\in\mathbb{R}^{3}. (4.4)

In order to make sense of (4.2) with no fear about the denominators vanishing, we need to consider the case cof(𝑻(𝒚))330\mathrm{cof}(\bm{T}(\bm{y}))_{33}\equiv 0 for all 𝒚3\bm{y}\in\mathbb{R}^{3} separately. Thus assume first it is the case. Observe that if one of the diagonal elements of 𝑻(𝒚)\bm{T}(\bm{y}) is identically zero, then by the positive semidefiniteness of 𝑻(𝒚)\bm{T}(\bm{y}) the elements in the same row and column of 𝑻(𝒚)\bm{T}(\bm{y}) must be identically zero too, thus f(𝒙𝒚)f(\bm{x}\otimes\bm{y}) becomes a 2×22\times 2 form, and its quasiconvexity automatically implies convexity. Assume now that all diagonal entries of 𝑻(𝒚)\bm{T}(\bm{y}) are nonzero positive semidefinite quadratic forms in 𝒚3.\bm{y}\in\mathbb{R}^{3}. By the positive semi-definiteness of cof(𝑻(𝒚)),\mathrm{cof}(\bm{T}(\bm{y})), we then have

cof(𝑻(𝒚))33=cof(𝑻(𝒚))32=cof(𝑻(𝒚))310,𝒚3.\mathrm{cof}(\bm{T}(\bm{y}))_{33}=\mathrm{cof}(\bm{T}(\bm{y}))_{32}=\mathrm{cof}(\bm{T}(\bm{y}))_{31}\equiv 0,\quad\bm{y}\in\mathbb{R}^{3}. (4.5)

The conditions (4.5) imply that the matrix obtained from 𝑻(𝒚)\bm{T}(\bm{y}) by removing the last row has rank at most one. We will obtain a more explicit form of 𝑻(𝒚)\bm{T}(\bm{y}) by means of the following representation lemma proven in [Reference].

Lemma 4.1.

Assume 𝐱d\bm{x}\in\mathbb{R}^{d} and 𝐀(𝐱)=(aij(𝐱)),\bm{A}(\bm{x})=(a_{ij}(\bm{x})), i=1,,m,j=1,,ni=1,\dots,m,\ j=1,\dots,n is an m×nm\times n matrix with polynomial coefficients, such that each entry aij(x)a_{ij}(x) is a homogeneous polynomial of degree 2p,2p, where m,n,d,p.m,n,d,p\in\mathbb{N}. If rank(𝐀(𝐱))1\mathrm{rank}(\bm{A}(\bm{x}))\leq 1 for all 𝐱d\bm{x}\in\mathbb{R}^{d}, then there exist homogeneous polynomials bi(𝐱)b_{i}(\bm{x}) and ci(𝐱),c_{i}(\bm{x}), such that aij(𝐱)=bi(𝐱)cj(𝐱),a_{ij}(\bm{x})=b_{i}(\bm{x})c_{j}(\bm{x}), for i=1,,m,j=1,,n.i=1,\dots,m,\ j=1,\dots,n.

If t11(𝒚)t_{11}(\bm{y}) is irreducible in the field of reals, then by Lemma 4.1 and the obtained rank conditions, the matrix 𝑻(𝒚)\bm{T}(\bm{y}) must have the form

[[1.5]PαPβPαPα2PαβPβPαβPQ],α,β,\begin{bmatrix}[1.5]P&\alpha P&\beta P\\ \alpha P&\alpha^{2}P&\alpha\beta P\\ \beta P&\alpha\beta P&Q\end{bmatrix},\quad\alpha,\beta\in\mathbb{R}, (4.6)

where PP and QQ are nonzero positive semidefinite quadratic forms and α,β\alpha,\beta\in\mathbb{R} with α0.\alpha\neq 0. In the same way if t11(𝒚)t_{11}(\bm{y}) is reducible in the field of reals, then it must be the square of a linear form, thus again by Lemma 4.1 and the obtained rank conditions the matrix 𝑻(𝒚)\bm{T}(\bm{y}) must have the form

[[1.5]l12l1l2l1l3l1l2l22l2l3l1l3l2l3Q],\begin{bmatrix}[1.5]l_{1}^{2}&l_{1}l_{2}&l_{1}l_{3}\\ l_{1}l_{2}&l_{2}^{2}&l_{2}l_{3}\\ l_{1}l_{3}&l_{2}l_{3}&Q\end{bmatrix}, (4.7)

where QQ is a nonzero positive semidefinite quadratic form and l1,l2,l3l_{1},l_{2},l_{3} are linear forms with l1l_{1} and l2l_{2} nonzero. In the situation of (4.6) we have that

0cof(𝑻(𝒚))11=α2P(𝒚)(Q(𝒚)β2P(𝒚)),0\leq\mathrm{cof}(\bm{T}(\bm{y}))_{11}=\alpha^{2}P(\bm{y})(Q(\bm{y})-\beta^{2}P(\bm{y})),

thus

Q(𝒚)β2P(𝒚)0,Q(\bm{y})-\beta^{2}P(\bm{y})\geq 0,

which implies, that the form Q(𝒚)β2P(𝒚)Q(\bm{y})-\beta^{2}P(\bm{y}) is convex, i.e.,

Q(𝒚)=β2P(𝒚)+R(𝒚),Q(\bm{y})=\beta^{2}P(\bm{y})+R(\bm{y}), (4.8)

where R(𝒚)R(\bm{y}) is convex. Consequently we get

f(𝒙𝒚)=P(𝒚)(x1+αx2+βx3)2+R(𝒚)x32f(\bm{x}\otimes\bm{y})=P(\bm{y})(x_{1}+\alpha x_{2}+\beta x_{3})^{2}+R(\bm{y})x_{3}^{2}

is polyconvex. In the situation of (4.7), similarly we get that

Q(𝒚)=l32(𝒚)+R(𝒚),Q(\bm{y})=l_{3}^{2}(\bm{y})+R(\bm{y}),

where RR is convex and thus

f(𝒙𝒚)=(l1(𝒚)x1+l2(𝒚)x2+l3(𝒚)x3)2+R(𝒚)x32,f(\bm{x}\otimes\bm{y})=(l_{1}(\bm{y})x_{1}+l_{2}(\bm{y})x_{2}+l_{3}(\bm{y})x_{3})^{2}+R(\bm{y})x_{3}^{2},

