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

On the Kronecker product of Schur functions of square shapes

Chenchen Zhao Department of Mathematics, University of Southern California, Los Angeles, CA 90089 [email protected]
Abstract.

Motivated by the Saxl conjecture and the tensor square conjecture, which states that the tensor squares of certain irreducible representations of the symmetric group contain all irreducible representations, we study the tensor squares of irreducible representations associated with square Young diagrams. We give a formula for computing Kronecker coefficients, which are indexed by two square partitions and a three-row partition, specifically one with a short second row and the smallest part equal to 1. We also prove the positivity of square Kronecker coefficients for particular families of partitions, including three-row partitions and near-hooks.

1. Introduction

Given partitions λ,μn\lambda,\mu\vdash n, we can decompose the internal product of Schur functions as

sλsμ=νng(λ,μ,ν)sν,s_{\lambda}\ast s_{\mu}=\sum_{\nu\vdash n}g(\lambda,\mu,\nu)s_{\nu},

where g(λ,μ,ν)g(\lambda,\mu,\nu) are the Kronecker coefficients. The Kronecker coefficients can also be interpreted as the multiplicities of an irreducible module of SnS_{n} in the tensor product of irreducible modules of SnS_{n} corresponding to λ\lambda and μ\mu. Therefore, the Kronecker coefficients are certainly non-negative integers, which naturally suggests that there may be a combinatorial interpretation of the coefficients. The problem of finding a non-negative combinatorial interpretation for the Kronecker coefficients was explicitly stated by Stanley in 2000 ([Sta00] Problem 10) as a major open problem in Algebraic Combinatorics. The Kronecker coefficients have recently gained prominence within the context of algebraic complexity theory, particularly in the realm of Geometric Complexity Theory (GCT). However, as addressed by Panova in [Pan23], despite the increasing interest in the problem, little progress has been made: The Kronecker product problem is still poorly understood, and deriving an explicit combinatorial formula to solve the Kronecker product remains as an outstanding open problem in the field of Algebraic Combinatorics.

The number of irreducible representations of the symmetric group SnS_{n} is equal to the number of conjugacy classes, which is the number of integer partitions of nn. Given μn\mu\vdash n, let 𝕊μ\mathbb{S}^{\mu} denote the Specht module of the symmetric group SnS_{n}, indexed by partition μ\mu. It is worth noting that these Specht modules provide us with a way to study the irreducible representations, with each representation being uniquely indexed by an integer partition (see e.g. [Sag91]).

In [Hei+12], Heide, Saxl, Tiep, and Zalesski proved that with a few exceptions, every irreducible character of a simple group of Lie type is a constituent of the tensor square of the Steinberg character. They conjectured that for n5n\geq 5, there is an irreducible character χ\chi of AnA_{n}, whose tensor square χχ\chi\otimes\chi contains every irreducible character as a constituent. The following is the symmetric group analog of this conjecture:

Conjecture 1.1 (Tensor Square Conjecture).

For every nn except 2,4,9, there exists an irreducible representation VV of the symmetric group SnS_{n} such that the tensor square VVV\otimes V contains every irreducible representation of SnS_{n} as a summand with positive multiplicity. In terms of the correspondence of partitions, there exists a partition λn\lambda\vdash n such that the Kronecker coefficient g(λ,λ,μ)g(\lambda,\lambda,\mu) is positive for any μn\mu\vdash n.

In 2012, Jan Saxl conjectured that all irreducible representations of SnS_{n} over \mathbb{C} occur in the decomposition of the tensor square of irreducible representation corresponding to the staircase shape partition [PPV16]. This conjecture is as follows:

Conjecture 1.2 (Saxl Conjecture).

Let ρm\rho_{m} denote the staircase partition of size n:=(m+12)n:=\binom{m+1}{2}. Then g(ρm,ρm,μ)>0g(\rho_{m},\rho_{m},\mu)>0 for every μn.\mu\vdash n.

Previous work made progress towards the Tensor Square Conjecture, and specifically towards the Saxl Conjecture, see e.g. [PPV16], [Ike15], [LS16], [Li21]. Attempts have also been made to understand the Kronecker coefficients from different aspects: combinatorial interpretations for some known special shapes, see e.g. [Rem89], [RW94], [BO05], [Bla17], [Liu17]; from the perspective of the computational complexity of computing or deciding positivity of the Kronecker coefficients, see e.g. [BI08], [PP17], [IMW17].

In 2020, Bessenrodt and Panova made the following conjecture concerned with the shape of partitions satisfying the tensor square conjecture:

Conjecture 1.3 ([Pan23], Bessenrodt-Panova 2020).

For every nn, there exists k(n)k(n) such that the tensor square of every self-conjugate partition whose Durfee size is at least k(n)k(n) and is not the k×kk\times k partition satisfies the Tensor Square Conjecture.

In [PPV16], Pak, Panova, and Vallejo suggested that caret partitions may satisfy the tensor square conjecture. Many of the arguments on staircase shape could also be adapted for caret shapes and chopped-square shapes.

Most approaches to proving the positivity of a certain family of Kronecker coefficients use the semigroup property, see Section 2, which relies on breaking the partition triple into smaller partitions. The minimal elements in this procedure are the rectangular shapes, and thus understanding Kronecker positivity in general starts from understanding Kronecker coefficients of rectangular shapes.

In this paper, we study the tensor squares of irreducible representations corresponding to square Young diagrams, denoted \scalerelgXm\mathord{\scalerel*{\Box}{gX}}_{m}. We show that the Kronecker coefficients g(\scalerelgXm,\scalerelgXm,μ)g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\mu) in the case where \scalerelgXm=(mm)\mathord{\scalerel*{\Box}{gX}}_{m}=(m^{m}) has square shape and μ=(m2k,k1,1)\mu=(m^{2}-k,k-1,1) vanish if and only if k4k\leq 4 when m5m\geq 5. We give an explicit formula for g(\scalerelgXm,\scalerelgXm,μ)g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\mu) when μ=(m2k,k1,1)\mu=(m^{2}-k,k-1,1) has a short second row:

Theorem 1.4 (Theorem 3.6).

Let f(k)f(k) be the number of partitions of kk with no parts equal to 1 or 2. Let 1(α)\ell_{1}(\alpha) denote the number of different parts of a partition α\alpha. Then for 2km2\leq k\leq m,

g(\scalerelgXm,\scalerelgXm,(m2k,k1,1))=αk1α1=α21(α)f(k).g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(m^{2}-k,k-1,1))=\sum_{\begin{subarray}{c}\alpha\vdash k-1\\ \alpha_{1}=\alpha_{2}\end{subarray}}\ell_{1}(\alpha)-f(k).

We completely prove the forward direction of the following conjecture and have partial work done towards the other direction, including showing the positivity of square Kronecker coefficients for three-row partitions and near-hooks.

Conjecture 1.5.

For m7m\geq 7, g(\scalerelgXm,\scalerelgXm,μ)=0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\mu)=0 if and only if μS\mu\in S or μS\mu^{\prime}\in S, where

S:={(m23,2,1),(m24,3,1),(m2j,1j)j{1,2,4,6}}.S:=\{(m^{2}-3,2,1),(m^{2}-4,3,1),(m^{2}-j,1^{j})\mid j\in\{1,2,4,6\}\}.

We state our main results as follows:

Theorem 1.6 (Corollary 3.6.1, 3.7.1, Theorem 4.2, 4.11).

For every integer m7m\geq 7, let μm2\mu\vdash m^{2} be a partition of length at most 33, we have g(\scalerelgXm,\scalerelgXm,μ)>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\mu)>0 if and only if μ{(m23,2,1),(m24,3,1),(m22,1,1),(m21,1)}.\mu\notin\{(m^{2}-3,2,1),(m^{2}-4,3,1),(m^{2}-2,1,1),(m^{2}-1,1)\}.

Theorem 1.7 (Corollary 5.7.1).

Let mm be an integer and assume that m20m\geq 20. Define near-hook partitions μi(k,m):=(m2ki,i,1k)\mu_{i}(k,m):=(m^{2}-k-i,i,1^{k}). Then for every i8i\geq 8, we have g(\scalerelgXm,\scalerelgXm,μi(k,m))>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\mu_{i}(k,m))>0 for all k0k\geq 0.

The rest of this paper is structured as follows. In Section 2, we equip the reader with some required background information and notations. In Section 3, we present the partitions that do not occur in tensor squares of square partitions. In Section 4 and Section 5, we present the results on the positivity of square Kronecker coefficients for certain families of partitions. In Section 6, we will discuss some additional remarks and related further research.

Acknowledgements

The author would like to thank her advisor, Greta Panova, for suggesting the problem and for helpful guidance and insightful discussions throughout the project.

2. Background

2.1. Partitions

A partition λ\lambda of nn, denoted as λn\lambda\vdash n, is a finite list of weakly decreasing positive integers a (λ1,,λk)(\lambda_{1},\dots,\lambda_{k}) such that i=1kλi=n\sum_{i=1}^{k}\lambda_{i}=n. Given a partition λ\lambda, the size |λ||\lambda| is defined to be i=1kλi\sum_{i=1}^{k}\lambda_{i}. The length of λ\lambda is defined to be the number of parts of the partition and we denote it by (λ).\ell(\lambda). We use P(n)P(n) to denote the set of all partitions of nn.

We associate each partition λ=(λ1,,λk)n\lambda=(\lambda_{1},\dots,\lambda_{k})\vdash n with a Young diagram, which is a left justified array of nn boxes with λi\lambda_{i} boxes in row i.i. Denote by λ\lambda^{\prime} the conjugate (or transpose) of a partition λ\lambda. For instance, below are the Young diagrams corresponding to partition λ=(5,3,2)\lambda=(5,3,2) and its transpose λ=(3,3,2,1,1)\lambda^{\prime}=(3,3,2,1,1).

\ytableausetupmathmode,boxsize=0.75em\ydiagram6,3,2\ydiagram2+3,2+3,2+2,2+1,2+1,2+1\ytableausetup{mathmode,boxsize=0.75em}\ydiagram{6,3,2}\ydiagram{2+3,2+3,2+2,2+1,2+1,2+1}

The Durfee size of a partition λ\lambda, denoted by d(λ)d(\lambda) is the number of boxes on the main diagonal of the Young diagram of λ.\lambda. For the sake of convenience, we will refer to the irreducible representation corresponding to λ\lambda be λ.\lambda.

Definition 2.1.

For m1m\geq 1, we define the square-shaped partition \scalerelgXmm2\mathord{\scalerel*{\Box}{gX}}_{m}\vdash m^{2} to be \scalerelgXm:=(mm)\mathord{\scalerel*{\Box}{gX}}_{m}:=(m^{m}).

For nn\in\mathbb{N}, we denote the symmetric group on nn symbols by SnS_{n}. Let λ,μn\lambda,\mu\vdash n. We say that λ\lambda dominates μ\mu, denoted by λμ\lambda\unrhd\mu, if i=1jλii=1jμi\sum_{i=1}^{j}\lambda_{i}\geq\sum_{i=1}^{j}\mu_{i} for all jj.

Let pk(a,b)p_{k}(a,b) denote the number of partitions of kk that fit into an a×ba\times b rectangle. We denote the number of partitions of kk that fit into an m×mm\times m square by Pk(m).P_{k}(m). Note that Pk(m)=pk(m,m).P_{k}(m)=p_{k}(m,m).

Given μn\mu\vdash n, let χμ\chi^{\mu} denote the irreducible character of the symmetric group SμS_{\mu} and let χμ[α]\chi^{\mu}[\alpha] denote the value of χμ(ω)\chi^{\mu}(\omega) on any permutation ω\omega of cycle type α.\alpha. The characters can be computed using the Murnaghan-Nakayama Rule (see e.g. [SF97] for more details about the rule).

Theorem 2.2 (Murnaghan-Nakayama Rule).

We have

χλ(α)=T(1)ht(T),\chi^{\lambda}(\alpha)=\sum_{T}(-1)^{ht(T)},

summed over all border-strip tableaux of shape λ\lambda and type α\alpha and ht(T)ht(T) is the sum of the heights of each border-strip minus (α)\ell(\alpha).

2.2. The Kronecker coefficients

When working over the field \mathbb{C}, the Specht modules are irreducible, and they form a complete set of irreducible representations of the symmetric group. Polytabloids associated with the standard Young tableaux form a basis for the Specht modules and hence, the Specht modules can be indexed by partitions. Given μn\mu\vdash n, let 𝕊μ\mathbb{S}^{\mu} denote the Specht module of the symmetric group SnS_{n}, indexed by partition μ\mu (see e.g. [Sag91] for more details on the construction of Specht modules).

The Kronecker coefficients g(μ,ν,λ)g(\mu,\nu,\lambda) are defined as the multiplicity of 𝕊λ\mathbb{S}^{\lambda} in the tensor product decomposition of 𝕊μ𝕊ν\mathbb{S}^{\mu}\otimes\mathbb{S}^{\nu}. In particular, for any μ,ν,λn\mu,\nu,\lambda\vdash n, we can write

𝕊μ𝕊ν=λn𝕊g(μ,ν,λ).\mathbb{S}^{\mu}\otimes\mathbb{S}^{\nu}=\oplus_{\lambda\vdash n}\mathbb{S}^{\oplus g(\mu,\nu,\lambda)}.

We can also write

χμχν=λng(μ,ν,λ)χλ,\chi^{\mu}\chi^{\nu}=\sum_{\lambda\vdash n}g(\mu,\nu,\lambda)\chi^{\lambda},

and it follows that

g(μ,ν,λ)=χμχν,χλ=1n!ω𝔖nχμ[ω]χν[ω]χλ[ω].g(\mu,\nu,\lambda)=\langle\chi^{\mu}\chi^{\nu},\chi^{\lambda}\rangle=\frac{1}{n!}\sum_{\omega\in\mathfrak{S}_{n}}\chi^{\mu}[\omega]\chi^{\nu}[\omega]\chi^{\lambda}[\omega].

It follows that the Kronecker coefficients have full symmetry over its three parameters μ,ν,λn\mu,\nu,\lambda\vdash n. Further, since 1n1^{n} is the sign representation, we have χμχ1n=χμ\chi^{\mu}\chi^{1^{n}}=\chi^{\mu^{\prime}} and therefore the Kronecker coefficients have the transposition property, namely

g(μ,ν,λ)=g(μ,ν,λ)=g(μ,ν,λ)=g(μ,ν,λ).g(\mu,\nu,\lambda)=g(\mu^{\prime},\nu^{\prime},\lambda)=g(\mu^{\prime},\nu,\lambda^{\prime})=g(\mu,\nu^{\prime},\lambda^{\prime}).

2.3. Symmetric functions

For main definitions and properties of symmetric functions, we refer to [SF97] Chapter 7. Let hλh_{\lambda} denote the homogeneous symmetric functions and sλs_{\lambda} denote the Schur functions. The Jacobi-Trudi Identity (see e.g. [SF97]) is a powerful tool in our work:

Theorem 2.3 (Jacobi-Trudi Indentity).

Let λ=(λ1,,λn)\lambda=(\lambda_{1},\dots,\lambda_{n}). Then

sλ=det(hλi+ji)1i,j,n and sλ=det(eλi+ji)1i,j,n.s_{\lambda}=\det(h_{\lambda_{i}+j-i})_{1\leq i,j,\leq n}\text{ and }s_{\lambda^{\prime}}=\det(e_{\lambda_{i}+j-i})_{1\leq i,j,\leq n}.

Let cμνλc_{\mu\nu}^{\lambda}, where |λ|=|μ|+|ν||\lambda|=|\mu|+|\nu|, denote the Littlewood-Richardson coefficients. Using the Hall inner product on symmetric functions, one can define the Littlewood-Richardson coefficients as

cμνλ=sλ,sμsν=sλ/μ,sν.c_{\mu\nu}^{\lambda}=\langle s_{\lambda},s_{\mu}s_{\nu}\rangle=\langle s_{\lambda/\mu},s_{\nu}\rangle.

Namely, the Littlewood-Richardson coefficients are defined to be the multiplicity of sλs_{\lambda} in the decomposition of sμsν.s_{\mu}\cdot s_{\nu}. It is well-known that the Littlewood-Richardson coefficients have a combinatorial interpretation in terms of certain semistandard Young tableaux (see e.g. [SF97], [Sag91]).

Using the Frobenius map, one can define the Kronecker product of symmetric functions as

sμsν=λng(μ,ν,λ)sλ.s_{\mu}\ast s_{\nu}=\sum_{\lambda\vdash n}g(\mu,\nu,\lambda)s_{\lambda}.

In [Lit58], Littlewood proved the following identity, which is used frequently in our calculations:

Theorem 2.4 (Littlewood’s Identity).

Let μ,ν,λ\mu,\nu,\lambda be partitions. Then

sμsνsλ=γ|μ|δ|ν|cγδλ(sμsγ)(sνsδ),s_{\mu}s_{\nu}\ast s_{\lambda}=\sum_{\gamma\vdash|\mu|}\sum_{\delta\vdash|\nu|}c_{\gamma\delta}^{\lambda}(s_{\mu}\ast s_{\gamma})(s_{\nu}\ast s_{\delta}),

where cγδλc_{\gamma\delta}^{\lambda} is the Littlewood-Richardson coefficient.

Another useful tool to simplify our calculations is Pieri’s rule:

Theorem 2.5 (Pieri’s rule).

Let μ\mu be a partition. Then

sμs(n)=λsλ,s_{\mu}s_{(n)}=\sum_{\lambda}s_{\lambda},

summed is over all partitions λ\lambda obtained from μ\mu by adding nn boxes, with no two added elements in the same column.

2.4. Semigroup property

Semigroup property, which was proved in [CHM07], has been used extensively to prove the positivity of some families of partitions.

