Startseite The Heisenberg product seen as a branching problem for connected reductive groups, stability properties
Artikel Öffentlich zugänglich

The Heisenberg product seen as a branching problem for connected reductive groups, stability properties

  • Maxime Pelletier EMAIL logo
Veröffentlicht/Copyright: 19. November 2019

Abstract

In this article, we study, in the context of complex representations of symmetric groups, some aspects of the Heisenberg product, introduced by Marcelo Aguiar, Walter Ferrer Santos and Walter Moreira in 2017. When applied to irreducible representations, this product gives rise to the Aguiar coefficients. We prove that these coefficients are in fact also branching coefficients for representations of connected complex reductive groups. This allows to use geometric methods already developed in a previous article, notably based on notions from geometric invariant theory, and to obtain some stability results on Aguiar coefficients, generalising some of the results concerning them given by Li Ying.

1 Introduction

The Heisenberg product was first introduced by Marcelo Aguiar, Walter Ferrer Santos and Walter Moreira in [1] in order to simplify and unify a diversity of related products and coproducts (e.g. Hadamard, Cauchy, Kronecker, induction, internal, external, Solomon, composition, Malvenuto–Reutenauer, convolution, …) defined on various objects: species, representations of symmetric groups, symmetric functions, endomorphisms of graded connected Hopf algebras, permutations, non-commutative symmetric functions, quasi-symmetric functions, … .

In this paper, we are only interested in one of these contexts where they thus defined this Heisenberg product, namely, the one of complex representations (every representation considered throughout the article will be a complex vector space) of symmetric groups. In this particular context, let us fix throughout the whole article two positive integers k and l. Then we denote by 𝔖k (resp. 𝔖l) the symmetric group formed by the permutations of the finite set {1,,k} (resp. {1,,l}), and the Heisenberg product can be defined on two complex representations V and W of 𝔖k and 𝔖l, respectively. It is denoted by VW and is a direct sum of representations of the groups 𝔖i for i from max(k,l) to k+l. We explain precisely the construction of this product in Section 2.1.

One interesting thing to notice about this product is that, when k=l, VW is a direct sum of representations of 𝔖k,𝔖k+1,,𝔖2k, and the term in this direct sum which is a representation of 𝔖k corresponds simply to the tensor product VW seen as a 𝔖k-module (with 𝔖k acting diagonally). This tensor product of representations of 𝔖k is sometimes referred to as the “Kronecker product” since it gives rise to the Kronecker coefficients when applied to irreducible 𝔖k-modules. As a consequence, the Heisenberg product extends – in a certain way – this so-called Kronecker product.

An important point that we use concerning the representation theory of the symmetric groups is that the irreducible complex representations of the group 𝔖k are known: they are in bijection with the partitions of the integer k, and one can moreover construct them. If λ is a partition of k (denoted λk), i.e. a finite non-increasing sequence (λ1λ2λn>0) of positive integers (called parts) whose sum (sometimes called the size of the partition, and denoted by |λ|) is k, there is an explicit construction – several, in fact – giving a representation of 𝔖k over the field of complex numbers, which happens to be irreducible. We denote this 𝔖k-module by Mλ. We do not detail the construction of Mλ here: two examples of such can for instance be found in [3, Chapter 4].

Considering that every complex representation of a symmetric group decomposes as a direct sum of irreducible ones, it is natural to seek to understand the Heisenberg product of two of the latter. If λ and μ are respectively partitions of k and l, the Heisenberg product MλMμ is a direct sum of 𝔖i-modules for i{max(k,l),,k+l}, and then every term in this sum decomposes as a direct sum of irreducible 𝔖i-modules:

MλMμ=i=max(k,l)k+lνiMνaλ,μν.

The multiplicities aλ,μν in these decompositions are non-negative integers which are called the Aguiar coefficients. They were introduced in [12] by Li Ying, who also proved interesting stability results concerning them. We recall these results in Section 2.2.

The fact is that Li Ying’s stability results look very much like similar results already proven concerning Kronecker coefficients. Let us recall that these particular coefficients are the multiplicities gλ,μ,ν arising in the decomposition

MλMμ=νkMνgλ,μ,ν,

where λ and μ are partitions of k. In [7], we gave some geometric methods allowing to prove stability results for those, as well as for some other similar coefficients. In fact, these techniques can be applied as soon as we are looking at branching coefficients for complex connected reductive groups: if G and G^ are two such groups and if we have a morphism GG^, the branching problem consists in seeing irreducible complex representations of G^ as G-modules via the previous morphism and in wondering how as such they decompose as a direct sum of irreducible ones. As a consequence, in Section 3.1, we express the Aguiar coefficients as such branching coefficients, obtaining the following result.

Theorem 1.1.

