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

Tropicalizing Principal Minors of Positive Definite Matrices

Abeer Al Ahmadieh School of Mathematics, Georgia Institute of Technology, Atlanta, GA, USA. [email protected] Felipe Rincón School of Mathematical Sciences, Queen Mary University of London, London, UK. [email protected] Cynthia Vinzant School of Mathematics, University of Washington, Seattle, WA, USA. [email protected]  and  Josephine Yu School of Mathematics, Georgia Institute of Technology, Atlanta, GA, USA. [email protected]
Abstract.

We study the tropicalization of the image of the cone of positive definite matrices under the principal minors map. It is a polyhedral subset of the set of MM-concave functions on the discrete nn-dimensional cube. We show it coincides with the intersection of the affine tropical flag variety with the submodular cone. In particular, any cell in the regular subdivision of the cube induced by a point in this tropicalization can be subdivided into base polytopes of realizable matroids. We also use this tropicalization as a guide to discover new algebraic inequalities among the principal minors of positive definite matrices of a fixed size.

1. Introduction

For an n×nn\times n matrix AA and subsets S,T[n]={1,2,,n}S,T\subset[n]=\{1,2,\dots,n\} of the same size, we will denote by A(S,T)A(S,T) the determinant of the submatrix of AA with rows indexed by SS and columns indexed by TT. For a square matrix AA, we use ASA_{S} to denote the principal minor A(S,S)A(S,S). A real symmetric or Hermitian matrix is called positive definite if all of its principal minors are positive.

The tropicalization of a subset 𝒮+N\mathcal{S}\subset\mathbb{R}_{+}^{N} is defined as the logarithmic limit set as in [Ale13]:

trop(𝒮)=limt{(logt(x1),,logt(xN)):x𝒮}.\operatorname{trop}(\mathcal{S})=\lim_{t\rightarrow\infty}\{(\log_{t}(x_{1}),\dots,\log_{t}(x_{N})):x\in\mathcal{S}\}.

It is a way to track the exponential behavior of a set as coordinates go to 0 or \infty. When 𝒮\mathcal{S} is semialgebraic, the tropicalization is a rational polyhedral fan that coincides with the closure of the image of 𝒮{{t}}\mathcal{S}_{\mathbb{R}\{\!\{t\}\!\}} under coordinate-wise valuation, where 𝒮{{t}}\mathcal{S}_{\mathbb{R}\{\!\{t\}\!\}} denotes the extension of 𝒮\mathcal{S} to the field of real Puiseux series. See Section 2.1 for further discussion.

The principal minors of an n×nn\times n positive definite matrix AA can be encoded by the degree-nn homogeneous polynomial in n+1n+1 variables

fA=det(diag(x1,,xn)+yA)=S[n]AS𝐱[n]\Sy|S|[x1,,xn,y],f_{A}=\det({\rm diag}(x_{1},\ldots,x_{n})+yA)=\sum_{S\subseteq[n]}A_{S}\cdot{\bf x}^{[n]\backslash S}y^{|S|}\in\mathbb{R}[x_{1},\dots,x_{n},y],

where 𝐱T:=iTxi{\bf x}^{T}:=\prod_{i\in T}x_{i}. This polynomial is stable [BB06], which implies that the tropicalization of the set of principal minors of positive definite matrices is a subset of the set of MM^{\natural}-concave functions on {0,1}n\{0,1\}^{n} [Brä10]. We see in Corollary 4.3 that for n6n\geq 6 not every MM^{\natural}-concave function arises this way from positive definite matrices. Our main result is a description of this tropicalization in terms of the affine tropical flag variety.

Theorem 4.1.

The tropicalization of the set of principal minors of n×nn\times n positive definite real symmetric (resp. Hermitian) matrices equals the intersection of the affine tropical flag variety over \mathbb{R} (resp. \mathbb{C}) with the cone of submodular functions on the hypercube {0,1}n\{0,1\}^{n}.

The study of the tropical flag variety was initiated by Haque in 20122012 [Haq12]. Recently, Brandt, Eur, and Zhang described it in terms of flag matroid polytope subdivisions [BEZ21]. The totally non-negative part of the tropical flag variety was studied in [Bor22].

An MM^{\natural}-concave function induces a regular subdivision of the unit cube {0,1}n\{0,1\}^{n} that coarsens a subdivision in which every cell is a dehomogenized matroid polytope, i.e., a 0/1 polytope with edges of the form eieje_{i}-e_{j} or eie_{i} contained in a layer {x[0,1]n:kixik+1}\{x\in[0,1]^{n}:k\leq\sum_{i}x_{i}\leq k+1\} for some kk. When the MM^{\natural}-concave function arises as the tropicalization of the principal minors of a positive definite matrix, we show that these matroid polytopes correspond to realizable matroids.

Theorem 4.5.

The regular subdivision of [0,1]n[0,1]^{n} induced by the tropicalization of the principal minors of a real symmetric (or Hermitian) positive definite matrix is a coarsening of a subdivision of [0,1]n[0,1]^{n} into dehomogenized matroid polytopes of matroids realizable over \mathbb{R} (or \mathbb{C}, respectively).

Understanding determinantal inequalities for positive definite matrices is a classical topic of study, see for example [JB93, CFB95], and it continues to be a topic of interest [HJ08, Cho16, JZC19, LS20, DWH22, BS23]. In [HJ08], for instance, Hall and Johnson use cone theoretic techniques to characterize the semigroup of ratios of products of principal minors over all positive definite matrices.

Any tropical polynomial inequality valid on the tropicalization of a semialgebraic set 𝒮\mathcal{S} can be lifted to a polynomial inequality valid on the set 𝒮\mathcal{S} itself [JSY22]. We use our description of the tropicalization of the set 𝒮\mathcal{S} of principal minors of positive definite matrices as a guide to understanding the possible polynomial inequalities valid on 𝒮\mathcal{S}; in particular, we study lifts of the MM^{\natural}-concavity constraints to polynomial inequalities on 𝒮\mathcal{S}.

Example 1.1.

Let PD3\mathrm{PD}_{3} be the set of 3×33\times 3 real symmetric positive definite matrices, and let 𝒮3+2[3]\mathcal{S}_{3}\subset\mathbb{R}_{+}^{2^{[3]}} be the image of PD3\mathrm{PD}_{3} under the map sending a matrix AA to its vector of principal minors (AS)S[3](A_{S})_{S\subset[3]}. Consider the set cone(𝒮3)={λxλ+ and x𝒮3}\operatorname{cone}(\mathcal{S}_{3})=\{\lambda\,x\mid\lambda\in\mathbb{R}_{+}\text{ and }x\in\mathcal{S}_{3}\}, where the coordinate indexed by the subset \emptyset is not required to be equal to 11. Theorem 4.1 says that trop(cone(𝒮3))\operatorname{trop}(\operatorname{cone}(\mathcal{S}_{3})) consists of all vectors w2[3]w\in\mathbb{R}^{2^{[3]}} that satisfy the submodular inequalities and the tropical incidence relation

(1) max(w1+w23,w2+w13,w3+w12) is attained at least twice.\max(w_{1}+w_{23},w_{2}+w_{13},w_{3}+w_{12})\text{ is attained at least twice}.

Consider now the projection 𝒫\mathcal{P} of the semialgebraic set cone(𝒮3)\operatorname{cone}(\mathcal{S}_{3}) onto the six coordinates indexed by the subsets 1,2,3,12,13,231,2,3,12,13,23. The set 𝒫\mathcal{P} is a full-dimensional semialgebraic set of +6\mathbb{R}_{+}^{6}; indeed, the six 1×11\times 1 and 2×22\times 2 principal minors of a positive definite matrix are algebraically independent. However, its tropicalization trop(𝒫)\operatorname{trop}(\mathcal{P}) satisfies the tropical equation (1), and thus it is only a 55-dimensional polyhedral subset of 6\mathbb{R}^{6}. This tropical equation cannot be lifted to an algebraic equation satisfied by 𝒫\mathcal{P}, but, as we explore in Section 5, it is instead a consequence of 𝒫\mathcal{P} satisfying the following three algebraic inequalities:

A1A23+A2A1312A3A12,A1A23+A3A1212A2A13,A2A13+A3A1212A1A23.A_{1}A_{23}+A_{2}A_{13}\geq\frac{1}{2}A_{3}A_{12},\quad A_{1}A_{23}+A_{3}A_{12}\geq\frac{1}{2}A_{2}A_{13},\quad A_{2}A_{13}+A_{3}A_{12}\geq\frac{1}{2}A_{1}A_{23}.

Other inequalities of this form with different coefficients, which also hold for all Lorentzian polynomials, are presented in Theorem 5.1. \diamond

Organization

The paper is organized as follows. We introduce relevant background and definitions in Section 2, including various characterizations of MM^{\natural}-concave functions and their relation to the affine flag Dressian. In Section 3, we discuss a connection between the tropicalization of the flag variety and a slice of the (n,2n)(n,2n) Grassmannian over arbitrary valuated fields. In Section 4, we specialize these results to real closed and algebraically closed fields of characteristic zero and show that the resulting sets coincide with the tropicalization of the positive definite cone under the principal minor map. In Section 5, we explore consequences for polynomial inequalities on the principal minors of positive definite matrices. Finally, in Section 6, we prove a technical lemma used in Section 3.

Acknowledgements

We are grateful to Jonathan Boretsky, H. Tracy Hall, Yassine El Maazouz, and Bernd Sturmfels for helpful discussions and insights. CV is partially supported by the NSF DMS grant #2153746. JY is partially supported by the NSF DMS grants #1855726 and #2348701.

2. Definitions and Background

2.1. Tropicalization

A (nonarchimedean) valuation on a field 𝒦\mathcal{K} is a map val:𝒦\operatorname{val}:\mathcal{K}^{*}\rightarrow\mathbb{R} satisfying

val(ab)=val(a)+val(b), and val(a+b)min(val(a),val(b)).\operatorname{val}(ab)=\operatorname{val}(a)+\operatorname{val}(b),\text{ and }\operatorname{val}(a+b)\geq\min(\operatorname{val}(a),\operatorname{val}(b)).

Here we denote 𝒦=𝒦{0}\mathcal{K}^{*}=\mathcal{K}\setminus\{0\}. The image Γ=val(𝒦)\Gamma=\operatorname{val}(\mathcal{K}^{*}) of val\operatorname{val} is an additive subgroup of \mathbb{R} called the value group. The valuation is non-trivial if Γ{0}\Gamma\neq\{0\}. In order to use the max convention, we will take ν:𝒦\nu:\mathcal{K}^{*}\to\mathbb{R} to be ν(a)=val(a)\nu(a)=-\operatorname{val}(a) so that

ν(ab)\displaystyle\nu(ab) =ν(a)+ν(b)\displaystyle=\nu(a)+\nu(b)
ν(a+b)\displaystyle\nu(a+b) max(ν(a),ν(b)).\displaystyle\leq\max(\nu(a),\nu(b)).

We use 𝕜\mathbbm{k} to denote the residue field of 𝒦\mathcal{K}, which is the quotient of the valuation ring 𝒪={a𝒦:val(a)0}\mathcal{O}=\{a\in\mathcal{K}:\operatorname{val}(a)\geq 0\} by its unique maximal ideal 𝔪={a𝒦:val(a)>0}\mathfrak{m}=\{a\in\mathcal{K}:\operatorname{val}(a)>0\}.

Throughout the paper, we use 𝒦\mathcal{K} to denote a field with a nontrivial valuation with an infinite residue field 𝕜\mathbbm{k}. The valuation has a splitting or a cross-section if there is a group homomorphism ϕ\phi from the value group Γ\Gamma to the multiplicative group 𝒦\mathcal{K}^{*} such that valϕ\operatorname{val}\circ\,\phi is the identity on Γ\Gamma. If 𝒦\mathcal{K} is real closed or algebraically closed, then there is a splitting by [AGS20, Lemma 2.4] and [MS21, Lemma 2.1.15]. The splitting ϕ\phi allows us to talk about the leading coefficient of an element a𝒦a\in\mathcal{K}^{*} as the element tval(a)a¯\overline{t^{-\operatorname{val}(a)}a} in the residue field, where t=ϕ(1)t=\phi(1). Typical examples include the fields of rational functions, Laurent series, or Puiseux series over a field 𝕜\mathbbm{k}. The Puiseux series field 𝕜{{t}}\mathbbm{k}\{\!\{t\}\!\} is real closed if 𝕜\mathbbm{k} is real closed and is algebraically closed if 𝕜\mathbbm{k} is algebraically closed of characteristic 0.

Given a subset 𝒮(𝒦)n\mathcal{S}\subset(\mathcal{K}^{*})^{n}, we define

ν(𝒮)={(ν(x1),,ν(xn)):x𝒮}.\nu(\mathcal{S})=\{(\nu(x_{1}),\ldots,\nu(x_{n})):x\in\mathcal{S}\}.

We will denote by \mathcal{R} a real closed field with a nontrivial valuation. Any real closed field has a unique total ordering \geq compatible with the field operations, where aba\geq b if aba-b is a complete square. We will always assume that the valuation is compatible with the order, namely, if a,ba,b\in\mathcal{R} with 0<a<b0<a<b, then ν(a)ν(b)\nu(a)\leq\nu(b). An example is the field of real Puiseux series {{t}}\mathbb{R}\{\!\{t\}\!\}, where a series is positive if its leading coefficient is positive.