For two partitions λ=(λ1,λ2,λk)\lambda=(\lambda_{1},\lambda_{2},\dots\lambda_{k}) and μ=(μ1,μ2,μl)\mu=(\mu_{1},\mu_{2},\dots\mu_{l}) with klk\leq l, the horizontal sum of λ\lambda and μ\mu is defined as λ+Hμ=μ+Hλ=(λ1+μ1,λ2+μ2,,λk+μk,μk+1,,μl)\lambda+_{H}\mu=\mu+_{H}\lambda=(\lambda_{1}+\mu_{1},\lambda_{2}+\mu_{2},\dots,\lambda_{k}+\mu_{k},\mu_{k+1},\dots,\mu_{l}). The vertical sum of two partitions can be defined analogously, by adding the column lengths instead of row lengths. We define the vertical sum λ+Vμ\lambda+_{V}\mu of two partitions λ\lambda and μ\mu to be (λ+Hμ)(\lambda^{\prime}+_{H}\mu^{\prime})^{\prime}.

Theorem 2.6 (Semigroup Property [CHM07]).

If g(λ1,λ2,λ3)>0g(\lambda^{1},\lambda^{2},\lambda^{3})>0 and g(μ1,μ2,μ3)>0g(\mu^{1},\mu^{2},\mu^{3})>0, then
g(λ1+Hμ1,λ2+Hμ2,λ3+Hμ3)>0.g(\lambda^{1}+_{H}\mu^{1},\lambda^{2}+_{H}\mu^{2},\lambda^{3}+_{H}\mu^{3})>0.

Corollary 2.6.1.

If g(λ1,λ2,λ3)>0g(\lambda^{1},\lambda^{2},\lambda^{3})>0 and g(μ1,μ2,μ3)>0g(\mu^{1},\mu^{2},\mu^{3})>0, then g(λ1+Vμ1,λ2+Vμ2,λ3+Hμ3)>0.g(\lambda^{1}+_{V}\mu^{1},\lambda^{2}+_{V}\mu^{2},\lambda^{3}+_{H}\mu^{3})>0.

Note that by induction, we can extend the semigroup property to an arbitrary number of partitions and a modified version of the semigroup property allows us to use an even number of vertical additions.

3. Missing partitions in tensor squares of square partitions

In this section, we will show the absence of partitions in the tensor squares of square partitions by discussing the occurrences of two special families of partitions. Note that it follows immediately that the square shape partitions does not satisfy the Tensor Square Conjecture.

3.1. Near two-row partitions (m2k,k1,1)(m^{2}-k,k-1,1)

Recall that we let Pk(m)P_{k}(m) denote the number of partitions of kk that fit into an m×mm\times m square and let n=m2n=m^{2}. The following lemma is proved in [PP13], see also [Val14].

Lemma 3.1 ([PP13],[Val14]).

For 1kn1\leq k\leq n, g(\scalerelgXm,\scalerelgXm,(nk,k))=Pk(m)Pk1(m)g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(n-k,k))=P_{k}(m)-P_{k-1}(m).

Let λ\lambda^{\ast} denotes the mnm^{n}-complement of λ\lambda with m=λ1m=\lambda_{1} and n=λ1.n=\lambda_{1}^{\prime}. We define a π\pi-rotation of a partition λ\lambda is the shape obtained by rotating λ\lambda by 180180^{\circ}. Following Thomas and Yong ([TY10]), let the mnm^{n}-shortness of λ\lambda denote the length of the shortest straight line segment of the path of length m+nm+n from the southwest to the northeast corner of m×nm\times n rectangle that separates λ\lambda from the π\pi-rotation of λ.\lambda^{\ast}.

Example 3.2.

Consider λ=(8,4,2,2,1)\lambda=(8,4,2,2,1), m=λ1=8m=\lambda_{1}=8 and n=λ1=5n=\lambda_{1}^{\prime}=5. Then λ=(7,6,6,4).\lambda^{\ast}=(7,6,6,4). The diagram below is a demonstration for the path of length m+nm+n from the southwest to the northeast corner of a 8×58\times 5 rectangle that separates (8,4,2,2,1)(8,4,2,2,1) from the π\pi-rotation of (7,6,6,4).(7,6,6,4). The shortest straight line segment of the blue path is 11. Therefore, the 858^{5}-shortness of (8,4,2,2,1)(8,4,2,2,1) is 11.

Example 3.3.

Now consider λ=(8,8,8,3,3)\lambda=(8,8,8,3,3), m=λ1=8m=\lambda_{1}=8 and n=λ1=5n=\lambda_{1}^{\prime}=5. Then λ=(5,5).\lambda^{\ast}=(5,5). From the diagram below, we can see the lengths of straight line segments of the blue path are 2,2,6,32,2,6,3, and hence the shortest straight line segment of the blue path is 22. Therefore, the 858^{5}-shortness of (8,8,8,3,3)(8,8,8,3,3) is 22.

For the following theorem, jointly due to Gutschwager, Thomas and Yong, we follow [DTK09]:

Theorem 3.4 ([Gut10], [TY10]).

The basic skew Schur function sλ/μs_{\lambda/\mu} is multiplicity-free if and only if at least one of the following is true:

  1. (i)

    μ\mu or λ\lambda^{\ast} is the zero partition 0;

  2. (ii)

    μ\mu or λ\lambda^{\ast} is a rectangle of mnm^{n}-shortness 1;

  3. (iii)

    μ\mu is a rectangle of mnm^{n}-shortness 2 and λ\lambda^{\ast} is a fat hook (or vice versa);

  4. (iv)

    μ\mu is a rectangle and λ\lambda^{\ast} is a fat hook of mnm^{n}-shortness 1 (or vice versa);

  5. (v)

    μ\mu and λ\lambda^{\ast} are rectangles;

where λ\lambda^{\ast} denotes the mnm^{n}-complement of λ\lambda with m=λ1m=\lambda_{1} and n=λ1.n=\lambda_{1}^{\prime}.

Corollary 3.4.1.

Let λm=(mm1,m1)\lambda_{m}=(m^{m-1},m-1) denote the chopped square partition of size m21.m^{2}-1. For every pair of partitions β\beta and μ\mu such that |β|+|μ|=m21|\beta|+|\mu|=m^{2}-1, cβμλm{0,1}.c_{\beta\mu}^{\lambda_{m}}\in\{0,1\}.

Proof.

Let λm\lambda_{m}^{\ast} denote the mmm^{m}-complement of λm\lambda_{m}. Then λm=(1).\lambda_{m}^{\ast}=(1). The lengths of straight line segments of the path from the southwest to the northeast corner that separates λm\lambda_{m} from λm\lambda_{m}^{\ast} are m1,1,1,m1m-1,1,1,m-1, and therefore the mmm^{m}-shortness of λm\lambda_{m}^{\ast} is 1. Let βkm21\beta\vdash k\leq m^{2}-1. Then, sλm/βs_{\lambda_{m}/\beta} is a basic skew Schur function as the difference between consecutive rows in λm\lambda_{m} is at most 1. By Theorem 3.4 A1, sλm/βs_{\lambda_{m}/\beta} is multiplicity-free, which implies that cβμλm{0,1}c_{\beta\mu}^{\lambda_{m}}\in\{0,1\} for any μm21k\mu\vdash m^{2}-1-k. ∎

Lemma 3.5.

Let 1(α)\ell_{1}(\alpha) denote the number of different parts of partition α\alpha. For 1km1\leq k\leq m,

βk1μm2kcβμλm=Pk1(m)+βk11(β).\sum_{\beta\vdash k-1}\sum_{\mu\vdash m^{2}-k}c_{\beta\mu}^{\lambda_{m}}=P_{k-1}(m)+\sum_{\beta\vdash k-1}\ell_{1}(\beta).
Proof.

Let 1km1\leq k\leq m. Since cβμλm=1>0c_{\beta\mu}^{\lambda_{m}}=1>0, partitions β,μλm\scalerelgXm\beta,\mu\subseteq\lambda_{m}\subseteq\mathord{\scalerel*{\Box}{gX}}_{m}. Let β\beta^{\ast} and λm\lambda_{m}^{\ast} denote the complements of β\beta and λm\lambda_{m} inside the m×mm\times m square, respectively. Since cβμλmc_{\beta\mu}^{\lambda_{m}} depends only on μ\mu and the skew partition λm/β\lambda_{m}/\beta, and the skew partitions λm/β\lambda_{m}/\beta and β/λm\beta^{\ast}/\lambda_{m}^{\ast} are identical when rotated, we have cβμλm=cλmμβ=c(1)μβc_{\beta\mu}^{\lambda_{m}}=c_{\lambda_{m}^{\ast}\mu}^{\beta^{\ast}}=c_{(1)\mu}^{\beta^{\ast}}. By the Pieri’s rule (Theorem 2.5), c(1)μβ=1c_{(1)\mu}^{\beta^{\ast}}=1 if and only if μ\mu is a partition obtained from β\beta^{\ast} by removing 1 element. Since the number of ways to obtain a partition by removing an element from β\beta^{\ast} is 1(β)\ell_{1}(\beta^{\ast}), we have

βk1μm2kcβμλm=βk1βλmμm2kc(1)μβ=βk1βλm1(β).\sum_{\beta\vdash k-1}\sum_{\mu\vdash m^{2}-k}c_{\beta\mu}^{\lambda_{m}}=\sum_{\begin{subarray}{c}\beta\vdash k-1\\ \beta\subseteq\lambda_{m}\end{subarray}}\sum_{\mu\vdash m^{2}-k}c_{(1)\mu}^{\beta^{\ast}}=\sum_{\begin{subarray}{c}\beta\vdash k-1\\ \beta\subseteq\lambda_{m}\end{subarray}}\ell_{1}(\beta^{\ast}).

Note that 1(β)=1(β)1\ell_{1}(\beta^{\ast})=\ell_{1}(\beta)-1 if β1=(β)=m\beta_{1}=\ell(\beta)=m; 1(β)=1(β)\ell_{1}(\beta^{\ast})=\ell_{1}(\beta) if exactly one of β1,(β)\beta_{1},\ell(\beta) is mm; otherwise, 1(β)=1(β)+1\ell_{1}(\beta^{\ast})=\ell_{1}(\beta)+1. Hence, when 1km1\leq k\leq m, we have

βk1βλm1(β)=Pk1(m)+βk11(β).\sum_{\begin{subarray}{c}\beta\vdash k-1\\ \beta\subseteq\lambda_{m}\end{subarray}}\ell_{1}(\beta^{\ast})=P_{k-1}(m)+\sum_{\beta\vdash k-1}\ell_{1}(\beta).

Proposition 3.6 (near two-row partitions).

Let 2km2\leq k\leq m. Let f(k)f(k) denote the number of partitions of kk with no parts equal to 1 or 2, and 1(α)\ell_{1}(\alpha) denote the number of different parts of partition α\alpha. Then

g(\scalerelgXm,\scalerelgXm,(nk,k1,1))=αk1α1=α21(α)f(k).g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(n-k,k-1,1))=\sum_{\begin{subarray}{c}\alpha\vdash k-1\\ \alpha_{1}=\alpha_{2}\end{subarray}}\ell_{1}(\alpha)-f(k).
Proof.

Letting sλs_{\lambda} denote the Schur function indexed by a partition λ\lambda, we have

g(\scalerelgXm,\scalerelgXm,(nk,k1,1))=s\scalerelgXm,s(nk,k1,1)s\scalerelgXm.g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(n-k,k-1,1))=\left\langle s_{\mathord{\scalerel*{\Box}{gX}}_{m}},s_{(n-k,k-1,1)}\ast s_{\mathord{\scalerel*{\Box}{gX}}_{m}}\right\rangle.

Observe that, by Pieri’s rule (Theorem 2.5), we have

s(nk,k1)s(1)=s(nk,k1,1)+s(nk+1,k1)+s(nk,k).s_{(n-k,k-1)}s_{(1)}=s_{(n-k,k-1,1)}+s_{(n-k+1,k-1)}+s_{(n-k,k)}.

Rewriting the above identity gives us that g(\scalerelgXm,\scalerelgXm,(nk,k1,1))g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(n-k,k-1,1)) can be interpreted as

s\scalerelgXm,(s(nk,k1)s(1))s\scalerelgXms\scalerelgXm,s(nk+1,k1)s\scalerelgXms\scalerelgXm,s(nk,k)s\scalerelgXm.\left\langle s_{\mathord{\scalerel*{\Box}{gX}}_{m}},(s_{(n-k,k-1)}s_{(1)})\ast s_{\mathord{\scalerel*{\Box}{gX}}_{m}}\right\rangle-\left\langle s_{\mathord{\scalerel*{\Box}{gX}}_{m}},s_{(n-k+1,k-1)}\ast s_{\mathord{\scalerel*{\Box}{gX}}_{m}}\right\rangle-\left\langle s_{\mathord{\scalerel*{\Box}{gX}}_{m}},s_{(n-k,k)}\ast s_{\mathord{\scalerel*{\Box}{gX}}_{m}}\right\rangle.

We first note that the last two terms give two Kronecker coefficients g(\scalerelgXm,\scalerelgXm,(nk+1,k1))g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(n-k+1,k-1)) and g(\scalerelgXm,\scalerelgXm,(nk,k))g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(n-k,k)). Notice that by Lemma 3.1, we have

g(\scalerelgXm,\scalerelgXm,(nk+1,k1))=Pk1(m)Pk2(m)g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(n-k+1,k-1))=P_{k-1}(m)-P_{k-2}(m)

and

g(\scalerelgXm,\scalerelgXm,(nk,k))=Pk(m)Pk1(m).g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(n-k,k))=P_{k}(m)-P_{k-1}(m).

By Littlewood’s Identity (Theorem 2.4),

(s(nk,k1)s(1))s\scalerelgXm\displaystyle(s_{(n-k,k-1)}s_{(1)})\ast s_{\mathord{\scalerel*{\Box}{gX}}_{m}} =γn1cγ,(1)\scalerelgXm(s(nk,k1)sγ)(s(1)s(1))\displaystyle=\sum_{\gamma\vdash{n-1}}c_{\gamma,(1)}^{\mathord{\scalerel*{\Box}{gX}}_{m}}(s_{(n-k,k-1)}\ast s_{\gamma})(s_{(1)}\ast s_{(1)})
=(s(nk,k1)sλm)(s(1)),\displaystyle=(s_{(n-k,k-1)}\ast s_{\lambda_{m}})(s_{(1)}),

as cγ,(1)\scalerelgXm=1c_{\gamma,(1)}^{\mathord{\scalerel*{\Box}{gX}}_{m}}=1 if γ=λm\gamma=\lambda_{m} and cγ,(1)\scalerelgXm=0c_{\gamma,(1)}^{\mathord{\scalerel*{\Box}{gX}}_{m}}=0 for all the other partitions of size n1n-1. Taking inner product with s\scalerelgXms_{\mathord{\scalerel*{\Box}{gX}}_{m}} on both sides, we have

s\scalerelgXm,(s(nk,k1)s(1))s\scalerelgXm\displaystyle\left\langle s_{\mathord{\scalerel*{\Box}{gX}}_{m}},(s_{(n-k,k-1)}s_{(1)})\ast s_{\mathord{\scalerel*{\Box}{gX}}_{m}}\right\rangle =s\scalerelgXm,(s(nk,k1)sλm)(s(1))\displaystyle=\left\langle s_{\mathord{\scalerel*{\Box}{gX}}_{m}},(s_{(n-k,k-1)}\ast s_{\lambda_{m}})(s_{(1)})\right\rangle
=s\scalerelgXm(1),(s(nk,k1)sλm)\displaystyle=\left\langle s_{\mathord{\scalerel*{\Box}{gX}}_{m}\setminus(1)},(s_{(n-k,k-1)}\ast s_{\lambda_{m}})\right\rangle
=sλm,s(nk,k1)sλm.\displaystyle=\langle s_{\lambda_{m}},s_{(n-k,k-1)}\ast s_{\lambda_{m}}\rangle.

By Littlewood’s Identity (2.4), Jacobi-Trudi Identity (2.3), together with Corollary 3.4.1, cμβλm{0,1}c_{\mu\beta}^{\lambda_{m}}\in\{0,1\}, we have

sλm,s(nk,k1)sλm\displaystyle\left\langle s_{\lambda_{m}},s_{(n-k,k-1)}\ast s_{\lambda_{m}}\right\rangle =βk1μnk(cμβλm)2αk2γnk+1(cαγλm)2\displaystyle=\sum_{\beta\vdash k-1}\sum_{\mu\vdash n-k}(c_{\mu\beta}^{\lambda_{m}})^{2}-\sum_{\alpha\vdash k-2}\sum_{\gamma\vdash n-k+1}(c_{\alpha\gamma}^{\lambda_{m}})^{2}
=βk1μnkcμβλmαk2γnk+1cαγλm.\displaystyle=\sum_{\beta\vdash k-1}\sum_{\mu\vdash n-k}c_{\mu\beta}^{\lambda_{m}}-\sum_{\alpha\vdash k-2}\sum_{\gamma\vdash n-k+1}c_{\alpha\gamma}^{\lambda_{m}}.