If V1 and V2 are finite-dimensional complex vector spaces (of large enough dimension), then the Aguiar coefficients are the branching coefficients for the groups G=GL(V1)×GL(V2) and G^=GL(V1(V1V2)V2), and the morphism φ:GG^ given by (g1,g2)φg1,g2, where

φg1,g2:V1(V1V2)V2V1(V1V2)V2,
u1+(iv1(i)v2(i))+u2g1(u1)+(ig1(v1(i))g2(v2(i)))+g2(u2).

As a consequence, we can use in Section 3.2 the same methods as in [7], and obtain some new stability results, generalising in part those of Li Ying. The main one is the following.

Theorem 1.2.

Let α, β and γ be three partitions such that, for all dZ>0, adα,dβdγ=1. Then, for every triple (λ,μ;ν) of partitions, the sequence

(aλ+dα,μ+dβν+dγ)d0

is constant for d0.

In fact, Li Ying obtained the same conclusion as in the previous theorem for the triple (α,β;γ)=((1),(1);(1)). We call the triples satisfying the same property “Aguiar-stable triples”, and we give – also in Section 3.2 – four other explicit examples of such ones.

Proposition 1.3.

The triples

((2),(1);(2)),((2),(1);(1,1)),((2),(1);(3))𝑎𝑛𝑑((2),(1);(2,1))

are all Aguiar-stable.

Finally, in Section 4, we discuss about what we call “stabilisation bounds”: if (α,β;γ) is Aguiar-stable, such a bound is a non-negative integer d0 (depending on partitions λ, μ and ν) such that the sequence (aλ+dα,μ+dβν+dγ)d0 is constant for dd0. The geometric methods from [7] can here also give means to compute some stabilisation bounds. We detail the computation for the Aguiar-stable triples ((1),(1);(1)), ((2),(1);(2)) and ((2),(1);(3)).

2 Definition and first properties of the Heisenberg product

2.1 Construction

Let us recall that we fixed, already in the introduction, two positive integers k and l and that we consider the two symmetric groups 𝔖k and 𝔖l.

Remark 2.1.

Notice that, for all non-negative integers a and b, 𝔖a×𝔖b can naturally be seen as a subgroup of 𝔖a+b, thanks to the injective group morphism

ιa,b:𝔖a×𝔖b𝔖a+b,
(σa,σb)i{σa(i)ifi{1,,a},a+σb(i-a)ifi{a+1,,a+b}.

Furthermore, for any non-negative integer a, 𝔖a can be considered as a subgroup of 𝔖a×𝔖a through the diagonal embedding Δa:𝔖a𝔖a×𝔖a.

Definition 2.2.

Let V and W be two (complex) representations of 𝔖k and 𝔖l, respectively. Let i{max(k,l),,k+l}. One has the following diagram:

We then consider VW, which is a representation of 𝔖k×𝔖l, and its restriction Res𝔖i-l×𝔖k+l-i×𝔖i-k𝔖k×𝔖l(VW) to a representation of 𝔖i-l×𝔖k+l-i×𝔖i-k. Finally, we define (VW)i as the representation induced to 𝔖i from

Res𝔖i-l×𝔖k+l-i×𝔖i-k𝔖k×𝔖l(VW).

It is then an 𝔖i-module, and the Heisenberg product of V and W is

VW=i=max(k,l)k+l(VW)i.

A remarkable thing proven in [1] is that this product is associative.

Definition 2.3.

Let λk and μl. The Heisenberg product between the associated irreducible representations of the symmetric group decomposes as

MλMμ=i=max(k,l)k+lνiMνaλ,μν.

The coefficients aλ,μν are called the Aguiar coefficients.

We adopt the convention that, if the weights of the partitions λ, μ and ν are not compatible to define an Aguiar coefficient (|ν|{max(|λ|,|μ|),,|λ|+|μ|}), then aλ,μν=0. Likewise, if V and W are respectively 𝔖k- and 𝔖l-modules and if i{max(k,l),,k+l} is a positive integer, we will say that (VW)i is the trivial 𝔖i-module {0}.

Remark 2.4.

As written earlier, the Heisenberg product extends the Kronecker one: when k=l, the lower term (VW)k of VW is just VW seen as a representation of 𝔖k. As a consequence, when the three partitions λ, μ and ν have the same size, the Aguiar coefficient aλ,μν coincides with the Kronecker coefficient gλ,μ,ν.

2.2 First stability results by Li Ying

In this paragraph, we recall some results from [12] concerning the Aguiar coefficients. One can first reformulate its main result (Theorem 2.3) as follows.

Theorem 2.5 (Ying).

Let λ and μ be two partitions, and let imax(|λ|,|μ|) be an integer. Then the decomposition of the Si+d-module (Mλ+(d)Mμ+(d))i+d stabilises when d3i-|λ|-|μ|-λ1-μ1+λ2+μ2. Moreover, the stabilisation begins exactly at this particular integer.