and is thus polyconvex. In what follows we will assume that all diagonal entries of both matrices 𝑻(𝒚)\bm{T}(\bm{y}) and cof(𝑻(𝒚))\mathrm{cof}(\bm{T}(\bm{y})) are nonzero. Note that this in particular implies that any of the three rows of 𝑻\bm{T} is a linear combination of the remaining two, as we have for the third row in (4.1). The rank condition (4.4) implies by Lemma 4.1 that the cofactor matrix cof(𝑻)\mathrm{cof}(\bm{T}) has the form

cof(𝑻)=[[1.5]c1d1c1d2c1d3c2d1c2d2c2d3c3d1c3d2c3d3],\mathrm{cof}(\bm{T})=\\ \begin{bmatrix}[1.5]c_{1}d_{1}&c_{1}d_{2}&c_{1}d_{3}\\ c_{2}d_{1}&c_{2}d_{2}&c_{2}d_{3}\\ c_{3}d_{1}&c_{3}d_{2}&c_{3}d_{3}\end{bmatrix}, (4.9)

for some homogeneous polynomials ci(𝒚)c_{i}(\bm{y}) and di(𝒚),d_{i}(\bm{y}), i=1,2,3i=1,2,3 and 𝒚3.\bm{y}\in\mathbb{R}^{3}. As all diagonal entries of cof(𝑻(𝒚))\mathrm{cof}(\bm{T}(\bm{y})) are polynomials of degree four, and deg(cidj)4\deg(c_{i}d_{j})\leq 4 for i,j=1,2,3,i,j=1,2,3, we must in fact have

deg(cidj)=4,for alli,j=1,2,3.\deg(c_{i}d_{j})=4,\quad\text{for all}\quad i,j=1,2,3. (4.10)

The cofactor matrix cof(𝑻(𝒚))\mathrm{cof}(\bm{T}(\bm{y})) must be positive semidefinite for all 𝒚3\bm{y}\in\mathbb{R}^{3} given that 𝑻(𝒚)\bm{T}(\bm{y}) is such, thus we get the set of inequalities

ci(𝒚)di(𝒚)0,for all𝒚3,i=1,2,3.c_{i}(\bm{y})d_{i}(\bm{y})\geq 0,\quad\text{for all}\quad\bm{y}\in\mathbb{R}^{3},\ \ i=1,2,3. (4.11)

Next we aim to come up with a more explicit form of cof(𝑻(𝒚))\mathrm{cof}(\bm{T}(\bm{y})) using the obtained conditions (4.9)-(4.11). To that end we consider the cases deg(c1)=0,1,2\deg(c_{1})=0,1,2 separately (note that the case deg(c1)>2\deg(c_{1})>2 implies deg(d1)<2\deg(d_{1})<2 and we can consider d1d_{1} instead of c1).c_{1}).
Case 1: deg(c1)=0\deg(c_{1})=0 .
Case 2: deg(c1)=1\deg(c_{1})=1 .
Case 3: deg(c1)=2\deg(c_{1})=2 .
Next we examine each case in detail.
Case 1. In this case we have from (4.10) that

ci(𝒚)=ci,deg(di)=4for alli=1,2,3,c_{i}(\bm{y})=c_{i}\in\mathbb{R},\quad\deg(d_{i})=4\quad\text{for all}\quad i=1,2,3,

which gives by (4.2) a(𝒚)=c1c3=aa(\bm{y})=\frac{c_{1}}{c_{3}}=a\in\mathbb{R} and b(𝒚)=c2c3=b.b(\bm{y})=-\frac{c_{2}}{c_{3}}=b\in\mathbb{R}. Consequently we get by (4.3)

f(𝒙𝒚)=t11(𝒚)(x1+ax3)2+2t12(𝒚)(x1+ax3)(x2+bx3)+t22(𝒚)(x2+bx3)2.f(\bm{x}\otimes\bm{y})=t_{11}(\bm{y})(x_{1}+ax_{3})^{2}+2t_{12}(\bm{y})(x_{1}+ax_{3})(x_{2}+bx_{3})+t_{22}(\bm{y})(x_{2}+bx_{3})^{2}. (4.12)

Terpstra has proven in [Reference] the following classical result.

Theorem 4.2 (Terpstra).

Any 2×n2\times n (or n×2n\times 2) quasiconvex quadratic form is polyconvex.

Note that Terpstra’s theorem implies that any 2×n2\times n (or n×2n\times 2) nonnegative biquadratic form Q(𝒙,𝒚)Q(\bm{x},\bm{y}) is in fact a sum of squares. Introducing the new independent variables X1=x1+ax3X_{1}=x_{1}+ax_{3} and X2=x2+bx3X_{2}=x_{2}+bx_{3} we have that the 2×32\times 3 biquadratic form

g(𝑿,𝒚)=t11(𝒚)X12+2t12(𝒚)X1X2+t22(𝒚)X22g(\bm{X},\bm{y})=t_{11}(\bm{y})X_{1}^{2}+2t_{12}(\bm{y})X_{1}X_{2}+t_{22}(\bm{y})X_{2}^{2}

is nonnegative, thus by Terpstra’s theorem above gg must be the sum of squares of 2-homogeneous forms that are linear combinations of Xiyj,X_{i}y_{j}, i.e., ff is polyconvex.
Case 2. On one hand we similarly have from (4.10) that

deg(ci)=1,deg(di)=3for alli=1,2,3.\deg(c_{i})=1,\quad\deg(d_{i})=3\quad\text{for all}\quad i=1,2,3. (4.13)