Putting the pieces together, we then have

g(\scalerelgXm,\scalerelgXm,(nk,k1,1))\displaystyle g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(n-k,k-1,1))
=\displaystyle=\,\, sλm,s(nk,k1)sλmg(\scalerelgXm,\scalerelgXm,(nk+1,k1))g(\scalerelgXm,\scalerelgXm,(nk,k))\displaystyle\left\langle s_{\lambda_{m}},s_{(n-k,k-1)}\ast s_{\lambda_{m}}\right\rangle-g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(n-k+1,k-1))-g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(n-k,k))
=\displaystyle=\,\, sλm,s(nk,k1)sλm(Pk1(m)Pk2(m))(Pk(m)Pk1(m))\displaystyle\langle s_{\lambda_{m}},s_{(n-k,k-1)}\ast s_{\lambda_{m}}\rangle-(P_{k-1}(m)-P_{k-2}(m))-(P_{k}(m)-P_{k-1}(m))
=\displaystyle=\,\, βk1μnkcμβλmαk2γnk+1cαγλm(Pk(m)Pk2(m))\displaystyle\sum_{\beta\vdash k-1}\sum_{\mu\vdash n-k}c_{\mu\beta}^{\lambda_{m}}-\sum_{\alpha\vdash k-2}\sum_{\gamma\vdash n-k+1}c_{\alpha\gamma}^{\lambda_{m}}-(P_{k}(m)-P_{k-2}(m))
=\displaystyle=\,\, Pk1(m)+βk11(β)(Pk2(m)+αk21(α))(Pk(m)Pk2(m))\displaystyle P_{k-1}(m)+\sum_{\beta\vdash k-1}\ell_{1}(\beta)-\left(P_{k-2}(m)+\sum_{\alpha\vdash k-2}\ell_{1}(\alpha)\right)-(P_{k}(m)-P_{k-2}(m))
=\displaystyle=\,\, βk11(β)αk21(α)(Pk(m)Pk1(m))\displaystyle\sum_{\beta\vdash k-1}\ell_{1}(\beta)-\sum_{\alpha\vdash k-2}\ell_{1}(\alpha)-(P_{k}(m)-P_{k-1}(m))
=\displaystyle=\,\, βk1β1=β21(β)+(βk1β1>β21(β)αk21(α))(Pk(m)Pk1(m))\displaystyle\sum_{\begin{subarray}{c}\beta\vdash k-1\\ \beta_{1}=\beta_{2}\end{subarray}}\ell_{1}(\beta)+\left(\sum_{\begin{subarray}{c}\beta\vdash k-1\\ \beta_{1}>\beta_{2}\end{subarray}}\ell_{1}(\beta)-\sum_{\alpha\vdash k-2}\ell_{1}(\alpha)\right)-(P_{k}(m)-P_{k-1}(m))
=\displaystyle=\,\, βk1β1=β21(β)+βk2β1=β21(Pk(m)Pk1(m))\displaystyle\sum_{\begin{subarray}{c}\beta\vdash k-1\\ \beta_{1}=\beta_{2}\end{subarray}}\ell_{1}(\beta)+\sum_{\begin{subarray}{c}\beta\vdash k-2\\ \beta_{1}=\beta_{2}\end{subarray}}1-(P_{k}(m)-P_{k-1}(m))
=\displaystyle=\,\, βk1β1=β21(β)(Pk(m)Pk1(m)βk2β1=β21)\displaystyle\sum_{\begin{subarray}{c}\beta\vdash k-1\\ \beta_{1}=\beta_{2}\end{subarray}}\ell_{1}(\beta)-\left(P_{k}(m)-P_{k-1}(m)-\sum_{\begin{subarray}{c}\beta\vdash k-2\\ \beta_{1}=\beta_{2}\end{subarray}}1\right)
=\displaystyle=\,\, αk1α1=α21(α)f(k).\displaystyle\sum_{\begin{subarray}{c}\alpha\vdash k-1\\ \alpha_{1}=\alpha_{2}\end{subarray}}\ell_{1}(\alpha)-f(k).

The following result, which provides a necessary and sufficient condition for a near two-row partition with a short second row to vanish in the tensor square of square partitions, follows from Theorem 3.6.

Corollary 3.6.1.

Let 2km2\leq k\leq m. Then g(\scalerelgXm,\scalerelgXm,(nk,k1,1))=0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(n-k,k-1,1))=0 if and only if k4k\leq 4.

Proof.

We can easily verify that αk1α1=α21(α)=f(k)\sum_{\begin{subarray}{c}\alpha\vdash k-1\\ \alpha_{1}=\alpha_{2}\end{subarray}}\ell_{1}(\alpha)=f(k) for k{2,3,4}k\in\{2,3,4\}. Then by Proposition 3.6, we conclude that g(\scalerelgXm,\scalerelgXm,(nk,k1,1))=0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(n-k,k-1,1))=0 when k4.k\leq 4.

Next, we consider the case when k5k\geq 5. We can establish an injection from the set of all partitions of kk whose parts are at least 33 to the set of partitions of k1k-1 whose first two parts are the same, that is from

S={βkβi{1,2} for all i}S=\{\beta\vdash k\mid\beta_{i}\notin\{1,2\}\text{ for all }i\}

to

T={αk1α1=α2}.T=\{\alpha\vdash k-1\mid\alpha_{1}=\alpha_{2}\}.

This injection is achieved by removing one box from the last row of βS\beta\in S and taking the transpose. When k5k\geq 5, it follows that αk1α1=α21(α)>|T||S|=f(k)\sum_{\begin{subarray}{c}\alpha\vdash k-1\\ \alpha_{1}=\alpha_{2}\end{subarray}}\ell_{1}(\alpha)>|T|\geq|S|=f(k). Hence, we conclude that g(\scalerelgXm,\scalerelgXm,(nk,k1,1))>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(n-k,k-1,1))>0. ∎

3.2. Hooks

The following results on hook positivity are due to Ikenmeyer and Panova:

Theorem 3.7 ([IP17]).

Let b7.b\geq 7. Assume that mbm\geq b. We have g((mbk,1k),b×m,b×m)>0g((mb-k,1^{k}),b\times m,b\times m)>0 for k[0,b21]{1,2,4,6,b22,b23,b25,b27}k\in[0,b^{2}-1]\setminus\{1,2,4,6,b^{2}-2,b^{2}-3,b^{2}-5,b^{2}-7\} and is 0 for all other values of k.k.

By Theorem 3.7 and results in the previous section, we prove the forward direction of Conjecture 1.5:

Corollary 3.7.1.

For m7m\geq 7, g(\scalerelgXm,\scalerelgXm,μ)=0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\mu)=0 if μS\mu\in S or μS\mu^{\prime}\in S, where

S={(m23,2,1),(m24,3,1),(m2j,1j)j{1,2,4,6}}.S=\{(m^{2}-3,2,1),(m^{2}-4,3,1),(m^{2}-j,1^{j})\mid j\in\{1,2,4,6\}\}.
Proof.

It follows directly from Theorem 3.7 and Corollary 3.6.1. ∎

4. Constituency of families of partitions of special shapes

In this section, we will discuss the constituency of three families of special shapes in tensor squares of square partitions, including two-row partitions, near two-row partitions, and three-row partitions.

4.1. Two-row partitions

The following Theorem shown in [PP14] is a generalization of Lemma 3.1 and it tells us how to compute the Kronecker coefficients of the form g(ml,ml,(lmk,k))g(m^{l},m^{l},(lm-k,k)).

Theorem 4.1 ([PP14]).

Let n=lmn=lm, τk=(nk,k)\tau_{k}=(n-k,k), where 0kn/20\leq k\leq n/2 and set p1(l,m)=0p_{-1}(l,m)=0. Then

g(ml,ml,τk)=pk(l,m)pk1(l,m).g(m^{l},m^{l},\tau_{k})=p_{k}(l,m)-p_{k-1}(l,m).

Furthermore, when l,m8l,m\geq 8, g(ml,ml,τk)>0g(m^{l},m^{l},\tau_{k})>0 when k2.k\geq 2.

Corollary 4.1.1.

Let m7m\geq 7. For any 1km221\leq k\leq m^{2}-2, g(\scalerelgXm,\scalerelgXm,(m2k,k))>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(m^{2}-k,k))>0.

Proof.

By direct computation using the formula in Theorem 4.1, we can verify that the statement holds for m=7m=7. By strict unimodality of qq-binomial coefficients as shown in [PP13], we can obtain positivity of the Kronecker coefficients of the form g(\scalerelgXm,\scalerelgXm,(m2k,k))g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(m^{2}-k,k)) for every m8.m\geq 8.

4.2. Near two-row partitions

We will first consider the occurrences of near two-row partitions (m2k,k1,1)(m^{2}-k,k-1,1) with a second row longer than m1m-1. The following is one of our main results and is proven by considering different cases depending on different values of kk and the parity of mm.

Theorem 4.2.

Let mm be an integer. For every m5m\geq 5, g(\scalerelgXm,\scalerelgXm,(m2k,k1,1))>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(m^{2}-k,k-1,1))>0 if and only if k5.k\geq 5.

The following is a well-known result on tensor square of 2×n2\times n rectangles from [RW94]:

Theorem 4.3 ([RW94]).

The Kronecker coefficient g((n,n),(n,n),μ)>0g((n,n),(n,n),\mu)>0 if and only if either (μ)4\ell(\mu)\leq 4 and all parts even or (μ)=4\ell(\mu)=4 and all parts odd.

When mm is even and 5km225\leq k\leq\frac{m^{2}}{2}, we decompose \scalerelgXm\mathord{\scalerel*{\Box}{gX}}_{m} as \scalerelgXm=(\scalerelgXm2+V(m2,m2))+H(2m)\mathord{\scalerel*{\Box}{gX}}_{m}=(\mathord{\scalerel*{\Box}{gX}}_{m-2}+_{V}(m-2,m-2))+_{H}(2^{m}). We can find a horizontal decomposition (m2k,k1,1)=μ1+Hμ2+Hμ3(m^{2}-k,k-1,1)=\mu^{1}+_{H}\mu^{2}+_{H}\mu^{3} where μ1\mu^{1} is a three-row partition with the second row longer than 4 and the third row equal to 1, and μ2\mu^{2} and μ3\mu^{3} are partitions of 2m2m and 2m42m-4 with all parts even. Then by induction and semigroup property, we have:

Proposition 4.4.

For every even number m6m\geq 6, g(\scalerelgXm,\scalerelgXm,(m2k,k1,1))>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(m^{2}-k,k-1,1))>0 for every m+1km22m+1\leq k\leq\frac{m^{2}}{2}.

Proof.

For an even integer m6m\geq 6, we can write m=2rm=2r where r3r\geq 3. We shall proceed by induction on rr. Based on computational evidence, we observe that g(\scalerelgX6,\scalerelgX6,(62k,k1,1))>0g(\mathord{\scalerel*{\Box}{gX}}_{6},\mathord{\scalerel*{\Box}{gX}}_{6},(6^{2}-k,k-1,1))>0 for every 7k187\leq k\leq 18.

Let r4r\geq 4. Assume the inductive hypothesis that g(\scalerelgX2(r1),\scalerelgX2(r1),(4(r1)2i,i1,1))g(\mathord{\scalerel*{\Box}{gX}}_{2(r-1)},\mathord{\scalerel*{\Box}{gX}}_{2(r-1)},(4(r-1)^{2}-i,i-1,1)) for any 2r1i2(r1)22r-1\leq i\leq 2(r-1)^{2}. Let 2(r+1)k2r2.2(r+1)\leq k\leq 2r^{2}. We can decompose the square partition with side length 2r2r as follows:

\scalerelgX2r=(\scalerelgX2(r1)+V(2r2,2r2))+H(22r).\mathord{\scalerel*{\Box}{gX}}_{2r}=\left(\mathord{\scalerel*{\Box}{gX}}_{2(r-1)}+_{V}(2r-2,2r-2)\right)+_{H}(2^{2r}).

Note that by Theorem 4.3 and the transposition property of Kronecker coefficients, we obtain that g((22r),(22r),(2(r+a),2(ra)))>0g((2^{2r}),(2^{2r}),(2(r+a),2(r-a)))>0 for any 0ar0\leq a\leq r, and g((2r2,2r2),(2r2,2r2),(2(r1+b),2(r1b)))>0g((2r-2,2r-2),(2r-2,2r-2),(2(r-1+b),2(r-1-b)))>0 for any 0br1.0\leq b\leq r-1.

Consider the following system of inequalities:

{4r2k2(r+a)k12(ra)4r2k2(r+a)2(r1+b)k12(ra)2(r1b)1k12(ra)2(r1b)5.\begin{cases}4r^{2}-k-2(r+a)\geq k-1-2(r-a)\\ 4r^{2}-k-2(r+a)-2(r-1+b)\geq k-1-2(r-a)-2(r-1-b)\geq 1\\ k-1-2(r-a)-2(r-1-b)\geq 5\end{cases}.

Suppose that 0ar0\leq a\leq r, 0br10\leq b\leq r-1 is a pair of solutions to the system. We define partition α(a,b):=(4r2k2(r+a)2(r1+b),k12(ra)2(r1b),1)\alpha(a,b):=(4r^{2}-k-2(r+a)-2(r-1+b),k-1-2(r-a)-2(r-1-b),1). By inductive hypothesis, together with Corollary 3.6.1, g(\scalerelgX2(r1),\scalerelgX2(r1),α(a,b))>0.g(\mathord{\scalerel*{\Box}{gX}}_{2(r-1)},\mathord{\scalerel*{\Box}{gX}}_{2(r-1)},\alpha(a,b))>0. Note that we can decompose the near two-row partition as

(4r2k,k1,1)=α(a,b)+H(2(r1+b),2(r1b)+H(2(r+a),2(ra)).(4r^{2}-k,k-1,1)=\alpha(a,b)+_{H}(2(r-1+b),2(r-1-b)+_{H}(2(r+a),2(r-a)).

Then by semigroup property (Theorem 2.6), g(\scalerelgX2r,\scalerelgX2r,(4r2k,k1,1))>0g\left(\mathord{\scalerel*{\Box}{gX}}_{2r},\mathord{\scalerel*{\Box}{gX}}_{2r},(4r^{2}-k,k-1,1)\right)>0. By the Principle of Mathematical Induction, the statement holds for all even m6.m\geq 6.

Hence, it suffices to show the system of inequalities has integral solutions 0ar,0br10\leq a\leq r,0\leq b\leq r-1. By simplifying and rearranging, we can further reduce this system of inequalities to:

{ar2k2+142r+2k2a+br2k2+14.\begin{cases}a\leq r^{2}-\frac{k}{2}+\frac{1}{4}\\ 2r+2-\frac{k}{2}\leq a+b\leq r^{2}-\frac{k}{2}+\frac{1}{4}.\end{cases}

Notice that when k(2r1)22k\leq\frac{(2r-1)^{2}}{2}, the values a=ra=r and b=max{r+2k2,0}b=\max\{\lceil r+2-\frac{k}{2}\rceil,0\} provide a feasible solution to the system. When (2r1)22k2r2\frac{(2r-1)^{2}}{2}\leq k\leq 2r^{2}, the values a=r2k2+14a=\lfloor r^{2}-\frac{k}{2}+\frac{1}{4}\rfloor and b=0b=0 provide a feasible solution to the system. ∎

Example 4.5.

Let m=6m=6 and k=10k=10. Diagrams below illustrate a way to decompose partitions \scalerelgX6\mathord{\scalerel*{\Box}{gX}}_{6} and (26,9,1)(26,9,1). Since g(\scalerelgX4,\scalerelgX4,(8,7,1))>0g(\mathord{\scalerel*{\Box}{gX}}_{4},\mathord{\scalerel*{\Box}{gX}}_{4},(8,7,1))>0, g(26,26,(10,2))>0g(2^{6},2^{6},(10,2))>0 by Theorem 4.3 and g((4,4),(4,4),(8))>0g((4,4),(4,4),(8))>0, we conclude that g(\scalerelgX6,\scalerelgX6,(26,9,1))>0g(\mathord{\scalerel*{\Box}{gX}}_{6},\mathord{\scalerel*{\Box}{gX}}_{6},(26,9,1))>0 by semigroup property. \ytableausetupboxsize = 8px

\ydiagram6,6,6,6,6,6\displaystyle\ydiagram{6,6,6,6,6,6}\hskip 14.22636pt \ydiagram26,9,1\displaystyle\ydiagram{26,9,1}
\ydiagram4,4,4,4,0,4,4\ydiagram1+2,1+2,1+2,1+2,1+2,1+2\displaystyle\ydiagram{4,4,4,4,0,4,4}\ydiagram{1+2,1+2,1+2,1+2,1+2,1+2}\hskip 14.22636pt \ydiagram8,7,1\ydiagram1+10,1+2\ydiagram1+8\displaystyle\ydiagram{8,7,1}\ydiagram{1+10,1+2}\ydiagram{1+8}

We will next prove the positivity of g(\scalerelgXm,\scalerelgXm,(m2k,k1,1))g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(m^{2}-k,k-1,1)) when mm is odd using the semigroup property.

Proposition 4.6.

For every odd integer m7m\geq 7 and k5k\geq 5 such that k(m1)2+12k\leq\frac{(m-1)^{2}+1}{2}, g(\scalerelgXm,\scalerelgXm,(m2k,k1,1))>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(m^{2}-k,k-1,1))>0.

Proof.

Let m7m\geq 7 and k5k\geq 5. Note that when k(m1)2+12k\leq\frac{(m-1)^{2}+1}{2}, we have (m2k)(k1)2m1(m^{2}-k)-(k-1)\geq 2m-1 and we can consider the decompositions

(m2k,k1,1)=(m2k2m+1,k1,1)+H(m1)+H(m)(m^{2}-k,k-1,1)=(m^{2}-k-2m+1,k-1,1)+_{H}(m-1)+_{H}(m)

and

\scalerelgXm=(\scalerelgXm1+V(m1))+H(1m).\mathord{\scalerel*{\Box}{gX}}_{m}=(\mathord{\scalerel*{\Box}{gX}}_{m-1}+_{V}(m-1))+_{H}(1^{m}).

Then by semigroup property and Proposition 4.4, we have g(\scalerelgXm,\scalerelgXm,(m2k,k1,1))>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(m^{2}-k,k-1,1))>0 in this case. ∎

Note that the previous proof only establishes the constituency of near two-row partitions with a relatively short second row in the tensor square of square partitions with an odd side length. Now we aim to demonstrate the constituency of near two-row partitions whose first part and second part have similar sizes. To accomplish this, we will first establish the constituency of an extreme case where the second row has a maximal length:

Lemma 4.7.

For every odd integer m3m\geq 3, g(\scalerelgXm,\scalerelgXm,(m212,m212,1))>0.g\left(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\left(\frac{m^{2}-1}{2},\frac{m^{2}-1}{2},1\right)\right)>0.

Proof.

We can write odd integers mm as m=2k+1m=2k+1, and we will proceed with a proof by induction on k1k\geq 1.

We can verify the statement directly for m{3,5,7}m\in\{3,5,7\} through direct computations. When k=4k=4, we have m=2k+1=9m=2k+1=9. In this case, the square partition \scalerelgX9\mathord{\scalerel*{\Box}{gX}}_{9} can be expressed as

\scalerelgX9=((55)+V(54))+H(49).\mathord{\scalerel*{\Box}{gX}}_{9}=((5^{5})+_{V}(5^{4}))+_{H}(4^{9}).