Finally, we use 𝒞\mathcal{C} to denote the degree-two field extension 𝒞=(𝕚)\mathcal{C}=\mathcal{R}(\mathbbm{i}) where 𝕚2=1\mathbbm{i}^{2}=-1. Since \mathcal{R} is real-closed, 𝒞\mathcal{C} is algebraically closed. This comes with complex conjugation a+𝕚b¯=a𝕚b\overline{a+\mathbbm{i}b}=a-\mathbbm{i}b where a,ba,b\in\mathcal{R}. The valuation on \mathcal{R} naturally extends to 𝒞\mathcal{C} by ν(a+𝕚b)=max{ν(a),ν(b)}\nu(a+\mathbbm{i}b)=\max\{\nu(a),\nu(b)\}.

For a semialgebraic subset 𝒮+n\mathcal{S}\subset\mathbb{R}_{+}^{n}, let 𝒮\mathcal{S}_{\mathcal{R}} be the semialgebraic subset of n\mathcal{R}^{n} defined by the same semialgebraic expression (a first-order formula in the language of ordered rings) defining 𝒮\mathcal{S}. By [Ale13, JSY22], the image of SS_{\mathcal{R}} under coordinate-wise valuation coincides with the tropicalization of 𝒮\mathcal{S} (up to closure):

trop(𝒮)=ν(𝒮)¯\operatorname{trop}(\mathcal{S})=\overline{\nu(\mathcal{S}_{\mathcal{R}})}

Here closure is taken in the Euclidean topology on n\mathbb{R}^{n}. In particular, this set does not depend on the choice of the extension \mathcal{R} or the choice of semialgebraic expression defining 𝒮\mathcal{S}. Moreover we have trop(𝒮)Γn=ν(𝒮)\operatorname{trop}(\mathcal{S})\cap\Gamma^{n}=\nu(\mathcal{S}_{\mathcal{R}}) [Ale13, Theorem  4.2], and trop(𝒮)\operatorname{trop}(\mathcal{S}) is a union of polyhedra [AGS20, Theorem 3.1].

If X()nX\subset(\mathbb{C}^{*})^{n} is an algebraic variety, then its image under coordinate-wise absolute-value is a semialgebraic subset |X||X| of +n\mathbb{R}_{+}^{n}. The tropicalization of XX can be defined as the tropicalization of |X||X|. We can similarly define X𝒞X_{\mathcal{C}} to be the algebraic variety of 𝒞n\mathcal{C}^{n} defined by the vanishing of all polynomials that vanish on XX. A part of the fundamental theorem of tropical geometry states that the logarithmic limit set of |X||X| coincides with the closure of the image of X𝒞X_{\mathcal{C}} under coordinate-wise valuation. That is, trop(X)=ν(X)¯\operatorname{trop}(X)=\overline{\nu(X)}. A proof can be found, for example, using [Ber71, Theorem 2] and [Gub13, Proposition 3.8].

A symmetric matrix over a real field \mathcal{R}, or a Hermitian matrix over 𝒞\mathcal{C}, is called positive definite or PD if all of its principal minors are positive. Thus the principal minors give a map from the set of n×nn\times n matrices to 2[n]\mathcal{R}^{2^{[n]}}, sending A(AS)S[n]A\mapsto(A_{S})_{S\subset[n]}. The image of the sets of symmetric and Hermitian matrices have been studied in [HS07, Oed11, LS09, AAV24a, AAV24b] and it has many applications in probability and statistics, see for instance [EM22]. Let PMn+()\operatorname{PM}^{+}_{n}(\mathcal{R}) and PMn+(𝒞)2[n]\operatorname{PM}^{+}_{n}(\mathcal{C})\subset\mathcal{R}^{2^{[n]}} denote the image of the symmetric and Hermitian PD cone respectively under this principal minor map. Because the image of the principal minors of positive definite matrices is a semialgebraic set, the theorems above imply that

trop(PMn+())=ν(PMn+())¯ and trop(PMn+())=ν(PMn+(𝒞))¯.\operatorname{trop}(\operatorname{PM}^{+}_{n}(\mathbb{R}))=\overline{\nu(\operatorname{PM}^{+}_{n}(\mathcal{R}))}\text{ and }\operatorname{trop}(\operatorname{PM}^{+}_{n}(\mathbb{C}))=\overline{\nu(\operatorname{PM}^{+}_{n}(\mathcal{C}))}.

2.2. Discrete Concavity

For a set S[n]S\subset[n] and an element i[n]i\in[n], let us use the shorthand SiSi for S{i}S\cup\{i\} and S\iS\backslash i for the set difference S\{i}S\backslash\{i\}. A function F:2[n]F:2^{[n]}\rightarrow\mathbb{R} is submodular if for all S,T[n]S,T\subseteq[n],

F(ST)+F(ST)F(S)+F(T).F(S\cap T)+F(S\cup T)\leq F(S)+F(T).

Submodularity is implied by the following “local” submodularity condition that for all S[n]S\subset[n] and distinct i,j[n]Si,j\in[n]\setminus S,

F(S)+F(Sij)F(Si)+F(Sj).F(S)+F(Sij)\leq F(Si)+F(Sj).

A function F:2[n]F:2^{[n]}\rightarrow\mathbb{R} induces two regular subdivisions of the unit cube [0,1]n[0,1]^{n} by lifting each of the vertices of the cube by height FF in a new dimension, and then projecting back to n\mathbb{R}^{n} either the upper or the lower convex hull of the lift. The alcoved triangulation of a cube is the triangulation of the unit cube [0,1]n[0,1]^{n} into n!n! simplices where each simplex is the convex hull of an edge path from (0,0,,0)(0,0,\dots,0) to (1,1,,1)(1,1,\dots,1) whose sum of coordinates is monotonously increasing along the path. Equivalently, it is given by dicing the cube with the type-AA braid arrangement hyperplanes xi=xjx_{i}=x_{j}. This triangulation is “dual” to the permutohedron. Lovász showed that a function is submodular if and only if the regular subdivision of the unit cube induced by the lower convex hull is a coarsening of the alcoved triangulation [Lov83]. In particular, the edges in the lower convex hull correspond to pairs of subsets S,T[n]S,T\subset[n] with STS\subset T. Thus the edges with directions eieje_{i}-e_{j} cannot appear in the lower hull and can only appear in the upper hull.

The notions of MM-convex and MM^{\natural}-convex functions were introduced by Murota and collaborators [Mur98]. MM-concave functions are also called valuated polymatroids. A function F:2[n]F:2^{[n]}\rightarrow\mathbb{R} is MM^{\natural}-concave if for all S,T[n]S,T\subseteq[n] and all iS\Ti\in S\backslash T, either

  • F(S)+F(T)F(S\i)+F(Ti)F(S)+F(T)\leq F(S\backslash i)+F(Ti), or

  • there exists jT\Sj\in T\backslash S such that F(S)+F(T)F(Sj\i)+F(Ti\j)F(S)+F(T)\leq F(Sj\backslash i)+F(Ti\backslash j).

We will see some more characterizations of MM^{\natural}-concave functions below in Proposition 2.2.

Our interest in MM^{\natural}-concave functions comes from the following fact.

Proposition 2.1.

The tropicalization of the image of the positive definite cone under the principal minor map is a subset of the set of MM^{\natural}-concave functions on {0,1}n\{0,1\}^{n}.

Proof.

If A{{t}}n×nA\in\mathbb{R}\{\!\{t\}\!\}^{n\times n} is a positive definite matrix then, using [BB06], the polynomial f=det(diag(x1,,xn)+A)=S[n]AS𝐱[n]\Sf=\det({\rm diag}(x_{1},\ldots,x_{n})+A)=\sum_{S\subseteq[n]}A_{S}{\bf x}^{[n]\backslash S} is stable and has positive coefficients, so (ν(AS))S[n](\nu(A_{S}))_{S\subseteq[n]} is an MM^{\natural}-concave function by [Brä10]. ∎

2.3. MM^{\natural}-concavity and valuated matroids

In this subsection we compile various characterizations of MM^{\natural}-concave functions in terms of valuated matroids and polytope subdivisions.

For integers 1kn11\leq k\leq n-1, the (affine) Dressian Dr(k,n)\operatorname{Dr}(k,n) is the polyhedral subset of (nk)\mathbb{R}^{\binom{n}{k}} consisting of all functions p:([n]k)p:\binom{[n]}{k}\to\mathbb{R} such that for any subsets S([n]k1)S\in\binom{[n]}{k-1} and T([n]k+1)T\in\binom{[n]}{k+1},

maxiTS(p(Si)+p(T\i))is attained at least twice.\max_{i\in T\setminus S}\,(p(Si)+p(T\backslash i))\quad\text{is attained at least twice}.

Points in the Dressian Dr(k,n)\operatorname{Dr}(k,n) are, up to scaling, in one-to-one correspondence with rank-kk valuated matroids on the ground set [n][n] or equivalently, kk-dimensional tropical linear spaces in n\mathbb{R}^{n}.

The affine flag Dressian FlDr(n)\operatorname{FlDr}(n) is the polyhedral subset of 2[n]\mathbb{R}^{2^{[n]}} consisting of all functions p:2[n]p:2^{[n]}\to\mathbb{R} such that for any subsets S,T[n]S,T\subset[n] with |S||T|2|S|\leq|T|-2,

maxiTS(p(Si)+p(T\i))is attained at least twice.\max_{i\in T\setminus S}(p(Si)+p(T\backslash i))\quad\text{is attained at least twice}.

When |S|=|T|2|S|=|T|-2 the conditions above are simply the Plücker relations on the points (p(S))S([n]k)(p(S))_{S\in\binom{[n]}{k}} with k=|S|+1k=|S|+1, while when |S|<|T|2|S|<|T|-2 they are incidence relations among the corresponding tropical linear spaces. Flag Dressians have been studied in [BEZ21].

For a function F:2[n]F:2^{[n]}\to\mathbb{R}, its multisymmetric lift is the function F^:([2n]n)\hat{F}:\binom{[2n]}{n}\to\mathbb{R} given by F^(T):=F(T[n])\hat{F}(T):=F(T\cap[n]). The homogenization of the kk to k+1k+1 layer is the function F~k:(n+1k+1)\tilde{F}_{k}:\binom{n+1}{k+1}\to\mathbb{R} given by

F~k(S):={F(S)if n+1SF(S\(n+1))if n+1S.\tilde{F}_{k}(S):=\begin{cases}F(S)&\text{if }n+1\notin S\\ F(S\backslash\!(n+1))&\text{if }n+1\in S.\end{cases}
Proposition 2.2.

For a function F:2[n]F:2^{[n]}\rightarrow\mathbb{R}, the following are equivalent.

  1. (1)

    The function FF is MM^{\natural}-concave.

  2. (2)

    The multisymmetric lift F^\hat{F} belongs to the Dressian Dr(n,2n)\operatorname{Dr}(n,2n).

  3. (3)

    The function FF is submodular and, for every 1kn21\leq k\leq n-2, the homogenized layer F~k\tilde{F}_{k} belongs to the Dressian Dr(k+1,n+1)\operatorname{Dr}(k+1,n+1).

  4. (4)

    The function FF is submodular and belongs to the affine flag dressian FlDr(n)\operatorname{FlDr}(n).

  5. (5)

    The function FF induces a regular subdivision of the unit cube [0,1]n[0,1]^{n} (via upper hull) in which every edge has direction eieje_{i}-e_{j} or eie_{i}.

Proof.

The equivalence (1)(2)(1)\iff(2) follows from [GRSU24, Proposition 1.4].

The equivalence (1)(3)(1)\iff(3) follows from [RvGP02, Theorem 10], which says that FF is MM^{\natural}-concave if and only if FF is submodular and for all S[n]S\subset[n] and i,j,k,l[n]Si,j,k,l\in[n]\setminus S, each of the maxima

max{F(Sij)+F(Sk),F(Sik)+F(Sj),F(Sjk)+F(Si)}\displaystyle\max\{F(Sij)+F(Sk),F(Sik)+F(Sj),F(Sjk)+F(Si)\}
max{F(Sij)+F(Sk),F(Sik)+F(Sj),F(Sjk)+F(Si)}\displaystyle\max\{F(Sij)+F(Sk\ell),F(Sik)+F(Sj\ell),F(Sjk)+F(Si\ell)\}

is attained at least twice.

The equivalence (3)(4)(3)\iff(4) follows from [BEZ21, Theorem 5.1.2].

The equivalence (1)(5)(1)\iff(5) holds more generally for functions on subsets of n\mathbb{Z}^{n}. By [Mur98, Theorem 6.30], a function is MM-concave on a finite subset of n\mathbb{Z}^{n} with a fixed coordinate sum if and only if every cell in the regular subdivision induced by the function via the upper hull is an MM-convex set. By [Mur98, Theorem 4.15] and [MUWY18, Lemma 2.3], MM-convex sets are integer points in generalized permutohedra, which are polytopes with edges in directions eieje_{i}-e_{j}. Projecting out one of the coordinates gives the desired result for MM^{\natural}-concave functions. ∎

Example 2.3.

Consider the matrix

A=(11111+t41+t311+t31+t2+t4)=BTB where B=(1110t2t00t2).A=\begin{pmatrix}1&1&1\\ 1&1+t^{4}&1+t^{3}\\ 1&1+t^{3}&1+t^{2}+t^{4}\end{pmatrix}=B^{T}B\ \ \text{ where }\ \ B=\begin{pmatrix}1&1&1\\ 0&t^{2}&t\\ 0&0&t^{2}\end{pmatrix}.