Remark 2.6.

In the case when the positive integer i<max(|λ|,|μ|), the stabilisation of the decomposition of (Mλ+(d)Mμ+(d))i+d is trivial: this module is {0} for any d0.

From Theorem 2.5, one can immediately deduce a stabilisation result for Aguiar coefficients. But this time the stabilisation bound obtained is not optimal, and indeed, Li Ying obtains a better one in [12, Corollary 5.2], refining this stability result for Aguiar coefficients, which can be reformulated in the following way.

Proposition 2.7 (Ying).

The sequence (aλ+(d),μ+(d)ν+(d))dZ0 stabilises for all partitions λ, μ and ν when

d12(3|ν|-|λ|-|μ|-λ1-μ1-ν1+λ2+μ2+ν2-1).

3 New stability results by geometric methods

3.1 The Aguiar coefficients as branching coefficients

In order to use on the Aguiar coefficients the same methods that we used in [7], we express these as branching coefficients for connected complex reductive groups. To do this, we use a fact given in [12, Lemma 3.3]: there is a remarkable expression of the Aguiar coefficients in terms of Littlewood–Richardson and Kronecker coefficients. We already defined Kronecker coefficients in the introduction; let us now recall a definition of Littlewood–Richardson coefficients.

If V is a finite-dimensional complex vector space, the irreducible polynomial GL(V)-modules are in one-to-one correspondence with the partitions of length at most dimV. For such a partition λ, we denote by 𝕊λV the corresponding irreducible representation of GL(V) (𝕊 is standard notation, denoting the Schur functor). Then the Littlewood–Richardson coefficients appear in the following situation.

Definition 3.1.

If λ and μ are two partitions of length at most dimV, the tensor product of 𝕊λV and 𝕊μV is naturally a representation of GL(V) and thus decomposes into a direct sum of irreducible ones in the following way:

𝕊λV𝕊μV=ν(ν)dimV𝕊νVcλ,μν.

The multiplicities cλ,μν are nonnegative integers called the Littlewood–Richardson coefficients.

Then the proposition (see [12, Lemma 3.3]) on which the proof of Theorem 2.5 is strongly based is the following.

Proposition 3.2 (Ying).

For all partitions λ, μ and ν,

aλ,μν=α,β,δ,η,ρ,τcα,βλcη,ρμgβ,η,δcα,δτcτ,ρν.

We are going to use the fact that the Kronecker and Littlewood–Richardson coefficients appear also in some classical results called “branching rules”.

Lemma 3.3 (Branching rules).

Let V1 and V2 be two finite-dimensional C-vector spaces. We have a morphism

GL(V1)×GL(V2)GL(V1V2),
(g1,g2)(iv1(i)v2(i)ig1(v1(i))g2(v2(i))).

Then, for any partition ν of length at most dimV1dimV2, Sν(V1V2) can be seen as a GL(V1)×GL(V2)-module via this morphism, and as such it decomposes in the following way:

𝕊ν(V1V2)=λ,μ(𝕊λV1𝕊μV2)gλ,μ,ν.

(The multiplicities are indeed the Kronecker coefficients.)

Likewise there is a morphism

GL(V1)×GL(V2)GL(V1V2),
(g1,g2)(v1+v2g1(v1)+g2(v2))

and, for any partition ν of length at most dimV1+dimV2,

𝕊ν(V1V2)=λ,μ(𝕊λV1𝕊μV2)cλ,μν

as representations of GL(V1)×GL(V2) once again. (The multiplicities are this time the Littlewood–Richardson coefficients.)

Proof.

Some proofs of these classical facts can be found in [9, Part 1] (using Schur–Weyl duality), (3.11) and (3.12). ∎

Using Proposition 3.2 and Lemma 3.3, we can then see the Aguiar coefficients as branching coefficients for connected complex reductive groups.

Theorem 3.4.

Let V1 and V2 be two finite-dimensional C-vector spaces, and denote G=GL(V1)×GL(V2) and G^=GL(V1(V1V2)V2). We consider the morphism φ:GG^, defined by (g1,g2)φg1,g2, where

φg1,g2:V1(V1V2)V2V1(V1V2)V2,
u1+(iv1(i)v2(i))+u2g1(u1)+(ig1(v1(i))g2(v2(i)))+g2(u2).

Then the Aguiar coefficients are the branching coefficients for this situation. In other words, if ν is a partition such that (ν)dim(V1(V1V2)V2), then

𝕊ν(V1(V1V2)V2)=λ,μ(𝕊λV1𝕊μV2)aλ,μν

(as representations of GL(V1)×GL(V2)).

Proof.