On the other hand we have from (4.11) that cic_{i} must divide did_{i} for each 1i3,1\leq i\leq 3, thus we get the form of cof(𝑻):\mathrm{cof}(\bm{T}):

cof(𝑻)=[[1.5]c12Pc1c2Qc1c3Rc1c2Pc22Qc2c3Rc1c3Pc2c3Qc32R],\mathrm{cof}(\bm{T})=\begin{bmatrix}[1.5]c_{1}^{2}P&c_{1}c_{2}Q&c_{1}c_{3}R\\ c_{1}c_{2}P&c_{2}^{2}Q&c_{2}c_{3}R\\ c_{1}c_{3}P&c_{2}c_{3}Q&c_{3}^{2}R\end{bmatrix}, (4.14)

where ci(𝒚)c_{i}(\bm{y}) are linear forms and P(𝒚),Q(𝒚),R(𝒚)P(\bm{y}),Q(\bm{y}),R(\bm{y}) are positive semidefinite quadratic forms. Again we get from (4.2) that a(𝒚)=c1(𝒚)c3(𝒚)a(\bm{y})=\frac{c_{1}(\bm{y})}{c_{3}(\bm{y})} and b(𝒚)=c2(𝒚)c3(𝒚),b(\bm{y})=-\frac{c_{2}(\bm{y})}{c_{3}(\bm{y})}, and utilizing (4.3):

f(𝒙𝒚)=1c32[t11(𝒚)(x1c3+x3c1)2+2t12(𝒚)(x1c3+x3c1)(x2c3x3c2)+t22(𝒚)(x2c3x3c2)2].f(\bm{x}\otimes\bm{y})=\frac{1}{c_{3}^{2}}\left[t_{11}(\bm{y})(x_{1}c_{3}+x_{3}c_{1})^{2}+2t_{12}(\bm{y})(x_{1}c_{3}+x_{3}c_{1})(x_{2}c_{3}-x_{3}c_{2})+t_{22}(\bm{y})(x_{2}c_{3}-x_{3}c_{2})^{2}\right]. (4.15)

The goal is to show that we can obtain necessary factorizations and abbreviations to end up with a 2×32\times 3 nonnegative biquadratic form that then must be polyconvex by Terpstra’s theorem. Consider the biquadratic form gg in the variables 𝒚\bm{y} and 𝑿=(X1,X2)\bm{X}=(X_{1},X_{2}) given by

g(𝑿,𝒚)=t11(𝒚)X12+2t12(𝒚)X1X2+t22(𝒚)X22.g(\bm{X},\bm{y})=t_{11}(\bm{y})X_{1}^{2}+2t_{12}(\bm{y})X_{1}X_{2}+t_{22}(\bm{y})X_{2}^{2}. (4.16)

We have that g(x1c3+x3c1,x2c3x3c2,𝒚)=f(𝒙𝒚)c320g(x_{1}c_{3}+x_{3}c_{1},x_{2}c_{3}-x_{3}c_{2},\bm{y})=\frac{f(\bm{x}\otimes\bm{y})}{c_{3}^{2}}\geq 0 for all 𝒙,𝒚3\bm{x},\bm{y}\in\mathbb{R}^{3} such that c3(𝒚)0,c_{3}(\bm{y})\neq 0, thus we obtain

g(x1c3+x3c1,x2c3x3c2,𝒚)0,for all𝒙,𝒚3,g(x_{1}c_{3}+x_{3}c_{1},x_{2}c_{3}-x_{3}c_{2},\bm{y})\geq 0,\quad\text{for all}\quad\bm{x},\bm{y}\in\mathbb{R}^{3}, (4.17)

by continuity as the set c3(𝒚)=0c_{3}(\bm{y})=0 is just a hyperplane in 3.\mathbb{R}^{3}. For any fixed 𝒚3\bm{y}\in\mathbb{R}^{3} such that c3(𝒚)0c_{3}(\bm{y})\neq 0 and for any fixed values X1,X2,X_{1},X_{2}, we can find values x1,x2,x3x_{1},x_{2},x_{3}\in\mathbb{R} such that x1c3+x3c1=X1x_{1}c_{3}+x_{3}c_{1}=X_{1} and x2c3x3c2=X2x_{2}c_{3}-x_{3}c_{2}=X_{2} (for instance take x1=X1c3(𝒚),x_{1}=\frac{X_{1}}{c_{3}(\bm{y})}, x2=X2c3(𝒚),x_{2}=\frac{X_{2}}{c_{3}(\bm{y})}, and x3=0x_{3}=0), thus again by continuity we get from (4.17) the condition

g(𝑿,𝒚)=t11(𝒚)X12+2t12(𝒚)X1X2+t22(𝒚)X220for all𝑿2,𝒚3.g(\bm{X},\bm{y})=t_{11}(\bm{y})X_{1}^{2}+2t_{12}(\bm{y})X_{1}X_{2}+t_{22}(\bm{y})X_{2}^{2}\geq 0\quad\text{for all}\quad\bm{X}\in\mathbb{R}^{2},\bm{y}\in\mathbb{R}^{3}. (4.18)

Hence gg is a 2×32\times 3 nonnegative biquadratic form, and by Terpstra’s theorem it must be a sum of squares of 2-homogeneous forms in Xiyj:X_{i}y_{j}:

g(𝑿,𝒚)=t11(𝒚)X12+2t12(𝒚)X1X2+t22(𝒚)X22=k=16(ak(𝒚)X1+bk(𝒚)X2)2,g(\bm{X},\bm{y})=t_{11}(\bm{y})X_{1}^{2}+2t_{12}(\bm{y})X_{1}X_{2}+t_{22}(\bm{y})X_{2}^{2}=\sum_{k=1}^{6}(a_{k}(\bm{y})X_{1}+b_{k}(\bm{y})X_{2})^{2}, (4.19)

where ak(𝒚)a_{k}(\bm{y}) and bk(𝒚)b_{k}(\bm{y}) are linear forms in 𝒚.\bm{y}. Thus we discover

f(𝒙𝒚)=1c32k=16(ak(x1c3+x3c1)+bk(x2c3x3c2))2=1c32k=16[c3(akx1+bkx2)+x3(akc1bkc2)]2.f(\bm{x}\otimes\bm{y})=\frac{1}{c_{3}^{2}}\sum_{k=1}^{6}(a_{k}(x_{1}c_{3}+x_{3}c_{1})+b_{k}(x_{2}c_{3}-x_{3}c_{2}))^{2}=\frac{1}{c_{3}^{2}}\sum_{k=1}^{6}[c_{3}(a_{k}x_{1}+b_{k}x_{2})+x_{3}(a_{k}c_{1}-b_{k}c_{2})]^{2}. (4.20)