The matrix AA is positive definite since A=BTBA=B^{T}B with det(B)0\det(B)\neq 0. For wS=val(AS)w_{S}=-\operatorname{val}(A_{S}), we have w=w1=w2=w3=0w_{\emptyset}=w_{1}=w_{2}=w_{3}=0, w12=4w_{12}=-4, w13=w23=2w_{13}=w_{23}=-2, and w123=8w_{123}=-8. The function SwSS\mapsto w_{S} is MM^{\natural}-concave and induces the regular subdivision of the cube shown in Figure 1. \diamond

w0=w_{0}= 0w1=w_{1}= 0w3=w_{3}= 0w13=w_{13}= 2-2w2=w_{2}= 0w12=w_{12}= 4-4w23=w_{23}= 2-2w123=w_{123}= 8-8
Figure 1. The regular subdivision induced by the matrix in Example 2.3.

We call a function F:2[n]F:2^{[n]}\to\mathbb{R} strictly submodular if

F(S)+F(Sij)<F(Si)+F(Sj)F(S)+F(Sij)<F(Si)+F(Sj)

for all S[n]S\subset[n] and distinct i,j[n]\Si,j\in[n]\backslash S.

Lemma 2.4.

Let F:2[n]F:2^{[n]}\to\mathbb{R} be MM^{\natural}-concave. The function FF is strictly submodular if and only if each cell of the regular subdivision of [0,1]n[0,1]^{n} induced by FF via upper hull is contained in {x[0,1]n:ki=1nxik+1}\{x\in[0,1]^{n}:k\leq\sum_{i=1}^{n}x_{i}\leq k+1\} for some k=0,,n1k=0,\ldots,n-1.

Proof.

(\Leftarrow) Suppose that every cell of the subdivision induced by FF has the desired form. Consider S[n]S\subset[n] with |S|=kn2|S|=k\leq n-2 and i,j[n]\Si,j\in[n]\backslash S. By assumption, no cell of the subdivision can contain both points eSe_{S} and eSije_{Sij}. Therefore, on the square with vertices eS,eSi,eSj,eSije_{S},e_{Si},e_{Sj},e_{Sij}, FF must induce the subdivision with an edge between eSie_{Si} and eSje_{Sj}. It follows that F(Si)+F(Sj)>F(S)+F(Sij)F(Si)+F(Sj)>F(S)+F(Sij).

(\Rightarrow) Suppose that FF is strictly submodular and suppose, for the sake of contradiction, that a cell CC of the subdivision induced by FF is not contained in a slice of the cube with the desired form. Let S[n]S\subset[n] be of minimum size |S|=k|S|=k with eSCe_{S}\in C. Let |x||x| denote ixi\sum_{i}x_{i}. By assumption, max{|x|:xC}k+2\max\{|x|:x\in C\}\geq k+2. Since eSe_{S} is a vertex of CC, there is some edge [eS,eS+v][e_{S},e_{S}+v] of CC for which |eS+v|>|eS||e_{S}+v|>|e_{S}|. By MM^{\natural}-concavity, this edge direction must be parallel to eie_{i} for some i[n]\Si\in[n]\backslash S. Then eSie_{Si} is also a vertex of CC. Since this does not achieve the maximum of |x||x| over CC, by the same reasoning, there exists some j[n]\(Si)j\in[n]\backslash(Si) for which eSijCe_{Sij}\in C. By MM^{\natural}-concavity, [eS,eSij][e_{S},e_{Sij}] cannot be an edge of CC, implying that eSjCe_{Sj}\in C as well. The only way for these four points to belong to the same cell CC is for F(Si)+F(Sj)=F(S)+F(Sij)F(Si)+F(Sj)=F(S)+F(Sij), contradicting the strict submodularity of FF. ∎

Remark 2.5.

The assumption of MM^{\natural}-concavity is necessary in Lemma 2.4. For example, consider the function F:2[4]F:2^{[4]}\to\mathbb{R} defined by F()=F([4])=6F(\emptyset)=F([4])=-6, F(i)=3F(i)=-3 for i1i\neq 1 and F(1)=0F(1)=0, F(S)=1F(S)=-1 for |S|=2|S|=2, and F(S)=3F(S)=-3 for |S|=3|S|=3 with S234S\neq 234 and F(234)=0F(234)=0. One can check that FF is strictly submodular but that the edge [e1,e234][e_{1},e_{234}] appears in the induced subdivision via the upper hull.

3. Tropical Grassmannians and Tropical Flag Varieties

In this section we will show a new relationship between tropical Grassmannians and tropical flag varieties, analogous to the equivalence (2)(4)(2)\iff(4) in Proposition 2.2.

The (complete) flag variety is a subvariety of a product of projective spaces k=1n(nk)1\prod_{k=1}^{n}\mathbb{P}^{\binom{n}{k}-1}. An element (pk)k[n](p_{k})_{k\in[n]} belongs to the flag variety if there is a complete flag of linear spaces {0}=L0L1L2Ln=𝒦n\{0\}=L_{0}\subset L_{1}\subset L_{2}\subset\ldots\subset L_{n}=\mathcal{K}^{n} so that pkp_{k} is the vector of Plücker coordinates of LkL_{k}. The (complete) affine flag variety is the variety Fl𝒦(n)𝒦2n=k=0nK(nk){\rm Fl}_{\mathcal{K}}(n)\subset\mathcal{K}^{2^{n}}=\prod_{k=0}^{n}K^{\binom{n}{k}} consisting of vectors (qS)S[n](q_{S})_{S\subseteq[n]} which, when considered as a sequence (q),(qS:|S|=1),(qS:|S|=2),,(qS:|S|=n1),(q[n])(q_{\varnothing}),(q_{S}:|S|=1),(q_{S}:|S|=2),\dots,(q_{S}:|S|=n-1),(q_{[n]}), belong to the flag variety. Each component (qS:|S|=k)(q_{S}:|S|=k) can be scaled independently by nonzero elements of 𝒦\mathcal{K}. In particular we allow (q)(q_{\varnothing}) and (q[n])(q_{[n]}) to take arbitrary values in 𝒦\mathcal{K}^{*}.

Consider the Plücker embedding of the partial flag variety

{(Lk,Lk+1)Gr𝒦(k,n)×Gr𝒦(k+1,n):LkLk+1}.\{(L_{k},L_{k+1})\in\operatorname{Gr}_{\mathcal{K}}(k,n)\times\operatorname{Gr}_{\mathcal{K}}(k+1,n):L_{k}\subset L_{k+1}\}.

Its image under the homogenizing map taking qSq_{S} to qS(n+1)q_{S(n+1)} if |S|=k|S|=k and to qSq_{S} if |S|=k+1|S|=k+1 coincides with the Plücker embedding of Gr𝒦(k+1,n+1)\operatorname{Gr}_{\mathcal{K}}(k+1,n+1). An analogous homogenization appeared in Proposition 2.2(3). Concretely, if LkL_{k} and Lk+1L_{k+1} are the span of the top kk and k+1k+1 rows of a (k+1)×n(k+1)\times n matrix MM, then the corresponding subspace in Gr𝒦(k+1,n+1)\operatorname{Gr}_{\mathcal{K}}(k+1,n+1) is obtained as the rowspan of the (k+1)×n(k+1)\times n matrix obtained by appending the column ek+1e_{k+1} to MM.

Let ν(Fl𝒦(n))Γ2[n]2[n]\nu({\rm Fl}_{\mathcal{K}}(n))\subset\Gamma^{2^{[n]}}\subset\mathbb{R}^{2^{[n]}} be the ν\nu values of the points in the affine flag variety Fl𝒦(n){\rm Fl}_{\mathcal{K}}(n), none of whose Plücker coordinates are zero. This set has lineality space containing the tropical scaling of each factor; that is, for any w:2[n]w:2^{[n]}\to\mathbb{R} in ν(Fl𝒦(n))\nu({\rm Fl}_{\mathcal{K}}(n)) and any vector of scalars λ=(λ0,,λn)Γn+1\lambda=(\lambda_{0},\ldots,\lambda_{n})\in\Gamma^{n+1}, the function λw:2[n]\lambda\cdot w:2^{[n]}\to\mathbb{R} given by

(2) (λw)(S)=λ|S|+w(S)(\lambda\cdot w)(S)=\lambda_{|S|}+w(S)

also belongs to ν(Fl𝒦(n))\nu({\rm Fl}_{\mathcal{K}}(n)). The following lemma says that the tropical flag variety ν(Fl𝒦(n))\nu({\rm Fl}_{\mathcal{K}}(n)) can be recovered from its intersection with the submodular cone using lineality.

Lemma 3.1.

For any w:2[n]w:2^{[n]}\to\mathbb{R}, there are tropical scalars λ=(λ0,,λn)Γn+1\lambda=(\lambda_{0},\ldots,\lambda_{n})\in\Gamma^{n+1} so that λw\lambda\cdot w is strictly submodular. Moreover, if ww is already submodular and Γ\Gamma is dense in \mathbb{R}, then λ\lambda can be chosen to be arbitrarily small.

Proof.

Let λ0=λ1=0\lambda_{0}=\lambda_{1}=0 and for each k=2,,nk=2,\ldots,n, inductively choose λk\lambda_{k} so that

λk<mini,jSS([n]k)w(S\i)+w(S\j)+2λk1w(S\{i,j})w(S)λk2,\lambda_{k}<\min_{\stackrel{{\scriptstyle S\in{\binom{[n]}{k}}}}{{i,j\in S}}}\ w(S\backslash i)+w(S\backslash j)+2\lambda_{k-1}-w(S\backslash\{i,j\})-w(S)-\lambda_{k-2},

which implies that λw\lambda\cdot w is strictly submodular.

If ww is already submodular, then w(S\i)+w(S\j)w(S\{i,j})w(S)0w(S\backslash i)+w(S\backslash j)-w(S\backslash\{i,j\})-w(S)\geq 0. We can choose λ0=λ1=0\lambda_{0}=\lambda_{1}=0 and λ2=ε\lambda_{2}=-\varepsilon where ε>0\varepsilon>0 is arbitrarily small. For k3k\geq 3 we can choose λk\lambda_{k} to satisfy 3λk1<λk<min{0,2λk1λk2}3\lambda_{k-1}<\lambda_{k}<\min\{0,2\lambda_{k-1}-\lambda_{k-2}\}. Inductively, this gives |λk|<3k1ε|\lambda_{k}|<3^{k-1}\varepsilon, thus λ\lambda can be arbitrarily small. ∎

Consider the linear subspace n\mathcal{L}_{n} of ([2n]n)\mathbb{R}^{\binom{[2n]}{n}} defined as

(3) n={p([2n]n):p(C)=p(D) whenever C[n]=D[n]}.\mathcal{L}_{n}=\left\{p\in\mathbb{R}^{\binom{[2n]}{n}}:p(C)=p(D)\text{ whenever }C\cap[n]=D\cap[n]\right\}.

The linear subspace n\mathcal{L}_{n} is isomorphic to 2[n]\mathbb{R}^{2^{[n]}} via the map

ϕ:n2[n]\phi:\mathcal{L}_{n}\xrightarrow[]{\ \ \cong\ \ }\mathbb{R}^{2^{[n]}}

sending any pnp\in\mathcal{L}_{n} to ϕ(p)2[n]\phi(p)\in\mathbb{R}^{2^{[n]}} given by ϕ(p)(S)=p(ST)\phi(p)(S)=p(S\cup T) with TT any subset of {n+1,,2n}\{n+1,\dots,2n\} of size n|S|n-|S|.

The equivalence (2)(4)(2)\iff(4) in Proposition 2.2 can be rephrased as saying

Dr(n,2n)n=FlDr(n)SUBMODn,{\rm Dr}(n,2n)\cap\mathcal{L}_{n}\ \ =\ \ {\rm FlDr}(n)\cap{\rm SUBMOD}_{n},

where SUBMODn{\rm SUBMOD}_{n} denotes the cone of submodular functions on {0,1}n\{0,1\}^{n} and we identify points in n\mathcal{L}_{n} with points in 2[n]\mathbb{R}^{2^{[n]}} as above. Our main result in this section, stated below, says that the realizable parts of these sets coincide. In the next section we show that these sets coincide with the tropicalization of the set of principal minors of positive definite matrices when 𝒦\mathcal{K} is equal to either \mathcal{R} or 𝒞\mathcal{C}. However, the result holds for more general fields 𝒦\mathcal{K} and may be of independent interest.

Theorem 3.2.

Let 𝒦\mathcal{K} be a field with a nonarchimedean valuation with a splitting and an infinite residue field. For any positive integer nn we have

ν(Gr𝒦(n,2n))n=ν(Fl𝒦(n))SUBMODn.\nu({\rm Gr}_{\mathcal{K}}(n,2n))\cap\mathcal{L}_{n}\ \ =\ \ \nu({\rm Fl}_{\mathcal{K}}(n))\cap{\rm SUBMOD}_{n}.

If 𝒦\mathcal{K} is algebraically closed or real closed, we thus have

(4) trop(Gr𝒦(n,2n))n=trop(Fl𝒦(n))SUBMODn.\operatorname{trop}({\rm Gr}_{\mathcal{K}}(n,2n))\cap\mathcal{L}_{n}\ \ =\ \ \operatorname{trop}({\rm Fl}_{\mathcal{K}}(n))\cap{\rm SUBMOD}_{n}.

Before providing the proof of our main result, we remark on the dimension of these polyhedral complexes.

Proposition 3.3.

If 𝒦\mathcal{K} is algebraically closed or real closed, the polyhedral complex in Equation (4) in Theorem 3.2 has dimension (n+12)+1\binom{n+1}{2}+1.

Proof.

The affine tropical flag variety trop(Fl𝒦(n))k=0n(nk)\operatorname{trop}({\rm Fl}_{\mathcal{K}}(n))\subset\prod_{k=0}^{n}\mathbb{R}^{\binom{n}{k}} is invariant under the tropical scaling of each factor as in Equation (2). Therefore, by Lemma 3.1, any polyhedral cone in the affine tropical flag variety has the same dimension as its intersection with the submodular cone. Thus it suffices to show that trop(Fl𝒦(n))\operatorname{trop}({\rm Fl}_{\mathcal{K}}(n)) has dimension (n+12)+1\binom{n+1}{2}+1.