Let ν be as above. Then, using the definition of the Littlewood–Richardson coefficients as well as Lemma 3.3,

𝕊ν(V1(V1V2)V2)=τ,ρcτ,ρν𝕊τ(V1(V1V2))𝕊ρV2=τ,ρ,α,δcτ,ρνcα,δτ𝕊αV1𝕊δ(V1V2)𝕊ρV2=τ,ρ,α,δ,β,ηcτ,ρνcα,δτgβ,η,δ𝕊αV1𝕊βV1𝕊ηV2𝕊ρV2=τ,ρ,α,δ,β,η,λcτ,ρνcα,δτgβ,η,δcα,βλ𝕊λV1𝕊ηV2𝕊ρV2=τ,ρ,α,δ,β,η,λ,μcτ,ρνcα,δτgβ,η,δcα,βλcη,ρμ𝕊λV1𝕊μV2=λ,μaλ,μν𝕊λV1𝕊μV2,

using Proposition 3.2. ∎

We can use this interpretation of the Aguiar coefficients to have a geometric point of view on them. For this reason, we have to recall a result known as the Borel–Weil theorem. Consider V a -vector space of dimension n>0. If B is a Borel subgroup of GL(V), then the complete flag variety

(V)={E1E2En-1for alli,dim(Ei)=i}ofV

is naturally isomorphic to GL(V)/B. Moreover, any n-tuple α=(α1,,αn) of integers defines uniquely a character of B, and we denote by α the associated one-dimensional complex representation of B. As a consequence, any partition λ of length at most n allows to define the fibre product λ=GL(V)×B-λ, which is a line bundle on GL(V)/B(V) on which GL(V) acts (by left multiplication). Then the Borel–Weil theorem states that H0((V),λ), its space of sections, is a GL(V)-module isomorphic to the dual of the irreducible representation 𝕊λV.

Corollary 3.5.

Let λ, μ, ν be three partitions. Taking V1 and V2 as in the previous theorem, we set

G=GL(V1)×GL(V2),
X=(V1)×(V2)×(V1(V1V2)V2),
=λμν*(G-linearised line bundle onX).

Then aλ,μν=dimH0(X,L)G.

Proof.

Using Theorem 3.4,

𝕊ν(V1(V1V2)V2)=λ,μ(𝕊λV1𝕊μV2)aλ,μν

as representations of G=GL(V1)×GL(V2). As a consequence, Schur’s lemma implies that

aλ,μν=dim((𝕊λV1)*(𝕊μV2)*𝕊ν(V1(V1V2)V2))G,

and we immediately get the conclusion by using the Borel–Weil theorem three times. ∎

3.2 Consequences and new examples of stable triples

Since the Aguiar coefficients can be expressed as dimH0(X,)G, for well-chosen G, X and (cf. Corollary 3.5), the same techniques used in [7] for Kronecker coefficients apply. The main notion we use comes from geometric invariant theory and is that of “semi-stable points”: if X is a projective algebraic variety (defined over ), on which a connected complex reductive group G acts, and if is a G-linearised line bundle on X, then the set of semi-stable points in X relative to is

Xss()={xXthere existsd>0andσH0(X,d)Gsuch thatσ(x)0}.

The set of unstable points relative to is its complement: Xus()=XXss(). This geometric point of view allows us to obtain the following.

Theorem 3.6.

Let α, β and γ be three partitions such that, for all dZ>0, adα,dβdγ=1. Then, for every triple (λ,μ;ν) of partitions, the sequence

(aλ+dα,μ+dβν+dγ)d0

is constant for d0.

Proof.

We give a sketch of the proof. For full details, see [7]. With Corollary 3.5, we can write, with V1 and V2 vector spaces of large enough dimension,

X=(V1)×(V2)×(V1(V1V2)V2),
G=GL(V1)×GL(V2),
=αβγ*,=λμν*,

for all d0,

aλ+dα,μ+dβν+dγ=dimH0(X,d)G.

Then a result by Victor Guillemin and Shlomo Sternberg (see [7, Proposition 2.4]) gives

H0(X,d)GH0(Xss(d),d)G.

Moreover, there exists d00 such that Xss(d)Xss() as soon as dd0 (see [7, Proposition 2.7]). Hence, when dd0,

H0(X,d)GH0(Xss(),d)G.

Now, since H0(X,d)G for any d>0, we can use a corollary of Luna’s étale slice theorem (see [7, Section 2.3]): Xss()G×HS with H a reductive subgroup of G and S a finite-dimensional -vector space on which H acts linearly. As a consequence, if dd0,

H0(X,d)GH0(G×HS,d)GH0(S,d)HH0(S,)H

since is trivial as an H-linearised line bundle on S (see [7, Proposition 2.8]). ∎

Definition 3.7.

