Skip to main content
Philosophical transactions. Series A, Mathematical, physical, and engineering sciences logoLink to Philosophical transactions. Series A, Mathematical, physical, and engineering sciences
. 2018 Sep 17;376(2131):20170428. doi: 10.1098/rsta.2017.0428

The U(n) Gelfand–Zeitlin system as a tropical limit of Ginzburg–Weinstein diffeomorphisms

A Alekseev 1,, J Lane 1, Y Li 1
PMCID: PMC6158378  PMID: 30224420

Abstract

In this paper, we show that the Ginzburg–Weinstein diffeomorphism Inline graphic of Alekseev & Meinrenken (Alekseev, Meinrenken 2007 J. Differential Geom. 76, 1–34. (10.4310/jdg/1180135664)) admits a scaling tropical limit on an open dense subset of Inline graphic. The target of the limit map is a product Inline graphic, where Inline graphic is the interior of a cone, T is a torus, and Inline graphic carries an integrable system with natural action-angle coordinates. The pull-back of these coordinates to Inline graphic recovers the Gelfand–Zeitlin integrable system of Guillemin & Sternberg (Guillemin, Sternberg 1983 J. Funct. Anal. 52, 106–128. (10.1016/0022-1236(83)90092-7)). As a by-product of our proof, we show that the Lagrangian tori of the Flaschka–Ratiu integrable system on the set of upper triangular matrices meet the set of totally positive matrices for sufficiently large action coordinates.

This article is part of the theme issue ‘Finite dimensional integrable systems: new trends and methods’.

Keywords: integrable systems, Poisson–Lie groups, tropicalization, Poisson geometry, Gelfand–Zeitlin

1. Introduction

One of the richest settings for the study of integrable systems is the dual vector space Inline graphic of a finite dimensional Lie algebra Inline graphic, equipped with its canonical Lie–Poisson structure. There are many important examples of integrable systems defined on Inline graphic, including spinning tops [1] and Mishchenko–Fomenko systems [2,3]. Systems on Inline graphic also give rise to collective integrable systems via moment maps [4,5], for instance, leading to complete integrability of the geodesic flow on certain homogeneous spaces [6].

Gelfand–Zeitlin systems, defined by Guillemin and Sternberg on the space of Hermitian matrices (interpreted as Inline graphic), are one of the most famous examples of such integrable systems [7]. Unlike Mishchenko–Fomenko systems, Gelfand–Zeitlin systems have natural global action-angle coordinates. This structure has led to results about the symplectic topology of coadjoint orbits [8,9]. Unfortunately, Gelfand–Zeitlin systems have only been defined for Inline graphic of types A, B and D.

Motivated by the problem of generalizing Gelfand–Zeitlin systems, we are naturally brought to the following question: What underlying structures give rise to Gelfand–Zeitlin systems? One answer to this question comes from the study of toric degenerations, which define integrable systems with global action-angle coordinates on coadjoint orbits and many other spaces [9,10]. In this paper, we make progress towards a new answer to this question by relating Gelfand–Zeitlin systems on Inline graphic to Ginzburg–Weinstein diffeomorphisms and upper cluster algebra structures on dual Poisson–Lie groups. We believe that this approach can be generalized and will give examples of new integrable systems on Lie algebra duals that have natural global action-angle coordinates.

In more detail, let K be a compact connected Poisson–Lie group (that is, a Lie group equipped with a multiplicative Poisson bracket). Poisson–Lie theory associates with K a dual Poisson–Lie group K* which is solvable. The Ginzburg–Weinstein theorem says that K* admits global linearization maps: Poisson isomorphisms Inline graphic called Ginzburg–Weinstein diffeomorphisms.

When K = U(n), it is possible to describe explicit Ginzburg–Weinstein diffeomorphisms [11]. The dual Poisson–Lie group U(n)* has a completely integrable system with global action-angle coordinates due to Flaschka–Ratiu [12]. On open dense subsets of Inline graphic and U(n)*, a Poisson isomorphism is given by the identity in global action-angle coordinates for the Gelfand–Zeitlin and Flaschka–Ratiu systems (the main result of [11] is that this isomorphism extends to all of Inline graphic and U(n)*).

Returning to the general case of compact connected K, equip Inline graphic with its Lie–Poisson structure Inline graphic, the dual Poisson–Lie group K* with its Poisson structure πK*, and fix a Ginzburg–Weinstein diffeomorphism Inline graphic. As Inline graphic is linear, for positive t the scaled Ginzburg–Weinstein diffeomorphism gwt(A) = gw(tA) is a Poisson isomorphism with respect to Inline graphic and the scaled Poisson structure K*.

It was recently shown that integrable systems can be constructed from K* by tropicalizing its Poisson structure, which means taking the scaling limit t → ∞ of K* [13,14]. More precisely, denoting n = rank(K) and Inline graphic, there exist t-dependent coordinate charts (coming from an upper cluster algebra structure on a double Bruhat cell)

1.

such that as t → ∞ the Poisson structure (Δt)*(K*) converges to a constant Poisson structure π, of rank m, on Inline graphic, where Inline graphic is the interior of a certain convex polyhedral cone and Tm is a torus. In terms of coordinates ζ on Inline graphic and φ on Tm, the constant Poisson structure π is of the form

1.

where πi,j are some constants (see theorem 3.1 and the following discussion and [13] for more details in the U(n) case. For the general case, see [14]). As π has rank m, the coordinates ζ are global action coordinates for an integrable system on Inline graphic (the coordinates φ differ from global angle coordinates by a linear transformation).

In general, we see that the composition

1. 1.1

is a Poisson isomorphism with respect to Inline graphic and (Δt)*(K*) for all positive finite t (on the open dense subset where the composition is defined). One may then make the following conjecture.

Conjecture 1.1 —

The limit as t → ∞ of the map (1.1) exists on an open dense subset Inline graphic and defines a Poisson isomorphism between Inline graphic, equipped with Inline graphic, and Inline graphic, equipped with π.

If this conjecture is true, then the limit defines a completely integrable system with global action-angle coordinates on an open dense subset of Inline graphic. Our main result is that conjecture 1.1 holds for K = U(n).

Theorem 1.2 —

For the Ginzburg–Weinstein diffeomorphism of [11] and coordinate charts Δt defined by the cluster coordinates used in [13], conjecture 1.1 is true. Moreover, the integrable system on the open dense subset Inline graphic obtained by pulling back action-angle coordinates from Inline graphic is the Gelfand–Zeitlin system (up to a linear transformation).