Equating the coefficients of x32x_{3}^{2} in the original form of ff and in (4.20) we get the key equality

k=16(akc1bkc2)2=c32t33.\sum_{k=1}^{6}(a_{k}c_{1}-b_{k}c_{2})^{2}=c_{3}^{2}t_{33}. (4.21)

The condition (4.21) in particular implies that ak(𝒚)c1(𝒚)bk(𝒚)c2(𝒚)=0a_{k}(\bm{y})c_{1}(\bm{y})-b_{k}(\bm{y})c_{2}(\bm{y})=0 for all k=1,2,,6k=1,2,\dots,6 whenever c3(𝒚)=0.c_{3}(\bm{y})=0. Since c3c_{3} is linear, this means that c3c_{3} divides akc1bkc2a_{k}c_{1}-b_{k}c_{2} for all k,k, thus we get

ak(𝒚)c1(𝒚)bk(𝒚)c2(𝒚)=hk(𝒚)c3(𝒚),k=1,2,,6,a_{k}(\bm{y})c_{1}(\bm{y})-b_{k}(\bm{y})c_{2}(\bm{y})=h_{k}(\bm{y})c_{3}(\bm{y}),\quad k=1,2,\dots,6, (4.22)

for some linear forms hk(𝒚).h_{k}(\bm{y}). Plugging the obtained forms of akc1+bkc2a_{k}c_{1}+b_{k}c_{2} back into (4.20), we obtain

f(𝒙𝒚)=k=16(akx1+bkx2+hkx3)2,f(\bm{x}\otimes\bm{y})=\sum_{k=1}^{6}(a_{k}x_{1}+b_{k}x_{2}+h_{k}x_{3})^{2}, (4.23)

i.e, ff is polyconvex.
Case 3. We have due to (4.6) that deg(di)=deg(ci)=2\deg(d_{i})=\deg(c_{i})=2 for all i=1,2,3.i=1,2,3. The following two cases are qualitatively different:
Case 3a: c3(y)c_{3}(\bm{y}) is indefinite.
Case 3b: c3(y)c_{3}(\bm{y}) is positive semidefinite.
Case 3a.
It is easy to verify that the steps (4.15)-(4.21) go through in this case too, thus we have

k=16(akc1bkc2)2=c32t33,\sum_{k=1}^{6}(a_{k}c_{1}-b_{k}c_{2})^{2}=c_{3}^{2}t_{33}, (4.24)

where aka_{k} and bkb_{k} are linear forms for k=1,2,,6.k=1,2,\dots,6. Equality (4.24) and positivity of both sides imply that

ak(𝒚)c1(𝒚)bk(𝒚)c2(𝒚)=0,wheneverc3(𝒚)=0,for some𝒚3.a_{k}(\bm{y})c_{1}(\bm{y})-b_{k}(\bm{y})c_{2}(\bm{y})=0,\quad\text{whenever}\quad c_{3}(\bm{y})=0,\quad\text{for some}\quad\bm{y}\in\mathbb{R}^{3}. (4.25)

Next we prove the following simple lemma.

Lemma 4.3.

Let Q(𝐲)Q(\bm{y}) be an indefinite quadratic form in the variable 𝐲3\bm{y}\in\mathbb{R}^{3} and let P(𝐲)P(\bm{y}) be a third order homogeneous polynomial. If P(𝐲)=0P(\bm{y})=0 whenever Q(𝐲)=0Q(\bm{y})=0 for some 𝐲3,\bm{y}\in\mathbb{R}^{3}, then QQ must divide P.P.

Proof of Lemma 4.3.

As QQ is indefinite, we can without loss of generality assume that it has one of the normal forms:

Q(𝒚)=y12y22,orQ(𝒚)=y12+y22y32.Q(\bm{y})=y_{1}^{2}-y_{2}^{2},\quad\text{or}\quad Q(\bm{y})=y_{1}^{2}+y_{2}^{2}-y_{3}^{2}.

In the first case we have Q=(y1y2)(y1+y2)Q=(y_{1}-y_{2})(y_{1}+y_{2}) and thus P(𝒚)=0P(\bm{y})=0 whenever one of the linear forms y1y2y_{1}-y_{2} or y1+y2y_{1}+y_{2} vanishes, thus obviously PP is divisible by both, and hence by their product too (by the unique factorization over the field of reals). In the case Q(𝒚)=y12+y22y32Q(\bm{y})=y_{1}^{2}+y_{2}^{2}-y_{3}^{2} we can separate the multiple of y32y_{3}^{2} within PP and write

P(𝒚)=(y12+y22y32)l(𝒚)+y3φ(y1,y2)+ψ(y1,y2),P(\bm{y})=(y_{1}^{2}+y_{2}^{2}-y_{3}^{2})l(\bm{y})+y_{3}\varphi(y_{1},y_{2})+\psi(y_{1},y_{2}), (4.26)

where ll is a linear form in 𝒚\bm{y}, and φ\varphi and ψ\psi are homogeneous forms in y1y_{1} and y2y_{2} of degree two and three, respectively. Now for any y1,y2y_{1},y_{2}\in\mathbb{R} such that y12+y22>0,y_{1}^{2}+y_{2}^{2}>0, by choosing y3=±y12+y22y_{3}=\pm\sqrt{y_{1}^{2}+y_{2}^{2}} we get the system

±y12+y22φ(y1,y2)+ψ(y1,y2)=0,\pm\sqrt{y_{1}^{2}+y_{2}^{2}}\cdot\varphi(y_{1},y_{2})+\psi(y_{1},y_{2})=0, (4.27)

which implies first ψ(y1,y2)=0\psi(y_{1},y_{2})=0 and then φ(y1,y2)=0,\varphi(y_{1},y_{2})=0, i.e., ψ\psi and φ\varphi identically vanish and thus P=Ql.P=Ql.