The flag variety has dimension (n2)\binom{n}{2}, as a dense subset of it can be parametrized by upper triangular matrices with ones on the diagonal. Thus the affine flag variety Fl𝒦(n){\rm Fl}_{\mathcal{K}}(n) has dimension (n+12)+1=(n2)+n+1\binom{n+1}{2}+1=\binom{n}{2}+n+1, accounting for the n+1n+1 extra dimensions from scaling. If 𝒦\mathcal{K} is algebraically closed, this implies that the tropicalization trop(Fl𝒦(n))\operatorname{trop}({\rm Fl}_{\mathcal{K}}(n)) is a pure polyhedral complex of the same dimension (n+12)+1\binom{n+1}{2}+1 by the Structure Theorem of Tropical Geometry.

If 𝒦\mathcal{K} is not algebraically closed, it is a priori possible that trop(Fl𝒦(n))\operatorname{trop}({\rm Fl}_{\mathcal{K}}(n)) is a (not necessarily pure) polyhedral complex of smaller dimension. However, when 𝒦\mathcal{K} is real closed, it was shown in [Bor22, Proposition 4.10trop4.10^{\rm trop}] that the tropicalization of the positive part Fl𝒦>0(n){\rm Fl}_{\mathcal{K}}^{>0}(n) of the flag variety, which is contained in trop(Fl𝒦(n))\operatorname{trop}({\rm Fl}_{\mathcal{K}}(n)), has dimension (n+12)+1\binom{n+1}{2}+1, by showing that the Marsh-Rietsch (bijective) parameterization 𝒦>0(n+12)+1Fl𝒦>0(n)\mathcal{K}_{>0}^{\binom{n+1}{2}+1}\to{\rm Fl}_{\mathcal{K}}^{>0}(n) tropicalizes to a (bijective) parametrization >0(n+12)+1trop(Fl𝒦>0(n))\mathbb{R}_{>0}^{\binom{n+1}{2}+1}\to\operatorname{trop}({\rm Fl}_{\mathcal{K}}^{>0}(n)). In [Bor22], the tropical flag variety is considered projectively, so the dimensions used there do not account for the n+1n+1 extra dimensions from scaling. The map easily extends to the affine setting. ∎

Proof of Theorem 3.2(\subseteq).

Since both ν(Gr𝒦(n,2n))n\nu({\rm Gr}_{\mathcal{K}}(n,2n))\cap\mathcal{L}_{n} and ν(Fl𝒦(n))SUBMODn\nu({\rm Fl}_{\mathcal{K}}(n))\cap{\rm SUBMOD}_{n} are closed under global tropical scaling (i.e., adding a constant function), we may assume all the functions w2[n]w\in 2^{[n]} under consideration have w()=0w(\varnothing)=0.

Let M𝒦n×2nM\in\mathcal{K}^{n\times 2n} be such that the valuation of its n×nn\times n minors belongs to n\mathcal{L}_{n}. The submatrix M([n],{n+1,,2n})M([n],\{n+1,\dots,2n\}) has valuation 0 by our assumption, so we can assume that MM has the form (BI)\begin{pmatrix}B&I\end{pmatrix} for some n×nn\times n matrix B{B} and the n×nn\times n identity matrix II. Since the valuation of the n×nn\times n minors of MM belongs to n\mathcal{L}_{n}, the valuation of the minor B(T,S){B}(T,S) is determined by SS independently of TT.

Consider the flag L1L2Ln=𝒦nL_{1}\subset L_{2}\subset\dots\subset L_{n}=\mathcal{K}^{n} where LkL_{k} is the span of the first kk rows of B{B}. Its Plücker coordinates are given by the minors (B({1,,k},S):S([n]k))({B}(\{1,\ldots,k\},S):S\in\binom{[n]}{k}), which are equal to the corresponding n×nn\times n minors of MM. This shows that the valuation of the n×nn\times n minors if MM belongs to ν(Fl𝒦(n))\nu({\rm Fl}_{\mathcal{K}}(n)). Moreover, since ν(Gr𝒦(n,2n))\nu({\rm Gr}_{\mathcal{K}}(n,2n)) is contained in the Dressian Dr(n,2n){\rm Dr}(n,2n), submodularity follows from the (5)(4)(5)\implies(4) part of Proposition 2.2. ∎

For the other inclusion, we need the following technical lemmas. Recall that ν\nu is the negative of the valuation.

Lemma 3.4.

Let B𝒦n×nB\in\mathcal{K}^{n\times n} be an upper triangular matrix such that the function 2[n]2^{[n]}\rightarrow\mathbb{R} given by Sν(B({1,,|S|},S))S\mapsto\nu(B(\{1,\ldots,|S|\},S)) is submodular. Then

(5) ν(B({1,,|S|},S))ν(B(T,S))\nu(B(\{1,\ldots,|S|\},S))\geq\nu(B(T,S))

for all S,T[n]S,T\subset[n] with |S|=|T||S|=|T|.

We defer the proof of Lemma 3.4 to Section 6. We will say that BB is top heavy if it is upper triangular and satisfies the condition (5).

Lemma 3.5.

Let 𝒦\mathcal{K} be a valued field with a splitting and an infinite residue field. Let B,C𝒦n×nB,C\in\mathcal{K}^{n\times n}, B~=CB\tilde{B}=CB and suppose CC is generic with entry-wise valuation zero. Then the valuation of a minor of B~\tilde{B} depends only on the choice of the columns, not on the choice of the rows. More precisely, for any subsets S,T[n]S,T\subseteq[n] of size |S|=|T|=k|S|=|T|=k,

ν(B~(T,S))=maxT([n]k)ν(B(T,S)).\nu(\tilde{B}(T,S))=\max_{T^{\prime}\in\binom{[n]}{k}}\nu(B(T^{\prime},S)).

In particular, the negative valuation of the n×nn\times n minors of the matrix (B~In)\begin{pmatrix}\tilde{B}&I_{n}\end{pmatrix} belongs to ν(Gr𝒦(n,2n))n\nu({\rm Gr}_{\mathcal{K}}(n,2n))\cap\mathcal{L}_{n}.

If, in addition, BB is a top heavy matrix as defined by (5) and CC is generic lower triangular with valuation zero for the nonzero entries, then

ν(B~(T,S))=ν(B([k],S))\nu(\tilde{B}(T,S))=\nu(B([k],S))

for all subsets S,T[n]S,T\subseteq[n] of the same size kk.

Here “generic” means that the leading coefficients of the entries of CC lie in a nonempty Zariski open set over the residue field. We are defining the leading coefficient using the splitting. If 𝒦\mathcal{K} is a Puiseux series field over a ground field 𝕜\mathbbm{k}, then we can take CC to be a generic matrix over 𝕜\mathbbm{k}.

Proof.

Consider two k×kk\times k submatrices of B~\tilde{B} on the same set of columns. We can transform one to the other by swapping two rows at a time, so let us assume that these two submatrices differ only at one row. Each row of B~\tilde{B} is a generic linear combination of rows of BB. By linearity of determinant, each of the two minors is a generic linear combination of the same nn minors where the row being swapped is replaced by each of the original rows. By genericity there is no cancellation of leading terms so the linear combinations must have the same valuation, as desired.

Now suppose BB is top heavy and CC is lower triangular. Then each row of B~\tilde{B} is a generic linear combination of a top-justified subset of rows of BB. When we repeatedly expand a k×kk\times k minor of B~\tilde{B} using the linearity of determinants, we see that its valuation agrees with the valuation of the top justified minor by the top-heaviness of BB. ∎

Proof of Theorem 3.2(\supseteq).

Suppose that the function w:2[n]w:2^{[n]}\to\mathbb{R} belongs to ν(Fl𝒦(n))SUBMODn{\nu}({\rm Fl}_{\mathcal{K}}(n))\cap{\rm SUBMOD}_{n}. We will show that it belongs to ν(Gr𝒦(n,2n))n\nu({\rm Gr}_{\mathcal{K}}(n,2n))\cap\mathcal{L}_{n}. Since both of these sets are closed under tropical scaling (adding a constant function to ww), we may assume that w()=0w(\varnothing)=0. Then

w(S)=ν(B([k],S))λk,w(S)=\nu(B([k],S))-\lambda_{k},

where BB is an upper triangular n×nn\times n matrix over 𝒦\mathcal{K} and λ1,λn\lambda_{1},\ldots\lambda_{n} are in the value group Γ\Gamma. The rows v1,,vnv_{1},\ldots,v_{n} of BB represent the flag

span{v1}span{v1,v2}span{v1,,vn}=𝒦n{\rm span}\{v_{1}\}\subset{\rm span}\{v_{1},v_{2}\}\subset\dots\subset{\rm span}\{v_{1},\ldots,v_{n}\}=\mathcal{K}^{n}

in the flag variety and λ1,λn\lambda_{1},\ldots\lambda_{n} results from scaling of the Plücker coordinates of each linear space in the flag. Since w(S)w(S) is finite for every SS, the minors B([k],[k])B([k],[k]) are nonsingular. After rescaling the rows v1tλ1v1v_{1}\rightarrow t^{\lambda_{1}}v_{1}, v2tλ2λ1v2v_{2}\rightarrow t^{\lambda_{2}-\lambda_{1}}v_{2}, v3tλ3λ2v3,,vntλnλn1vnv_{3}\rightarrow t^{\lambda_{3}-\lambda_{2}}v_{3},\dots,v_{n}\rightarrow t^{\lambda_{n}-\lambda_{n-1}}v_{n} we may assume that

w(S)=ν(B([k],S)).w(S)=\nu(B([k],S)).

By Lemma 3.4, BB is top heavy. Let B~=LB\tilde{B}=LB where LL is a generic lower triangular matrix as in Lemma 3.5, so for any S[n]S\subset[n] with |S|=k|S|=k

w(S)=ν(B([k],S))=ν(B~(T,S))w(S)=\nu(B([k],S))=\nu(\tilde{B}(T,S))

for all T([n]k)T\in{\binom{[n]}{k}} by Lemma 3.5. Since w()=0w(\varnothing)=0, the vector ww coincides with the ν\nu values of the n×nn\times n minors of the matrix (B~I)(\tilde{B}~{}I) under the identification of 2[n]\mathbb{R}^{2^{[n]}} with the linear space n\mathcal{L}_{n} as in (3), and so ww belongs to ν(Gr𝒦(n,2n))n\nu({\rm Gr}_{\mathcal{K}}(n,2n))\cap\mathcal{L}_{n}. ∎

Remark 3.6.

El Maazouz [EM22] defines the entropy map associated to a lattice Λ=B𝒪n𝒦n\Lambda=B\mathcal{O}^{n}\subset\mathcal{K}^{n} to be the function on {0,1}n\{0,1\}^{n} defined by Imin|J|=|I|val(det(B(I,J)))I\mapsto\min_{|J|=|I|}\operatorname{val}(\det(B(I,J))), which coincides with the formula in Lemma 3.5 above up to sign and transpose. Interestingly, in [EM22], this function is primarily considered over fields 𝒦\mathcal{K} with finite residue field, for which Lemma 3.5 does not apply. For fixed nn, one should be able to amend the arguments above to apply when the residue field 𝕜\mathbbm{k} is finite but sufficiently large.

Example 3.7 (Example 2.3 continued).

Consider the matrix BB from Example 2.3. One can check that the function Sν(B([|S|],S)S\mapsto\nu(B([|S|],S) on {0,1}3\{0,1\}^{3} is submodular. Since BB is also upper triangular, it follows from Lemma 3.4 that BB is top heavy. Consider a “generic” lower triangular matrix

C=(100110211) with B~=CB=(11111+t21+t22+t22+t+t2).C=\begin{pmatrix}1&0&0\\ 1&1&0\\ 2&1&1\end{pmatrix}\ \ \text{ with }\ \ \tilde{B}=CB=\begin{pmatrix}1&1&1\\ 1&1+t^{2}&1+t\\ 2&2+t^{2}&2+t+t^{2}\end{pmatrix}.

Then by the Cauchy-Binet identity,

B~(23,23)\displaystyle\tilde{B}(23,23) =C(23,12)B(12,23)+C(23,13)B(13,23)+C(23,23)B(23,23)\displaystyle=C(23,12)B(12,23)+C(23,13)B(13,23)+C(23,23)B(23,23)
=(1)(tt2)+(1)(t2)+(1)(t4)=t+2t2+t4.\displaystyle=(-1)(t-t^{2})+(1)(t^{2})+(1)(t^{4})=-t+2t^{2}+t^{4}.

In particular, ν(B~(23,23))=1\nu(\tilde{B}(23,23))=-1 is equal to the maximum ν\nu value of a minor B(ij,23)B(ij,23), which is achieved by ij=12ij=12, since BB is top-heavy. Similarly, one can check that ν(B~(ij,23))=1\nu(\tilde{B}(ij,23))=-1 for every pair ijij. \diamond

4. Tropicalizing principal minors of positive definite matrices

In this section we show that the tropicalization of the image of the positive definite cone under the principal minor map coincides with the subsets of tropical Grassmannians and tropical flag varieties studied in the previous section. As before we will use \mathcal{R} and 𝒞\mathcal{C} to denote a real closed field and an algebraically closed field respectively, with nontrivial nonarchimedean valuation.

Theorem 4.1.