Theorem 1.2 is illustrated by the following example of U(2) where the map gwt can be made completely explicit. For U(n), n≥3, the map gwt does not admit a tractable closed formula in standard matrix coordinates.

Example 1.3 —

If Inline graphic is identified with Hermitian 2 × 2 matrices, with coordinates

graphic file with name rsta20170428-e4.jpg

and U(2)* is identified with upper triangular matrices, with positive diagonal entries,

graphic file with name rsta20170428-e5.jpg

then the t-scaling of the Ginzburg–Weinstein diffeomorphism of [11] is given by the formula

graphic file with name rsta20170428-e6.jpg

where Inline graphic, z = ρ eiθ, and t > 0. The action coordinates of the Gelfand–Zeitlin system are

graphic file with name rsta20170428-e7.jpg

and they satisfy ‘interlacing inequalities’ λ(2)1λ(1)1λ(2)2. The functions λ(2)1, λ(2)2 are Casimir functions and the function λ(1)1 generates a S1 action on the coadjoint orbits (which are 2-spheres).

In coordinates (ζ(1)1, ζ(2)1, ζ(2)2, φ(2)1) on Inline graphic, the chart Δt is given by

graphic file with name rsta20170428-e8.jpg

The open dense subset of Inline graphic is the set where the interlacing inequalities are strict. For A in this subset, one computes using the interlacing inequalities that

graphic file with name rsta20170428-e9.jpg

while ζ(1)1○gwt = λ(1)1/2 and ζ(2)2○gwt = (λ(2)1 + λ(2)2)/2 for all t > 0. Thus,

graphic file with name rsta20170428-e10.jpg

The organization of this paper is as follows. In §2, we recall the definition of action-angle coordinates for the Gelfand–Zeitlin system on Inline graphic, some background from the theory of Poisson–Lie groups, and the explicit formula for the Ginzburg–Weinstein diffeomorphisms of [11]. Section 3 describes the cluster coordinates on U(n)* that were used in [13] to tropicalize the Poisson bracket on U(n)*, and explains their relation to the Gelfand–Zeitlin functions via the maps gwt. In §4, we introduce matrix factorization coordinates on U(n)* and the tropical estimation results from [15], which are the main ingredients for proving convergence. Finally, §§5 and 6 are dedicated to the proof of theorem 1.2, which is divided into two propositions. Convergence of action coordinates is proved in proposition 5.1, and convergence of angle coordinates is proved in proposition 6.1. Part of the proof of proposition 6.1 involves showing that the Flaschka–Ratiu tori meet the set of totally positive matrices for sufficiently large values of action variables. A list of notation used throughout the paper is provided in table ??.

Table 1.

List of notation.

Inline graphic The set of n × n Hermitian matrices
Inline graphic The set of positive definite n × n Hermitian matrices
Inline graphic The subset of Inline graphic where all interlacing inequalities are strict
λ(k)i Gelfand–Zeitlin function on Inline graphic
ψ(k)i Angle coordinates for the Gelfand–Zeitlin system on Inline graphic
(k)i The sum λ(k)1 +  · s + λ(k)i
L The Gelfand–Zeitlin map with coordinates ℓ(k)i
ΔI,J Minor consisting of rows and columns I, J⊆{1, …, n}
Δ(k)i Minor ΔI,J with I = {n − k + 1, …, n − k + i} and J = {n − i, …, n}
ζ(k)i, φ(k)i Defined by Inline graphic
Δt The t-dependent coordinate chart defined by ζ(k)i, φ(k)i
m(k)i Tropical Gelfand–Zeitlin functions obtained by tropicalizing the polynomials (4.1)
Inline graphic The tropical Gelfand–Zeitlin map with coordinates m(k)i
Inline graphic The t-dependent Gelfand–Zeitlin map on AN
Inline graphic The interior of the cone defined by rhombus inequalities (2.4)
Inline graphic The subset of Inline graphic defined by inequalities (3.1)

2. Gelfand–Zeitlin and Ginzburg–Weinstein

In the first part of this section, we recall the definition of the classical Gelfand–Zeitlin system, along with several details about action-angle coordinates and the Gelfand–Zeitlin cone. In the second part of this section, we briefly recall the theory of Poisson–Lie groups, the Ginzburg–Weinstein theorem, and the explicit Ginzburg–Weinstein diffeomorphism of [11].

(a). The classical Gelfand–Zeitlin system

For any Lie group K with Lie algebra Inline graphic, the dual vector space Inline graphic is endowed with a linear Poisson bracket called the Lie–Poisson structure, defined by the formula

(a).

for Inline graphic and smooth functions Inline graphic.

Let K = U(n), the group of unitary n × n matrices, and let Inline graphic denote the set of Hermitian n × n matrices. We can identify Inline graphic via the non-degenerate bilinear form (X, Y ) = tr(XY ). With this identification, the Gelfand–Zeitlin functions on Inline graphic are the functions Inline graphic, 1 ≤ i ≤ k ≤ n, defined so that for Inline graphic,

(a).

are the ordered eigenvalues of the k × k principal submatrix A(k) in the bottom-right corner of A (figure 1).

Figure 1.

Figure 1.

Bottom-right principal submatrices of A.

The Gelfand–Zeitlin functions satisfy ‘interlacing inequalities’,

(a). 2.1

and the image of the Gelfand–Zeitlin map Inline graphic, defined by the Gelfand–Zeitlin functions, is the polyhedral cone defined by inequalities (2.1), called the Gelfand–Zeitlin cone.

Let Inline graphic denote the open dense subset of Inline graphic where all inequalities (2.1) are strict (this will be the set Inline graphic in theorem 1.2, cf. conjecture 1.1). The Gelfand–Zeitlin functions are smooth on Inline graphic and define global action coordinates for a completely integrable system: the functions λ(n)1, …, λ(n)n are a complete set of Casimir functions, and the functions λ(k)i, 1 ≤ i ≤ k < n, generate

(a).

commuting Hamiltonian S1-actions, whose orbits coincide with the joint level-sets of the Gelfand–Zeitlin functions [7].