Furthermore, we can write

(40,40,1)=(12,12,1)+H(10,10)+H(18,18).(40,40,1)=(12,12,1)+_{H}(10,10)+_{H}(18,18).

By assumption, we have g((55),(55),(12,12,1))>0g((5^{5}),(5^{5}),(12,12,1))>0. Using computer software, we can verify the positivity of g((54),(54),(10,10))g((5^{4}),(5^{4}),(10,10)) and g((49),(49),(18,18))g((4^{9}),(4^{9}),(18,18)). Therefore, by the semigroup property, we conclude that g(\scalerelgX9,\scalerelgX9,(40,40,1))>0g\left(\mathord{\scalerel*{\Box}{gX}}_{9},\mathord{\scalerel*{\Box}{gX}}_{9},(40,40,1)\right)>0.

Now let k5k\geq 5 and m=2k+1m=2k+1. By the inductive hypothesis, we assume that

g(\scalerelgXm,\scalerelgXm,(m212,m212,1))>0,g\left(\mathord{\scalerel*{\Box}{gX}}_{m^{\prime}},\mathord{\scalerel*{\Box}{gX}}_{m^{\prime}},\left(\frac{{m^{\prime}}^{2}-1}{2},\frac{{m^{\prime}}^{2}-1}{2},1\right)\right)>0,

holds for all m=2k+1<2k+1m^{\prime}=2k^{\prime}+1<2k+1. We can express \scalerelgXm\mathord{\scalerel*{\Box}{gX}}_{m} as

\scalerelgXm=(((m8)(m8))+V((m8)8))+H(8m).\mathord{\scalerel*{\Box}{gX}}_{m}=(((m-8)^{(m-8)})+_{V}((m-8)^{8}))+_{H}(8^{m}).

Furthermore, we have

(m212,m212,1)=((m8)212,(m8)212,1)+H(4(m8),4(m8))+H(4m,4m).\left(\frac{m^{2}-1}{2},\frac{m^{2}-1}{2},1\right)=\left(\frac{(m-8)^{2}-1}{2},\frac{(m-8)^{2}-1}{2},1\right)+_{H}(4(m-8),4(m-8))+_{H}(4m,4m).

Using Theorem 4.1, we know that g((8m),(8m),(4m,4m))>0g((8^{m}),(8^{m}),(4m,4m))>0. In the case of k=5k=5, where m=11m=11, we can directly compute and show the positivity of g((m8)8,(m8)8,(4(m8),4(m8)))g((m-8)^{8},(m-8)^{8},(4(m-8),4(m-8))). For k{6,7}k\in\{6,7\}, we can use the semigroup property and Theorem 4.3 to establish the positivity of g((m8)8,(m8)8,(4(m8),4(m8)))g((m-8)^{8},(m-8)^{8},(4(m-8),4(m-8))) since g((8,8),(8,8),(8,8))>0g((8,8),(8,8),(8,8))>0. For k8k\geq 8, the positivity of g((m8)8,(m8)8,(4(m8),4(m8)))g((m-8)^{8},(m-8)^{8},(4(m-8),4(m-8))) follows from Theorem 4.1. Additionally, by the inductive hypothesis, we have

g(\scalerelgXm8,\scalerelgXm8,((m8)212,(m8)212,1))>0.g\left(\mathord{\scalerel*{\Box}{gX}}_{m-8},\mathord{\scalerel*{\Box}{gX}}_{m-8},\left(\frac{(m-8)^{2}-1}{2},\frac{(m-8)^{2}-1}{2},1\right)\right)>0.

By the semigroup property (2.6), we conclude that

g(\scalerelgXm,\scalerelgXm,(m212,m212,1))>0,g\left(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\left(\frac{m^{2}-1}{2},\frac{m^{2}-1}{2},1\right)\right)>0,

which completes the induction. ∎

Corollary 4.7.1.

For every pair of odd integers l,m11l,m\geq 11, g(m×l,m×l,(ml12,ml12,1))>0.g(m\times l,m\times l,(\frac{ml-1}{2},\frac{ml-1}{2},1))>0.

Proof.

By Lemma 4.7, we know that g(\scalerelgXm,\scalerelgXm,(m212,m212,1))>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(\frac{m^{2}-1}{2},\frac{m^{2}-1}{2},1))>0 for any odd integer m3.m\geq 3.

Let m,lm,l be odd integers. Without loss of generality, assume that mlm\geq l. If |ml|0mod4|m-l|\equiv 0\mod 4, then we can write the square partition of shape m×lm\times l as \scalerelgXl+V(l(ml))\mathord{\scalerel*{\Box}{gX}}_{l}+_{V}(l^{(m-l)}). Since mlm-l is a multiple of 44, by Lemma 4.8 and the semigroup property, we conclude that g(m×l,m×l,(ml12,ml12,1))>0.g(m\times l,m\times l,(\frac{ml-1}{2},\frac{ml-1}{2},1))>0. If |ml|2mod4|m-l|\equiv 2\mod 4, we can write m×lm\times l as m×l=10×l+V(m10)×lm\times l=10\times l+_{V}(m-10)\times l. Note then (m10l)0mod4(m-10-l)\equiv 0\mod 4, and by Theorem 4.1, g(10×l,10×l,(5l,5l))>0.g(10\times l,10\times l,(5l,5l))>0. Hence, by semigroup property, we conclude that g(m×l,m×l,(ml12,ml12,1))>0g(m\times l,m\times l,(\frac{ml-1}{2},\frac{ml-1}{2},1))>0 for any odd integers m,l11.m,l\geq 11.

Lemma 4.8.

For every integer m2m\geq 2, g(m4,m4,(2m,2m))>0.g(m^{4},m^{4},(2m,2m))>0.

Proof.

If mm is even, it follows from Theorem 4.3. If mm is odd, we first note that with the help of the computer, one can check that g(34,34,(6,6))>0g(3^{4},3^{4},(6,6))>0. Then we can decompose the partition m4m^{4} as m4=34+H(m3)4m^{4}=3^{4}+_{H}(m-3)^{4}. Since m3m-3 is even, we have g((m3)4,(m3)4,(2m6.2m6))>0g((m-3)^{4},(m-3)^{4},(2m-6.2m-6))>0. By semigroup property, we can conclude that g(m4,m4,(2m,2m))>0.g(m^{4},m^{4},(2m,2m))>0.

Lemma 4.9.

For every odd integer m3m\geq 3, g(\scalerelgXm,\scalerelgXm,(m2+12,m232,1))>0.g\left(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\left(\frac{m^{2}+1}{2},\frac{m^{2}-3}{2},1\right)\right)>0.

Proof.

We can check by direct computation that the statement holds for m=3m=3 and m=5.m=5. Let m7m\geq 7. Suppose that the statement holds for odd numbers less than m.m. Consider the decomposition \scalerelgXm=(\scalerelgXm4+V(m4)4)+H(4m4)\mathord{\scalerel*{\Box}{gX}}_{m}=(\mathord{\scalerel*{\Box}{gX}}_{m-4}+_{V}(m-4)^{4})+_{H}(4^{m-4}). Since g(4m4,4m4,(2m,2m))>0g(4^{m-4},4^{m-4},(2m,2m))>0 and g((m4)4,(m4)4,(2m8,2m8))>0g((m-4)^{4},(m-4)^{4},(2m-8,2m-8))>0 by Lemma 4.8, by semigroup property, we have that g(\scalerelgXm,\scalerelgXm,(m2+12,m232,1))>0g\left(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\left(\frac{m^{2}+1}{2},\frac{m^{2}-3}{2},1\right)\right)>0. By induction, g(\scalerelgXm,\scalerelgXm,(m2+12,m232,1))>0g\left(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\left(\frac{m^{2}+1}{2},\frac{m^{2}-3}{2},1\right)\right)>0 for any odd integer m3m\geq 3. ∎

We will use Lemma 4.7 and Lemma 4.9 as ingredients to establish the positivity in the case where mm is an odd integer and the first part and second part of the near two-row partition are of similar sizes.

Proposition 4.10.

For every odd integer m7m\geq 7 and k5k\geq 5 such that k(m1)22+1k\geq\frac{(m-1)^{2}}{2}+1, g(\scalerelgXm,\scalerelgXm,(m2k,k1,1))>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(m^{2}-k,k-1,1))>0.

Proof.

We shall prove the statement by induction on odd integers m7m\geq 7. Note that we can check by semigroup property and computer that the statement holds for m=7m=7. Let m9m\geq 9 be an odd integer. Suppose that the statement holds for m2m-2. Consider the decomposition that \scalerelgXm=(\scalerelgXm2+V(m2,m2))+H(2m)\mathord{\scalerel*{\Box}{gX}}_{m}=(\mathord{\scalerel*{\Box}{gX}}_{m-2}+_{V}(m-2,m-2))+_{H}(2^{m}). Let a:=(m2k)(k1)a:=(m^{2}-k)-(k-1). Since k(m1)22+1k\geq\frac{(m-1)^{2}}{2}+1, we have (m2k)(k1)2m2.(m^{2}-k)-(k-1)\leq 2m-2. We will discuss three cases as follows.

  1. Case 1:

    If a=0a=0, by Lemma 4.7, we know that g(\scalerelgXm,\scalerelgXm,(m212,m212,1))>0.g\left(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\left(\frac{m^{2}-1}{2},\frac{m^{2}-1}{2},1\right)\right)>0.

  2. Case 2:

    If a=2a=2, by Lemma 4.9, we know that g(\scalerelgXm,\scalerelgXm,(m2+12,m232,1))>0.g\left(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\left(\frac{m^{2}+1}{2},\frac{m^{2}-3}{2},1\right)\right)>0.

  3. Case 3:

    If a>0a>0 and a0mod4a\equiv 0\mod 4, consider the following decomposition of (m2k,k1,1)(m^{2}-k,k-1,1):

    ((m2)212,(m2)212,1)+H(m+1+2x,m12x)+(m1+2y,m32y),\left(\frac{(m-2)^{2}-1}{2},\frac{(m-2)^{2}-1}{2},1\right)+_{H}(m+1+2x,m-1-2x)+(m-1+2y,m-3-2y),

    where xm12,ym32x\leq\frac{m-1}{2},y\leq\frac{m-3}{2} are non-negative integers such that 4(x+y+1)=a.4(x+y+1)=a. By Lemma 4.7, Theorem 4.3 and semigroup property, we can conclude that g(\scalerelgXm,\scalerelgXm,(m2k,k1,1))>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(m^{2}-k,k-1,1))>0 in this case.

  4. Case 4:

    If a>2a>2 and a2mod4a\equiv 2\mod 4, consider the following decomposition of (m2k,k1,1)(m^{2}-k,k-1,1):

    ((m2)2+12,(m2)232,1)+H(m+1+2x,m12x)+(m1+2y,m32y),\left(\frac{(m-2)^{2}+1}{2},\frac{(m-2)^{2}-3}{2},1\right)+_{H}(m+1+2x,m-1-2x)+(m-1+2y,m-3-2y),

    where xm12,ym32x\leq\frac{m-1}{2},y\leq\frac{m-3}{2} are non-negative integers such that 4(x+y+1)=a2.4(x+y+1)=a-2. By the inductive hypothesis, Theorem 4.3 and semigroup property, we can conclude that g(\scalerelgXm,\scalerelgXm,(m2k,k1,1))>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(m^{2}-k,k-1,1))>0 in this case.

We now put the above pieces together to prove Theorem 4.2.

Proof of Theorem 4.2.

One can check by direct computation that the proposition holds for m=5m=5 and m=7m=7. Then the statement follows directly from Corollary 3.6.1, Proposition 4.4, Proposition 4.6 and Proposition 4.10. ∎

4.3. Three-row partitions

Next, we consider the case when μ\mu is a three-row partitions with μ32.\mu_{3}\geq 2. Below is one of our main results. We will prove it by discussing different cases according to the parity of mm and different values of kk.

Theorem 4.11.

For every odd integer m5m\geq 5, g(\scalerelgXm,\scalerelgXm,μ)>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\mu)>0 for any three-row partition λm2\lambda\vdash m^{2} with μ32\mu_{3}\geq 2.

Below are some results that will be used to prove the positivity of g(\scalerelgXm,\scalerelgXm,μ)g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\mu) when mm is an even integer.

Proposition 4.12.

Let l,k,ml,k,m be positive integers such that lk=m2lk=m^{2}. Then g(\scalerelgXm,\scalerelgXm,kl)>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},k^{l})>0 if lkl\mid k.

Proof.

Let l,k,ml,k,m be positive integers such that lk=m2lk=m^{2}. Suppose that lk.l\mid k. Then, l2m2l^{2}\mid m^{2} and hence lml\mid m. It follows that we can decompose \scalerelgXm\mathord{\scalerel*{\Box}{gX}}_{m} as

\scalerelgXm=+H(+V\scalerelgXl).\mathord{\scalerel*{\Box}{gX}}_{m}=\sum_{+_{H}}\left(\sum_{+_{V}}\mathord{\scalerel*{\Box}{gX}}_{l}\right).

By semigroup property, we can conclude that g(\scalerelgXm,\scalerelgXm,kl)>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},k^{l})>0

Corollary 4.12.1.

If mm is a multiple of 33, then g(\scalerelgXm,\scalerelgXm,(m23,m23,m23))>0.g\left(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\left(\frac{m^{2}}{3},\frac{m^{2}}{3},\frac{m^{2}}{3}\right)\right)>0.

Proof.

It follows from Proposition 4.12. ∎

Lemma 4.13.

For k3k\geq 3, g(3k,3k,k3)=g(k3,k3,k3)>0.g(3^{k},3^{k},k^{3})=g(k^{3},k^{3},k^{3})>0.

Proof.

With the help of the computer, we can check that g(k3,k3,k3)>0g(k^{3},k^{3},k^{3})>0 for k{3,4,5}.k\in\{3,4,5\}. For any k6k\geq 6, we can write k=3j+rk=3j+r for some non-negative integers j,rj,r such that r{0,4,5}r\in\{0,4,5\}. Then, we can write the partition (k,k,k)(k,k,k) as a horizontal sum of jj square partitions of side length 33, and the rectangular partition (r,r,r)(r,r,r). The generalized semigroup property shows that g(k3,k3,k3)>0g(k^{3},k^{3},k^{3})>0 for k6k\geq 6. Furthermore, by the transposition property, we have g(3k,3k,k3)=g(k3,k3,k3)>0g(3^{k},3^{k},k^{3})=g(k^{3},k^{3},k^{3})>0 for k3.k\geq 3.

Lemma 4.14, 4.15 and 4.16 will be used in the proof of Proposition 4.17. These specific cases are addressed individually due to their different decomposition approach, setting them apart from the remaining cases of the proposition’s proof.

Lemma 4.14.

The Kronecker coefficient g(\scalerelgXm,\scalerelgXm,(m2+23,m213,m213))>0g\left(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\left(\frac{m^{2}+2}{3},\frac{m^{2}-1}{3},\frac{m^{2}-1}{3}\right)\right)>0 for any positive integer m5m\geq 5 such that m2mod3m\equiv 2\mod 3.

Proof.

For any positive integer m5m\geq 5 such that m2mod3m\equiv 2\mod 3, we can write m=3r+2m=3r+2 for some r1r\geq 1. We will prove the proposition by induction on rr. When r=1r=1, 3r+2=53r+2=5 and with the help of the computer, we can check that g(\scalerelgX5,\scalerelgX5,(52+23,5213,5213))>0.g\left(\mathord{\scalerel*{\Box}{gX}}_{5},\mathord{\scalerel*{\Box}{gX}}_{5},\left(\frac{5^{2}+2}{3},\frac{5^{2}-1}{3},\frac{5^{2}-1}{3}\right)\right)>0. Let r2r\geq 2. Assume the statement is true for r1r-1. We can decompose \scalerelgX3r+2\mathord{\scalerel*{\Box}{gX}}_{3r+2} as

\scalerelgX3r+2=(\scalerelgX3r1+V(3r1))+H(33r+2),\mathord{\scalerel*{\Box}{gX}}_{3r+2}=\left(\mathord{\scalerel*{\Box}{gX}}_{3r-1}+_{V}(3r-1)\right)+_{H}(3^{3r+2}),

and we can decompose the partition ((3r+2)2+23,(3r+2)213,(3r+2)213)\left(\frac{(3r+2)^{2}+2}{3},\frac{(3r+2)^{2}-1}{3},\frac{(3r+2)^{2}-1}{3}\right) as

((3r+2)2+23,(3r+2)213,(3r+2)213)\displaystyle\left(\frac{(3r+2)^{2}+2}{3},\frac{(3r+2)^{2}-1}{3},\frac{(3r+2)^{2}-1}{3}\right) =((3r1)2+23,(3r1)213,(3r1)213)\displaystyle=\left(\frac{(3r-1)^{2}+2}{3},\frac{(3r-1)^{2}-1}{3},\frac{(3r-1)^{2}-1}{3}\right)
+H(3r1,3r1,3r1)\displaystyle+_{H}(3r-1,3r-1,3r-1)
+H(3r+2,3r+2,3r+2).\displaystyle+_{H}(3r+2,3r+2,3r+2).

Then, by the inductive hypothesis, Lemma 4.13 and semigroup property, we can conclude that g(\scalerelgX3r+2,\scalerelgX3r+2,((3r+2)2+23,(3r+2)213,(3r+2)213))>0g\left(\mathord{\scalerel*{\Box}{gX}}_{3r+2},\mathord{\scalerel*{\Box}{gX}}_{3r+2},\left(\frac{(3r+2)^{2}+2}{3},\frac{(3r+2)^{2}-1}{3},\frac{(3r+2)^{2}-1}{3}\right)\right)>0. Thus, by the principle of mathematical induction, g(\scalerelgXm,\scalerelgXm,(m2+23,m213,m213))>0g\left(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\left(\frac{m^{2}+2}{3},\frac{m^{2}-1}{3},\frac{m^{2}-1}{3}\right)\right)>0 for every positive integer m5m\geq 5 such that m2mod3m\equiv 2\mod 3. ∎

Lemma 4.15.