For any positive integer nn and a field 𝒦=\mathcal{K}=\mathcal{R} or 𝒞\mathcal{C}, we have

ν(cone(PMn+(𝒦))=ν(Gr𝒦(n,2n))n=ν(Fl𝒦(n))SUBMODn,\nu({\rm cone}(\operatorname{PM}^{+}_{n}(\mathcal{K}))\ \ =\ \ \nu({\rm Gr}_{\mathcal{K}}(n,2n))\cap\mathcal{L}_{n}\ \ =\ \ \nu({\rm Fl}_{\mathcal{K}}(n))\cap{\rm SUBMOD}_{n},

where cone(PMn+(𝒦))={λa:λ𝒦,aPMn+(𝒦)}{\rm cone}(\operatorname{PM}^{+}_{n}(\mathcal{K}))=\{\lambda a:\lambda\in\mathcal{K},a\in\operatorname{PM}^{+}_{n}(\mathcal{K})\}. Taking closures gives

trop(cone(PMn+(𝒦))=trop(Gr𝒦(n,2n))n=trop(Fl𝒦(n))SUBMODn.\operatorname{trop}({\rm cone}(\operatorname{PM}^{+}_{n}(\mathcal{K}))\ \ =\ \ \operatorname{trop}({\rm Gr}_{\mathcal{K}}(n,2n))\cap\mathcal{L}_{n}\ \ =\ \ {\rm trop}({\rm Fl}_{\mathcal{K}}(n))\cap{\rm SUBMOD}_{n}.
Proof.

Let 𝒦=\mathcal{K}=\mathcal{R} or 𝒞\mathcal{C}, and let 𝕜\mathbbm{k} denote its residue field. The equality of the last two sets follows from Theorem 3.2. We will now show that the first two sets are equal.

(\subseteq) Let AA be a positive definite matrix over 𝒦\mathcal{K}. Then A=BBA=B^{*}B for some B𝒦n×nB\in\mathcal{K}^{n\times n}. By Cauchy-Binet,

A(S,S)=T([n]k)B(S,T)B(T,S)=T([n]k)B(T,S)¯B(T,S).A(S,S)=\sum_{T\in\binom{[n]}{k}}B^{*}(S,T)B(T,S)=\sum_{T\in\binom{[n]}{k}}\overline{B(T,S)}B(T,S).

Since all of the terms in the first sum are nonnegative, the tropicalization (ν\nu values) of the sum is the tropical sum (maximum) of the tropicalization of the summands. Moreover, for any a𝒦a\in\mathcal{K}, ν(a)=ν(a¯)\nu(a)=\nu(\overline{a}), so ν(a¯a)=2ν(a)\nu(\overline{a}a)=2\nu(a). All together this gives

(6) ν(A(S,S))=maxT([n]k)2ν(B(T,S)).\nu(A(S,S))=\max_{T\in\binom{[n]}{k}}2\nu(B(T,S)).

Let U𝕜n×nU\in\mathbbm{k}^{n\times n} be generic and let B~=UB\tilde{B}=UB. Then it follows from (6) and Lemma 3.5 that

val(A(S,S))=2val(B~(T,S))\operatorname{val}(A(S,S))=2\operatorname{val}(\tilde{B}(T,S))

for any T[n]T\subset[n] with |T|=|S||T|=|S|. The values ν(B~(T,S))\nu(\tilde{B}(T,S)) are tropicalizations of n×nn\times n minors of the n×2nn\times 2n matrix (B~I)(\tilde{B}~{}I), and they lie in the linear subspace n\mathcal{L}_{n}. Thus

12val(A(S,S))trop(Gr𝒦(n,2n))n.\frac{1}{2}\operatorname{val}(A(S,S))\in\operatorname{trop}({\rm Gr}_{\mathcal{K}}(n,2n))\cap\mathcal{L}_{n}.

However trop(Gr𝒦(n,2n))\operatorname{trop}({\rm Gr}_{\mathcal{K}}(n,2n)) and n\mathcal{L}_{n} are invariant under scaling, so we also have

val(A(S,S))trop(Gr𝒦(n,2n))n.\operatorname{val}(A(S,S))\in\operatorname{trop}({\rm Gr}_{\mathcal{K}}(n,2n))\cap\mathcal{L}_{n}.

(\supseteq) Suppose M𝒦n×2nM\in\mathcal{K}^{n\times 2n} is such that the ν\nu values of the n×nn\times n minors of MM belong to n\mathcal{L}_{n}. We may assume that the valuation of the minor M([n],{n+1,,2n})M([n],\{n+1,\ldots,2n\}) is zero, and that MM has the form (BI)\begin{pmatrix}B&I\end{pmatrix} for some n×nn\times n matrix BB. Then the valuation of the minors B(T,S)B(T,S) are independent of TT. Consider the positive definite matrix A=BTBA=B^{T}B. Then for any S,T[n]S,T\subset[n] with |S|=|T|=k|S|=|T|=k, by (6) we have

ν(B(T,S))=12maxT([n]k)2ν(B(T,S))=12ν(A(S,S)),\nu(B(T,S))=\frac{1}{2}\max_{T^{\prime}\in\binom{[n]}{k}}2\nu(B(T^{\prime},S))=\frac{1}{2}\nu(A(S,S)),

completing the proof. ∎

Corollary 4.2.

For 𝒦=\mathcal{K}=\mathcal{R} or 𝒞\mathcal{C}, the set trop(PMn+(𝒦))\operatorname{trop}(\operatorname{PM}_{n}^{+}(\mathcal{K})) has dimension (n+12)\binom{n+1}{2}.

Proof.

The set cone(PMn+(𝒦))\operatorname{cone}(\operatorname{PM}_{n}^{+}(\mathcal{K})) is obtained from PMn+(𝒦)\operatorname{PM}_{n}^{+}(\mathcal{K}) by global scaling. Its dimension and the dimension of its tropicalization is therefore one more than the respective dimensions for PMn+(𝒦)\operatorname{PM}_{n}^{+}(\mathcal{K}). The result then follows from Proposition 3.3. ∎

Corollary 4.3.

The set of tropicalized principal minors of n×nn\times n positive definite matrices is contained in the set of MM^{\natural}-concave functions on 2[n]2^{[n]} with value zero on \varnothing. This containment is an equality for n5n\leq 5, but it is strict for n6n\geq 6.

Proof.

By Theorem 4.1, the set of tropicalized principal minors of n×nn\times n positive definite matrices coincides with the tropical flag variety inside the submodular cone. The MM^{\natural}-concave functions coincide with the flag Dressian by Proposition 2.2. By Section 5 of [BEZ21], the tropical flag variety is contained in the flag Dressian; this containment is an equality for n5n\leq 5 and is strict for n6n\geq 6. This is still the case after intersecting with the submodular cone by Lemma 3.1. ∎

If w2[n]w\in\mathbb{R}^{2^{[n]}}, for each k[n]k\in[n] let wkw_{k} denote the restriction of ww to subsets of size kk. As discussed in Section 3, the fact that ν(PMn+(𝒦))ν(Fl𝒦(n))\nu(\operatorname{PM}^{+}_{n}(\mathcal{K}))\subset\nu({\rm Fl}_{\mathcal{K}}(n)) implies the following corollary.

Corollary 4.4.

For wν(PMn+(𝒦))w\in\nu(\operatorname{PM}^{+}_{n}(\mathcal{K})), the homogenization of the restriction (wk,wk+1)(w_{k},w_{k+1}) belongs to ν(Gr𝒦(k+1,n+1))\nu(\operatorname{Gr}_{\mathcal{K}}(k+1,n+1)) for any 0kn10\leq k\leq n-1.

Recall from Lemma 2.4 that if a function w:2[n]w:2^{[n]}\rightarrow\mathbb{R} is MM^{\natural}-concave and strictly submodular, then each cell in the regular subdivision of {0,1}n\{0,1\}^{n} induced by ww via upper hull is contained in a layer {x[0,1]n:kixik+1}\{x\in[0,1]^{n}:k\leq\sum_{i}x_{i}\leq k+1\} for some kk. Even when wν(PMn+(𝒦))w\in\nu(\operatorname{PM}^{+}_{n}(\mathcal{K})) is not strictly submodular, we can perturb ww and get a subdivision where every cell is contained in one of these layers.

Theorem 4.5.

The regular subdivision of [0,1]n[0,1]^{n} induced by the tropicalization of the principal minors of any positive definite matrix is a coarsening of a subdivision of [0,1]n[0,1]^{n} into dehomogenized realizable matroid polytopes. Specifically, for points in ν(PMn+())\nu(\operatorname{PM}^{+}_{n}(\mathcal{R})) or ν(PMn+(𝒞))\nu(\operatorname{PM}^{+}_{n}(\mathcal{C})), the matroids are realizable over \mathbb{R} or \mathbb{C}, respectively.

Proof.

Let wν(PMn+(𝒦))w\in\nu(\operatorname{PM}^{+}_{n}(\mathcal{K})) for 𝒦=\mathcal{K}=\mathcal{R} or 𝒞\mathcal{C}. By Theorem 4.1, ww belongs to ν(Fl𝒦(n))SUBMODn\nu({\rm Fl}_{\mathcal{K}}(n))\cap{\rm SUBMOD}_{n}. By Lemma 3.1, for arbitrarily small λΓn+1\lambda\in\Gamma^{n+1}, the point λw\lambda\cdot w is strictly submodular. It also belongs to ν(Fl𝒦(n))\nu({\rm Fl}_{\mathcal{K}}(n)) and, in particular, is MM^{\natural}-concave. By Lemma 2.4, each cell of the subdivision induced by λw\lambda\cdot w is a subset of {x[0,1]n:kixik+1}\{x\in[0,1]^{n}:k\leq\sum_{i}x_{i}\leq k+1\} for some kk. Because λ\lambda was taken arbitarily small, the subdivision induced by λw\lambda\cdot w is a refinement of that induced by ww. We conclude by checking that each cell induced by λw\lambda\cdot w is a dehomogenized realizable matroid polytope.

Let w=(wk,wk+1)w^{\prime}=(w^{\prime}_{k},w^{\prime}_{k+1}) denote the restriction of λw\lambda\cdot w to subsets of size kk and k+1k+1. Since λwν(Fl𝒦(n))\lambda\cdot w\in\nu({\rm Fl}_{\mathcal{K}}(n)), the homogenization of ww^{\prime} is an element of ν(Gr𝒦(k+1,n+1))\nu(\operatorname{Gr}_{\mathcal{K}}(k+1,n+1)) as described in Section 3. In particular, any cell in the subdivision of {x[0,1]n+1:ixi=k+1}\{x\in[0,1]^{n+1}:\sum_{i}x_{i}=k+1\} induced by the homogenization of ww^{\prime} is the matroid polytope of a matroid realizable over the residue field of 𝒦\mathcal{K}. For 𝒦=\mathcal{K}=\mathcal{R} or 𝒞\mathcal{C}, it follows that the matroid is realizable over \mathbb{R} or \mathbb{C}, respectively. Dehomogenizing, i.e., dropping the coordinate xn+1x_{n+1}, gives the result. ∎

5. Inequalities from MM^{\natural}-concavity

By the Lifting Lemma [JSY22], for any semi-algebraic set SS, any tropical polynomial inequality valid on trop(S)\operatorname{trop}(S) can be lifted to a usual polynomial inequality valid on SS. In this section we use the theory of Lorentzian polynomials [BH20, AGV18] to derive new inequalities on principal minors on positive definite matrices that lift the MM^{\natural}-concavity inequalities satisfied by the tropicalization, based on the fact that the polynomial f=det(diag(x1,,xn)+A)f=\det({\rm diag}(x_{1},\ldots,x_{n})+A) is Lorentzian for any positive definite matrix AA. Furthermore, we provide examples to show that these inequalities are tight.

For D={d1,,dk}[n]D=\{d_{1},\ldots,d_{k}\}\subset[n], let Df\partial^{D}f denote the partial derivatives of ff with respect to the variables xd1,,xdkx_{d_{1}},\ldots,x_{d_{k}}. For S[n]S\subset[n], let ASA_{S} denote the principal minor indexed on the rows and columns by SS. To simplify notation, we write Sij=S{i,j}Sij=S\cup\{i,j\}. We first derive inequalities on the coefficients of arbitrary Lorentzian polynomials.

Theorem 5.1.

Let f=S[n]cS𝐱[n]\Sy|S|f=\sum_{S\subseteq[n]}c_{S}\,{\bf x}^{[n]\backslash S}y^{|S|} be a Lorentzian polynomial in the variables x1,,xn,yx_{1},\dots,x_{n},y. If n4n\geq 4, then for any S[n]{1,2,3,4}S\subseteq[n]\setminus\{1,2,3,4\} and any rr\in\mathbb{R}, the coefficients of ff satisfy

(r+1)cS14cS23+r(r+1)cS13cS24rcS12cS34(r+1)\,c_{S14}c_{S23}+r(r+1)\,c_{S13}c_{S24}\geq r\,c_{S12}c_{S34}

and all analogous inequalities obtained by permutations of indices. Similarly, if n3n\geq 3, then for any S[n]{1,2,3}S\subseteq[n]\setminus\{1,2,3\} and any rr\in\mathbb{R}, the coefficients satisfy

(r+1)cS1cS23+r(r+1)cS2cS13rcS3cS12(r+1)\,c_{S1}c_{S23}+r(r+1)\,c_{S2}c_{S13}\geq r\,c_{S3}c_{S12}

and all analogous inequalities obtained by permutations of indices.

Moreover, both of these inequalities are tight for the subset of Lorentzian polynomials whose coefficients cSc_{S} are the principal minors ASA_{S} of a positive semidefinite matrix AA.

We will make use of the following lemma.

Lemma 5.2.