Angle coordinates on Inline graphic corresponding to the global action coordinates λ(k)i are defined by choosing a Lagrangian section σ of the Gelfand–Zeitlin map and defining ψ(k)i(p) = 0 for all p∈Im(σ). The Gelfand–Zeitlin functions are invariant under the transpose map AAT, which is an anti-Poisson involution of Inline graphic. The fixed point set of the transpose map is the set Sym(n) of (real) symmetric n × n matrices and the intersection Inline graphic is a union of Lagrangian submanifolds of Inline graphic that are images of sections of the Gelfand–Zeitlin map. Thus, we may fix global angle coordinates for the Gelfand–Zeitlin system by choosing a connected component of Sym0(n). In these coordinates, the bivector of the Lie–Poisson bracket on Inline graphic has the form

(a). 2.2

In this paper, it will be convenient to introduce the following notation. For all 1 ≤ i ≤ k ≤ n, let

(a). 2.3

The functions ℓ(k)i satisfy a list of inequalities equivalent to (2.1),

(a). 2.4

with the convention that ℓ(k)0 = 0 (these inequalities can be visualized as rhombi in a triangular tableau, see [16, figs. 4 and 5]). We denote the interior of the cone defined by (2.4) by Inline graphic. The set Inline graphic is the image under a linear transformation of the interior of the Gelfand–Zeitlin cone. In coordinates ℓ(k)i, ψ(k)i, the Poisson bivector (2.2) is no longer diagonal; it has a ‘lower-triangular’ form,

(a). 2.5

where the coefficients c(k)i,j are determined by (2.2) and (2.3).

(b). Poisson–Lie groups and Ginzberg–Weinstein diffeomorphisms

First introduced by Drinfel'd [17] and Semenov-Tian-Shansky [18], a Poisson–Lie group is a Lie group G equipped with a Poisson bivector π such that group multiplication is a Poisson map. The linearization of π at identity of G is a 1-cocycle on Inline graphic with respect to the adjoint representation. This cocycle defines a Lie bracket on Inline graphic and endows the pair Inline graphic with the structure of a Lie bialgebra. The dual Poisson–Lie group of G is the connected, simply connected Poisson–Lie group whose Lie bialgebra is Inline graphic. More details can be found in [19,20].

Let K be a connected compact Lie group. There are two natural Poisson–Lie group structures on K, which in turn define two dual Poisson–Lie groups. We now explain these structures in more detail.

First, let Inline graphic, and fix an Iwasawa decomposition G = KAN, Inline graphic, relative to a choice of maximal torus TK and positive roots. With these choices, Inline graphic and Inline graphic is the direct sum of positive root spaces. Let B( · , · ) be a non-degenerate, K-invariant, bilinear form on Inline graphic, and let Inline graphic be its complexification. The imaginary part of Inline graphic defines a non-degenerate pairing between Inline graphic and Inline graphic which identifies Inline graphic and endows Inline graphic with the Lie algebra structure of Inline graphic. The pair Inline graphic is a Lie bialgebra, which defines a Poisson–Lie group structure on K such that the dual Poisson–Lie group K* is identified with AN.

Second, any Lie group has a trivial Poisson–Lie group structure when equipped with the zero Poisson bivector. The dual Poisson–Lie group of (K, 0) is Inline graphic equipped with the Lie–Poisson structure defined in the previous subsection.

The linearization of the Poisson bivector of K* at the group unit equals the Poisson bivector of Inline graphic. Therefore, K* and Inline graphic are isomorphic as Poisson manifolds in a neighbourhood of their group units by local normal forms for Poisson manifolds [21]. In fact, this isomorphism extends globally.

Theorem 2.1 (Ginzburg–Weinstein Theorem [20]) —

The Poisson manifolds Inline graphic and (K*, πK*) are Poisson isomorphic.

Such Poisson isomorphisms are called Ginzburg–Weinstein isomorphisms/diffeomorphisms. Note that Inline graphic is abelian, whereas K* is not, so Ginzburg–Weinstein diffeomorphisms cannot be group homomorphisms, and Inline graphic and K* cannot be isomorphic as Poisson–Lie groups.

There are several proofs of the Ginzburg–Weinstein theorem in the literature: the original proof [20] is an existence proof using a cohomology calculation, the proof in [1] gives Ginzburg–Weinstein diffeomorphisms as flows of certain Moser vector fields, the proof in [22] is by integration of a nonlinear PDE of a classical dynamical r-matrix, and the proof in [23] uses the Stokes data of an ODE on a disc with an irregular singular point in the centre.

For the remainder of this section, fix K = U(n) and take the Iwasawa decomposition GLn = KAN, where A is the set of diagonal matrices with positive real entries and N is the set of upper triangular unipotent matrices. Let Inline graphic denote the set of positive definite n × n Hermitian matrices. The map

(b). 2.6

is a diffeomorphism (with inverse given by Gaussian decomposition). It was observed by Flaschka & Ratiu [12] that the functions ln(λ(k)i) define a completely integrable system on Inline graphic (equipped with the Poisson structure h*πK*). This system was related to the Gelfand–Zeitlin system on Inline graphic by [11] who proved the following theorem.

Theorem 2.2 ([11]) —

There exists a Poisson isomorphism Inline graphic such that

  • (i) γ intertwines the Gelfand–Zeitlin functions

    graphic file with name rsta20170428-e20.jpg 2.7
  • (ii) γ intertwines the Gelfand–Zeitlin torus actions on Inline graphic and Inline graphic.

  • (iii) For any connected component Inline graphic, Inline graphic.

  • (iv) γ is equivariant with respect to the conjugation action of T⊆U(n).

  • (v) γ(A + uI) = euγ(A).

  • (vi) Inline graphic.

Remark 2.3 —

The map γ is uniquely determined by properties (i)–(iii). The map h−1γ is a Ginzburg–Weinstein diffeomorphism. One should note that [11] use a slightly different h, but the composition h−1γ remains Poisson. See remark 3.3.

3. Cluster coordinates on dual Poisson–Lie groups

Let K = U(n) and K* = AN as in the previous section. We define coordinates on an open dense subset of AN following [13]: for all 1 ≤ i ≤ k ≤ n, let Δ(k)i denote the minor which is the determinant of the solid i × i submatrix formed by intersecting rows n − k + 1 to n − k + i and the last i columns (figure 2).

Figure 2.

Figure 2.

The minors Δ(k)i.

The minor Δ(k)i is a principal minor if and only if i = k, in which case it takes values in Inline graphic, otherwise it takes values in Inline graphic. Let m = n(n − 1)/2; the number of such minors with i < k. Together, the Δ(k)i's define a map

3.

whose restriction to the open dense subset where Δ(k)i≠0 is a coordinate chart (diffeomorphism). On this subset, the equations

3.

define a t-dependent, polar coordinate chart

3.

where Inline graphic, and Inline graphic is equipped with coordinates (ζ, φ) defined by the equations above. The diffeomorphisms Δt define Poisson structures

3.

on Inline graphic.