For any positive integer m7m\geq 7 such that m1mod3m\equiv 1\mod 3, the Kronecker coefficients g(\scalerelgXm,\scalerelgXm,λ)>0g\left(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\lambda\right)>0 for λ\lambda in the set

{(m2+53,m213,m243),(m2+53,m2+53,m2103),(m2+53,m2+23,m273)}.\left\{\left(\frac{m^{2}+5}{3},\frac{m^{2}-1}{3},\frac{m^{2}-4}{3}\right),\left(\frac{m^{2}+5}{3},\frac{m^{2}+5}{3},\frac{m^{2}-10}{3}\right),\\ \left(\frac{m^{2}+5}{3},\frac{m^{2}+2}{3},\frac{m^{2}-7}{3}\right)\right\}.
Proof.

For any positive integer m7m\geq 7 such that m1mod3m\equiv 1\mod 3, we can write m=3r+1m=3r+1 for some r2r\geq 2. We will prove the proposition by induction on rr. When r=2r=2, 3r+1=73r+1=7, and with the help of the computer, we can verify the statement holds true for r=2.r=2. Let r3r\geq 3, and assume that the statement is true for r1r-1. We can decompose \scalerelgXm(r)\mathord{\scalerel*{\Box}{gX}}_{m}(r) as

\scalerelgXm(r)=(\scalerelgXm(r1)+V(m(r1)3))+H3m(r),\mathord{\scalerel*{\Box}{gX}}_{m}(r)=\left(\mathord{\scalerel*{\Box}{gX}}_{m(r-1)}+_{V}(m(r-1)^{3})\right)+_{H}3^{m(r)},

and we can decompose the partition (m(r)2+i3,m(r)2+j3,m(r)2+k3)\left(\frac{m(r)^{2}+i}{3},\frac{m(r)^{2}+j}{3},\frac{m(r)^{2}+k}{3}\right) as

(m(r)2+i3,m(r)2+j3,m(r)2+k3)\displaystyle\left(\frac{m(r)^{2}+i}{3},\frac{m(r)^{2}+j}{3},\frac{m(r)^{2}+k}{3}\right) =(m(r1)2+i3,m(r1)2+j3,m(r1)2+k3)\displaystyle=\left(\frac{m(r-1)^{2}+i}{3},\frac{m(r-1)^{2}+j}{3},\frac{m(r-1)^{2}+k}{3}\right)
+H(m(r1),m(r1),m(r1))\displaystyle+_{H}(m(r-1),m(r-1),m(r-1))
+H(m(r),m(r),m(r)),\displaystyle+_{H}(m(r),m(r),m(r)),

where (i,j,k){(5,1,4),(5,5,10),(5,2,7)}.(i,j,k)\in\{(5,-1,-4),(5,5,-10),(5,2,-7)\}. Then, by the inductive hypothesis, Lemma 4.13 and semigroup property, we have g(\scalerelgXm(r),\scalerelgXm(r),(m(r)2+i3,m(r)2+j3,m(r)2+k3))>0g\left(\mathord{\scalerel*{\Box}{gX}}_{m(r)},\mathord{\scalerel*{\Box}{gX}}_{m(r)},\left(\frac{m(r)^{2}+i}{3},\frac{m(r)^{2}+j}{3},\frac{m(r)^{2}+k}{3}\right)\right)>0, where (i,j,k){(5,1,4),(5,5,10),(5,2,7)}(i,j,k)\in\{(5,-1,-4),(5,5,-10),(5,2,-7)\}, for any positive integer m7m\geq 7 such that m1mod3m\equiv 1\mod 3

Lemma 4.16.

The Kronecker coefficient g(\scalerelgXm,\scalerelgXm,(m2+33,m23,m233))>0g\left(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\left(\frac{m^{2}+3}{3},\frac{m^{2}}{3},\frac{m^{2}-3}{3}\right)\right)>0 and
g(\scalerelgXm,\scalerelgXm,(m2+33,m2+33,m263))g\left(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\left(\frac{m^{2}+3}{3},\frac{m^{2}+3}{3},\frac{m^{2}-6}{3}\right)\right) for any positive integer m6m\geq 6 such that m0mod3m\equiv 0\mod 3.

Proof.

For any positive integer m6m\geq 6 such that m0mod3m\equiv 0\mod 3, we can write m(r)=3rm(r)=3r for some r2r\geq 2. We will prove the proposition by induction on rr. When r=2r=2, m(r)=6m(r)=6, and with the help of the computer, we can verify the statement holds true for r=2.r=2. Let r3r\geq 3, and assume that the statement is true for r1r-1. We can decompose \scalerelgXm(r)\mathord{\scalerel*{\Box}{gX}}_{m}(r) as

\scalerelgXm(r)=(\scalerelgXm(r1)+V(m(r1)3))+H3m(r),\mathord{\scalerel*{\Box}{gX}}_{m}(r)=\left(\mathord{\scalerel*{\Box}{gX}}_{m(r-1)}+_{V}(m(r-1)^{3})\right)+_{H}3^{m(r)},

and we can decompose the partition (m(r)2+33,m(r)23,m(r)233)\left(\frac{m(r)^{2}+3}{3},\frac{m(r)^{2}}{3},\frac{m(r)^{2}-3}{3}\right) as

(m(r)2+33,m(r)23,m(r)233)\displaystyle\left(\frac{m(r)^{2}+3}{3},\frac{m(r)^{2}}{3},\frac{m(r)^{2}-3}{3}\right) =(m(r1)2+33,m(r1)23,m(r1)233)\displaystyle=\left(\frac{m(r-1)^{2}+3}{3},\frac{m(r-1)^{2}}{3},\frac{m(r-1)^{2}-3}{3}\right)
+H(m(r1),m(r1),m(r1))\displaystyle+_{H}(m(r-1),m(r-1),m(r-1))
+H(m(r),m(r),m(r)).\displaystyle+_{H}(m(r),m(r),m(r)).

Then, by the inductive hypothesis, Lemma 4.13 and semigroup property, we can conclude that g(\scalerelgXm(r),\scalerelgXm(r),(m(r)2+33,m(r)23,m(r)233))>0.g\left(\mathord{\scalerel*{\Box}{gX}}_{m(r)},\mathord{\scalerel*{\Box}{gX}}_{m(r)},\left(\frac{m(r)^{2}+3}{3},\frac{m(r)^{2}}{3},\frac{m(r)^{2}-3}{3}\right)\right)>0. By a completely analogous argument, we can show that g(\scalerelgXm,\scalerelgXm,(m2+33,m2+33,m263))g\left(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\left(\frac{m^{2}+3}{3},\frac{m^{2}+3}{3},\frac{m^{2}-6}{3}\right)\right) for any positive integer m6m\geq 6 such that m0mod3m\equiv 0\mod 3

When mm is even, we decompose \scalerelgXm\mathord{\scalerel*{\Box}{gX}}_{m} as (\scalerelgXm2+V(m,m))+H(2m)(\mathord{\scalerel*{\Box}{gX}}_{m-2}+_{V}(m,m))+_{H}(2^{m}). By analyzing various cases based on the values and parities of λ1,λ2,λ3\lambda_{1},\lambda_{2},\lambda_{3}, we are able to prove the following result:

Proposition 4.17.

For every even number m6m\geq 6, g(\scalerelgXm,\scalerelgXm,λ)>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\lambda)>0 for any three-row partition λm2\lambda\vdash m^{2} with λ32\lambda_{3}\geq 2.

Proof.

If λm2\lambda\vdash m^{2} is a three-row rectangular partition, then 3m3\mid m. By Corollary 4.12.1, we conclude that g(\scalerelgXm,\scalerelgXm,λ)>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\lambda)>0. Now we assume that λ\lambda is not a rectangular partition.

For any even integer m6m\geq 6, we can write m=2rm=2r where r3r\geq 3. We will prove this statement by induction on rr. First, consider the base case r=3r=3. In this case, m=6m=6, and we can check that g(\scalerelgX6,\scalerelgX6,λ)>0g(\mathord{\scalerel*{\Box}{gX}}_{6},\mathord{\scalerel*{\Box}{gX}}_{6},\lambda)>0 for every three-row partition λ36\lambda\vdash 36 with λ32\lambda_{3}\geq 2 with the help of computer.

Next, let r4r\geq 4 and assume the statement holds for r1r-1. We will prove it for rr. By the inductive hypothesis, we assume that g(\scalerelgX2(r1),\scalerelgX2(r1),λ)>0g(\mathord{\scalerel*{\Box}{gX}}_{2(r-1)},\mathord{\scalerel*{\Box}{gX}}_{2(r-1)},\lambda)>0 for any three-row partition λ4(r1)2\lambda\vdash 4(r-1)^{2} with λ32\lambda_{3}\geq 2. Note that we can decompose \scalerelgX2r\mathord{\scalerel*{\Box}{gX}}_{2r} as

\scalerelgX2r=(\scalerelgX2(r1)+V(2r2,2r2))+H(22r).\mathord{\scalerel*{\Box}{gX}}_{2r}=(\mathord{\scalerel*{\Box}{gX}}_{2(r-1)}+_{V}(2r-2,2r-2))+_{H}(2^{2r}).

By Theorem 4.3,

g((22r),(22r),(2a,2b,2(2rab)))=g((2r,2r),(2r,2r),(2a,2b,2(2rab)))>0g((2^{2r}),(2^{2r}),(2a,2b,2(2r-a-b)))=g((2r,2r),(2r,2r),(2a,2b,2(2r-a-b)))>0

for all integers a,ba,b satisfying that 02rabba2r0\leq 2r-a-b\leq b\leq a\leq 2r, and

g((2r2,2r2),(2r2,2r2),(2x,2y,2(2r2xy)))>0g((2r-2,2r-2),(2r-2,2r-2),(2x,2y,2(2r-2-x-y)))>0

for all integers x,yx,y such that 02r2xyyx2r2.0\leq 2r-2-x-y\leq y\leq x\leq 2r-2.

Let τ:=(2u,2v,(8r4)2u2v)\tau:=(2u,2v,(8r-4)-2u-2v) be a non-rectangular three-row partition of 8r48r-4 with all parts even. Then, we can write (2u,2v,2w)(2u,2v,2w) as a horizontal sum of partitions (2a,2b,2(2rab))4r(2a,2b,2(2r-a-b))\vdash 4r and (2x,2y,2(2r2xy))4r4(2x,2y,2(2r-2-x-y))\vdash 4r-4, where a=u2a=\lceil\frac{u}{2}\rceil, b=v2b=\lceil\frac{v}{2}\rceil, x=u2x=\lfloor\frac{u}{2}\rfloor and y=v2y=\lfloor\frac{v}{2}\rfloor. Hence, it suffices to show that we can rewrite a non-rectangular three-row partition of m2m^{2} as a horizontal sum of a three-row partition of (m2)2(m-2)^{2} appearing in the tensor square of \scalerelgXm2\mathord{\scalerel*{\Box}{gX}}_{m-2} and a non-rectangular three-row partition τ8r4\tau\vdash 8r-4 whose parts are all even. We will consider the following cases for the partition λ=(λ1,λ2,λ3)\lambda=(\lambda_{1},\lambda_{2},\lambda_{3}) with λ32\lambda_{3}\geq 2.

  • Case 1:

    λ2λ34r2\lambda_{2}-\lambda_{3}\geq 4r-2. In this case, we can write λ\lambda as a horizontal sum of (4r2,4r2)(4r-2,4r-2) and a partition (λ14r+2,λ24r+2,λ3)(\lambda_{1}-4r+2,\lambda_{2}-4r+2,\lambda_{3}).

  • Case 2:

    λ2λ3<4r2\lambda_{2}-\lambda_{3}<4r-2 and λ1λ28r44λ2λ32\lambda_{1}-\lambda_{2}\geq 8r-4-4\lfloor\frac{\lambda_{2}-\lambda_{3}}{2}\rfloor. If these conditions hold, we can define τ=(8r42λ2λ32,2λ2λ32)8r4\tau=(8r-4-2\lfloor\frac{\lambda_{2}-\lambda_{3}}{2}\rfloor,2\lfloor\frac{\lambda_{2}-\lambda_{3}}{2}\rfloor)\vdash 8r-4. Then, we can write λ\lambda as a horizontal sum of τ\tau and a three-row partition of (m2)2(m-2)^{2}.

  • Case 3:

    λ2λ3<4r2\lambda_{2}-\lambda_{3}<4r-2 and λ1λ2<8r44λ2λ32\lambda_{1}-\lambda_{2}<8r-4-4\lfloor\frac{\lambda_{2}-\lambda_{3}}{2}\rfloor. In this case, we observe that 2(λ2λ3)+(λ1λ2)<8r42(\lambda_{2}-\lambda_{3})+(\lambda_{1}-\lambda_{2})<8r-4 if λ2λ3\lambda_{2}-\lambda_{3} is even, and 2(λ2λ3)+(λ1λ2)<8r22(\lambda_{2}-\lambda_{3})+(\lambda_{1}-\lambda_{2})<8r-2 if λ2λ3\lambda_{2}-\lambda_{3} is odd. Therefore, we can conclude that λ3(m2)23\lambda_{3}\geq\lfloor\frac{(m-2)^{2}}{3}\rfloor under the given conditions. We further consider the following subcases:

    1. (1)

      If 3(m2)23\mid(m-2)^{2}, then we can write (m2)2=3k(m-2)^{2}=3k for some kk even.

      1. (a)

        If λ\lambda has all parts even, then consider τ=(λ1k,λ2k,λ3k)\tau=\left(\lambda_{1}-k,\lambda_{2}-k,\lambda_{3}-k\right).

      2. (b)

        If the parities of λ1,λ2,λ3\lambda_{1},\lambda_{2},\lambda_{3} are odd, odd, even, respectively, and λ1>λ2\lambda_{1}>\lambda_{2},
        consider τ=(λ1k1,λ2k1,λ3k+2)\tau=\left(\lambda_{1}-k-1,\lambda_{2}-k-1,\lambda_{3}-k+2\right).

      3. (c)

        If the parities of λ1,λ2,λ3\lambda_{1},\lambda_{2},\lambda_{3} are odd, odd, even, respectively, and λ1=λ2\lambda_{1}=\lambda_{2}, then we must have λ2λ35\lambda_{2}-\lambda_{3}\geq 5 as otherwise m2m^{2} or m2+1m^{2}+1 is a multiple of 33, which is impossible. Consider τ=(λ1k1,λ2k1,λ3k+2)\tau=\left(\lambda_{1}-k-1,\lambda_{2}-k-1,\lambda_{3}-k+2\right).

      4. (d)

        If the parities of λ1,λ2,λ3\lambda_{1},\lambda_{2},\lambda_{3} are odd, even, odd, respectively,
        consider τ=(λ1k1,λ2k,λ3k+1)\tau=\left(\lambda_{1}-k-1,\lambda_{2}-k,\lambda_{3}-k+1\right).

      5. (e)

        If the parities of λ1,λ2,λ3\lambda_{1},\lambda_{2},\lambda_{3} are even, odd, odd, respectively, and λ2>λ3\lambda_{2}>\lambda_{3},
        consider τ=(λ1k2,λ2k+1,λ3k+1)\tau=\left(\lambda_{1}-k-2,\lambda_{2}-k+1,\lambda_{3}-k+1\right).

      6. (f)

        If the parities of λ1,λ2,λ3\lambda_{1},\lambda_{2},\lambda_{3} are even, odd, odd, respectively, λ2=λ3\lambda_{2}=\lambda_{3}, and λ1λ25\lambda_{1}-\lambda_{2}\geq 5, we consider τ=(λ1k2,λ2k+1,λ3k+1)\tau=\left(\lambda_{1}-k-2,\lambda_{2}-k+1,\lambda_{3}-k+1\right). (Note that λ1λ23\lambda_{1}-\lambda_{2}\neq 3 as m2mod3m\equiv 2\mod 3.

      7. (g)

        If the parities of λ1,λ2,λ3\lambda_{1},\lambda_{2},\lambda_{3} are even, odd, odd, respectively, λ2=λ3\lambda_{2}=\lambda_{3}, and λ1λ2=1\lambda_{1}-\lambda_{2}=1, then by Lemma 4.14, we can prove the positivity of
        g(\scalerelgXm,\scalerelgXm,(m2+23,m213,m213))g\left(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\left(\frac{m^{2}+2}{3},\frac{m^{2}-1}{3},\frac{m^{2}-1}{3}\right)\right).

    2. (2)

      If (m2)21mod3(m-2)^{2}\equiv 1\bmod 3, then we can write (m2)2=3k+1(m-2)^{2}=3k+1 for some odd integer kk.

      1. (a)

        If λ\lambda has all parts even, consider τ=(λ1k1,λ2k1,λ3k+1)\tau=\left(\lambda_{1}-k-1,\lambda_{2}-k-1,\lambda_{3}-k+1\right).

      2. (b)

        If the parities of λ1,λ2,λ3\lambda_{1},\lambda_{2},\lambda_{3} are odd, odd, even, respectively, λ1λ2>2\lambda_{1}-\lambda_{2}>2 or λ2λ3>1\lambda_{2}-\lambda_{3}>1 ,
        consider τ=(λ1k2,λ2k,λ3k+1)\tau=\left(\lambda_{1}-k-2,\lambda_{2}-k,\lambda_{3}-k+1\right).

      3. (c)

        If the parities of λ1,λ2,λ3\lambda_{1},\lambda_{2},\lambda_{3} are odd, odd, even, respectively, λ1=λ2+2=λ3+3\lambda_{1}=\lambda_{2}+2=\lambda_{3}+3, then m1mod3m\equiv 1\mod 3. By Lemma 4.15, we can obtain the positivity of g(\scalerelgXm,\scalerelgXm,(m2+53,m213,m243))g\left(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\left(\frac{m^{2}+5}{3},\frac{m^{2}-1}{3},\frac{m^{2}-4}{3}\right)\right).