A triple (α,β;γ) of partitions such that aα,βγ0 and that, for every triple (λ,μ;ν) of partitions, (aλ+dα,μ+dβν+dγ)d0 is constant for d0 is said to be Aguiar-stable.

With Theorem 3.6, we re-obtain immediately Li Ying’s result on the stabilisation of the Aguiar coefficients (minus the stabilisation bound), which can be reformulated as follows.

Corollary 3.8.

The triple ((1),(1);(1)) is Aguiar-stable.

Proof.

For all d>0, according to Remark 2.4, a(d),(d)(d)=g(d),(d),(d)=1. ∎

Remark 3.9.

On a more general note, the same reasoning shows that every stable triple (i.e. the same as Aguiar-stable but in the sense of Kronecker coefficients) is Aguiar-stable. For results producing stable triples, see [10, 4, 11, 6].

We can also give some new explicit examples of “small” Aguiar-stable triples.

Proposition 3.10.

The triples

((2),(1);(2)),((2),(1);(1,1)),((2),(1);(3))𝑎𝑛𝑑((2),(1);(2,1))

are all Aguiar-stable triples.

Proof.

Let us write the proof in detail for ((2),(1);(2)), for instance. The three other ones work similarly. Let d>0. Then

ad(2),d(1)d(2)=α,ρ,τ,β,η,δcα,β(2d)cη,ρ(d)gβ,η,δcα,δτcτ,ρ(2d).

But the Littlewood–Richardson rule (see for instance [2, Section 5]) shows that the coefficient cα,β(2d) is zero unless α and β have only one part, and |α|+|β|=2d (and then this coefficient is 1). As a consequence,

ad(2),d(1)d(2)=ρ,τ,η,δ,n0,2dcη,ρ(d)g(n),η,δc(2d-n),δτcτ,ρ(2d).

The same is true for the coefficient cη,ρ(d) and the partitions η and ρ. So

ad(2),d(1)d(2)=τ,δ,n0,2d,m0,dg(n),(d-m),δc(2d-n),δτcτ,(m)(2d).

And then the Kronecker coefficient g(n),(d-m),δ is zero unless n=d-m. Moreover, if this is verified, g(n),(n),δ is zero unless δ=(n) (and then this coefficient is 1). Hence

ad(2),d(1)d(2)=τ,n0,dc(2d-n),(n)τcτ,(d-n)(2d).

The coefficient c(2d-n),(n)τ is then zero unless |τ|=2d. Furthermore, the other coefficient cτ,(d-n)(2d) is zero unless |τ|=2d-d+n=d+n. So

ad(2),d(1)d(2)=τ2dc(2d-d),(d)τcτ,(d-d)(2d)=τ2dc(d),(d)τcτ,(0)(2d).

Finally, this product is zero unless τ=(2d) (by the Littlewood–Richardson rule, for instance). Thus

ad(2),d(1)d(2)=c(d),(d)(2d)c(2d),(0)(2d)=1,

and ((2),(1);(2)) is Aguiar-stable by Theorem 3.6. ∎

4 Some explicit stabilisation bounds

Definition 4.1.

When (α,β;γ) is an Aguiar-stable triple, a stabilisation bound for (α,β;γ) is, for any triple (λ,μ;ν) of partitions, an integer d00 (depending on λ, μ and ν) such that aλ+dα,μ+dβν+dγ is constant for dd0.

In this section, we are going to compute stabilisation bounds for three examples of Aguiar-stable triples: ((1),(1);(1)), ((2),(1);(2)), and ((2),(1);(3)). The proof of Theorem 3.6 gives us a sufficient condition to obtain them: let us fix from now on an Aguiar-stable triple (α,β;γ) (we will specialise this triple later) and a triple (λ,μ;ν) of partitions. We also consider vector spaces V1 and V2 as before (of dimension at least 2), and denote V=V1(V1V2)V2 such that

aα,βγ(=1)=dimH0(X,)Gandaλ,μν=dimH0(X,)G,

with

G=GL(V1)×GL(V2),X=(V1)×(V2)×(V),
=αβγ*and=λμν*.

We fix finally a basis e¯=(e1,,en1) of V1 and a basis f¯=(f1,,fn2) of V2. We now know that every d00 such that, for all dd0,

Xss(d)Xss()

is a stabilisation bound.

One important tool for our computation is a numerical criterion of semi-stability known as the Hilbert–Mumford criterion.

Definition 4.2.

Let Y be a projective variety defined over , on which a connected complex reductive group H acts, and 𝒩 an H-linearised line bundle on Y. Let yY, and let τ be a one-parameter subgroup of H (denoted τX*(H)). Since Y is projective, limt0τ(t).y exists. We denote it by z. This point is fixed by the image of τ, and so * acts via τ on the fibre 𝒩z. Then there exists an integer μ𝒩(y,τ) such that, for all t* and z~𝒩z,

τ(t).z~=t-μ𝒩(y,τ)z~.