Consequently, applying the lemma for the pairs akc1bkc2a_{k}c_{1}-b_{k}c_{2} and c3c_{3} we get

akc1bkc2=c3hk,k=1,2,,6,a_{k}c_{1}-b_{k}c_{2}=c_{3}h_{k},\quad k=1,2,\dots,6, (4.28)

for some linear forms hkh_{k}. Owing back to the form of ff in (4.20) we arrive at

f(𝒙𝒚)=k=16(akx1+bkx2+hkx3)2,f(\bm{x}\otimes\bm{y})=\sum_{k=1}^{6}(a_{k}x_{1}+b_{k}x_{2}+h_{k}x_{3})^{2},

utilizing (4.28), i.e., ff is polyconvex.
Case 3b. We can assume without loss of generality that all forms cic_{i} and did_{i} are semidefinite (positive or negative). We divide this case further into two possible cases.
Case 3ba: c3(y)c_{3}(\bm{y}) is reducible in the field of reals.
Case 3bb: c3(y)c_{3}(\bm{y}) is irreducible in the field of reals.
Case 3ba
. As c3c_{3} is semidefinite and reducible, it must be a multiple of the square of a linear form, i.e, c3(𝒚)=σl2(𝒚),c_{3}(\bm{y})=\sigma l^{2}(\bm{y}), where σ{1,1}\sigma\in\{-1,1\} and l(𝒚)l(\bm{y}) is linear. Again, we can easily verify that the steps (4.15)-(4.21) go through in this case too, thus we have

k=16(akc1bkc2)2=c32t33=l4t33,\sum_{k=1}^{6}(a_{k}c_{1}-b_{k}c_{2})^{2}=c_{3}^{2}t_{33}=l^{4}t_{33}, (4.29)

where aka_{k} and bkb_{k} are linear forms for k=1,2,,6.k=1,2,\dots,6. Equality (4.29), the linearity of l,l, and positivity of both parts of the equality imply that all the 33-homogeneous forms akc1bkc2a_{k}c_{1}-b_{k}c_{2} contain a factor of ll for all k.k. After factoring an l2l^{2} out in (4.29) we get for the same reason that all of the forms akc1bkc2a_{k}c_{1}-b_{k}c_{2} contain a factor of l2,l^{2}, i.e., c3.c_{3}. Denoting akc1bkc2=c3hka_{k}c_{1}-b_{k}c_{2}=c_{3}h_{k} for k=1,2,,6k=1,2,\dots,6 we again end up with the form

f(𝒙𝒚)=k=16(akx1+bkx2+hkx3)2,f(\bm{x}\otimes\bm{y})=\sum_{k=1}^{6}(a_{k}x_{1}+b_{k}x_{2}+h_{k}x_{3})^{2},

by (4.20), hence ff is polyconvex.
Case 3ba. We assume without loss of generality that cic_{i} and did_{i} are all irreducible and semidefinite for i=1,2,3.i=1,2,3. From the fact that the cofactor matrix cof(𝑻)\mathrm{cof}(\bm{T}) is symmetric, we have the set of equalities

cidj=cjdifor alli,j=1,2,3.c_{i}d_{j}=c_{j}d_{i}\quad\text{for all}\quad i,j=1,2,3. (4.30)

In the case when c1c_{1} and d1d_{1} are linearly independent, we get from the equality c1di=d1ci,c_{1}d_{i}=d_{1}c_{i}, i=2,3,i=2,3, from the irreducibility of the factors in it, and from the unique factorization of homogeneous polynomials in the field of reals, that c2=αc1c_{2}=\alpha c_{1} and c3=βc1c_{3}=\beta c_{1} for some nonzero α,β.\alpha,\beta\in\mathbb{R}. Again, the steps (4.15)-(4.21) go through and thus (4.20) yields the polyconvex form for f:f:

f(𝒙𝒚)=k=16[akx1+bkx2+x3β(akαbk)]2.f(\bm{x}\otimes\bm{y})=\sum_{k=1}^{6}[a_{k}x_{1}+b_{k}x_{2}+\frac{x_{3}}{\beta}(a_{k}-\alpha b_{k})]^{2}.

In what follows we can assume that cic_{i} and did_{i} are linearly dependent for all i=1,2,3.i=1,2,3. After a change of sign in all of cic_{i} and did_{i} if necessary, multiplying all of did_{i} by the same factor we can assume that all of cic_{i} are irreducible, positive semidefnite, and

di=cifor alli=1,2,3.d_{i}=c_{i}\quad\text{for all}\quad i=1,2,3. (4.31)

Recall next that by (4.9) we have

c32+t122=t11t22,c22+t132=t11t33,c12+t232=t22t33,c_{3}^{2}+t_{12}^{2}=t_{11}t_{22},\qquad c_{2}^{2}+t_{13}^{2}=t_{11}t_{33},\qquad c_{1}^{2}+t_{23}^{2}=t_{22}t_{33}, (4.32)

where we aim at examining the above conditions by factoring in the complex field as

(t12+ic3)(t12ic3)=t11t22.(t_{12}+ic_{3})(t_{12}-ic_{3})=t_{11}t_{22}. (4.33)

Because each of the factors above is a second order homogeneous polynomial in 𝒚,\bm{y}, the factors on the right hand side are purely real and the factors on the left hand side have nonzero imaginary parts, then again by the unique factorization in the field of complex numbers, we discover that both multipliers t12+ic3t_{12}+ic_{3} and t12ic3t_{12}-ic_{3} must be reducible, and thus both t11t_{11} and t22t_{22} must split into the product of two linear forms with complex coefficients as well, i.e.,

t11=(u1+iu1)(v1iv1),t22=(u2+iu2)(v2iv2),t_{11}=(u_{1}+iu_{1}^{\prime})(v_{1}-iv_{1}^{\prime}),\qquad t_{22}=(u_{2}+iu_{2}^{\prime})(v_{2}-iv_{2}^{\prime}), (4.34)

where u1,v1,u1,v1u_{1},v_{1},u_{1}^{\prime},v_{1}^{\prime} are real linear forms. From the fact that t11t_{11} is real we get u1v1=v1u1,u_{1}v_{1}^{\prime}=v_{1}u_{1}^{\prime}, which eventually implies from the unique factorization and positive semidefiniteness of t11t_{11} that we can choose u1,v1,u1,v1u_{1},v_{1},u_{1}^{\prime},v_{1}^{\prime} so that u1=v1u_{1}=v_{1} and u1=v1,u_{1}^{\prime}=v_{1}^{\prime}, which also gives t11=(u1+iu1)(u1iu1).t_{11}=(u_{1}+iu_{1}^{\prime})(u_{1}-iu_{1}^{\prime}). This observation for all k=1,2,3k=1,2,3 will lead to the forms

tkk=uk2+uk2,k=1,2,3.t_{kk}=u_{k}^{2}+u_{k}^{\prime 2},\quad k=1,2,3. (4.35)