      4. (d)

        If the parities of λ1,λ2,λ3\lambda_{1},\lambda_{2},\lambda_{3} are odd, odd, even, respectively, λ1=λ2\lambda_{1}=\lambda_{2} and λ2λ3=3\lambda_{2}-\lambda_{3}=3, then 3m3\mid m. By Lemma 4.16, we can obtain the positivity of g(\scalerelgXm,\scalerelgXm,(m2+33,m2+33,m263))g\left(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\left(\frac{m^{2}+3}{3},\frac{m^{2}+3}{3},\frac{m^{2}-6}{3}\right)\right) by semigroup property.

      5. (e)

        If the parities of λ1,λ2,λ3\lambda_{1},\lambda_{2},\lambda_{3} are odd, odd, even, respectively, λ1=λ2\lambda_{1}=\lambda_{2} and λ2λ3=5\lambda_{2}-\lambda_{3}=5, then m1mod3m\equiv 1\mod 3. By Lemma 4.15, we can obtain the positivity of g(\scalerelgXm,\scalerelgXm,(m2+53,m2+53,m2103))g\left(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\left(\frac{m^{2}+5}{3},\frac{m^{2}+5}{3},\frac{m^{2}-10}{3}\right)\right).

      6. (f)

        If the parities of λ1,λ2,λ3\lambda_{1},\lambda_{2},\lambda_{3} are odd, odd, even, respectively, λ1=λ2\lambda_{1}=\lambda_{2} and λ2λ37\lambda_{2}-\lambda_{3}\geq 7, consider τ=(λ1k2,λ2k2,λ3k+3)\tau=\left(\lambda_{1}-k-2,\lambda_{2}-k-2,\lambda_{3}-k+3\right).

      7. (g)

        If the parities of λ1,λ2,λ3\lambda_{1},\lambda_{2},\lambda_{3} are odd, even, odd, respectively, and λ2λ3>3\lambda_{2}-\lambda_{3}>3 or λ1λ2>1\lambda_{1}-\lambda_{2}>1, consider τ=(λ1k2,λ2k1,λ3k+2)\tau=\left(\lambda_{1}-k-2,\lambda_{2}-k-1,\lambda_{3}-k+2\right).

      8. (h)

        If the parities of λ1,λ2,λ3\lambda_{1},\lambda_{2},\lambda_{3} are odd, even, odd, respectively, and λ1=λ2+1=λ3+4\lambda_{1}=\lambda_{2}+1=\lambda_{3}+4, then m1mod3m\equiv 1\mod 3. By Lemma 4.15, we can obtain the positivity of g(\scalerelgXm,\scalerelgXm,(m2+53,m2+23,m273))g\left(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\left(\frac{m^{2}+5}{3},\frac{m^{2}+2}{3},\frac{m^{2}-7}{3}\right)\right).

      9. (i)

        If the parities of λ1,λ2,λ3\lambda_{1},\lambda_{2},\lambda_{3} are odd, even, odd, respectively, λ2λ3=1\lambda_{2}-\lambda_{3}=1 and λ1λ2<5\lambda_{1}-\lambda_{2}<5, then λ1=λ2+1\lambda_{1}=\lambda_{2}+1. Note that if λ1=λ2+3\lambda_{1}=\lambda_{2}+3, it implies that m22mod3m^{2}\equiv 2\bmod 3, which is impossible. Thus, 3m23\mid m^{2}, and we can obtain the positivity of g(\scalerelgXm,\scalerelgXm,(m2+33,m23,m233))g\left(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\left(\frac{m^{2}+3}{3},\frac{m^{2}}{3},\frac{m^{2}-3}{3}\right)\right) by Lemma 4.16.

      10. (j)

        If the parities of λ1,λ2,λ3\lambda_{1},\lambda_{2},\lambda_{3} are even, odd, odd, respectively and λ2>λ3\lambda_{2}>\lambda_{3}, consider τ=(λ1k1,λ2k,λ3k)\tau=\left(\lambda_{1}-k-1,\lambda_{2}-k,\lambda_{3}-k\right).

      11. (k)

        If the parities of λ1,λ2,λ3\lambda_{1},\lambda_{2},\lambda_{3} are even, odd, odd, respectively, λ2=λ3\lambda_{2}=\lambda_{3} and λ1λ29\lambda_{1}-\lambda_{2}\geq 9, consider τ=(λ1k5,λ2k+2,λ3k+2)\tau=\left(\lambda_{1}-k-5,\lambda_{2}-k+2,\lambda_{3}-k+2\right).

      12. (l)

        If the parities of λ1,λ2,λ3\lambda_{1},\lambda_{2},\lambda_{3} are even, odd, odd, respectively, λ2=λ3\lambda_{2}=\lambda_{3} and λ1λ2{1,7}\lambda_{1}-\lambda_{2}\in\{1,7\}, then m1mod3m\equiv 1\mod 3 and we can prove the positivity of g(\scalerelgXm,\scalerelgXm,λ)g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\lambda) by a similar argument as in the proof of Lemma 4.15.

      13. (m)

        If the parities of λ1,λ2,λ3\lambda_{1},\lambda_{2},\lambda_{3} are even, odd, odd, respectively, λ2=λ3\lambda_{2}=\lambda_{3} and λ1λ2=3\lambda_{1}-\lambda_{2}=3, then m0mod3m\equiv 0\mod 3. We can show the positivity of g(\scalerelgXm,\scalerelgXm,λ)g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\lambda) by a similar argument as in the proof of Lemma 4.16.

For each of the cases above, τ\tau is a non-rectangular three-row partition of 8r48r-4 with all parts even, and we can write λ\lambda as a horizontal sum of τ\tau and a three-row partition with a long third-row of (m2)2(m-2)^{2}. Then, by the semigroup property and the inductive hypothesis, we can conclude that the statement holds true for rr. By induction, we therefore know that for m6m\geq 6, g(\scalerelgXm,\scalerelgXm,λ)>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\lambda)>0 for any three-row partition λm2\lambda\vdash m^{2} with λ32\lambda_{3}\geq 2. ∎

Next, we will prove the positivity of g(\scalerelgXm,\scalerelgXm,μ)g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\mu) when mm is an odd integer.

Proposition 4.18.

For every odd integer m5m\geq 5, g(\scalerelgXm,\scalerelgXm,μ)>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\mu)>0 for any three-row partition λm2\lambda\vdash m^{2} with μ32m1\mu_{3}\geq 2m-1.

Proof.

Let m5m\geq 5 be an odd integer. With the help of the computer, we can verify the statement when m{5,7}.m\in\{5,7\}. Now consider the case where m9.m\geq 9. Note that we can decompose \scalerelgXm\mathord{\scalerel*{\Box}{gX}}_{m} as

\scalerelgXm=(\scalerelgXm3+V(m3)3)+H3m,\mathord{\scalerel*{\Box}{gX}}_{m}=(\mathord{\scalerel*{\Box}{gX}}_{m-3}+_{V}(m-3)^{3})+_{H}3^{m},

and μ\mu as

μ=(m3)3+H(m3)+H(μ12m+3,μ22m+3,μ32m+3).\mu=(m-3)^{3}+_{H}(m^{3})+_{H}(\mu_{1}-2m+3,\mu_{2}-2m+3,\mu_{3}-2m+3).

Notice that g(\scalerelgXm3,\scalerelgXm3,(μ12m+3,μ22m+3,μ32m+3))>0g(\mathord{\scalerel*{\Box}{gX}}_{m-3},\mathord{\scalerel*{\Box}{gX}}_{m-3},(\mu_{1}-2m+3,\mu_{2}-2m+3,\mu_{3}-2m+3))>0 by Theorem 4.17, g((m3)3,(m3)3,(m3)3)>0g((m-3)^{3},(m-3)^{3},(m-3)^{3})>0 and g(3m,3m,m3)>0g(3^{m},3^{m},m^{3})>0 by Lemma 4.13. Then by semigroup property, we have g(\scalerelgXm,\scalerelgXm,μ)>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\mu)>0 for every three-row partition λm2\lambda\vdash m^{2} with μ32m1\mu_{3}\geq 2m-1. ∎

Proposition 4.19.

For any odd integer m5m\geq 5, g(\scalerelgXm,\scalerelgXm,μ)>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\mu)>0 for any three-row partition μm2\mu\vdash m^{2} with 2μ32m22\leq\mu_{3}\leq 2m-2.

Proof.

We can verify that the statement holds for m{5,7,9}m\in\{5,7,9\} by semigroup property, together with the help of the computer.

Let m5m\geq 5. Notice that when μ1μ22m1\mu_{1}-\mu_{2}\geq 2m-1, we can decompose \scalerelgXm\mathord{\scalerel*{\Box}{gX}}_{m} as \scalerelgXm=(\scalerelgXm1+V(m1))+H1m\mathord{\scalerel*{\Box}{gX}}_{m}=(\mathord{\scalerel*{\Box}{gX}}_{m-1}+_{V}(m-1))+_{H}1^{m} and μ\mu as μ=(m)+H(m1)+H(μ12m+1,μ2,μ3).\mu=(m)+_{H}(m-1)+_{H}(\mu_{1}-2m+1,\mu_{2},\mu_{3}). By Theorem 4.17 and the semigroup property, we conclude that g(\scalerelgXm,\scalerelgXm,μ)>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\mu)>0 for any three-row partition μm2\mu\vdash m^{2} with 2μ32m22\leq\mu_{3}\leq 2m-2 and μ1μ22m1\mu_{1}-\mu_{2}\geq 2m-1.

We shall prove the statement by induction. Let m11m\geq 11 be an odd integer. Suppose the statement is true for any odd integer less than mm. Let μm2\mu\vdash m^{2} be a three-row partition such that μ1μ22m2\mu_{1}-\mu_{2}\leq 2m-2 and 2μ32m22\leq\mu_{3}\leq 2m-2.

  1. Case 1:

    If μ1μ2{0,1}\mu_{1}-\mu_{2}\in\{0,1\}, consider the decomposition \scalerelgXm=(\scalerelgXm4+V(m4)4)+H4m\mathord{\scalerel*{\Box}{gX}}_{m}=(\mathord{\scalerel*{\Box}{gX}}_{m-4}+_{V}(m-4)^{4})+_{H}4^{m}. If (m4)23μ3(m-4)^{2}\geq 3\mu_{3}, we can decompose μ\mu as

    μ=μ1+H(2m,2m)+H(2m8,2m8),\mu=\mu^{1}+_{H}(2m,2m)+_{H}(2m-8,2m-8),

    where μ1:=((m4)2μ32,(m4)2μ32,μ3).\mu^{1}:=\left(\left\lceil\frac{(m-4)^{2}-\mu_{3}}{2}\right\rceil,\left\lfloor\frac{(m-4)^{2}-\mu_{3}}{2}\right\rfloor,\mu_{3}\right). Otherwise, we know that μ316\mu_{3}\geq 16, and we can decompose μ\mu as

    μ=μ2+H(2m2,2m2,4)+H(2m8,2m8),\mu=\mu^{2}+_{H}(2m-2,2m-2,4)+_{H}(2m-8,2m-8),

    where μ2:=((m4)2μ32+2,(m4)2μ32+2,μ34)\mu^{2}:=\left(\left\lceil\frac{(m-4)^{2}-\mu_{3}}{2}\right\rceil+2,\left\lfloor\frac{(m-4)^{2}-\mu_{3}}{2}\right\rfloor+2,\mu_{3}-4\right). Then we have the following:

    • By inductive hypothesis, we have g(\scalerelgXm4,\scalerelgXm4,μ1)>0g\left(\mathord{\scalerel*{\Box}{gX}}_{m-4},\mathord{\scalerel*{\Box}{gX}}_{m-4},\mu^{1}\right)>0 and g(\scalerelgXm4,\scalerelgXm4,μ2)>0.g\left(\mathord{\scalerel*{\Box}{gX}}_{m-4},\mathord{\scalerel*{\Box}{gX}}_{m-4},\mu^{2}\right)>0.

    • By semigroup property, Theorem 4.3 and the fact that g((3,3,3,3),(3,3,3,3),(6,6))>0g((3,3,3,3),(3,3,3,3),(6,6))>0, we have g((m4)4,(m4)4,(2m8,2m8))>0g((m-4)^{4},(m-4)^{4},(2m-8,2m-8))>0, g(4m,4m,(2m,2m))>0g(4^{m},4^{m},(2m,2m))>0, and g(4m,4m,(2m2,2m2,4))>0.g(4^{m},4^{m},(2m-2,2m-2,4))>0.

    Hence, by semigroup property, we can conclude that g(\scalerelgXm,\scalerelgXm,μ)>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\mu)>0 when μ1μ2{0,1}.\mu_{1}-\mu_{2}\in\{0,1\}.

  2. Case 2:

    If μ1μ2{2,3}\mu_{1}-\mu_{2}\in\{2,3\}, consider the decomposition \scalerelgXm=(\scalerelgXm4+V(m4)4)+H4m\mathord{\scalerel*{\Box}{gX}}_{m}=(\mathord{\scalerel*{\Box}{gX}}_{m-4}+_{V}(m-4)^{4})+_{H}4^{m}. If (m4)23(μ32)(m-4)^{2}\geq 3(\mu_{3}-2), we can decompose μ\mu as

    μ=μ1+H(2m,2m2,2)+H(2m8,2m8),\mu=\mu^{1}+_{H}(2m,2m-2,2)+_{H}(2m-8,2m-8),

    where μ1:=((m4)2μ32+1,(m4)2μ32+1,μ32)\mu^{1}:=\left(\left\lceil\frac{(m-4)^{2}-\mu_{3}}{2}\right\rceil+1,\left\lfloor\frac{(m-4)^{2}-\mu_{3}}{2}\right\rfloor+1,\mu_{3}-2\right). Otherwise, it implies that μ318\mu_{3}\geq 18 and we can decompose μ\mu as

    μ=μ2+H(2m,2m2,2)+H(2m10,2m10,4),\mu=\mu^{2}+_{H}(2m,2m-2,2)+_{H}(2m-10,2m-10,4),

    where μ2:=((m4)2μ32+3,(m4)2μ32+3,μ36).\mu^{2}:=\left(\left\lceil\frac{(m-4)^{2}-\mu_{3}}{2}\right\rceil+3,\left\lfloor\frac{(m-4)^{2}-\mu_{3}}{2}\right\rfloor+3,\mu_{3}-6\right). Then we have the following:

    • By inductive hypothesis and Theorem 4.2, we have g(\scalerelgXm4,\scalerelgXm4,μ1)>0g\left(\mathord{\scalerel*{\Box}{gX}}_{m-4},\mathord{\scalerel*{\Box}{gX}}_{m-4},\mu^{1}\right)>0 and g(\scalerelgXm4,\scalerelgXm4,μ2)>0.g\left(\mathord{\scalerel*{\Box}{gX}}_{m-4},\mathord{\scalerel*{\Box}{gX}}_{m-4},\mu^{2}\right)>0.

    • By semigroup property and Theorem 4.3, we have g((m4)4,(m4)4,(2m8,2m8))>0g((m-4)^{4},(m-4)^{4},(2m-8,2m-8))>0, g((m4)4,(m4)4,(2m10,2m10,4))>0g((m-4)^{4},(m-4)^{4},(2m-10,2m-10,4))>0, and g(4m,4m,(2m,2m2,2))>0g(4^{m},4^{m},(2m,2m-2,2))>0.

    Hence by semigroup property, we can conclude that g(\scalerelgXm,\scalerelgXm,μ)>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\mu)>0 when μ1μ2{2,3}\mu_{1}-\mu_{2}\in\{2,3\}.

  3. Case 3:

    If a:=μ1μ24a:=\mu_{1}-\mu_{2}\geq 4, consider the decomposition \scalerelgXm=(\scalerelgXm2+V(m2,m2))+H2m\mathord{\scalerel*{\Box}{gX}}_{m}=(\mathord{\scalerel*{\Box}{gX}}_{m-2}+_{V}(m-2,m-2))+_{H}2^{m} and we will decompose μ\mu as

    μ\displaystyle\mu =((m2)2μ32+δ(a),(m2)2μ32δ(a),μ3)\displaystyle=\left(\left\lceil\frac{(m-2)^{2}-\mu_{3}}{2}\right\rceil+\delta(a),\left\lfloor\frac{(m-2)^{2}-\mu_{3}}{2}\right\rfloor-\delta(a),\mu_{3}\right)
    +H(m+1+2x,m12x)\displaystyle\quad+_{H}(m+1+2x,m-1-2x)
    +H(m1+2y,m32y),\displaystyle\quad+_{H}(m-1+2y,m-3-2y),

    where δ(a):={0if a0 or 1mod41if a2 or 3mod4\delta(a):=\begin{cases}0&\text{if }a\equiv 0\text{ or }1\mod 4\\ 1&\text{if }a\equiv 2\text{ or }3\mod 4\end{cases} and x,yx,y are non-negative integers such that 4(x+y+1)=4a44(x+y+1)=4\left\lfloor\frac{a}{4}\right\rfloor. Then we have the following:

    • By inductive hypothesis,

      g(\scalerelgXm2,\scalerelgXm2,((m2)2μ32+δ(a),(m2)2μ32δ(a),μ3))>0.g\left(\mathord{\scalerel*{\Box}{gX}}_{m-2},\mathord{\scalerel*{\Box}{gX}}_{m-2},\left(\left\lceil\frac{(m-2)^{2}-\mu_{3}}{2}\right\rceil+\delta(a),\left\lfloor\frac{(m-2)^{2}-\mu_{3}}{2}\right\rfloor-\delta(a),\mu_{3}\right)\right)>0.
    • By Theorem 4.3, we have g((2m2,2m2),(2m2,2m2),(m1+2y,m32y))>0g((2m-2,2m-2),(2m-2,2m-2),(m-1+2y,m-3-2y))>0 and g(2m,2m,(m+1+2x,m12x))>0.g(2^{m},2^{m},(m+1+2x,m-1-2x))>0.