The Hilbert–Mumford criterion can then be stated as follows (see, for example, [8, Lemma 2]).

Proposition 4.3 (Hilbert–Mumford criterion).

In the settings of the previous definition, if in addition N is semi-ample, then

yYss(𝒩)μ𝒩(y,τ)0for allτX*(H).

Following this property, a one-parameter subgroup τ such that μ𝒩(y,τ)>0 will be said to be “destabilising” for y (relative to 𝒩).

We will then begin the computation by considering the projection

π:XX¯=(V1)×(V2)×(V*),
((W1,i)i,(W2,i)i,(Wi)i)(W1,1,W2,1,{φV*kerφ=Wn1n2+n1+n2-1}).

We also denote by ¯ the ample line bundle on X¯ whose pull-back by π is , by e¯*=(e1*,,en1*) and f¯*=(f1*,,fn2*) the dual bases of e¯ and f¯, respectively, and set n=min(n1,n2).

Proposition 4.4.

Set

φn=i=1nei*fi*V1*V2*(V1V2)*.

The G-orbit O0 of x¯0=(Ce1,Cf1,C(e1*+en1*+f1*+fn2*+φn))X¯ is open in X¯. Moreover, if we denote respectively by O1, O2 and O3 the G-orbits in X¯ of

x¯1=(e1,f2,(e1*+en1*+f2*+fn2*+φn)),
x¯2=(e1,f1,(en1*+f1*+fn2*+φn)),
x¯3=(e1,f1,(e1*+en1*+fn2*+φn)),

then

𝒪1¯𝒪2¯𝒪3¯={(v1,v2,(φ1V1*+φ2V2*+φ(V1V2)*))X¯φ1(v1)φ2(v2)φ(v1v2)=0}.

In addition, among {O1,O2,O3}, no orbit is contained in the closure of another one.

Proof.

We consider an element (v1,v2,(φ1+φ2+φ))X¯. Like the proof of [7, Proposition 3.3], we are only interested in the orbits of maximal dimension or of dimension just below that. Then, considering the usual isomorphism

(V1V2)*Hom(V1,V2*),

we say that φ corresponds to a linear map φ:V1V2*, on which G acts by conjugation. As a consequence, we only need to consider the case when φ is of maximal rank (n, that is) since all the orbits with φ of lower rank will be contained in the closure of such an orbit.

Thus we rather consider an element x¯=(v1,v2,(φ1+φ2+φn))X¯, with φn=i=1nei*fi*V1*V2*(V1V2)*, corresponding to a linear map φn:V1V2*. Then the linear maps φ1,φ2,φn,φn, together with the vectors v1 and v2, give some vector subspaces of V1, V2 and V1V2, whose relative positions will give us descriptions of the orbits we are interested in:

  1. in V1: v1, kerφn(φn)-1((v2)) and kerφ1,

  2. in V2: v2, kerφnt(φnt)-1((v1)) and kerφ2,

  3. in V1V2: v1v2 and kerφn.

Then we see that there is an open orbit 𝒪0, characterised by

  1. φ1(v1)0, φn(v1)0, kerφnkerφ1 (or rather (φn)-1((v2))kerφ1 if n=n1) and kerφ1(φn)-1((v2)),

  2. φ2(v2)0, φnt(v2)0, kerφntkerφ2 (or (φnt)-1((v1))kerφ2 if n=n2) and kerφ2(φnt)-1((v1)),

  3. φn(v1v2)0.

And the point x¯0 given above verifies all these conditions.

Finally, the subset

{(v1,v2,(φ1+φ2+φ))X¯φ1(v1)φ2(v2)φ(v1v2)=0}

can be written as 𝒪1¯𝒪2¯𝒪3¯ for three orbits 𝒪1, 𝒪2 and 𝒪3, characterised by the same equations as 𝒪0 except for

  1. φn(v1v2)=0 for 𝒪1,

  2. φ1(v1)=0 for 𝒪2,

  3. φ2(v2)=0 for 𝒪3.

Then it is easy to check that x¯1𝒪1, x¯2𝒪2 and x¯3𝒪3. ∎

4.1 Murnaghan case and comparison with the results by Li Ying

We deal in this section with the case of (α,β;γ)=((1),(1);(1)), which is the one treated by Li Ying in [12]. Since the triple ((1),(1),(1)) is also a stable triple for Kronecker coefficients, well-studied and first observed by Francis Murnaghan in [5]; we often refer to it as the “Murnaghan case”. In that case,

¯=𝒪(1)𝒪(1)𝒪(1).