Also, it is straightforward that examining all the equalities in (4.32) like in the steps (4.33)-(4.34) we get the possible forms for tij:t_{ij}:

tij=uiuj+σijuiuj,1i<j3,t_{ij}=u_{i}u_{j}+\sigma_{ij}u_{i}^{\prime}u_{j}^{\prime},\quad 1\leq i<j\leq 3, (4.36)

where σij{1,1},\sigma_{ij}\in\{-1,1\}, and σij=σji\sigma_{ij}=\sigma_{ji} for 1i<j31\leq i<j\leq 3. In the case when an even number of σij,\sigma_{ij}, (ij)(i\neq j) are negative (which means zero or two of them), then by changing the sign of one of the linear forms ui,u_{i}^{\prime}, we can assume without loss of generality that all σij\sigma_{ij} are positive. Hence taking into account the forms of tijt_{ij} in (4.35) and (4.36), we get

f(𝒙𝒚)\displaystyle f(\bm{x}\otimes\bm{y}) =i,j=13xixjtij(𝒚)\displaystyle=\sum_{i,j=1}^{3}x_{i}x_{j}t_{ij}(\bm{y})
=(x1u1(𝒚)+x2u2(𝒚)+x3u3(𝒚))2+(x1u1(𝒚)+x2u2(𝒚)+x3u3(𝒚))2,\displaystyle=(x_{1}u_{1}(\bm{y})+x_{2}u_{2}(\bm{y})+x_{3}u_{3}(\bm{y}))^{2}+(x_{1}u_{1}^{\prime}(\bm{y})+x_{2}u_{2}^{\prime}(\bm{y})+x_{3}u_{3}^{\prime}(\bm{y}))^{2},

i.e., ff is polyconvex. It remains to analysis the case when an odd number of σij,\sigma_{ij}, (ij)(i\neq j) are negative, in which case we can again assume without loss of generality that σij=1\sigma_{ij}=-1 for all 1i<j3.1\leq i<j\leq 3. In this situation we need to further analyze the off-diagonal elements of the cofactor matrix in (4.9). Recall that we have the following identities:

{tkk=uk2+uk2,k=1,2,3,tij=uiujuiuj,1i<j3,ck=uiuj+ujui,{i,j,k}={1,2,3}.\begin{cases}t_{kk}=u_{k}^{2}+u_{k}^{\prime 2},&\quad k=1,2,3,\\ t_{ij}=u_{i}u_{j}-u_{i}^{\prime}u_{j}^{\prime},&\quad 1\leq i<j\leq 3,\\ c_{k}=u_{i}u_{j}^{\prime}+u_{j}u_{i}^{\prime},&\quad\{i,j,k\}=\{1,2,3\}.\end{cases} (4.37)

Consequently, given the values of the entries of 𝑻,\bm{T}, and the values of cic_{i} in (4.37), for the off-diagonal elements of the cofactor matrix cof(𝑻)\mathrm{cof}(\bm{T}) we get the following set of identities:

{u1(u2u3u2u3)=u1(u2u3+u2u3),u2(u1u3u1u3)=u2(u1u3+u1u3),u3(u1u2u1u2)=u3(u1u2+u1u2).\begin{cases}u_{1}^{\prime}(u_{2}^{\prime}u_{3}^{\prime}-u_{2}u_{3})=u_{1}(u_{2}u_{3}^{\prime}+u_{2}^{\prime}u_{3}),\\ u_{2}^{\prime}(u_{1}^{\prime}u_{3}^{\prime}-u_{1}u_{3})=u_{2}(u_{1}u_{3}^{\prime}+u_{1}^{\prime}u_{3}),\\ u_{3}^{\prime}(u_{1}^{\prime}u_{2}^{\prime}-u_{1}u_{2})=u_{3}(u_{1}u_{2}^{\prime}+u_{1}^{\prime}u_{2}).\\ \end{cases} (4.38)

Note first that if the linear forms u1u_{1} and u1u_{1}^{\prime} are collinear, i.e., u1=λu1u_{1}^{\prime}=\lambda u_{1} for some λ,\lambda\in\mathbb{R}, then we can separate the variable x1u1x_{1}u_{1} from the bi-quadratic form f(𝒙𝒚)f(\bm{x}\otimes\bm{y}) by extracting a perfect square, ending up with anther nonnegative form g, depending only on the vector variables (x2,x3)(x_{2},x_{3}) and 𝒚.\bm{y}. Hence by Terpstra’s theorem (Theorem 4.2), the form gg, being 2×3,2\times 3, must be a sum of squares, thus so must be ff too, i.e., ff is polyconvex. In what follows we assume that none of the pairs (ui,ui)(u_{i},u_{i}^{\prime}) is linearly dependent, in particular none of the forms uiu_{i} and uiu_{i}^{\prime} is identically zero for i=1,2,3.i=1,2,3. Observe that as u1u_{1} and u1u_{1}^{\prime} are linear forms, then from the first equality in (4.38) we get that u2u3u2u3u_{2}^{\prime}u_{3}^{\prime}-u_{2}u_{3} must be divisible by u1u_{1} and u2u3+u2u3u_{2}u_{3}^{\prime}+u_{2}^{\prime}u_{3} must be divisible by u1.u_{1}^{\prime}. Let u2u3u2u3=v1u1,u_{2}^{\prime}u_{3}^{\prime}-u_{2}u_{3}=v_{1}u_{1}, thus we also have u2u3+u2u3=v1u1,u_{2}u_{3}^{\prime}+u_{2}^{\prime}u_{3}=v_{1}u_{1}^{\prime}, for some linear form v1=v1(𝒚).v_{1}=v_{1}(\bm{y}). By analogous observations for the second and third identities in (4.38) we arrive at the set of equalities:

{u2u3u2u3=v1u1,u2u3+u2u3=v1u1,u1u3u1u3=v2u2,u1u3+u1u3=v2u2,u1u2u1u2=v3u3,u1u2+u1u2=v3u3,\begin{cases}u_{2}^{\prime}u_{3}^{\prime}-u_{2}u_{3}=v_{1}u_{1},\quad u_{2}u_{3}^{\prime}+u_{2}^{\prime}u_{3}=v_{1}u_{1}^{\prime},\\ u_{1}^{\prime}u_{3}^{\prime}-u_{1}u_{3}=v_{2}u_{2},\quad u_{1}u_{3}^{\prime}+u_{1}^{\prime}u_{3}=v_{2}u_{2}^{\prime},\\ u_{1}^{\prime}u_{2}^{\prime}-u_{1}u_{2}=v_{3}u_{3},\quad u_{1}u_{2}^{\prime}+u_{1}^{\prime}u_{2}=v_{3}u_{3}^{\prime},\\ \end{cases} (4.39)

for some linear forms v1,v2v_{1},v_{2}, and v3.v_{3}. Multiplying the first identity by u2,u_{2}^{\prime}, the second identity by u2u_{2} in the first row of (4.39) and adding the obtained equalities together we get

u3(u22+u22)=v1(u1u2+u1u2).u_{3}^{\prime}(u_{2}^{2}+u_{2}^{\prime 2})=v_{1}(u_{1}u_{2}^{\prime}+u_{1}^{\prime}u_{2}).

Consequently, utilizing the second identity in the third row of (4.39), and keeping in mind that the form u3u_{3}^{\prime} is nonzero, we derive from the last equality:

u22+u22=v1v3.u_{2}^{2}+u_{2}^{\prime 2}=v_{1}v_{3}.

Finally, note that the lest equality is equivalent to

(u2+iu2)(u2iu2)=v1v3,(u_{2}+iu_{2}^{\prime})(u_{2}-iu_{2}^{\prime})=v_{1}v_{3},

which by the unique factorization is possible if and only if the forms u2u_{2} and u2u_{2}^{\prime} are collinear, i.e., we are back in the first situation and hence ff is polyconvex. This completes the proof of Theorem 2.2