    By semigroup property, we can conclude that g(\scalerelgXm,\scalerelgXm,μ)>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\mu)>0 when μ1μ24.\mu_{1}-\mu_{2}\geq 4.

Hence, by induction, for any odd integer m5m\geq 5, we have g(\scalerelgXm,\scalerelgXm,μ)>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\mu)>0 for any three-row partition μm2\mu\vdash m^{2} with 2μ32m22\leq\mu_{3}\leq 2m-2. ∎

We now have all the ingredients to prove our main theorem.

proof of Theorem 4.11.

With the help of computer and semigroup property, we check that for m{7,9,11,13,15,17}m\in\{7,9,11,13,15,17\}, g(\scalerelgXm,\scalerelgXm,μ)>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\mu)>0 for any three-row partition μm2\mu\vdash m^{2} with 2μ32m22\leq\mu_{3}\leq 2m-2 and μ1μ22m2\mu_{1}-\mu_{2}\leq 2m-2, as shown in the appendix. The result then follows from Proposition 4.17, Proposition 4.18 and Proposition 4.19. ∎

5. Constituency of near-hooks (m2ki,i,1k)(m^{2}-k-i,i,1^{k})

In this section, we will discuss sufficient conditions for near-hooks to be constituents in tensor squares of square partitions

In their work [IP17], Ikenmeyer and Panova employed induction and the semigroup property to demonstrate the constituency of near-hooks with a second row of at most 66 in the tensor square of a rectangle with large side lengths.

Theorem 5.1 ([IP17] Corollary 4.6).

Fix wh7w\geq h\geq 7. We have that g(λ,h×w,h×w)>0g(\lambda,h\times w,h\times w)>0 for all λ=(hwj|ρ|,1j+Hρ)\lambda=(hw-j-|\rho|,1^{j}+_{H}\rho) with ρ\rho\neq\emptyset and |ρ|6|\rho|\leq 6 for all j[1,h2Rρ]j\in[1,h^{2}-R_{\rho}] where Rρ=|ρ|+ρ1+1R_{\rho}=|\rho|+\rho_{1}+1, except in the following cases: λ{(hw3,2,1),(hwh2+3,2,1h25),(hw4,3,1),(hwh2+3,2,2,1h27)}.\lambda\in\{(hw-3,2,1),(hw-h^{2}+3,2,1^{h^{2}-5}),(hw-4,3,1),(hw-h^{2}+3,2,2,1^{h^{2}-7})\}.

The positivity of certain classes of near-hooks can be directly derived from Theorem 5.1.

Corollary 5.1.1.

Let m7m\geq 7. For all μi(k,m)=(m2ki,i,1k)\mu_{i}(k,m)=(m^{2}-k-i,i,1^{k}) with i[2,7]i\in[2,7] and k[k\in[, g(\scalerelgXm,\scalerelgXm,μi(k,m))>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\mu_{i}(k,m))>0 except in the following cases: (1) i=2i=2 with k=1k=1 or k=m25k=m^{2}-5, (2) i=3i=3 and k=1k=1.

We present an alternative proof approach for finding sufficient conditions for two classes of near-hooks to be constituents in tensor squares of square partitions using a tool aimed specifically at the Saxl conjecture was developed in [PPV16] as follows:

Theorem 5.2 ([PPV16] Main Lemma).

Let μ=μ\mu=\mu^{\prime} be a self-conjugate partition of nn, and let ν=(2μ11,2μ23,2μ35,)n\nu=(2\mu_{1}-1,2\mu_{2}-3,2\mu_{3}-5,\dots)\vdash n be the partition whose parts are lengths of the principal hooks of μ\mu. Suppose χλ[ν]0\chi^{\lambda}[\nu]\neq 0 for some λn\lambda\vdash n. Then χλ\chi^{\lambda} is a constituent of χμχμ.\chi^{\mu}\bigotimes\chi^{\mu}.

Let μi(k,m):=(m2ki,i,1k)\mu_{i}(k,m):=(m^{2}-k-i,i,1^{k}) and αm=(2m1,2m3,,1).\alpha_{m}=(2m-1,2m-3,\dots,1). By Theorem 5.2, that |χμ(k,m)(αm)|0|\chi^{\mu(k,m)}(\alpha_{m})|\neq 0 would imply g(\scalerelgXm,\scalerelgXm,μi(k,m))>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\mu_{i}(k,m))>0. In particular, we will discuss the number of rim-hook tableaux of shape μ2(k,m)=(m2k2,2,1k)\mu_{2}(k,m)=(m^{2}-k-2,2,1^{k}) and weight αm\alpha_{m}.

To use the Murnaghan-Nakayama rule to compute the characters, we consider the construction of an arbitrary rim-hook tableau of shape μ2(k,m)\mu_{2}(k,m) and weight (1,3,,2m1).(1,3,\dots,2m-1). Observe that the 11-hook can only be placed at the upper left corner, and there are three ways to place the 33-hook, as illustrated in the following diagrams:

\ytableausetup

smalltableaux \ytableaushort13\none,33, \none * 10,2,1,1,1,1,1,1           \ytableaushort1333\none,\none,\none * 10,2,1,1,1,1,1,1           \ytableaushort1\none,3\none,3\none,3\none * 10,2,1,1,1,1,1,1

Let PR(m)(k)P_{R(m)}(k) denote the number of partitions of kk whose parts are distinct odd integers from the set

R(m)={5,7,,2m1}.{R(m)}=\{5,7,\dots,2m-1\}.

Observing the diagrams above, we can deduce that the height of any rim-hook tableau with the shape μ2(k,m)\mu_{2}(k,m) and weight (1,3,,2m1)(1,3,\dots,2m-1) is always an odd number. From left to right, the quantities of rim-hook tableaux corresponding to the three diagrams are PR(m)(k)P_{R(m)}(k), PR(m)(k+2)P_{R(m)}(k+2), and PR(m)(k2)P_{R(m)}(k-2), respectively. Thus, by Murnaghan-Nakayama rule, we therefore have

χ(nk2,2,1k)(2m1,2m3,,3,1)=PR(m)(k)PR(m)(k+2)PR(m)(k2).\chi^{(n-k-2,2,1^{k})}(2m-1,2m-3,\dots,3,1)=-P_{R(m)}(k)-P_{R(m)}(k+2)-P_{R(m)}(k-2).

Thus, g(λm,λm,(nk2,2,1k))>0g(\lambda_{m},\lambda_{m},(n-k-2,2,1^{k}))>0 if PR(m)(k)+PR(m)(k+2)+PR(m)(k2)>0P_{R(m)}(k)+P_{R(m)}(k+2)+P_{R(m)}(k-2)>0, which is equivalent to that

max{PR(m)(k),PR(m)(k+2),PR(m)(k2)}>0.\max\{P_{R(m)}(k),P_{R(m)}(k+2),P_{R(m)}(k-2)\}>0.
Lemma 5.3.

Let m8m\geq 8 be fixed, 0km240\leq k\leq m^{2}-4 and let

NK2(m)={1,2,4,6,8,m212,m210,m28,m26,m25}.NK_{2}(m)=\{1,2,4,6,8,m^{2}-12,m^{2}-10,m^{2}-8,m^{2}-6,m^{2}-5\}.

Then PR(m)(k)+PR(m)(k+2)+PR(m)(k2)>0P_{R(m)}(k)+P_{R(m)}(k+2)+P_{R(m)}(k-2)>0 if and only if kNK2(m)k\notin NK_{2}(m).

Proof.

By directly checking the values for k{1,2,4,6,8}k\in\{1,2,4,6,8\}, we find that

PR(m)(k)+PR(m)(k+2)+PR(m)(k2)=0P_{R(m)}(k)+P_{R(m)}(k+2)+P_{R(m)}(k-2)=0

holds true. Note the sum of elements in R(m)R(m) is m24m^{2}-4 and therefore PR(m)(k)=PR(m)(m24k)P_{R(m)}(k)=P_{R(m)}(m^{2}-4-k). It follows that if k{m212,m210,m28,m26,m25}k\in\{m^{2}-12,m^{2}-10,m^{2}-8,m^{2}-6,m^{2}-5\}, then PR(m)(k)+PR(m)(k+2)+PR(m)(k2)=0.P_{R(m)}(k)+P_{R(m)}(k+2)+P_{R(m)}(k-2)=0.

We shall prove the other direction by induction on mm. We can check the statement is true for m=8.m=8. Now, assuming that the statement is true for m8m\geq 8, we will show that it holds true for m+1.m+1. Due to the symmetry of PR(m+1)(k)P_{R(m+1)}(k), it suffices to demonstrate that PR(m+1)(k)+PR(m+1)(k+2)+PR(m+1)(k2)>0P_{R(m+1)}(k)+P_{R(m+1)}(k+2)+P_{R(m+1)}(k-2)>0 for any k[(m+1)242]{1,2,4,6,8}k\in[\lceil\frac{(m+1)^{2}-4}{2}\rceil]\setminus\{1,2,4,6,8\}.

Since R(m)R(m+1)R(m)\subset R(m+1) by construction, we can assert that

PR(m+1)(k)+PR(m+1)(k+2)+PR(m+1)(k2)PR(m)(k)+PR(m)(k+2)+PR(m)(k2)>0P_{R(m+1)}(k)+P_{R(m+1)}(k+2)+P_{R(m+1)}(k-2)\geq P_{R(m)}(k)+P_{R(m)}(k+2)+P_{R(m)}(k-2)>0

for any k[m24]NK2(m)k\in[m^{2}-4]\setminus NK_{2}(m) by inductive hypothesis. It is easy to see that (m+1)242<m212\lceil\frac{(m+1)^{2}-4}{2}\rceil<m^{2}-12 when m6m\geq 6.

Then by the inductive hypothesis, it follows that PR(m+1)(k)+PR(m+1)(k+2)+PR(m+1)(k2)>0P_{R(m+1)}(k)+P_{R(m+1)}(k+2)+P_{R(m+1)}(k-2)>0 for any kNK2(m+1)k\notin NK_{2}(m+1), which completes the induction. ∎

Theorem 5.4.

Let m8m\geq 8 be fixed and 0km240\leq k\leq m^{2}-4. Then, g(\scalerelgXm,\scalerelgXm,μ2(k,m))>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\mu_{2}(k,m))>0 if and only if k{1,m25}k\notin\{1,m^{2}-5\}.

Proof.

(\Rightarrow) If k=1k=1, by Corollary 3.6.1, g(\scalerelgXm,\scalerelgXm,(m23,2,1))=0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(m^{2}-3,2,1))=0. If k=m25k=m^{2}-5, by the transposition property, g(\scalerelgXm,\scalerelgXm,(m2k2,2,1k))=0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(m^{2}-k-2,2,1^{k}))=0.

(\Leftarrow) If kNK2(m)k\notin NK_{2}(m), the result follows from Theorem 5.2, Murnaghan-Nakayama rule, and Lemma 5.3. If k=2,4,6,8k=2,4,6,8, we consider the decomposition (m24,2,12)=(21,2,1,1)+H(m225)(m^{2}-4,2,1^{2})=(21,2,1,1)+_{H}(m^{2}-25), (m26,2,14)=(19,2,14)+H(m225)(m^{2}-6,2,1^{4})=(19,2,1^{4})+_{H}(m^{2}-25), (m28,2,16)=(17,2,16)+H(m225)(m^{2}-8,2,1^{6})=(17,2,1^{6})+_{H}(m^{2}-25), (m210,2,18)=(15,2,18)+H(m225)(m^{2}-10,2,1^{8})=(15,2,1^{8})+_{H}(m^{2}-25), respectively. Then by the semigroup property, it follows that g(\scalerelgXm,\scalerelgXm,(m2k,2,1k))>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(m^{2}-k,2,1^{k}))>0 for k{2,4,6,8}k\in\{2,4,6,8\}. By the transposition property of Kronecker coefficients, g(\scalerelgXm,\scalerelgXm,(m2k,2,1k))>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(m^{2}-k,2,1^{k}))>0 for k{m26,m28,m210,m212}k\in\{m^{2}-6,m^{2}-8,m^{2}-10,m^{2}-12\}. ∎

Similarly, there are only two ways to place the 1-hook and 3-hook into a rim-hook tableau of shape μ3(k,m)\mu_{3}(k,m) and weight αm\alpha_{m}, as illustrated below.

\ytableausetup

smalltableaux \ytableaushort13\none,33, \none * 10,3,1,1,1,1,1,1           \ytableaushort1333\none,\none,\none * 10,3,1,1,1,1,1,1

Therefore, we have

χ(nk3,3,1k)(2m1,2m3,,3,1)=PR(m)(k)+PR(m)(k+3).\chi^{(n-k-3,3,1^{k})}(2m-1,2m-3,\dots,3,1)=P_{R(m)}(k)+P_{R(m)}(k+3).

It follows that g(\scalerelgXm,\scalerelgXm,(nk3,3,1k))>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(n-k-3,3,1^{k}))>0 if PR(m)(k)+PR(m)(k+3)>0P_{R(m)}(k)+P_{R(m)}(k+3)>0.

Lemma 5.5.

Let m5m\geq 5 be fixed, 0km270\leq k\leq m^{2}-7 and let

NK3(m):={1,3,m210,m28}.NK_{3}(m):=\{1,3,m^{2}-10,m^{2}-8\}.

Then PR(m)(k)+PR(m)(k+3)>0P_{R(m)}(k)+P_{R(m)}(k+3)>0 if and only if kNK3(m)k\notin NK_{3}(m).

Proof.

By directly checking the values for k{1,3}k\in\{1,3\}, we find that PR(m)(k)+PR(m)(k+3)=0P_{R(m)}(k)+P_{R(m)}(k+3)=0 holds true. Note the sum of elements in R(m)R(m) is m24m^{2}-4 and therefore PR(m)(k)=PR(m)(m24k)P_{R(m)}(k)=P_{R(m)}(m^{2}-4-k). It follows that if k{m28,m210}k\in\{m^{2}-8,m^{2}-10\}, then PR(m)(k)+PR(m)(k+3)=0P_{R(m)}(k)+P_{R(m)}(k+3)=0.

We shall prove the other direction by induction on mm. It is easy to check that the statement is true for m=7.m=7. Now, assuming that the statement is true for m7m\geq 7, we will show that it is also true for m+1.m+1. Due to the symmetry of PR(m+1)(k)P_{R(m+1)}(k), it suffices to demonstrate that PR(m+1)(k)+PR(m+1)(k+3)>0P_{R(m+1)}(k)+P_{R(m+1)}(k+3)>0 for any k[(m+1)272]{1,3}k\in[\lfloor\frac{(m+1)^{2}-7}{2}\rfloor]\setminus\{1,3\}.

Since R(m)R(m+1)R(m)\subset R(m+1) by construction, we can assert that

PR(m+1)(k)+PR(m+1)(k+3)PR(m)(k)+PR(m)(k+3)>0P_{R(m+1)}(k)+P_{R(m+1)}(k+3)\geq P_{R(m)}(k)+P_{R(m)}(k+3)>0

for any k[m24]NK3(m)k\in[m^{2}-4]\setminus NK_{3}(m) by inductive hypothesis. We can verify that (m+1)272<m210\lfloor\frac{(m+1)^{2}-7}{2}\rfloor<m^{2}-10 when m5m\geq 5. Then by the inductive hypothesis, we conclude that PR(m)(k)+PR(m)(k+3)>0P_{R(m)}(k)+P_{R(m)}(k+3)>0 for any kNK3(m+1)k\notin NK_{3}(m+1), which completes the induction. ∎

Theorem 5.6.

Let m7m\geq 7 be fixed and 0km260\leq k\leq m^{2}-6. Then, g(\scalerelgXm,\scalerelgXm,μ3(k,m))>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\mu_{3}(k,m))>0 if and only if k1k\neq 1.

Proof.

(\Rightarrow) If k=1k=1, by Corollary 3.6.1, g(\scalerelgXm,\scalerelgXm,(m2k3,3,1k))=0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(m^{2}-k-3,3,1^{k}))=0.

(\Leftarrow) Now assume that k{0,2,3,,m26}k\in\{0,2,3,\dots,m^{2}-6\}. If k=3k=3, we can decompose the partition (m26,3,13)(m^{2}-6,3,1^{3}) as (3,3,1,1,1)+H(m29)(3,3,1,1,1)+_{H}(m^{2}-9). Since g((3,3,3),(3,3,3),(3,3,1,1,1))>0g((3,3,3),(3,3,3),(3,3,1,1,1))>0, by semigroup property, it follows that g(\scalerelgXm,\scalerelgXm,(m26,3,13))>0.g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(m^{2}-6,3,1^{3}))>0. If k=m210k=m^{2}-10, we can decompose the partition (7,3,1k)(7,3,1^{k}) as (7,3,16)+V(1(m216))(7,3,1^{6})+_{V}(1^{(m^{2}-16)}). Since g((4,4,4,4),(4,4,4,4),(7,3,16))>0g((4,4,4,4),(4,4,4,4),(7,3,1^{6}))>0, by semigroup property,it follows that g(\scalerelgXm,\scalerelgXm,(7,3,1k))>0.g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(7,3,1^{k}))>0. If k=m28k=m^{2}-8, we can decompose the partition (5,3,1k)(5,3,1^{k}) as (5,3,18)+V(1(m216))(5,3,1^{8})+_{V}(1^{(m^{2}-16)}). Since g((4,4,4,4),(4,4,4,4),(5,3,18))>0g((4,4,4,4),(4,4,4,4),(5,3,1^{8}))>0, by semigroup property, it follows that g(\scalerelgXm,\scalerelgXm,(5,3,1k))>0.g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(5,3,1^{k}))>0.

If k=m26k=m^{2}-6, g(\scalerelgXm,\scalerelgXm,(3,3,1k))=g(\scalerelgXm,\scalerelgXm,(m24,2,2))>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(3,3,1^{k}))=g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(m^{2}-4,2,2))>0 by Theorem 4.11. If k<m26k<m^{2}-6 and kNK3(m)k\notin NK_{3}(m), the result follows from Theorem 5.2, Murnaghan-Nakayama rule, and Lemma 5.5. ∎

Next, we will discuss the constituency of near-hooks with a second row of length of at least 8.