Let Inline graphic be the interior of the cone defined by inequalities (2.4), replacing ℓ(k)i with ζ(k)i, and for all δ > 0, define Inline graphic as the subset where

3. 3.1

Theorem 3.1 ([13]) —

For t > 0 and Inline graphic,

graphic file with name rsta20170428-e26.jpg 3.2

where C is the number of columns that Δ(k)i and Δ(p)q have in common, R is the number of rows that Δ(k)i and Δ(p)q have in common, and ε(x) = x/|x| with ε(0) = 0. Note that δ is constant in the big-O notation O(e−δt).

This theorem prompts the definition of a Poisson manifold Inline graphic where π is the constant Poisson structure obtained by taking the t → ∞ limit of πt, with coefficients given by (3.2). The coordinates ζ(n)1, …, ζ(n)n are a complete set of Casimir functions for π. By [13, theorem 7], there exists a Poisson isomorphism from Inline graphic to Inline graphic, linear with respect to coordinates ℓ, ψ as in §2, of the form

3. 3.3

where B is the integer matrix defining an automorphism of Tm, uniquely determined so that the map is Poisson. Expanding this formula, the isomorphism is of the form

3. 3.4

where ψ(p)q is higher than ψ(k)i if p < k or p = k and q > i.

Remark 3.2 —

Theorem 3.1 has been generalized in [14]. Given an arbitrary complex semisimple Lie group G and a reduced word for the longest element w0 of the Weyl group, the coordinate algebra Inline graphic of the double Bruhat cell Ge,w0 = BB − w0B − carries an upper cluster algebra structure along with a cluster seed [24]. In the coordinates provided by the cluster seed, the Poisson structure on the dual Poisson–Lie group K* (Inline graphic) admits a ‘partial tropicalization’: a t-deformation such that the t = ∞ limit is an integrable system Inline graphic, where Inline graphic is the interior of an extended string-cone and π is a constant Poisson bracket. Setting G = GLn and taking the standard reduced word for w0, the cluster seed is given by the minors Δ(k)i and one recovers theorem 3.1.

Given a n × n matrix b, and subsets I, J⊆{1, …, n} with |I| = |J|, let ΔI,J(b) denote the determinant of the submatrix with rows I and columns J. For any 1 ≤ i ≤ n, by the Cauchy–Binet formula,

3. 3.5

where λj(bb*) denotes the eigenvalues of bb* (in our notation, these are the Gelfand–Zeitlin functions λ(n)j).

We define a family of Ginzburg–Weinstein diffeomorphisms

3. 3.6

where γ is the Ginzburg–Weinstein diffeomorphism of [11], and h is as defined in §2 (see equation (2.6) and theorem 2.2). For all t > 0,

3.

Denote bt = gwt(A). Combining (3.5) and (2.7), for all 1 ≤ i ≤ n,

3. 3.7

Note that if b is upper triangular, the lower right k × k submatrix (bb*)(k) = b(k)(b(k))*, so equation (3.7) remains valid when the eigenvalues λj are replaced by the Gelfand–Zeitlin functions λ(k)j, and we take I, J⊆{n − k + 1, …, n}.

Remark 3.3 —

If we take Inline graphic instead, then equation (3.7) does not hold for the Gelfand–Zeitlin functions λ(k)j, because in general the bottom right k × k submatrix of Inline graphic is not equal to Inline graphic.

Using the generalized Plücker relations, every minor ΔI,J can be written as a Laurent polynomial in the Δ(k)i's (this is a particular example of the Laurent phenomenon in the theory of cluster algebras, see [25]). We can therefore write each |ΔI,J|2 as a Laurent polynomial PI,J (in variables Δ(k)i and Inline graphic)

3. 3.8

Thus, equations (3.7) can be written in the form

3. 3.9

Example 3.4 —

For n = 3, equation (3.9) corresponding to j = 1, k = 3 is

graphic file with name rsta20170428-e35.jpg

For t large, as λ(k)i satisfies the interlacing inequalities, the dominant term on the left side is e(3)1. We will see in §4 that if Inline graphic, the dominant term on the right side is e2(3)1 = |Δ(3)1|2.

4. Matrix factorizations and tropical Gelfand–Zeitlin

In this section, we rephrase the results of [16], which uses planar networks, in the language of matrix factorizations.

(a). Matrix factorizations

Let Eij denote the matrix with (i, j) entry equal to 1 and all other entries 0. For a complex number z, denote Inline graphic. Any word i = (i1, …, ik) in the alphabet {1, …, n − 1} defines a map

(a).

where N⊆GLn is the subgroup of upper triangular unipotent matrices. We also define

(a).

where A⊆GLn is the subgroup of diagonal matrices with positive real diagonal entries.

Recall that the Weyl group of GLn is Sn generated by simple reflections s1, …, sn−1. The length l(w) of wSn is the smallest integer k such that w can be written as the product of k simple reflections. A word i = (i1, …, ik) is reduced if l(w) = k, where w = si1 · ssik. Let w0 be the longest element of Sn with length l(w0) = m. The standard reduced word for w0 is

(a).

For any reduced word i of w0, the map, or matrix factorization,

(a).

defines a chart, when restricted to Inline graphic. Here, we write (x, z) as shorthand for the tuple (x1, …, xn, z1, …, zm). When i is the standard reduced word i0, we write az0 for simplicity.

Example 4.1 —

For GL3 and the standard reduced word i0 = (1, 2, 1), we have:

graphic file with name rsta20170428-e40.jpg

Remark 4.2 —

More generally, if G is a reductive group with a choice of positive roots, and i = (i1, …, im) is a reduced word for the longest element w0 of the Weyl group, one can define the maps ei(z) using a Chevalley basis, and a chart azi as above. See [26].

Matrix factorization coordinates on AN⊆GLn can be represented by planar networks; weighted planar graphs, oriented left to right, with n ‘source’ vertices on the left, and n ‘sink’ vertices on the right (see [16] for more details). The matrix factorization az0 is represented by the standard planar network, Γ, with n horizontal edges connecting the sources to the sinks (both labelled by 1, …, n as in figure 3), and n(n − 1)/2 non-horizontal edges, arranged as in figure 3 for the case n = 3. The non-horizontal edges (which correspond to ei(z)'s in the matrix factorization) are labelled with the z's and the horizontal edges are labelled at the sources with the x's, as in figure 3. The (i, j) entry of az0(x, z) equals

(a).

where (i, j) is the set of directed paths in Γ from source i to sink j, e is an edge contained in γ, and w(e) is the weight assigned to e (weights not written on Γ are the multiplicative identity).

Figure 3.

Figure 3.

The planar network representing the matrix factorization in example 4.1.