Reference

  • [1] G. Allaire and R.V. Kohn. Optimal lower bounds on the elastic energy of a composite made from two non-well-ordered isotropic materials, Quarterly of applied mathematics, vol. LII, 311-333 (1994)
  • [2] E. Artin. Über die Zerlegung definiter Funktionen in Quadrate, Abhandlungen aus dem Mathematischen Seminar der Universität Hamburg. 5 (1): 100–115, 1927.
  • [3] J. M. Ball. Convexity conditions and existence theorems in nonlinear elasticity, Arch. Ration. Mech. Anal, 63, 337-403 (1976).
  • [4] J. M. Ball and R. D. James. Fine phase mixtures as minimizers of energy, Arch. Rational Mech. Anal., 100(1):13–52, 1987.
  • [5] B. Benešová and M. Kružík. Weak lower semicontinuity of integral functionals and applications. SIAM Rev., 59(4):703–766, 2017.
  • [6] G. Blekherman. Nonnegative Polynomials and Sums of Squares, Journal of the AMS, 25, 2012, 617–635.
  • [7] G. Blekherman, G. G. Smith and M. Velasco. Sums of squares and varieties of minimal degree, Journal of the AMS, 29 (2016) 893–913.
  • [8] G. Blekherman, R. Sinn, G. Smith, M. Velasco. Sums of Squares: A Real Projective Story", Notices of the AMS, arXiv:2101.05773.
  • [9] G. Blekherman. A Brief Introduction to Sums of Squares, Proceedings of Symposia in Applied Mathematics, AMS.
  • [10] O. Boussaid, C. Kreisbeck, and A. Schlömerkemper. Characterizations of Symmetric Polyconvexity, Archive for Rational Mechanichs and Analysis, vol. 234, pp. 417–451(2019).
  • [11] A. Buckley and K. S̆ivic. Nonnegative biquadratic forms with maximal number of zeros, preprint, https://arxiv.org/pdf/1611.09513.pdf
  • [12] A. Cherkaev. Variational Methods for Structural Optimization, Springer Applied Mathematical Sciences, Vol. 140 (2000).
  • [13] A. V. Cherkaev and L. V. Gibiansky. The exact coupled bounds for effective tensors of electrical and magnetic properties of two-component two-dimensional composites, Proceedings of the Royal Society of Edinburgh. Section A, Mathematical and Physical Sciences, 122 (1992), pp. 93–125.
  • [14] A. V. Cherkaev and L. V. Gibiansky. Coupled estimates for the bulk and shear moduli of a two-dimensional isotropic elastic composite, Journal of the Mechanics and Physics of Solids, 41 (1993), pp. 937–980.
  • [15] S. J. Cho. Generalized Choi maps in three-dimensional matrix algebra, Linear algebra and its applications, 171: 213–224 (1992).
  • [16] M.-D. Choi. Positive semidefinite biquadratic forms, Linear algebra and its applications, 12, 95–100 (1975).
  • [17] M.-D. Choi and T.-Y. Lam. Extremal positive semidefinite forms, Math. Ann., 231, 1–18 (1977).
  • [18] B. Dacorogna. Direct methods in the calculus of variations. Springer Applied Mathematical Sciences, Vol. 78, 2nd Edition (2008).
  • [19] I. Fonseca and S. Müller. A–quasiconvexity, lower semicontinuity, and Young measures. SIAM J. Math. Anal., 30(6):1355–1390, 1999.
  • [20] D. Harutyunyan. A note on the extreme points of the cone of quasiconvex quadratic forms with orthotropic symmetry, Journal of Elasticity, 140, pp. 79–93 (2020).
  • [21] D. Harutyunyan and G. W. Milton. Explicit examples of extremal quasiconvex quadratic forms that are not polyconvex, Calculus of Variations and Partial Differential Equations , October 2015, Vol. 54, Iss. 2, pp. 1575-1589.
  • [22] D. Harutyunyan and G. W. Milton. On the relation between extremal elasticity tensors with orthotropic symmetry and extremal polynomials, Archive for Rational Mechanics and Analysis, Vol. 223, Iss. 1, pp 199-212, 2017.
  • [23] D. Harutyunyan and G. W. Milton. Towards characterization of all 3×33\times 3 extremal quasiconvex quadratic forms, Communications of Pure and Applied Mathematics, Vol. 70, Iss. 11, Nov. 2017, pp. 2164-2190.
  • [24] J. W. Helton, S. A. McCullough, and V. Vinnikov. Noncommutative convexity arises from linear matrix inequalities, J. Funct. Anal. 240 (2006), 105–191.
  • [25] D. Hilbert. Über die Darstellung definiter Formen als Summen von Formenquadraten, Math. Ann., 32 (1888), 342–350.
  • [26] J. Hou, Ch.-K. Li, Y.-T. Poon, X. Qi, and N.-S. Sze. A new criterion and a special class of kk-positive maps, Linear Algebra and its Applications 470 (2015) 51-69.
  • [27] H. Kang, E. Kim, and G. W. Milton. Sharp bounds on the volume fractions of two materials in a two-dimensional body from electrical boundary measurements: the translation method, Calculus of Variations and Partial Differential Equations, 45, 367-401 (2012).
  • [28] H. Kang and G. W. Milton. Bounds on the volume fractions of two materials in a three dimensional body from boundary measurements by the translation method, SIAM Journal on Applied Mathematics, 73, 475–492 (2013).
  • [29] H. Kang, G. W. Milton, and J.-N. Wang. Bounds on the Volume Fraction of the Two-Phase Shallow Shell Using One Measuremen. Journal of Elasticity, 114, 41-53 (2014).
  • [30] X. Li and W. Wu. A class of generalized positive linear maps on matrix algebras, Linear algebra and its applications 439 (2013) 2844-2860.
  • [31] R. V. Kohn and R. Lipton. Optimal bounds for the effective energy of a mixture of isotropic, incompressible, elastic materials. Archive for Rational Mechanics and Analysis, 102, 331–350 (1988).
  • [32] P. Marcellini. Quasiconvex quadratic forms in two dimensions. Appl. Math. Optim. 11(2), 183–189, 1984.
  • [33] G. W. Milton. On characterizing the set of positive effective tensors of composites: The variational method and the translation method, Communications on Pure and Applied Mathematics, Vol. XLIII, 63-125 (1990).
  • [34] G. W. Milton. The Theory of Composites, vol. 6 of Cambridge Monographs on Applied and Computational Mathematics, Cambridge University Press, Cambridge, United Kingdom, 2002
  • [35] G. W. Milton. Sharp inequalities which generalize the divergence theorem: an extension of the notion of quasi-convexity, Proceedings Royal Society A 469, 20130075 (2013).
  • [36] G. W. Milton and L. H. Nguyen. Bounds on the volume fraction of 2-phase, 2-dimensional elastic bodies and on (stress, strain) pairs in composites. Comptes Rendus Mécanique, 340, 193-204 (2012).
  • [37] C. B. Morrey. Quasiconvexity and the lower semicontinuity of multiple integrals, Pacific Journal of Mathematics 2, 25-53 (1952)
  • [38] C. B. Morrey. Multiple integrals in the calculus of variations, Springer–Verlag, Berlin, 1966.
  • [39] S. Müller. Variational models for microstructure and phase transitions. In Calculus of variations and geometric evolution problems (Cetraro, 1996), volume 1713 of Lecture Notes in Math., pp. 85–210. Springer, Berlin, 1999.
  • [40] F. Murat and L. Tartar. Calcul des variations et homogénísation. (French) [Calculus of variation and homogenization], in Les méthodes de l’homogénéisation: théorie et applications en physique, volume 57 of Collection de la Direction des études et recherches d’Electricité de France, pages 319-369, Paris, 1985, Eyrolles, English translation in Topics in the Mathematical Modelling of Composite Materials, pages 139-173, ed. by A. Cherkaev and R. Kohn, ISBN 0-8176-3662-5.
  • [41] R. Quarez. On the real zeros of positive semidefinite biquadratic forms, Comm. Algebra, 43 (2015), 1317–1353.
  • [42] R. Quarez. Symmetric determinantal representation of polynomials, Linear Algebra and its Applications 436 (2012), 3642–3660.
  • [43] B. Reznick. On Hilbert’s construction of positive polynomials, preprint, available at: https://arxiv.org/pdf/0707.2156v1.pdf.
  • [44] C. Scheiderer. Sums of squares of polynomials with rational coefficients. J. Eur. Math. Soc. 18(7):1495–1513, 2016.
  • [45] D. Serre. Condition de Legendre-Hadamard: Espaces de matrices de rang1\neq 1. (French) [Legendre-Hadamard condition: Space of matrices of rank1\neq 1], Comptes Rendus de l’Académie des sciences, Paris 293, 23-26 (1981).
  • [46] A. Stefan and A. Welters. A Short Proof of the Symmetric Determinantal Representation of Polynomials, preprint, https://arxiv.org/abs/2101.03589
  • [47] E. Stormer. Separable states and the structural physical approximation of a positive map, Journal of Functional Alanysis 264 (2013) 2197-2205.
  • [48] V. Šverák. New examples of quasiconvex functions. Arch. Ration. Mech. Anal., 119(4):293–300, 1992.
  • [49] V. Šverák. Rank-one convexity does not imply quasiconvexity. Proc. Roy. Soc. Edinburgh Sect. A, 120(1-2):185–189, 1992.
  • [50] L. Tartar. Compensated compactness and applications to partial differential equations, in Nonlinear Analysis and Mechanics, Heriot-Watt Symposium, Volume IV, edited by R. J. Knops, volume 39 of Research Notes in Mathematics, pages 136-212, London, 1979, Pitman Publishing Ltd.
  • [51] F. J. Terpstra. Die Darstellung biquadratischer Formen als Summen von Quadraten mit Anwendung auf die Variationsrechnung, Mathematische Annalen, 116 (1938), 166–180.
  • [52] L. Van Hove. Sur l’extension de la condition de Legendre du calcul des variations aux intégrales multiples á plusieurs functions inconnues, Nederl. Akad. Wetensch. Proc. 50 (1947), 18-23.
  • [53] L. Van Hove. Sur le signe de la variation seconde des intégrales multiples á plusieurs functions inconnues, Acad. Roy. Belgique Cl. Sci. Mém. Coll. 24 (1949), 68.
  • [54] K. Zhang. The structure of rank-one convex quadratic forms on linear elastic strains. Proc. Roy. Soc. Edinburgh Sect. A, 133(1):213–224, 2003.