If q=c34x1x2+c24x1x3+c23x1x4+c14x2x3+c13x2x4+c12x3x4q=c_{34}x_{1}x_{2}+c_{24}x_{1}x_{3}+c_{23}x_{1}x_{4}+c_{14}x_{2}x_{3}+c_{13}x_{2}x_{4}+c_{12}x_{3}x_{4} is Lorentzian, then for any rr\in\mathbb{R}

(r+1)c14c23+r(r+1)c13c24rc12c34.(r+1)\,c_{14}c_{23}+r(r+1)\,c_{13}c_{24}\geq r\,c_{12}c_{34}.
Proof.

By the definition of Lorentzian polynomials, the coefficients cijc_{ij} are all nonnegative and the symmetric matrix

Q=2q=(0c34c24c23c340c14c13c24c140c12c23c13c120)Q=\nabla^{2}q=\begin{pmatrix}0&c_{34}&c_{24}&c_{23}\\ c_{34}&0&c_{14}&c_{13}\\ c_{24}&c_{14}&0&c_{12}\\ c_{23}&c_{13}&c_{12}&0\end{pmatrix}

representing the quadratic form qq has at most one positive eigenvalue. It follows that det(Q)0\det(Q)\leq 0.

The determinant of QQ is equal to the discriminant of the quadratic polynomial

(7) c13c24r2+(c13c24+c14c23c12c34)r+c14c23=(r+1)c14c23+r(r+1)c13c24rc12c34c_{13}c_{24}\,r^{2}+(c_{13}c_{24}+c_{14}c_{23}-c_{12}c_{34})\,r+c_{14}c_{23}=(r+1)\,c_{14}c_{23}+r(r+1)\,c_{13}c_{24}-r\,c_{12}c_{34}

with respect to rr.

Since the extremal coefficients c13c24c_{13}c_{24} and c14c23c_{14}c_{23} and nonnegative and the discriminant is det(Q)0\det(Q)\leq 0, it follows that the quadratic polynomial (7) is nonnegative for all rr\in\mathbb{R}. ∎

Proof of Theorem 5.1.

We first show that the first inequality implies the second. If ff is Lorentzian, then so is its polarization

Pol(f)=S[n]cS(n|S|)𝐱[n]\Se|S|(y1,,yn),{\rm Pol}(f)=\sum_{S\subseteq[n]}\frac{c_{S}}{\binom{n}{|S|}}{\bf x}^{[n]\backslash S}e_{|S|}(y_{1},\ldots,y_{n}),

where ej(y1,,yn)e_{j}(y_{1},\ldots,y_{n}) is the jjth elementary symmetric polynomial in y1,,yny_{1},\ldots,y_{n}; see [BH20, Prop. 3.1]. For any set S[n]{1,2,3}S\subseteq[n]\setminus\{1,2,3\} with |S|=k1n3|S|=k-1\leq n-3 and i,j{1,2,3}i,j\in\{1,2,3\}, the coefficient of 𝐱[n]\Sij𝐲[k1]{\bf x}^{[n]\backslash Sij}{\bf y}^{[k-1]} in Pol(f){\rm Pol}(f) is (nk1)1cSij\binom{n}{k-1}^{-1}c_{Sij} and the coefficient of 𝐱[n]\Si𝐲[k]{\bf x}^{[n]\backslash Si}{\bf y}^{[k]} in Pol(f){\rm Pol}(f) is (nk)1cSi\binom{n}{k}^{-1}c_{Si}. Assuming the first inequality holds for all Lorentzian polynomials and applying it to Pol(f){\rm Pol}(f) with the variables (x1,x2,x3,yk)(x_{1},x_{2},x_{3},y_{k}) in place of (x1,x2,x3,x4)(x_{1},x_{2},x_{3},x_{4}) and {xi:iS}{y1,,yk1}\{x_{i}:i\in S\}\cup\{y_{1},\ldots,y_{k-1}\} in place of {xi:iS}\{x_{i}:i\in S\} gives the desired inequality.

Now we show the first inequality. For general D[n]D\subseteq[n], Df=ScS𝐱[n]\(SD)y|S|\partial^{D}f=\sum_{S}c_{S}{\bf x}^{[n]\backslash(S\cup D)y^{|S|}} where the sum is taken over S[n]S\subseteq[n] for which SD=S\cap D=\emptyset. Now we fix S[n]\{1,2,3,4}S\subseteq[n]\backslash\{1,2,3,4\} and let D=[n]\(S{1,2,3,4})D=[n]\backslash(S\cup\{1,2,3,4\}). Consider the polynomial gg obtained by specializing the derivative y|S|+2Df\partial^{|S|+2}_{y}\partial^{D}f to y=0y=0 and xi=0x_{i}=0 for iSi\in S:

g=(y|S|+2Df)|{y=0,xi=0iS}.g=\left(\partial^{|S|+2}_{y}\partial^{D}f\right)|_{\{y=0,\ x_{i}=0\ \forall i\in S\}}.

Taking derivatives and specializing variables to zero preserves Lorentzian-ity by [BH20, Thm. 2.30], so gg is Lorentzian. One can check that gg is given by

g=(|S|+2)!(c34Sx1x2+c24Sx1x3+c23Sx1x4+c14Sx2x3+c13Sx2x4+c12Sx3x4).g=(|S|+2)!(c_{34S}x_{1}x_{2}+c_{24S}x_{1}x_{3}+c_{23S}x_{1}x_{4}+c_{14S}x_{2}x_{3}+c_{13S}x_{2}x_{4}+c_{12S}x_{3}x_{4}).

The desired inequality then holds by Lemma 5.2.

Finally, in Examples 5.4 and 5.5 we will see that these inequalities are tight for the minors of positive semidefinite matrices when n=4n=4 and 33, respectively, and S=S=\emptyset. Appending identity matrices of arbitrary size shows that these inequalities are tight for general nn and SS. ∎

Corollary 5.3.

Let AA be an n×nn\times n positive semidefinite matrix. If n4n\geq 4, then for any S[n]\{1,2,3,4}S\subseteq[n]\backslash\{1,2,3,4\} and any rr\in\mathbb{R},

(r+1)AS14AS23+r(r+1)AS13AS24rAS12AS34(r+1)\,A_{S14}A_{S23}+r(r+1)\,A_{S13}A_{S24}\geq r\,A_{S12}A_{S34}

Similarly, if n3n\geq 3, then for any S[n]\{1,2,3}S\subseteq[n]\backslash\{1,2,3\} and any rr\in\mathbb{R},

(r+1)AS1AS23+r(r+1)AS2AS13rAS3AS12.(r+1)\,A_{S1}A_{S23}+r(r+1)\,A_{S2}A_{S13}\geq r\,A_{S3}A_{S12}.

Moreover these inequalities are tight for any rr\in\mathbb{R}.

Proof.

The polynomial f=det(diag(x1,,xn)+yA)=S[n]AS𝐱[n]\Sy|S|f=\det({\rm diag}(x_{1},\ldots,x_{n})+yA)=\sum_{S\subseteq[n]}A_{S}{\bf x}^{[n]\backslash S}y^{|S|} is stable [BB06] and hence Lorentzian [BH20]. The result then follows from Theorem 5.1. ∎

Example 5.4.

For rr\in\mathbb{R}, consider the matrix

A=(111r12r+22r11r+2r22r+2r2+3r1r2r1r2+3r12r22r+1).A=\left(\begin{array}[]{cccc}1&1&1&r\\ 1&2&-r+2&2r-1\\ 1&-r+2&r^{2}-2r+2&-r^{2}+3r-1\\ r&2r-1&-r^{2}+3r-1&2r^{2}-2r+1\end{array}\right).

One can check that AA is a positive semidefinite matrix of rank two. For example, all of the principal minors are sums of squares in rr. Moreover, the 2×22\times 2 minors of AA satisfy

(r+1)A14A23+r(r+1)A13A24=rA12A34,(r+1)A_{14}A_{23}+r(r+1)A_{13}A_{24}=rA_{12}A_{34},

showing the inequalities are tight. \diamond

Example 5.5.

Interestingly, there is not a parametrized family of 3×33\times 3 positive semidefinite matrices for which the second inequality of Theorem 5.1 holds with equality. The image of the set of 3×33\times 3 positive semidefinite matrices under the map A(A1A23,A2A13,A3A12)A\mapsto(A_{1}A_{23},A_{2}A_{13},A_{3}A_{12}) is not closed. We see that the equality is only attained in the limit.

Fix rr\in\mathbb{R} and consider the following parameterized collection of 3×33\times 3 matrices:

A=(1+ε1λ(r+1)21λ1λ(r+1)21+ε1λr21λ1λr21+ε).A=\begin{pmatrix}1+\varepsilon&\sqrt{1-\lambda(r+1)^{2}}&\sqrt{1-\lambda}\\ \sqrt{1-\lambda(r+1)^{2}}&1+\varepsilon&\sqrt{1-\lambda r^{2}}\\ \sqrt{1-\lambda}&\sqrt{1-\lambda r^{2}}&1+\varepsilon\end{pmatrix}.

For 0<λ<min{1/(r+1)2,1/r2,1}0<\lambda<\min\{1/(r+1)^{2},1/r^{2},1\} and ε>0\varepsilon>0, this matrix has real entries and positive 1×11\times 1 and 2×22\times 2 principal minors. Moreover, for fixed ε>0\varepsilon>0, the limit of det(A)\det(A) as λ0\lambda\to 0 is ε2(3+ε)\varepsilon^{2}(3+\varepsilon). Thus for sufficiently small λ>0\lambda>0, the matrix AA is positive definite. One can compute that

(r+1)A1A23+r(r+1)A2A13rA3A12=(1+r+r2)ε(1+ε)(2+ε).(r+1)\,A_{1}A_{23}+r(r+1)\,A_{2}A_{13}-r\,A_{3}A_{12}=(1+r+r^{2})\varepsilon(1+\varepsilon)(2+\varepsilon).

In particular, as ε0\varepsilon\to 0, the limit is zero and the desired inequality becomes tight. \diamond

Remark 5.6.

One can check that the linear inequalities in Theorem 5.1 cut out a quadratic cone. More precisely, a point (x,y,z)3(x,y,z)\in\mathbb{R}^{3} satisfies the inequality (r+1)x+r(r+1)yrz(r+1)x+r(r+1)y\geq rz for all rr\in\mathbb{R} if and only if 2(xy+xz+yz)x2+y2+z22(xy+xz+yz)\geq x^{2}+y^{2}+z^{2} and x+y+z0x+y+z\geq 0.

6. Proof of the technical lemma

This section is dedicated to the proof of Lemma 3.4, which was used in the proof of Theorem 3.2. Throughout this section we take B𝒦n×nB\in\mathcal{K}^{n\times n} to be an upper triangular matrix whose upper justified minors det(B({1,,|S|},S))\det(B(\{1,\ldots,|S|\},S)) are all nonzero. We consider the function

(8) w(T,S)=ν(det(B(T,S)))w(T,S)=\nu(\det(B(T,S)))

which is defined on pairs (T,S)(T,S) of subsets of the same size and takes values in {}\mathbb{R}\cup\{-\infty\}.

Before the proof, we introduce some notation. We define the following partial order on ([n]k)\binom{[n]}{k}. Given S,T([n]k)S,T\in\binom{[n]}{k}, we say that

ST if |{sS:stj}|j for all j=1,,kS\preceq T\quad\text{ if }\quad\left|\{s\in S:s\leq t_{j}\}\right|\geq j\text{ for all }j=1,\ldots,k

where t1<t2<<tkt_{1}<t_{2}<\ldots<t_{k} are the elements of TT. In particular, since BB is upper triangular, det(B(T,S))=0\det(B(T,S))=0 whenever STS\prec T. We call I[n]I\subset[n] an interval if it is a collection of consecutive numbers. That is, whenever i,kIi,k\in I and i<j<ki<j<k we have jIj\in I. We use [a][a] denote the interval {1,,a}\{1,\ldots,a\}.

Proposition 6.1.

The function ww in (8) satisfies the following:

  • (i)

    w({1,..,|S|},S)w(\{1,..,|S|\},S)\neq-\infty for all SS.

  • (ii)

    w(T,S)=w(T,S)=-\infty whenever STS\prec T.

  • (iii)

    for any interval II with max(I)min(TS)\max(I)\leq\min(T\cup S), w(IT,IS)=iIw(i,i)+w(T,S)w(I\cup T,I\cup S)=\sum_{i\in I}w(i,i)+w(T,S).

  • (iv)

    if max(ST)t\max(S\cup T)\leq t and max(ST)s\max(S\cup T)\leq s, then w(Tt,Ss)=w(T,S)+w(t,s)w(Tt,Ss)=w(T,S)+w(t,s).

  • (v)

    for any SS, the function Tw(T,S)T\mapsto w(T,S) satisfies the 3-term tropical Plücker relations.

Proof.

Item (i) holds by assumption. Whenever STS\prec T, det(B(T,S))=0\det(B(T,S))=0 and w(T,S)=w(T,S)=-\infty, since BB is upper triangular, showing (ii). For (iii), the conditions on I,S,TI,S,T imply that det(B(IT,IS))=(iIbii)det(B(T,S))).\det(B(I\cup T,I\cup S))=\left(\prod_{i\in I}b_{ii}\right)\cdot\det(B(T,S))). Taking valuations of both sides proves the claim. Under the assumptions of (iv), B(t,u)=0B(t,u)=0 for all uSu\in S. If tst\leq s, then the determinant of B(Tt,Ss)B(Tt,Ss) factors as product of det(B(T,S))\det(B(T,S)) and B(t,s)B(t,s). Otherwise, both det(B(Tt,Ss))\det(B(Tt,Ss)) and B(t,s)B(t,s) will be zero. Finally, for any fixed SS, the values w(T,S)w(T,S) are the maximal minors of an n×|S|n\times|S| matrix, which therefore satisfy the three-term Plücker relations. ∎

To prove Lemma 3.4, it therefore suffices to show the following.