The planar network representation of a matrix factorization gives minors of azi(x, z) a combinatorial interpretation. Recall from the previous section that ΔI,J denotes the minor with rows I and columns J, |I| = |J| = i. By the Lindström Lemma [13,27],

(a).

where (I, J) is the set of i-multipaths from I to J; a union of i disjoint directed paths with sources in I and sinks in J. For the minors Δ(k)i, there is exactly one such multipath (figure 4).

Figure 4.

Figure 4.

Multipaths appearing in the minors of example 4.3. (a) The minor Δ(3)2 and (b) the minor Δ1,2.

Example 4.3 —

Continuing from the previous example,

graphic file with name rsta20170428-e43.jpg

As illustrated by the example,

Theorem 4.4 ([26, theorem 5.8]) —

For any reduced word i, ΔI,J(azi(x, z)) is a polynomial in the coordinates x, z with positive coefficients. Moreover, the minors Δ(k)i(azi(x, z)) are all monomials.

One should note that [26, theorem 5.8] is considerably more general.

(b). Tropical Gelfand–Zeitlin functions

A polynomial Inline graphic is positive if all the coefficients are positive. The tropicalization of a positive polynomial Inline graphic is the piecewise linear function

(b).

If the variables x are all real then, substituting xi = etwi, this equals the limit

(b).

If the variables are complex, then we may still define Inline graphic as above, and—on an open dense set—this equals the function obtained by substituting Inline graphic and evaluating the limit as before (the limit does not equal Inline graphic on the subset where leading terms are cancelled due to their complex arguments).

By theorem 4.4, the minors ΔI,J are positive polynomials in matrix factorization coordinates azi. Thus, for all 1 ≤ i ≤ k ≤ n, the polynomials

(b). 4.1

(which are simply the right sides of equations (3.7)) are positive and can be tropicalized. As the p(k)i(x, z) are related, by equation (3.7), to the logarithmic Gelfand–Zeitlin functions on AN, their tropicalizations are called the tropical Gelfand–Zeitlin functions and denoted m(k)i. One should note that our definition of m(k)i differs from that of m(k)i in [15] by a factor of 2.

Theorem 4.5 ([16, theorem 2]) —

For any choice of matrix factorization coordinates azi, the tropical Gelfand–Zeitlin functions satisfy the rhombus inequalities

graphic file with name rsta20170428-e47.jpg 4.2

Taken together, the tropical Gelfand–Zeitlin functions define a piecewise linear map, called the tropical Gelfand–Zeitlin map,

(b).

Theorem 4.6 ([16, theorem 3]) —

If Inline graphic is defined using the standard reduced word i0, then Inline graphic (the closure of Inline graphic).

As Inline graphic is piecewise linear, Inline graphic decomposes into polyhedral chambers where Inline graphic is linear. There is a unique chamber, which we denote by Inline graphic, where the rank of Inline graphic is n + m, and for this chamber, Inline graphic [16]. For all δ > 0, [15] define Wδ to be the set of all Inline graphic such that,

(b).

and for any two disjoint subsets α, β (the edge set of the standard planar network Γ defined in §3),

(b).

Let azi(tw, ϕ) denote azi(x, z) with the substitutions xi = etwi, Inline graphic. The condition |w(α) − w(β)| > δ guarantees that there is only one leading term in the polynomials p(k)i(azi(tw, ϕ)), which means that the tropicalization m(k)i is equal to the tropical limit described at the beginning of the section. This is the content of the following proposition.

Proposition 4.7 ([16, proposition 2]) —

Fix δ > 0 and let wWδ. For any reduced word i of w0, there is a constant C such that for all t≥1 and any ϕTm,

graphic file with name rsta20170428-e51.jpg

for all 1 ≤ i ≤ k ≤ n.

Let Inline graphic denote the map with coordinates given by

(b).

Proposition 4.8 ([16, proposition 5]) —

For every δ > 0, there exists t0 > 0 such that for all tt0, the following statement holds: for all Inline graphic Inline graphic there exists w ∈ Wδ/2 and ϕTm such that

graphic file with name rsta20170428-e53.jpg

We end this section with one last detail from [16] that is crucial in the proof of proposition 5.1. For all Inline graphic, the maximum in the definition of m(k)i(w) is the sum corresponding to the monomial Δ(k)i(az0(w, ϕ)) (see also Appendix C of [13]). Thus, for all Inline graphic,

(b). 4.3

5. Convergence of action coordinates

In this section we prove,

Proposition 5.1 —

For all δ > 0 and 1 ≤ i ≤ k ≤ n, there exists a constant C and t0≥0 such that tt0 implies

graphic file with name rsta20170428-e55.jpg 5.1

for all Inline graphic.

Proof. —

Let δ > 0, and fix Inline graphic. Denote gwt(A) = bt. By definition of ζ(k)i and theorem 2.2,

graphic file with name rsta20170428-e56.jpg 5.2

for all t > 0.

As Inline graphic, by proposition 4.8, there exists t0 > 0 such that for all tt0, there exists wWδ/2 and ϕTm such that bt = az0(tw, ϕ). As wWδ/2, we can combine equations (5.2), (4.3) and proposition 4.7 to conclude that there exists C≥0 such that

graphic file with name rsta20170428-e57.jpg

which completes the proof. ▪

6. Convergence of angle coordinates

Recall from §2 that a choice of connected component of Inline graphic determines a choice of angle coordinates ψ for the Gelfand–Zeitlin system. Recall also that (up to a linear transformation), the functions φ are angle coordinates on Inline graphic with respect to π. In this section, we prove there exists a choice of angle coordinates ψ for the Gelfand–Zeitlin system so that,

Proposition 6.1 —

On Inline graphic, for all 1 ≤ i < k ≤ n,

graphic file with name rsta20170428-e58.jpg

The sum on the right is a linear combination of the coordinates ψ as in equations (3.3) and (3.4) (in particular, ‘higher terms’ refers to the ordering of angle coordinates defined after equation (3.4)).

In §6a, we show that the Hamiltonian vector fields of the Flaschka–Ratiu system on AN converge as t → ∞ to the Hamiltonian vector fields of the action coordinates ζ. In §6b, we prove that for t sufficiently large, the fibres of the Flaschka–Ratiu system intersect the chamber AN + of matrices in AN such that the minors Δ(k)i > 0 (one may think of this as the chamber of ‘totally positive’ matrices in AN). These facts are combined in §6c to prove proposition 6.1 and theorem 1.2.