Moreover, since dimH0(X¯,¯d)G=1 for any d>0 and since

(v1,v2,(φ1+φ2+φ))X¯φ(v1v2)

gives a non-zero G-invariant section of ¯ on X¯,

X¯us(¯)={(v1,v2,(φ1+φ2+φ))X¯φ(v1v2)=0}.

Thus, according to Proposition 4.4 (and its proof),

X¯us(¯)=𝒪1¯=G.x¯1¯.

In the group G=GL(V1)×GL(V2), we consider the maximal torus T of elements whose matrices in the bases e¯ and f¯ are diagonal. Let us give a practical notation for one-parameter subgroups of T: such a subgroup is of the form

τ:*T,
t((ta1ta2tan1),(tb1tb2tbn2)),

with a1,,an1, b1,,bn2 integers, and will be denoted by

τ=(a1,,an1b1,,bn2).

Then we see that the one-parameter subgroup τ1=(1,0,,00,1,0,,0) of T (hence of G) is destabilising for x¯1: μ¯(x¯1,τ1)=1. Moreover, a – not so complicated – calculation (see [7, Section 3.2.4] for details of a completely similar computation) yields

maxxπ-1(x¯1)(-μ(x,τ1))=-λ1-μ1+ν1+2ν2+k=1n1+n2-1νk+2.

It follows that:

Theorem 4.5.

The sequence (aλ+(d),μ+(d)ν+(d))dZ0 is constant when

d-λ1-μ1+ν1+2ν2+k=1n1+n2-1νk+2.

Proof.

Set

d0=-λ1-μ1+ν1+2ν2+k=1n1+n2-1νk+2.

Then, for all xπ-1(x¯1) and all d>d0,

μd(x,τ1)=μ(x,τ1)+dμ¯(x¯1,τ1)=μ(x,τ0)+d>0

because d>d0-μ(x,τ1). Thus, by the Hilbert–Mumford criterion, we have xXus(d), and π-1(G.x¯1)Xus(d) since π is G-equivariant. As a consequence, since π-1(G.x¯1) is dense in Xus() (because G.x¯1 is dense in X¯us(¯)) and since Xus(d) is closed,

Xus()Xus(d).

That shows why the sequence (aλ+(d),μ+(d)ν+(d))d0 is constant when d>d0. To justify the fact that it is constant as soon as d=d0, we use an argument of quasipolynomiality detailed in [7, Section 3.4]. ∎

Recovering Li Ying’s bound with our method

This is possible by choosing a different one-parameter subgroup destabilising x¯1: if we consider the one-parameter subgroup

τ1=(2,0,1,,10,2,1,,1),

it destabilises x¯1: μ¯(x¯1,τ1)=2. Moreover,

maxxπ-1(x¯1)(-μ(x,τ1))=-2λ1-λ3--λn1-2μ1-μ3--μn2+2ν1+4ν2+3(ν3++νn1+n2-2)+2(νn1+n2-1++νn1n2-n1-n2+5)+νn1n2-n1-n2+6++νn1n2+n1+n2-3,

which gives even a slight improvement of Li Ying’s bound for “long” partitions ν (i.e. of length >n1+n2-2), according to the previous expression of this bound (see Proposition 2.7).

Examples

The bound of Theorem 4.5 is, for instance, 15 for the triple

((7,3),(5,4,2);(6,6,5,4)),

whereas Li Ying’s (cf. Corollary 2.7) is 18 (note that the improved bound obtained right above is 17). But the contrary is also possible: for the triple

((3,3,3),(4,3,2,1);(5,4,1)),

Li Ying’s bound is 4 whereas ours is 7.

4.2 Two other cases

For (α,β;γ)=((2),(1);(2))

Then ¯=𝒪(2)𝒪(1)𝒪(2) and a non-zero G-invariant section of ¯ on X¯ is given by

(φ1+φ2+φ)X¯φ1(v1)φ(v1v2).

As a consequence,

X¯us(¯)={(v1,v2,(φ1+φ2+φ))X¯φ1(v1)φ(v1v2)=0}=𝒪1¯𝒪2¯,

thanks to Proposition 4.4 and its proof.

Then we take the same τ1 as before to destabilise x¯1 (still μ¯(x¯1,τ1)=1), and τ2=(1,0,,0-1,0,,0) which destabilises x¯2: μ¯(x¯2,τ2)=1. Finally,

maxxπ-1(x¯2)(-μ(x,τ2))=-λ1+μ1+k=1n2νk+1-k=1n1νn1n2+n1+n2+1-k.

Theorem 4.6.

The sequence (aλ+(2d),μ+(d)ν+(2d))dZ0 is constant when

d-λ1+max(-μ1+ν1+2ν2+k=1n1+n2-1νk+2,μ1+k=1n2νk+1-k=1n1νn1n2+n1+n2+1-k).

Proof.