Proposition 5.7.

For every i8i\geq 8, we have g(\scalerelgXm,\scalerelgXm,μi(k,m))>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\mu_{i}(k,m))>0 for all m20m\geq 20 and k0k\geq 0.

Proof.

Let i8i\geq 8 be fixed. Suppose that m20m\geq 20.

If k7m+9ik\geq 7m+9-i, we can decompose the transpose of partition μi(k,m)\mu_{i}(k,m), that is (k+2,2i1,1m22ik)(k+2,2^{i-1},1^{m^{2}-2i-k}) as (k+2,2i1,1m22ik)=(k1,1i1)+H(k+2k1,1m2ik1)(k+2,2^{i-1},1^{m^{2}-2i-k})=(k_{1},1^{i-1})+_{H}(k+2-k_{1},1^{m^{2}-i-k-1}) where k1=7mi+1k_{1}=7m-i+1. Since k7m+9ik\geq 7m+9-i, we have k+2k110k+2-k_{1}\geq 10. Then by Theorem 3.7, we have g(7m,7m,(k1,1i1))>0g(7^{m},7^{m},(k_{1},1^{i-1}))>0 and g((m7)m,(m7)m,(k+2k1,1m2ik1))>0g((m-7)^{m},(m-7)^{m},(k+2-k_{1},1^{m^{2}-i-k-1}))>0. We can use the Semigroup property to add the partition triples, which implies that g(\scalerelgXm,\scalerelgXm,(k+2,2i1,1m22ik))>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},(k+2,2^{i-1},1^{m^{2}-2i-k}))>0. Then by the transposition property, we have g(\scalerelgXm,\scalerelgXm,μi(k,m))>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\mu_{i}(k,m))>0.

If k7m+8ik\leq 7m+8-i, 2i+k2(7m+8)15m2i+k\leq 2(7m+8)\leq 15m, we consider the decomposition \scalerelgXm=\scalerelgXm1+V(m1mm1)+H((mm1)m)\mathord{\scalerel*{\Box}{gX}}_{m}=\mathord{\scalerel*{\Box}{gX}}_{m_{1}}+_{V}(m_{1}^{m-m_{1}})+_{H}((m-m_{1})^{m}), where m1=k+8m_{1}=\left\lceil\sqrt{k+8}\,\right\rceil. Since m1k+8m_{1}\leq\left\lceil\sqrt{k+8}\,\right\rceil and m16m\geq 16, we have m12(k+8)k+8m2k2im_{1}^{2}-(k+8)\leq k+8\leq m^{2}-k-2i, which implies that m2m12i+4i4m^{2}-m_{1}^{2}-i+4\geq i-4. Moreover, since k7mk\leq 7m and m20m\geq 20, we have m1m8m_{1}\leq m-8. We will show that there exists a decomposition μi(k,m)=μ4(k,m1)+H(a+d1,a)+H(b+d2,b)\mu_{i}(k,m)=\mu_{4}(k,m_{1})+_{H}(a+d_{1},a)+_{H}(b+d_{2},b) such that (a+d1,a)m1(mm1)(a+d_{1},a)\vdash m_{1}(m-m_{1}), (b+d2,b)m(mm1)(b+d_{2},b)\vdash m(m-m_{1}) and a+b=i4a+b=i-4. We consider the following two cases:

  1. Case 1:

    If mm is odd, then m(mm1)m(m-m_{1}) and m2m122(i4)m^{2}-m_{1}^{2}-2(i-4) always have the same parity. If m2m122(i4)=m(mm1)2m^{2}-m_{1}^{2}-2(i-4)=m(m-m_{1})-2, let d2=m(mm1)4d_{2}=m(m-m_{1})-4 and it follows that b=2b=2; otherwise, let d2=min(m(mm1),m2m122(i4))d_{2}=\min(m(m-m_{1}),m^{2}-m_{1}^{2}-2(i-4)). It is easy to check that a1a\neq 1 and b1b\neq 1 in this case.

  2. Case 2:

    If mm is even, then m1(mm1)m_{1}(m-m_{1}) and m2m122(i4)m^{2}-m_{1}^{2}-2(i-4) always have the same parity. If m2m122(i4)=m1(mm1)2m^{2}-m_{1}^{2}-2(i-4)=m_{1}(m-m_{1})-2, let d1=m1(mm1)4d_{1}=m_{1}(m-m_{1})-4 and it follows that a=2a=2; otherwise, let d1=min(m1(mm1),m2m122(i4))d_{1}=\min(m_{1}(m-m_{1}),m^{2}-m_{1}^{2}-2(i-4)). It is easy to check that a1a\neq 1 and b1b\neq 1 in this case.

By Corollary 5.1.1, we have g(\scalerelgXm1,\scalerelgXm1,μ4(k,m1)>0g(\mathord{\scalerel*{\Box}{gX}}_{m_{1}},\mathord{\scalerel*{\Box}{gX}}_{m_{1}},\mu_{4}(k,m_{1})>0. Since m,m1,mm18m,m_{1},m-m_{1}\geq 8, by Theorem 4.1, we can conclude that g((m1mm1),(m1mm1),(a+d1,a))>0g((m_{1}^{m-m_{1}}),(m_{1}^{m-m_{1}}),(a+d_{1},a))>0 and g(((mm1)m),((mm1)m),(b+d2,b))>0g(((m-m_{1})^{m}),((m-m_{1})^{m}),(b+d_{2},b))>0. Then, adding the partition triples horizontally by semigroup property, we can conclude that g(\scalerelgXm,\scalerelgXm,μi(k,m))>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\mu_{i}(k,m))>0 for every m20.m\geq 20.

Corollary 5.7.1.

For every i8i\geq 8, we have g(\scalerelgXm,\scalerelgXm,μi(k,m))>0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\mu_{i}(k,m))>0 for all m20m\geq 20 and k0k\geq 0.

Proof.

It follows directly from Corollary 5.1.1 and Proposition 5.7. ∎

6. Additional Remarks

6.1.

We have proved the positivity of Kronecker coefficients indexed by pairs of rectangular Young diagrams and certain three-row partitions of special shapes. We could further use the result of square Kronecker coefficients to investigate the behavior of tensor squares of irreducible representations for rectangular Young diagrams and explore the positivity properties for specific families of rectangular partitions.

6.2.

Since the decomposition of a rectangular partition can only be achieved by writing it as a horizontal or vertical sum of two rectangular partitions, it limits the application of the semigroup property. For partitions with more rows or larger Durfee size, there are instances where the semigroup property fails to prove positivity. A specific example is the Kronecker coefficient g(\scalerelgXm,\scalerelgXm,((m+1)m1,1))g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},((m+1)^{m-1},1)). Due to the partition shapes involved, the only valid method to decompose them to satisfy number-theoretical conditions is as follows: \scalerelgXm=((m1)m)+H(1m)=(mm1)+V(m)\mathord{\scalerel*{\Box}{gX}}_{m}=((m-1)^{m})+_{H}(1^{m})=(m^{m-1})+_{V}(m) and ((m+1)m1,1)=(mm1)+H(1m).((m+1)^{m-1},1)=(m^{m-1})+_{H}(1^{m}). However, g(1m,1m,1m)=g(m,m,1m)=0g(1^{m},1^{m},1^{m})=g(m,m,1^{m})=0, indicating that we cannot rely on this approach to prove positivity. This demonstrates the limitations of the semigroup property in certain cases.

When mm is even, we can establish through a recursive argument that there exists no rim-hook tableau of shape ((m+1)m1,1)((m+1)^{m-1},1) with type αm\alpha_{m}. Note that there is a unique arrangement for both the (2m1)(2m-1)-hook and the (2m3)(2m-3)-hook. These two longest rim-hooks invariably occupy the skew-shape ((m+1)m1,1)/((m1)m3,1)((m+1)^{m-1},1)/((m-1)^{m-3},1), as depicted in the diagram below. Then the problem is reduced to a search for a rim-hook tableau with shape ((m1)m3,1)((m-1)^{m-3},1) and type αm2\alpha_{m-2}. By iterating this process, we know that a rim-hook tableau with shape ((m+1)m1,1)((m+1)^{m-1},1) and type αm\alpha_{m} exists if and only if a rim-hook tableau with shape (5,4)(5,4) and type (5,3,1)(5,3,1) can be found. Therefore, there does not exist a rim-hook tableau of shape ((m+1)m1,1)((m+1)^{m-1},1) and type αm\alpha_{m}, which implies that χ((m+1)m1,1)(αm)=0\chi^{((m+1)^{m-1},1)}(\alpha_{m})=0 by Murnaghan-Nakayama Rule. Hence, the character approach (Theorem 5.2) is also not applicable in this case.

\ydiagram

[*(white)]5,5,5,1 *[ *(white)∙]7,6,6,6 *[*(gray)]7,7,7,7,7,1

Appendix A Missing partitions in tensor square of square with a small side length

With the help of computer, we find all partitions λm2\lambda\vdash m^{2} such that g(\scalerelgXm,\scalerelgXm,λ)=0g(\mathord{\scalerel*{\Box}{gX}}_{m},\mathord{\scalerel*{\Box}{gX}}_{m},\lambda)=0 for m=4,5,6,7m=4,5,6,7:

  • g(\scalerelgX4,\scalerelgX4,λ)=0g(\mathord{\scalerel*{\Box}{gX}}_{4},\mathord{\scalerel*{\Box}{gX}}_{4},\lambda)=0 if and only if λ\lambda or λ\seqsplit{(15,1),(14,1,1),(13,2,1),(12,3,1),(12,1,1,1,1),(11,5),(10,1,1,1,1,1,1),(9,7),(8,7,1),(8,2,1,1,1,1,1,1),(7,7,2),(7,5,4)}\lambda^{\prime}\in\seqsplit{\{(15,1),(14,1,1),(13,2,1),(12,3,1),(12,1,1,1,1),(11,5),(10,1,1,1,1,1,1),(9,7),(8,7,1),(8,2,1,1,1,1,1,1),(7,7,2),(7,5,4)\}};

  • g(\scalerelgX5,\scalerelgX5,λ)=0g(\mathord{\scalerel*{\Box}{gX}}_{5},\mathord{\scalerel*{\Box}{gX}}_{5},\lambda)=0 if and only if λ\lambda or λ\seqsplit{(24,1),(23,1,1),(22,2,1),(21,3,1),(21,1,1,1,1),(19,1,1,1,1,1,1),(14,1,1,1,1,1,1,1,1,1,1,1)}\lambda^{\prime}\in\seqsplit{\{(24,1),(23,1,1),(22,2,1),(21,3,1),(21,1,1,1,1),(19,1,1,1,1,1,1),(14,1,1,1,1,1,1,1,1,1,1,1)\}};

  • g(\scalerelgX6,\scalerelgX6,λ)=0g(\mathord{\scalerel*{\Box}{gX}}_{6},\mathord{\scalerel*{\Box}{gX}}_{6},\lambda)=0 if and only if λ\lambda or λ\seqsplit{(35,1),(34,1,1),(33,2,1),(32,3,1),(32,1,1,1,1),(30,1,1,1,1,1,1),(23,113),(19,17)}\lambda^{\prime}\in\seqsplit{\{(35,1),(34,1,1),(33,2,1),(32,3,1),(32,1,1,1,1),(30,1,1,1,1,1,1),}{(23,1^{13}),(19,17)\}};

  • g(\scalerelgX7,\scalerelgX7,λ)=0g(\mathord{\scalerel*{\Box}{gX}}_{7},\mathord{\scalerel*{\Box}{gX}}_{7},\lambda)=0 if and only if λ\lambda or λ\seqsplit{{(48,1),(47,1,1),(46,2,1),(45,3,1),(45,1,1,1,1),(43,1,1,1,1,1,1)}.\lambda^{\prime}\in\seqsplit{\{\{(48,1),(47,1,1),(46,2,1),(45,3,1),(45,1,1,1,1),(43,1,1,1,1,1,1)\}}.

References

  • [BI08] Peter Bürgisser and Christian Ikenmeyer “The complexity of computing Kronecker coefficients” In FPSAC’08 - 20th International Conference on Formal Power Series and Algebraic Combinatorics DMTCS Proceedings vol., 2008 DOI: 10.46298/dmtcs.3622
  • [Bla17] Jonah Blasiak “Kronecker coefficients for one hook shape” In Séminaire Lotharingien de Combinatoire 77, 2017
  • [BO05] Cristina M. Ballantine and Rosa C. Orellana “On the Kronecker Product s(n-p,p) * sλ\lambda In Electron. J. Comb. 12, 2005
  • [CHM07] Matthias Christandl, Aram W Harrow and Graeme Mitchison “Nonzero Kronecker Coefficients and What They Tell us about Spectra” In Communications in Mathematical Physics 270.3, 2007, pp. 575–585 DOI: 10.1007/s00220-006-0157-3
  • [DTK09] Donna Q.. Dou, Robert L. Tang and Ronald C. King “A hive model determination of multiplicity-free Schur function products and skew Schur functions”, 2009 arXiv:0901.0186 [math.CO]
  • [Gut10] Christian Gutschwager “On Multiplicity-Free Skew Characters and the Schubert Calculus” In Annals of Combinatorics 14.3 Springer ScienceBusiness Media LLC, 2010, pp. 339–353 DOI: 10.1007/s00026-010-0063-4
  • [Hei+12] Gerhard Heide, Jan Saxl, Pham Huu Tiep and Alexandre E. Zalesski “Conjugacy action, induced representations and the Steinberg square for simple groups of Lie type” In Proceedings of the London Mathematical Society 106.4 Wiley, 2012, pp. 908–930 DOI: 10.1112/plms/pds062
  • [Ike15] Christian Ikenmeyer “The Saxl conjecture and the dominance order” In Discrete Mathematics 338.11 Elsevier BV, 2015, pp. 1970–1975 DOI: 10.1016/j.disc.2015.04.027
  • [IMW17] Christian Ikenmeyer, Ketan D. Mulmuley and Michael Walter “On vanishing of Kronecker coefficients” In computational complexity 26.4 Springer ScienceBusiness Media LLC, 2017, pp. 949–992 DOI: 10.1007/s00037-017-0158-y
  • [IP17] Christian Ikenmeyer and Greta Panova “Rectangular Kronecker coefficients and plethysms in geometric complexity theory” In Advances in Mathematics 319, 2017, pp. 40–66 DOI: https://doi.org/10.1016/j.aim.2017.08.024
  • [Li21] Xin Li “Saxl Conjecture for triple hooks” In Discrete Mathematics 344.6, 2021, pp. 112340 DOI: https://doi.org/10.1016/j.disc.2021.112340
  • [Lit58] Dudley E. Littlewood “Products and plethysms of characters with orthogonal, symplectic and symmetric groups” In Canadian Journal of Mathematics 10, 1958, pp. 17–32
  • [Liu17] Ricky Ini Liu “A simplified Kronecker rule for one hook shape” In Proceedings of the American Mathematical Society 145.9 American Mathematical Society, 2017, pp. pp. 3657–3664 URL: https://www.jstor.org/stable/90013039
  • [LS16] Sammy Luo and Mark Sellke “The Saxl conjecture for fourth powers via the semigroup property” In Journal of Algebraic Combinatorics 45.1 Springer ScienceBusiness Media LLC, 2016, pp. 33–80 DOI: 10.1007/s10801-016-0700-z
  • [Pan23] Greta Panova “Complexity and asymptotics of structure constants”, 2023 arXiv:2305.02553 [math.CO]
  • [PP13] Igor Pak and Greta Panova “Strict unimodality of q-binomial coefficients” In Comptes Rendus Mathematique 351.11-12 Elsevier BV, 2013, pp. 415–418 DOI: 10.1016/j.crma.2013.06.008
  • [PP14] Igor Pak and Greta Panova “Unimodality via Kronecker products” In Journal of Algebraic Combinatorics 40.4, 2014, pp. 1103–1120
  • [PP17] Igor Pak and Greta Panova “On the complexity of computing Kronecker coefficients” In computational complexity 26.1, 2017, pp. 1–36
  • [PPV16] Igor Pak, Greta Panova and Ernesto Vallejo “Kronecker products, characters, partitions, and the tensor square conjectures” In Advances in Mathematics 288, 2016, pp. 702–731 DOI: https://doi.org/10.1016/j.aim.2015.11.002
  • [Rem89] Jeffrey B. Remmel “A formula for the Kronecker products of Schur functions of hook shapes” In Journal of Algebra 120, 1989, pp. 100–118
  • [RW94] Jeffrey B. Remmel and Tamsen Whitehead “On the Kronecker product of Schur functions of two row shapes” In Bulletin of the Belgian Mathematical Society - Simon Stevin 1.5 The Belgian Mathematical Society, 1994, pp. 649–683 DOI: 10.36045/bbms/1103408635
  • [Sag91] Bruce E. Sagan “The symmetric group - representations, combinatorial algorithms, and symmetric functions.”, Wadsworth & Brooks / Cole mathematics series Wadsworth, 1991, pp. I–XV\bibrangessep1–197
  • [SF97] R.P. Stanley and S. Fomin “Enumerative Combinatorics: Volume 2”, Cambridge Studies in Advanced Mathematics Cambridge University Press, 1997 URL: https://books.google.com/books?id=2QOtugEACAAJ
  • [Sta00] Richard P. Stanley “Positivity problems and conjectures in algebraic combinatorics” In Mathematics: Frontiers and Perspectives Providence, RI: American Mathematical Society, 2000, pp. 295–319
  • [TY10] Hugh Thomas and Alexander Yong “Multiplicity-Free Schubert Calculus” In Canadian Mathematical Bulletin 53.1 Cambridge University Press, 2010, pp. 171–186 DOI: 10.4153/CMB-2010-032-x
  • [Val14] Ernesto Vallejo “A diagrammatic approach to Kronecker squares” In Journal of Combinatorial Theory, Series A 127, 2014, pp. 243–285 DOI: https://doi.org/10.1016/j.jcta.2014.06.002