(a). Convergence of Hamiltonian vector fields

For I⊆{1, …, k}, let LI be the linear function so that Inline graphic (cf. equation (2.3)). Recall also that for I, J⊆{n − k + 1, …, n}, |I| = |J| = j, there exists a Laurent polynomial PI,J so that

(a).

(see equation (3.8)). For all 1 ≤ j ≤ k ≤ n, rearrange equation (3.9) to define functions (for fixed finite t)

(a). 6.1

where we have substituted Inline graphic into each PI,J (see example 3.4).

By theorem 2.2, the functions Inline graphic solve the system of equations

(a).

The first step in the proof of proposition 6.1 is to apply the Implicit Function Theorem to this system of equations to get asymptotic control on the partial derivatives of ℓ with respect to ζ, φ.

For both sums on the right side of equation (6.1), there is a single term that dominates exponentially for large t. In the first sum, recall that if Inline graphic, and I≠{1, …, i},

(a). 6.2

so the term et(k)j dominates for large t. In the second sum, if Inline graphic, and t > 0 is sufficiently large, then by proposition 4.8, there exists wWδ and ϕTm such that Δt(az0(tw, ϕ)) = (ζ, φ). Unpacking the proof of [15, Proposition 2], for I, J⊆{n − k + 1, …, n}, |I| = |J| = j, not both equal to {n − k + 1, …, n − k + i},

(a). 6.3

so the term |Δ(k)j|2 = e2(k)j dominates the sum for large t. In other words, the j × j minors of the bottom right k × k submatrix are dominated exponentially by the corner minor Δ(k)j. The reader may find it useful to work this out for a couple examples, or compare with example 3.4.

In what follows, we order the coordinates

(a).

and similarly for ζ and φ.

Lemma 6.2 —

For t > 0 sufficiently large,

graphic file with name rsta20170428-e65.jpg

where δk,p and δi,q are Kronecker delta functions.

Note that as the functions f(k)j do not involve ζ(p)'s and φ(p)'s for p > k, the partial derivatives

(a).

Proof. —

For finite t, let Inline graphic with coordinates

graphic file with name rsta20170428-e67.jpg

as defined in equation (6.1). We will apply the Implicit Function Theorem at solution of f(ℓ, ζ, φ) = 0 with Inline graphic fixed, and

graphic file with name rsta20170428-e68.jpg

Note that if Inline graphic, then by proposition 5.1, Inline graphic for t sufficiently large.

Order the coordinates ℓ, ζ, φ as above. The matrix of partial derivatives of f with respect to ℓ is the (n + m) × (n + m) block diagonal matrix

graphic file with name rsta20170428-e69.jpg

where f(k) = (f(k)1, …, f(k)k) and ℓ(k) = (ℓ(k)1, …, ℓ(k)k). If Inline graphic, then by equation (6.2),

graphic file with name rsta20170428-e70.jpg

where A(k) is a k × k matrix whose entries are O(e). Thus, for t sufficiently large, Df is invertible. By the Implicit Function Theorem,

graphic file with name rsta20170428-e71.jpg 6.4

The function f(k) does not depend on ζ(k′)'s for k′ > k. Therefore, the (n + m) × (n + m) matrix

graphic file with name rsta20170428-e72.jpg

is the block upper-triangular, as is the (n + m) × m matrix (∂f/∂φ). As Inline graphic, by equation (6.3), for each i ≤ k,

graphic file with name rsta20170428-e73.jpg

where for k = i, B(k,k) = I + B(k) is k × k with the entries of matrix B(k) in O(e), and for k > i, B(k,i) is a k × i matrix with entries in O(e). Similarly, for each i ≤ k,

graphic file with name rsta20170428-e74.jpg

where C(k,i)t is a k × (i − 1) matrix with entries in O(t−1e) (we absorb a factor of t−1 into C(k,i)t to simplify the equation on the next line). Thus,

graphic file with name rsta20170428-e75.jpg

(note that the ordering of the diagonal entries is the same as for ℓ and f above). Plugging this into equation (6.4), we have

graphic file with name rsta20170428-e76.jpg

where A is the block diagonal matrix whose diagonal blocks are A(k). The matrix (I + A)−1 = I + O(e), and the result can now be read from the form of this matrix. ▪

In our chart Inline graphic, the Hamiltonian vector field of ℓ(k)i(ζ, φ) with respect to πt is

(a).

and the Hamiltonian vector field of ζ(k)i with respect to π is Xζ(k)i = {ζ(k)i, − }.

Lemma 6.3 —

For all 1 ≤ i ≤ k ≤ n and δ > 0, X(k)i converges uniformly to 2Xζ(k)i on Inline graphic.

Proof. —

In matrix notation, the Hamiltonian vector field

graphic file with name rsta20170428-e78.jpg

Combining theorem 3.1 and lemma 6.2, for Inline graphic and t sufficiently large,

graphic file with name rsta20170428-e79.jpg

The result follows by proposition 5.1, because et(2ζ(k)i − ℓ(k)i) → 1 as t → ∞. ▪

(b). Totally positive matrices and fibres of the Flaschka–Ratiu system

Let Inline graphic be the set of matrices in AN with real entries. The hypersurfaces defined by equations Δ(k)i = 0 divide Inline graphic into chambers. The chamber of ‘totally positive’ matrices is

(b).

The restrictions of the functions φ(k)i to Inline graphic, defined where Δ(k)≠0, take values in {0, π}. Each chamber of Inline graphic is a joint level set of these functions. By the description above, AN + is the joint level set where every φ(k)i = 0.

The set Inline graphic is also divided into chambers by the Flaschka–Ratiu systems (i.e. by the hypersurfaces where singular values ln(λ(k)i(bb*)) collide). By theorem 2.2, each of these chambers equals Inline graphic, for a connected component Inline graphic of Inline graphic.

Although the two chamber structures are different, in lemma 6.6 we prove there is a connected component Inline graphic so that for arbitrary δ > 0 and t sufficiently large, the subset Inline graphic is contained in AN + . Here, Inline graphic. As we will see in the next subsection, this implies there is a choice of angle coordinates ψ for the Gelfand–Zeitlin system such that for sufficiently large t, points in the Lagrangian section where ψ = 0 are sent by the scaled Ginzburg–Weinstein map to points in the Lagrangian section where φ = 0.

Lemma 6.4 —

The map Inline graphic gives a 1-1 correspondence between chambers of Inline graphic and coordinates ϕ∈{0, π}m, i.e. the chambers of Inline graphic equal the images Inline graphic for fixed ϕ∈ {0, π}m, and these images are distinct.

This lemma implies that the chambers of Inline graphic are connected.