Theorem 6.2.

Let ww be any function defined on pairs (T,S)(T,S) of subsets of [n][n] of the same size and taking values in {}\mathbb{R}\cup\{-\infty\}. If ww satisfies the properties (i)-(v) in Proposition 6.1 and the function F(S)=w({1,..,|S|},S)F(S)=w(\{1,..,|S|\},S) is submodular, then

w({1,,|S|},S)w(T,S)w(\{1,\ldots,|S|\},S)\geq w(T,S)

for all S,T[n]S,T\subseteq[n] with |S|=|T||S|=|T|.

For the remainder of the section, we assume ww is an arbitrary function satisfying these properties and for which F(S)=w({1,..,|S|},S)F(S)=w(\{1,..,|S|\},S) is submodular. Our goal is to prove this theorem. We first prove it in the case |S|=1|S|=1.

Lemma 6.3.

w(j,k)w(j+1,k)w(j,k)\geq w(j+1,k) for all 1j,kn1\leq j,k\leq n.

Proof.

If j+1>kj+1>k, then w(j+1,k)=w(j+1,k)=-\infty by property (ii), and the inequality holds trivially. So we can assume that j+1kj+1\leq k. Let II denote the interval {1,,j1}\{1,\ldots,j-1\}. By submodularity,

w(Ij,Ij)+w(Ij,Ik)\displaystyle w(Ij,Ij)+w(Ij,Ik) =F(Ij)+F(Ik)\displaystyle=F(Ij)+F(Ik)
F(I)+F(I{j,k})\displaystyle\geq F(I)+F(I\cup\{j,k\})
=w(I,I)+w(I{j,j+1},I{j,k}).\displaystyle=w(I,I)+w(I\cup\{j,j+1\},I\cup\{j,k\}).

Using property (iv), this gives that

iIw(i,i)+w(j,j)+iIw(i,i)+w(j,k)iIw(i,i)+iIw(i,i)+w(j,j)+w(j+1,k).\sum_{i\in I}w(i,i)+w(j,j)+\sum_{i\in I}w(i,i)+w(j,k)\geq\sum_{i\in I}w(i,i)+\sum_{i\in I}w(i,i)+w(j,j)+w(j+1,k).

Canceling out diagonal terms gives w(j,k)w(j+1,k)w(j,k)\geq\ w(j+1,k). ∎

For an interval II, we use 1+I1+I to denote the shifted interval {1+i:iI}\{1+i:i\in I\}.

Lemma 6.4.

Let II be an interval and subset S[n]S\subseteq[n] of the same size with ISI\preceq S. Then

w(I,S)w(1+I,S)w(I\max(I),S\max(S))w((1+I)\max(1+I),S\max(S))0.w(I,S)-w(1+I,S)\geq w(I\backslash\!\max(I),S\backslash\!\max(S))-w((1+I)\backslash\!\max(1+I),S\backslash\!\max(S))\geq 0.
Proof.

Let s=max(S)s=\max(S), i=min(I)i=\min(I), and j=max(I)j=\max(I). Then [i1]I=[j][i-1]\cup I=[j].

First we assume that i<min(S)i<\min(S), then [i]S=ϕ[i]\cap S=\phi. For the left inequality, we use the submodularity inequality

F([i1]S)+F([i]S\s)F([i1]S\s)+F([i]S).F([i-1]\cup S)+F([i]\cup S\backslash s)\geq F([i-1]\cup S\backslash s)+F([i]\cup S).

Using property (iv) and canceling our diagonal terms k=1i1w(k,k)+k=1iw(k,k)\sum_{k=1}^{i-1}w(k,k)+\sum_{k=1}^{i}w(k,k), this inequality becomes

w(I,S)+w((I+1)\(j+1)),S\s)w(I\j,S\s)+w(1+I,S).w(I,S)+w((I+1)\backslash(j+1)),S\backslash s)\geq w(I\backslash j,S\backslash s)+w(1+I,S).

The claim follows after rearranging terms.

If min(S)=i\min(S)=i, then let JJ be an interval such that min(J)=i\min(J)=i and J=ISJ=I\cap S. Let I=I\JI^{\prime}=I\backslash J and S=S\JS^{\prime}=S\backslash J. Then, by the previous part, the claim holds for II^{\prime} and SS^{\prime}, that is

w(I,S)w(1+I,S)w(I\max(I),S\max(S))w((1+I)\max(1+I),S\max(S)).w(I^{\prime},S^{\prime})-w(1+I^{\prime},S^{\prime})\geq w(I^{\prime}\backslash\!\max(I^{\prime}),S\backslash\!\max(S))-w((1+I^{\prime})\backslash\!\max(1+I^{\prime}),S^{\prime}\backslash\!\max(S^{\prime})).

By (iii), w(I,S)=k=imax(J)w(k,k)+w(I,S)w(I,S)=\sum_{k=i}^{\max(J)}w(k,k)+w(I^{\prime},S^{\prime}) and by (iv) we have w(1+I,S)=k=imax(J)w(k+1,k)+w(1+I,S)w(1+I,S)=\sum_{k=i}^{\max(J)}w(k+1,k)+w(1+I^{\prime},S^{\prime}). Thus, adding k=imax(J)w(k,k)w(k+1,k)\sum_{k=i}^{\max(J)}w(k,k)-w(k+1,k) to both sides of the above inequality proves the result.

The right inequality holds by induction on |S||S|. The base case |I\max(I)|=1|I\backslash\max(I)|=1 is covered by Lemma 6.3. The inductive step is given by the left inequality. ∎

Lemma 6.5.

Let II be an interval and a subset S[n]S\subseteq[n] of the same size. For any t>max(I)t>\max(I) such that (It\max(I))S(It\backslash\!\max(I))\preceq S,

w(I,S)w(It\max(I),S)).w(I,S)\geq w(It\backslash\!\max(I),S)).
Proof.

Let s=max(S)s=\max(S), s=max(S\s)s^{\prime}=\max(S\backslash\!s), i=min(I)i=\min(I), and j=max(I)j=\max(I).

(t=st=s) By submodularity of FF, we have

F([i1]S)+F([s])F([i1]S\s)+F([s]s).F([i-1]\cup S)+F([s^{\prime}])\geq F([i-1]\cup S\backslash s)+F([s^{\prime}]s).

Cancelling diagonal terms k=1i1w(k,k)+k=1sw(k,k)\sum_{k=1}^{i-1}w(k,k)+\sum_{k=1}^{s^{\prime}}w(k,k) on both sides gives

w(I,S)w(I\j,S\s)+w(s+1,s).w(I,S)\geq w(I\backslash j,S\backslash s)+w(s^{\prime}+1,s).

By Lemma 6.3, w(s+1,s)w(s,s)w(s^{\prime}+1,s)\geq w(s,s). Then, using property (iv), we find

w(I,S)w(I\j,S\s)+w(s,s)=w((Is\j),S).w(I,S)\geq w(I\backslash j,S\backslash s)+w(s,s)=w((Is\backslash j),S).

(t<st<s) We induct on |S||S| and then tmin(I)t-\min(I). The case |S|=1|S|=1 is Lemma 6.3. If tmin(I)=1t-\min(I)=1, then the inequality holds also by Lemma 6.3. We therefore assume |S|2|S|\geq 2 and tmin(I)>1t-\min(I)>1.

By property (v), the following maximum is attained at least twice

max{w(Iab\{i,j},S))+w(Ics\{i,j},S)},\max\left\{w(Iab\backslash\{i,j\},S))+w(Ics\backslash\{i,j\},S)\right\},

where the max is taken over all assignments {a,b,c}={i,j,t}\{a,b,c\}=\{i,j,t\}. We break the argument into two cases, depending on which terms attain this maximum.

(Case 1) If the term with {a,b}={i,j}\{a,b\}=\{i,j\} attains the maximum then

w(I,S)+w(Its\{i,j},S)w(It\j,S)+w(Is\i,S).w(I,S)+w(Its\backslash\{i,j\},S)\geq w(It\backslash j,S)+w(Is\backslash i,S).

Rearranging terms gives

w(I,S))w(It\j,S)\displaystyle w(I,S))-w(It\backslash j,S) w(Is\i,S)w(Its\{i,j},S)\displaystyle\geq w(Is\backslash i,S)-w(Its\backslash\{i,j\},S)
=w(I\i),S\s)w(It\{i,j},S\s)0.\displaystyle=w(I\backslash i),S\backslash s)-w(It\backslash\{i,j\},S\backslash s)\geq 0.

The last inequality follows by induction on |S||S|.

(Case 2) If the term with {a,b}={i,j}\{a,b\}=\{i,j\} does not attain the maximum, then the other two terms are equal. That is,

w(It\j,S)+w(Is\i,S)=w(Is\j,S)+w(It\i,S).w(It\backslash j,S)+w(Is\backslash i,S)=w(Is\backslash j,S)+w(It\backslash i,S).

Rearranging terms and assuming that i<min(S)i<\min(S) gives

w(It\j,S)w(It\i,S)\displaystyle w(It\backslash j,S)-w(It\backslash i,S) =w(Is\j,S)w(Is\i,S)\displaystyle=w(Is\backslash j,S)-w(Is\backslash i,S)
=w(I\j,S\s)w(I\i,S\s)\displaystyle=w(I\backslash j,S\backslash s)-w(I\backslash i,S\backslash s)
w(I,S)w(1+I,S)\displaystyle\leq w(I,S)-w(1+I,S)
w(I,S)w(((1+I)t\(j+1),S)\displaystyle\leq w(I,S)-w(((1+I)t\backslash(j+1),S)
=w(I,S)w(It\i,S).\displaystyle=w(I,S)-w(It\backslash i,S).

The first inequality follows from Lemma 6.4 and the fact that (1+I)\max(1+I)=I\i(1+I)\backslash\!\max(1+I)=I\backslash i, and the last inequality follows by induction on tmin(I)t-\min(I), since tmin(1+I)<tmin(I)t-\min(1+I)<t-\min(I). We have 1+IS1+I\preceq S since i<min(S)i<\min(S). Simplifying the final inequality gives

w(It\j,S)w(I,S),w(It\backslash j,S)\leq w(I,S),

as desired.

Now if i=min(S)i=\min(S), then let JJ be an interval such that min(J)=i\min(J)=i and J=ISJ=I\cap S. Let I=I\JI^{\prime}=I\backslash J and S=S\JS^{\prime}=S\backslash J. Then, by the previous part, the claim holds for II^{\prime} and SS^{\prime}, that is

w(It\j,S)w(I,S).w(I^{\prime}t\backslash j,S^{\prime})\leq w(I^{\prime},S^{\prime}).

By (iii), w(I,S)=k=imax(J)w(k,k)+w(I,S)w(I,S)=\sum_{k=i}^{\max(J)}w(k,k)+w(I^{\prime},S^{\prime}). Thus, adding k=imax(J)w(k,k)\sum_{k=i}^{\max(J)}w(k,k) to both sides of the above inequality finishes the proof. ∎

Lemma 6.6.

Let S,T[n]S,T\subseteq[n] with TST\preceq S. Let II be the interval with min(I)=min(T)\min(I)=\min(T) and |I|=|S|=|T||I|=|S|=|T|. Then w(I,S)w(T,S)w(I,S)\geq w(T,S).

Proof.

First we assume that max(S)max(T)\max(S)\neq\max(T) and we proceed by induction on |T||T|. For |T|=1|T|=1, I=T={min(T)}I=T=\{\min(T)\} so the statement holds trivially. Therefore we can assume |S|=|T|2|S|=|T|\geq 2.

Consider the first consecutive run of elements of TT. That is, let JJ be the largest interval with min(J)=min(T)\min(J)=\min(T) and JTJ\subseteq T. We also induct on |T||J||T|-|J|. If |J|=|T||J|=|T|, then I=J=TI=J=T and the inequality holds trivially. The case |T||J|=1|T|-|J|=1 follows from Lemma 6.5. Therefore we may suppose |T||J|2|T|-|J|\geq 2. Let j=max(J)j=\max(J). We will show by induction on |T||T| that

(9) w(T,S)maxtT\Jw((T(j+1)\t,S).w(T,S)\leq\max_{t\in T\backslash J}w((T(j+1)\backslash t,S).

Since (T(j+1)\t)(T(j+1)\backslash t) has a longer initial consecutive sequence than TT, it follows by induction on |T||J||T|-|J| that w((T(j+1)\t,S)w(I,S)w((T(j+1)\backslash t,S)\leq w(I,S), implying w(T,S)w(I,S)w(T,S)\leq w(I,S), as desired.

Let s=max(S)s=\max(S). Suppose that U={x,y}U=\{x,y\} attains the maximum

maxUw(T(j+1)\U,S\s)\max_{U}w(T(j+1)\backslash U,S\backslash s)

taken over subsets UT\JU\subseteq T\backslash J with |U|=2|U|=2. By property (v), the maximum of

max{a,b,c}={j+1,x,y}w((Tab\{x,y}),S)+w((Tcs\{x,y}),S)\max_{\{a,b,c\}=\{j+1,x,y\}}w((Tab\backslash\{x,y\}),S)+w((Tcs\backslash\{x,y\}),S)

is attained at least twice. In particular, after possibly relabeling x,yx,y, we have

w(T,S)+w((Ts(j+1)\{x,y}),S)w((T(j+1)\y),S)+w(Ts\x,S).w(T,S)+w((Ts(j+1)\backslash\{x,y\}),S)\leq w((T(j+1)\backslash y),S)+w(Ts\backslash x,S).