Completely similar to Theorem 4.5. ∎

For (α,β;γ)=((2),(1);(3))

Then ¯=𝒪(2)𝒪(1)𝒪(3) and a non-zero G-invariant section of ¯ on X¯ is given by

(φ1+φ2+φ)X¯φ1(v1)φ1(v1)φ2(v2).

As a consequence,

X¯us(¯)={(v1,v2,(φ1+φ2+φ))X¯φ1(v1)φ2(v2)=0}=𝒪2¯𝒪3¯.

Then we take the same τ2 as before to destabilise x¯2 (still μ¯(x¯2,τ2)=1), and τ3=(-3,-2,,-21,0,,0,-2) which destabilises x¯3: μ¯(x¯3,τ3)=1. Finally,

maxxπ-1(x¯3)(-μ(x,τ3))=3λ1+2k=1n1-1λk+1-μ1+2μ2-2ν1+ν2-k=1n1-1νn2+k-2k=1n1n2-n1-n2+2νn1+n2-1+k-3k=1n2-1νn1n2+1+k-4k=1n1-1νn1n2+n2+k-5νn1n2.

Theorem 4.7.

The sequence (aλ+(2d),μ+(d)ν+(3d))dZ0 is constant when

dmax(-λ1+μ1+k=1n2νk+1-k=1n1νn1n2+n1+n2+1-k,maxxπ-1(x¯3)(-μ(x,τ3))).

Proof.

Once again similar to Theorem 4.5. ∎


Communicated by Britta Späth


Award Identifier / Grant number: ANR-15-CE40-0012

Funding statement: Supported by the French ANR (ANR project ANR-15-CE40-0012).

Acknowledgements

A large part of the research work concerning this article was carried out while working at the Institut Camille Jordan, CNRS UMR 5208, Université de Lyon. I would especially like to thank Nicolas Ressayre for pointing out the paper [12], and for interesting discussions and advice during the preparation of this one.

References

[1] M. Aguiar, W. Ferrer Santos and W. Moreira, The Heisenberg Product: From Hopf algebras and species to symmetric functions, São Paulo J. Math. Sci. 11 (2017), no. 2, 261–311. 10.1007/s40863-017-0072-xSuche in Google Scholar

[2] W. Fulton, Young Tableaux, London Math. Soc. Stud. Texts 35, Cambridge University, Cambridge, 1997. Suche in Google Scholar

[3] V. Lakshmibai and J. Brown, Flag Varieties - An Interplay of Geometry, Combinatorics, and Representation Theory, Texts Read. Math. 53, Hindustan Book, New Delhi, 2009. Suche in Google Scholar

[4] L. Manivel, On the asymptotics of Kronecker coefficients, J. Algebraic Combin. 42 (2015), no. 4, 999–1025. 10.1007/s10801-015-0614-1Suche in Google Scholar

[5] F. D. Murnaghan, The analysis of the Kronecker product of irreducible representations of the symmetric group, Amer. J. Math. 60 (1938), no. 3, 761–784. 10.2307/2371610Suche in Google Scholar

[6] M. Pelletier, Production of faces of the Kronecker cone containing only stable triples, preprint (2018), https://arxiv.org/abs/1805.10218. Suche in Google Scholar

[7] M. Pelletier, A geometric approach to the stabilisation of certain sequences of Kronecker coefficients, Manuscripta Math. 158 (2019), no. 1–2, 235–271. 10.1007/s00229-018-1021-4Suche in Google Scholar

[8] N. Ressayre, Geometric invariant theory and the generalized eigenvalue problem, Invent. Math. 180 (2010), no. 2, 389–441. 10.1007/s00222-010-0233-3Suche in Google Scholar

[9] S. V. Sam and A. Snowden, Introduction to twisted commutative algebras, preprint (2012), https://arxiv.org/abs/1209.5122. Suche in Google Scholar

[10] J. R. Stembridge, Generalized stability of Kronecker coefficients, preprint (2014), http://www.math.lsa.umich.edu/~jrs/papers/kron.pdf. Suche in Google Scholar

[11] E. Vallejo, Stability of Kronecker coefficients via discrete tomography, preprint (2014), http://arxiv.org/abs/1408.6219. 10.1016/j.disc.2020.111817Suche in Google Scholar

[12] L. Ying, Stability of the Heisenberg product on symmetric functions, Sém. Lothar. Combin. 80(B) (2018), Article ID 42. 10.1007/s00026-022-00591-6Suche in Google Scholar

Received: 2019-04-10
Revised: 2019-10-11
Published Online: 2019-11-19
Published in Print: 2020-03-01

© 2020 Walter de Gruyter GmbH, Berlin/Boston

Heruntergeladen am 25.9.2025 von https://www.degruyterbrill.com/document/doi/10.1515/jgth-2019-0059/html
Button zum nach oben scrollen