Proof. —

We can prescribe the chamber of Inline graphic containing az0(x, z) = a(x)z0(z) by prescribing the z's inductively. We ignore the matrix factor a(x), as it is always positive. As shown in figure 5, label the coordinates z1, …, zm by

graphic file with name rsta20170428-e81.jpg

The minor Δ(2)1(z0(z)) = z1,2, so we can prescribe its sign by prescribing the sign of z1,2. Assume we have prescribed the signs of the minors Δ(p)q(z0(z)) by prescribing the coordinates zq,p, for all 1 ≤ q < p < k. By Lindström lemma, for all 1 ≤ i < k,

graphic file with name rsta20170428-e82.jpg

Thus, we can prescribe the signs of Δ(k)i(z0(z)) by prescribing the signs of zi,k. ▪

Figure 5.

Figure 5.

Label of variables.

Combining lemma 6.4 and proposition 4.8, for all δ > 0, there exists a t0 > 0 such that, for all tt0 and Inline graphic, there exists wWδ/2 and ϕTm, so that

(b).

is contained whichever chamber of Inline graphic we please, by choosing ϕ appropriately. Thus, the fibre Inline graphic intersects every chamber of Inline graphic at least once.

Lemma 6.5 —

For all δ > 0, there exists t0 > 0 such that for all tt0 and Inline graphic, the fibre Inline graphic intersects every chamber of Inline graphic exactly once.

Proof. —

Fix δ > 0 and let t0 as in proposition 4.8. The fibre L−1(ℓ) intersects Sym(n) at exactly 2m points (this follows directly from linear algebra). By theorem 2.2, Inline graphic, and Inline graphic if and only if x∈Sym(n). Thus, the fibres of Inline graphic intersect Inline graphic at exactly 2m points. As there are 2m chambers in Inline graphic, and Inline graphic intersects every chamber of Inline graphic at least once, this completes the proof. ▪

Next, we show that there is a unique connected component Inline graphic whose elements are sent to AN + by gwt for t sufficiently large. For a connected component Inline graphic, let Inline graphic.

Lemma 6.6 —

There is a unique connected component Inline graphic such that for all δ > 0, there exists a t0≥0 such that for all tt0 and Inline graphic,

graphic file with name rsta20170428-e84.jpg

Proof. —

Fix δ > 0 arbitrary. By lemma 6.5, there exists t0≥0 such that for all tt0 and Inline graphic, Inline graphic intersects every chamber of Inline graphic exactly once.

Consider the set

graphic file with name rsta20170428-e85.jpg

where Inline graphic (recall L is the Gelfand–Zeitlin map with coordinates ℓ(k)i). By lemma 6.5, the image of this set under L equals Inline graphic.

If there exists disjoint open subsets A, B⊆Symδ(n), so that

graphic file with name rsta20170428-e86.jpg

then

graphic file with name rsta20170428-e87.jpg

As the restriction of L to Inline graphic is open as a map to Inline graphic, L(A) and L(B) are open. As for all Inline graphic, Inline graphic intersects AN+ exactly once, the sets L(A) and L(B) are disjoint. This implies that Inline graphic is not connected, which is a contradiction (it is convex). Thus, the set gw−1t0(AN + )∩Symδ(n) is connected, and must be contained in a connected component Inline graphic.

The result follows for all tt0, since Inline graphic and Inline graphic for all t≥1. Explicitly, given tt0 and Inline graphic, we have by definition of gw that

graphic file with name rsta20170428-e88.jpg

 ▪

(c). Proof of proposition 6.1 and theorem 1.2

Proof of proposition 6.1. —

Fix an arbitrary regular fibre of the Gelfand–Zeitlin system, L−1(ℓ0). As Inline graphic, Inline graphic for some δ > 0. We will show the functions φ(k)i○Δt○gwt converge on L−1(ℓ0) by showing that they converge at a point in L−1(ℓ0) and their derivatives converge uniformly on L−1(ℓ0).

Let Inline graphic be the unique connected component described in lemma 6.6. Choose angle coordinates ψ for the Gelfand–Zeitlin system so that Inline graphic. Let x be the unique point in Inline graphic that has coordinates ψ(x) = 0. By lemma 6.6, for all t sufficiently large, gwt(x)∈AN + . Thus, for all t sufficiently large and 1 ≤ q < p ≤ n,

graphic file with name rsta20170428-e89.jpg

which equals the value of the linear combination ψ(p−1)q + ‘higher terms’ at x for our choice of angle coordinates.

Second, for all AL−1(ℓ0), 1 ≤ i ≤ k, and for t sufficiently large,

graphic file with name rsta20170428-e90.jpg 6.5

By lemma 6.3, this converges uniformly on any Inline graphic to the constant function

graphic file with name rsta20170428-e91.jpg 6.6

As Δt○gwt(A) is contained in Inline graphic for t sufficiently large, this completes the proof. ▪

Proof of theorem 1.2. —

As observed in the proof of [13, theorem 7], there is a unique Poisson isomorphism from Inline graphic equipped with Inline graphic to Inline graphic equipped with π such that

graphic file with name rsta20170428-e92.jpg

and

graphic file with name rsta20170428-e93.jpg

Combining propositions 5.1 and 6.1, we see that in coordinates ζ, φ on Inline graphic the map (1.1) converges on Inline graphic to this Poisson isomorphism. This is the Gelfand–Zeitlin system, up to a linear change of coordinates. ▪

7. Conclusion

Recall from §4 that the tropicalization of a positive polynomial p(x), in complex variables, is equal, on an open dense set, to the ‘tropical’ limit

7.

where we have substituted Inline graphic. The scaling limit of Ginzburg–Weinstein diffeomor- phisms studied in this paper may be viewed as a non-abelian ‘tropical limit’: Ginzburg–Weinstein diffeomorphisms can be written as the composition

7.

where Inline graphic, and f(A) = Ad*Ψ(A)A, Inline graphic, is the flow of a Moser vector field on Inline graphic [1,11]. Using this factorization, we can rewrite the limit of proposition 5.1 in the more suggestive form,

7.

where p = p(k)i is a positive polynomial in matrix factorization coordinates, and we have suppressed the diffeomorphism PAN.

We hope that a more general theory of non-abelian tropical limits including Ginzburg–Weinstein maps for compact Lie groups other than U(n) will emerge.

Acknowledgments

The authors would like to thank B. Hoffman, M. Podkopaeva and A. Szenes for interesting and productive discussions.

Data accessibility

This work does not have any supporting data.

Author's contributions

All authors contributed equally to the main results. J.L. and Y.L. drafted the manuscript. All authors read and approved the manuscript.