By property (iii), we can cancel the diagonal terms w(s,s)w(s,s) and rearrange to get

w((T(j+1)\{x,y}),S\s)w(T\x,S\s)w((T(j+1)\y),S)w(T,S).w((T(j+1)\backslash\{x,y\}),S\backslash s)-w(T\backslash x,S\backslash s)\leq w((T(j+1)\backslash y),S)-w(T,S).

By induction on |T||T|, we have that

w(T\x,S\s)maxtw((T(j+1)\{x,t}),S\s)w(T\backslash x,S\backslash s)\leq\max_{t}w((T(j+1)\backslash\{x,t\}),S\backslash s)

where tt runs over (T\J)\x(T\backslash J)\backslash x. By assumption on {x,y}\{x,y\}, this maximum is attained by t=yt=y, giving that

w(T\x,S\s)w(T(j+1)\{x,y},S\s).w(T\backslash x,S\backslash s)\leq w(T(j+1)\backslash\{x,y\},S\backslash s).

Together with the inequality from above, we get that

0w(T(j+1)\{x,y},S\s)w(T\x,S\s))w(T(j+1)\y,S))w(T,S),0\leq w(T(j+1)\backslash\{x,y\},S\backslash s)-w(T\backslash x,S\backslash s))\leq w(T(j+1)\backslash y,S))-w(T,S),

showing that

w(T,S)w(T(j+1)\y,S))maxtT\Jw(T(j+1)\t,S).w(T,S)\leq w(T(j+1)\backslash y,S))\leq\max_{t\in T\backslash J}w(T(j+1)\backslash t,S).

If max(S)=max(T)=s\max(S)=\max(T)=s, let JJ be the interval such that max(J)=max(S)=max(T)\max(J)=\max(S)=\max(T), JTSJ\subset T\cap S and it is of maximum length. We induct on the size of JJ. If J=1J=1, that is if max(S\s)max(T\s)\max(S\backslash s)\neq\max(T\backslash s), then let I=I\max(I)I^{\prime}=I\backslash\!\max(I), S=S\sS^{\prime}=S\backslash s, and T=T\sT^{\prime}=T\backslash s. Then by the previous part w(I,S)w(T,S)w(I^{\prime},S^{\prime})\geq w(T^{\prime},S^{\prime}). Using Lemma 6.5, we get

w(I,S)\displaystyle w(I,S) w(Is\max(I),S)\displaystyle\geq w(Is\backslash\!\max(I),S)
=w(Is,S)\displaystyle=w(I^{\prime}\!s,S)
=w(I,S)+w(s,s)\displaystyle=w(I^{\prime},S^{\prime})+w(s,s)
w(T,S)+w(s,s)\displaystyle\geq w(T^{\prime},S^{\prime})+w(s,s)
=w(T,S).\displaystyle=w(T,S).

For the inductive step, we use similar argument as the one above and this finishes the proof. ∎

Now we are ready to complete the proof of this section.

Proof of Theorem 6.2.

Let S,T[n]S,T\subseteq[n]. By Lemma 6.6, w(T,S)w(I,S)w(T,S)\leq w(I,S), where II is the interval with min(I)=min(T)\min(I)=\min(T) and |I|=|T||I|=|T|. By Lemma 6.4, w(1+J,S)w(J,S)w(1+J,S)\leq w(J,S) for every interval JJ. Inducting then gives w(k+J,S)w(J,S)w(k+J,S)\leq w(J,S) for arbitrary k>0k\in\mathbb{Z}_{>0}. In particular, for J={1,,|S|}J=\{1,\ldots,|S|\} and k=min(T)1k=\min(T)-1, this gives

w(T,S)w(I,S)w({1,,|S|},S),w(T,S)\leq w(I,S)\leq w(\{1,\ldots,|S|\},S),

as desired. ∎

Remark 6.7.

One might hope for inequalities of the form w(T,S)w(T,S)w(T,S)\geq w(T^{\prime},S) when TTT\succeq T^{\prime}, but these do not hold in general. For example, consider the matrix

B=(111101230011+t0001).B=\begin{pmatrix}1&1&1&1\\ 0&1&2&3\\ 0&0&1&1+t\\ 0&0&0&1\\ \end{pmatrix}.

This matrix is upper triangular and all of its upper justified minors are nonzero and have valuation 0. Therefore the function FF is identically 0 and thus submodular. However w({1,3},{3,4}))=1w(\{1,3\},\{3,4\}))=-1, whereas w({2,3},{3,4})=0w(\{2,3\},\{3,4\})=0.

Example 6.8 (n=8,k=4n=8,k=4).

To illustrate the strategy of the proof above, consider S={5,6,7,8}S=\{5,6,7,8\}, and T={1,4,5,7}T=\{1,4,5,7\}. Here J={1}J=\{1\}. Inequality (9) states that

w(1457,S)max{w(1257,S),w(1247,S),w(1245,S)}.w(1457,S)\leq\max\{w(1257,S),w(1247,S),w(1245,S)\}.

Let us show this inequality in the case that w(127,567)w(127,567) achieves the maximum among {w(124,567),w(125,567),w(127,567)}\{w(124,567),w(125,567),w(127,567)\}. In the proof above, this corresponds to {x,y}={4,5}\{x,y\}=\{4,5\}. By the tropical Plücker relations, the maximum

max{a,b,c}={2,4,5}{w(17ab,S)+w(178c,S)} is attained at least twice,\max_{\{a,b,c\}=\{2,4,5\}}\{w(17ab,S)+w(178c,S)\}\text{ is attained at least twice,}

from which we can conclude that

w(1457,S)+w(1278,S)w(127a,S)+w(178b,S)w(1457,S)+w(1278,S)\leq w(127a,S)+w(178b,S)

for some assignment of {a,b}={4,5}\{a,b\}=\{4,5\}. Then w(xyz8,S)=w(xyz,S)+w(8,8)w(xyz8,S)=w(xyz,S^{\prime})+w(8,8) for any subset {x,y,z}[7]\{x,y,z\}\subset[7] where S={5,6,7}S^{\prime}=\{5,6,7\}. The inequality above give

w(127a,S)w(1457,S)\displaystyle w(127a,S)-w(1457,S) w(1278,S)w(178b,S)\displaystyle\geq w(1278,S)-w(178b,S)
=w(127,S)w(17b,S).\displaystyle=w(127,S^{\prime})-w(17b,S^{\prime}).

By induction, w(17b,S)w(17b,S^{\prime}) is bounded above by max{w(127,S),w(12b,S)}\max\{w(127,S^{\prime}),w(12b,S^{\prime})\}, which is bounded above by w(127,S)w(127,S^{\prime}), by assumption. Therefore this difference is nonnegative, giving

w(1457,S)w(127a,S)maxx,y{4,5,7}w(12xy,S).w(1457,S)\leq w(127a,S)\leq\max_{x,y\in\{4,5,7\}}w(12xy,S).

From this, we can continue by induction. Using (9) again gives

w(1257,S)\displaystyle w(1257,S) max{w(1237,S),w(1235,S)},\displaystyle\leq\max\{w(1237,S),w(1235,S)\},
w(1247,S)\displaystyle w(1247,S) max{w(1237,S),w(1234,S)},\displaystyle\leq\max\{w(1237,S),w(1234,S)\},
w(1245,S))\displaystyle w(1245,S)) max{w(1235,S),w(1234,S)}.\displaystyle\leq\max\{w(1235,S),w(1234,S)\}.

Finally, by Lemma 6.5, we see that these are all bounded above by w(1234,S)w(1234,S). That is, for x,y,z4x,y,z\geq 4

maxx,y,z4w(1xyz,S)maxx,y4w(12xy,S)maxx4w(123y,S)w(1234,S).\max_{x,y,z\geq 4}w(1xyz,S)\leq\max_{x,y\geq 4}w(12xy,S)\leq\max_{x\geq 4}w(123y,S)\leq w(1234,S).

\diamond

References

  • [AAV24a] Abeer Al Ahmadieh and Cynthia Vinzant. Characterizing principal minors of symmetric matrices via determinantal multiaffine polynomials. Journal of Algebra, 638:255–278, 2024.
  • [AAV24b] Abeer Al Ahmadieh and Cynthia Vinzant. Determinantal representations and the image of the principal minor map. International Mathematics Research Notices, 2024(10):8930–8958, 2024.
  • [AGS20] Xavier Allamigeon, Stéphane Gaubert, and Mateusz Skomra. Tropical spectrahedra. Discrete & Computational Geometry, 63:507–548, 2020.
  • [AGV18] Nima Anari, Shayan Oveis Gharan, and Cynthia Vinzant. Log-concave polynomials, entropy, and a deterministic approximation algorithm for counting bases of matroids. In 2018 IEEE 59th Annual Symposium on Foundations of Computer Science (FOCS), pages 35–46. IEEE, 2018.
  • [Ale13] Daniele Alessandrini. Logarithmic limit sets of real semi-algebraic sets. Advances in Geometry, 13(1):155–190, 2013.
  • [BB06] Julius Borcea and Petter Brändén. Applications of stable polynomials to mixed determinants: Johnson’s conjectures, unimodality, and symmetrized fischer products. Duke Mathematical Journal, 143, 08 2006.
  • [Ber71] George M Bergman. The logarithmic limit-set of an algebraic variety. Transactions of the American Mathematical Society, 157:459–469, 1971.
  • [BEZ21] Madeline Brandt, Christopher Eur, and Leon Zhang. Tropical flag varieties. Advances in Mathematics, 384:107695, 2021.
  • [BH20] Petter Brändén and June Huh. Lorentzian polynomials. Annals of Mathematics, 192(3):821–891, 2020.
  • [Bor22] Jonathan Boretsky. Totally nonnegative tropical flags and the totally nonnegative flag dressian. arXiv preprint arXiv:2208.09128, 2022.
  • [Brä10] Petter Brändén. Discrete concavity and the half-plane property. SIAM Journal on Discrete Mathematics, 24(3):921–933, 2010.
  • [BS23] Phillip Braun and Hristo Sendov. On the hadamard-fischer inequality, the inclusion-exclusion formula, and bipartite graphs. Linear Algebra and its Applications, 668:64–92, 2023.
  • [CFB95] J Orestes Cerdeira, Isabel Faria, and Paulo Bárcia. Establishing determinantal inequalities for positive-definite matrices. Discrete applied mathematics, 63(1):13–24, 1995.
  • [Cho16] Daeshik Choi. Determinantal inequalities of positive definite matrices. Math. Inequal. Appl, 19(1):167–172, 2016.
  • [DWH22] Sheng Dong, Qingwen Wang, and Lei Hou. Determinantal inequalities for block hadamard product and khatri-rao product of positive definite matrices. AIMS Mathematics, 7(6):9648–9655, 2022.
  • [EM22] Yassine El Maazouz. The Gaussian entropy map in valued fields. Algebr. Stat., 13(1):1–18, 2022.
  • [GRSU24] Jeffrey Giansiracusa, Felipe Rincón, Victoria Schleis, and Martin Ulirsch. Log-concavity for independent sets of valuated matroids. arXiv preprint arXiv:2407.05808, 2024.
  • [Gub13] Walter Gubler. A guide to tropicalizations. Algebraic and combinatorial aspects of tropical geometry, 589:125–189, 2013.
  • [Haq12] Mohammad Moinul Haque. Tropical incidence relations, polytopes, and concordant matroids. arXiv preprint arXiv:1211.2841, 2012.
  • [HJ08] H Tracy Hall and Charles R Johnson. Bounded ratios of products of principal minors of positive definite matrices. arXiv preprint arXiv:0806.2645, 2008.
  • [HS07] Olga Holtz and Bernd Sturmfels. Hyperdeterminantal relations among symmetric principal minors. Journal of Algebra, 316(2):634–648, 2007.
  • [JB93] Charles R Johnson and Wayne W Barrett. Determinantal inequalities for positive definite matrices. Discrete Mathematics, 119(1-3):97–106, 1993.
  • [JSY22] Philipp Jell, Claus Scheiderer, and Josephine Yu. Real tropicalization and analytification of semialgebraic sets. International Mathematics Research Notices, 2022(2):928–958, 2022.
  • [JZC19] Xiaoyu Jiang, Yanpeng Zheng, and Xiaoting Chen. Extending a refinement of koteljanskiĭ’s inequality. Linear Algebra and its Applications, 574:252–261, 2019.
  • [Lov83] László Lovász. Submodular functions and convexity. Mathematical Programming The State of the Art: Bonn 1982, pages 235–257, 1983.
  • [LS09] Shaowei Lin and Bernd Sturmfels. Polynomial relations among principal minors of a 4×\times 4-matrix. Journal of Algebra, 322(11):4121–4131, 2009.
  • [LS20] Minghua Lin and Gord Sinnamon. Revisiting a sharpened version of hadamard’s determinant inequality. Linear Algebra and its Applications, 606:192–200, 2020.
  • [MS21] Diane Maclagan and Bernd Sturmfels. Introduction to tropical geometry, volume 161. American Mathematical Society, 2021.
  • [Mur98] Kazuo Murota. Discrete convex analysis. Mathematical Programming, 83:313–371, 1998.
  • [MUWY18] Fatemeh Mohammadi, Caroline Uhler, Charles Wang, and Josephine Yu. Generalized permutohedra from probabilistic graphical models. SIAM J. Discrete Math., 32(1):64–93, 2018.
  • [Oed11] Luke Oeding. Set-theoretic defining equations of the variety of principal minors of symmetric matrices. Algebra & Number Theory, 5(1):75–109, 2011.
  • [RvGP02] Hans Reijnierse, Anita van Gellekom, and Jos AM Potters. Verifying gross substitutability. Economic Theory, 20:767–776, 2002.