Competing interests

The authors declare that they have no competing interests.

Funding

Our work was supported in part by the project MODFLAT of the European Research Council (ERC), by the grant nos. 178794 and 178828 of the Swiss National Science Foundation (SNSF) and by the NCCR SwissMAP of the SNSF.

References

  • 1.Alekseev A. 1997. On Poisson actions of compact Lie groups on symplectic manifolds. J. Differential Geom. 45, 241–256. ( 10.4310/jdg/1214459796) [DOI] [Google Scholar]
  • 2.Mishchenko AS, Fomenko AT. 1987. Euler equations on finite-dimensional Lie groups. Izv. Acad. Nauk SSSR, Ser. matem. 42, 396–415. (1978) (Russian); English translation: Math. USSR-Izv. 12 (1978), No. 2, pp. 371–389. [Google Scholar]
  • 3.Sadetov ST. 2004. A proof of the Mishchenko-Fomenko conjecture. Dokl. Akad. Nauk 397, 751–754. [Google Scholar]
  • 4.Guillemin V, Sternberg S. 1980. The moment map and collective motion. Ann. Phys. 127, 220–253. ( 10.1016/0003-4916(80)90155-4) [DOI] [Google Scholar]
  • 5.Guillemin V, Sternberg S. 1983. On collective complete integrability according to the method of Thimm. Ergodic Theory Dyn. Syst. 3, 219–230. ( 10.1017/s0143385700001930) [DOI] [Google Scholar]
  • 6.Thimm A. 1981. Integrable geodesic flows on homogeneous spaces. Ergodic Theory Dyn. Syst. 1, 495–517. ( 10.1017/S0143385700001401) [DOI] [Google Scholar]
  • 7.Guillemin V, Sternberg S. 1983. The Gelfand-Cetlin system and quantization of the complex flag manifolds. J. Funct. Anal. 52, 106–128. ( 10.1016/0022-1236(83)90092-7) [DOI] [Google Scholar]
  • 8.Pabiniak P. 2014. Gromov width of non-regular coadjoint orbits of U(n), SO(2n) and SO(2n + 1). Math. Res. Lett. 21, 187–205. ( 10.4310/MRL.2014.v21.n1.a15) [DOI] [Google Scholar]
  • 9.Nishinou T, Nohara Y, Ueda K. 2010. Toric degenerations of Gelfand-Cetlin systems and potential functions. Adv. Math. 224, 648–706. ( 10.1016/j.aim.2009.12.012) [DOI] [Google Scholar]
  • 10.Harada M, Kaveh K. 2015. Integrable systems, toric degenerations and Okounkov bodies. Invent. Math. 202, 927–985. ( 10.1007/s00222-014-0574-4) [DOI] [Google Scholar]
  • 11.Alekseev A, Meinrenken E. 2007. Ginzburg-Weinstein via Gelfand-Zeitlin. J. Differential Geom. 76, 1–34. ( 10.4310/jdg/1180135664) [DOI] [Google Scholar]
  • 12.Flaschka H, Ratiu T. 1995. A convexity theorem for Poisson actions of compact Lie groups. IHES preprint available at http://preprints.cern.ch.
  • 13.Alekseev A, Davydenkova I. 2014. Inequalities from Poisson brackets. Indag. Math. (N.S.) 25, 846–871. ( 10.1016/j.indag.2014.07.003) [DOI] [Google Scholar]
  • 14.Alekseev A, Berenstein A, Hoffman B, Li Y. 2017 Poisson structures and potentials. (http://arxiv.org/abs/1709.09281)
  • 15.Alekseev A, Podkopaeva M, Szenes A. 2017. A symplectic proof of the Horn inequalities. Adv. Math 318, 711–736. ( 10.1016/j.aim.2017.01.006) [DOI] [Google Scholar]
  • 16.Alekseev A, Podkopaeva M, Szenes A. 2017. The Horn problem and planar networks. Adv. Math 318, 678–710. ( 10.1016/j.aim.2017.01.019) [DOI] [Google Scholar]
  • 17.Drinfel'd VG. 1983. Hamiltonian structures on Lie groups, Lie bialgebras and the geometric meaning of the classical Yang - Baxter equations. Soviet Math. Dokl 27, 68–71. [Google Scholar]
  • 18.Semenov-Tian-Shansky MA. 1985. Dressing transformations and Poisson Lie group actions. Publ. RIMS, Kyoto University 21, 1237–1260. ( 10.2977/prims/1195178514) [DOI] [Google Scholar]
  • 19.Lu J-H. 1990. Multiplicative and affine Poisson structures on Lie groups. PhD thesis, University of California, Berkeley.
  • 20.Ginzburg VL, Weinstein A. 1992. Lie–Poisson structure on some Poisson Lie groups. J. Am. Math. Soc. 5, 445–445. ( 10.1090/S0894-0347-1992-1126117-8) [DOI] [Google Scholar]
  • 21.Conn J. 1985. Normal forms for smooth Poisson structures. Ann. Math. 121, 565–593. ( 10.2307/1971210) [DOI] [Google Scholar]
  • 22.Enriquez B, Etingof P, Marshall I. 2005. Comparison of Poisson structures and Poisson–Lie dynamical r-matrices. Int. Math. Res. Not. 36, 2183–2198. [Google Scholar]
  • 23.Boalch P. 2001. Stokes matrices, Poisson Lie groups and Frobenius manifolds. Invent. Math. 146, 479–506. ( 10.1007/s002220100170) [DOI] [Google Scholar]
  • 24.Berenstein A, Fomin S, Zelevinsky A. 2005. Cluster algebras III: upper bounds and double Bruhat cells. Duke Math. J. 126, 1–52. ( 10.1215/S0012-7094-04-12611-9) [DOI] [Google Scholar]
  • 25.Fomin S, Zelevinsky A. 2002. Cluster algebras. I. Foundations. J. Am. Math. Soc. 15, 497–530. ( 10.1090/S0894-0347-01-00385-X) [DOI] [Google Scholar]
  • 26.Berenstein A, Zelevinsky A. 2001. Tensor product multiplicities, canonical bases and totally positive varieties. Invent. Math. 143, 77–128. ( 10.1007/s002220000102) [DOI] [Google Scholar]
  • 27.Fomin S, Zelevinsky A. 2000. Total positivity: tests and parametrizations. Math. Intelligencer 22, 23–33. ( 10.1007/BF03024444) [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

This work does not have any supporting data.


Articles from Philosophical transactions. Series A, Mathematical, physical, and engineering sciences are provided here courtesy of The Royal Society

RESOURCES