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

The sign character of the triagonal fermionic coinvariant ring

John Lentfer Department of Mathematics
University of California, Berkeley, CA, USA
[email protected]
Abstract.

We determine the trigraded multiplicity of the sign character of the triagonal fermionic coinvariant ring Rn(0,3)R_{n}^{(0,3)}. As a corollary, this proves a conjecture of Bergeron (2020) that the dimension of the sign character of Rn(0,3)R_{n}^{(0,3)} is n2n+1n^{2}-n+1. We also give an explicit formula for double hook characters in the diagonal fermionic coinvariant ring Rn(0,2)R_{n}^{(0,2)}, and discuss methods towards calculating the sign character of Rn(0,4)R_{n}^{(0,4)}. Finally, we give a multigraded refinement of a conjecture of Bergeron (2020) that the dimension of the sign character of the (1,3)(1,3)-bosonic-fermionic coinvariant ring Rn(1,3)R_{n}^{(1,3)} is 12F3n\frac{1}{2}F_{3n}, where FnF_{n} is a Fibonacci number.

1. Introduction

The diagonal coinvariant ring Rn(2,0):=[x1,,xn,y1,,yn]/[x1,,xn,y1,,yn]+𝔖nR_{n}^{(2,0)}:=\mathbb{Q}[x_{1},\ldots,x_{n},y_{1},\ldots,y_{n}]/\allowbreak\langle\mathbb{Q}[x_{1},\ldots,x_{n},y_{1},\ldots,y_{n}]_{+}^{\mathfrak{S}_{n}}\rangle was introduced by Haiman in 1994 [9], and since then has been studied extensively. Its defining ideal is generated by all polynomials in [x1,,xn,y1,,yn]\mathbb{Q}[x_{1},\ldots,x_{n},y_{1},\ldots,y_{n}], with no constant term, which are invariant under the diagonal action of 𝔖n\mathfrak{S}_{n}:

(1) σp(x1,,xn,y1,,yn)=p(xσ(1),,xσ(n),yσ(1),,yσ(n)).\sigma\cdot p(x_{1},\ldots,x_{n},y_{1},\ldots,y_{n})=p(x_{\sigma(1)},\ldots,x_{\sigma(n)},y_{\sigma(1)},\ldots,y_{\sigma(n)}).

Haiman found the dimension, bigraded Hilbert series, and bigraded Frobenius series of Rn(2,0)R_{n}^{(2,0)} [10].

There has been much recent interest (see [3, 2, 1, 17]) in studying a more general class of coinvariant rings Rn(k,j)R_{n}^{(k,j)} with kk sets of nn commuting (bosonic) variables 𝒙n:={x1,,xn}\bm{x}_{n}:=\{x_{1},\ldots,x_{n}\}, 𝒚n:={y1,,yn}\bm{y}_{n}:=\{y_{1},\ldots,y_{n}\}, 𝒛n:={z1,,zn}\bm{z}_{n}:=\{z_{1},\ldots,z_{n}\}, etc., and jj sets of nn anticommuting (fermionic) variables 𝜽n:={θ1,,θn}\bm{\theta}_{n}:=\{\theta_{1},\ldots,\theta_{n}\}, 𝝃n:={ξ1,,ξn}\bm{\xi}_{n}:=\{\xi_{1},\ldots,\xi_{n}\}, 𝝆n:={ρ1,,ρn}\bm{\rho}_{n}:=\{\rho_{1},\ldots,\rho_{n}\}, etc. We define the (k,j)(k,j)-bosonic-fermionic coinvariant ring by

(2) Rn(k,j):=[𝒙n,𝒚n,𝒛n,k,𝜽n,𝝃n,𝝆n,j]/[𝒙n,𝒚n,𝒛n,k,𝜽n,𝝃n,𝝆n,j]+𝔖n,R_{n}^{(k,j)}:=\mathbb{Q}[\underbrace{\bm{x}_{n},\bm{y}_{n},\bm{z}_{n},\ldots}_{k},\underbrace{\bm{\theta}_{n},\bm{\xi}_{n},\bm{\rho}_{n},\ldots}_{j}]/\langle\mathbb{Q}[\underbrace{\bm{x}_{n},\bm{y}_{n},\bm{z}_{n},\ldots}_{k},\underbrace{\bm{\theta}_{n},\bm{\xi}_{n},\bm{\rho}_{n},\ldots}_{j}]_{+}^{\mathfrak{S}_{n}}\rangle,

where its defining ideal is generated by all polynomials in [𝒙n,𝒚n,𝒛n,k,𝜽n,𝝃n,𝝆n,j]\mathbb{Q}[\underbrace{\bm{x}_{n},\bm{y}_{n},\bm{z}_{n},\ldots}_{k},\underbrace{\bm{\theta}_{n},\bm{\xi}_{n},\bm{\rho}_{n},\ldots}_{j}], without constant term, which are invariant under the diagonal action of the symmetric group 𝔖n\mathfrak{S}_{n}, given by permuting the indices of the variables. Commuting variables commute with all variables. Anticommuting variables anticommute with all anticommuting variables. That is, θiθj=θjθi\theta_{i}\theta_{j}=-\theta_{j}\theta_{i} for all i,ji,j, and mixed products between different sets of fermionic variables likewise anticommute. Note that this implies that θi2=0\theta_{i}^{2}=0.

We recall the definitions of the multigraded Hilbert and Frobenius series of Rn(k,j)R_{n}^{(k,j)} (see for example [2]). For fixed integers k,j0k,j\geq 0, Rn(k,j)R_{n}^{(k,j)} decomposes as a direct sum of multihomogenous components, which are 𝔖n\mathfrak{S}_{n}-modules:

(3) Rn(k,j)=r1,,rk,s1,,sj0(Rn(k,j))r1,,rk,s1,,sj.R_{n}^{(k,j)}=\bigoplus_{r_{1},\ldots,r_{k},s_{1},\ldots,s_{j}\geq 0}(R_{n}^{(k,j)})_{r_{1},\ldots,r_{k},s_{1},\ldots,s_{j}}.

We denote the multigraded Hilbert series by

Hilb\displaystyle\operatorname{Hilb} (Rn(k,j);q1,,qk;u1,,uj)\displaystyle(R_{n}^{(k,j)};q_{1},\ldots,q_{k};u_{1},\ldots,u_{j})
(4) :=r1,,rk,s1,,sj0dim((Rn(k,j))r1,,rk,s1,,sj)q1r1qkrku1s1ujsj,\displaystyle:=\sum_{r_{1},\ldots,r_{k},s_{1},\ldots,s_{j}\geq 0}\dim\left((R_{n}^{(k,j)})_{r_{1},\ldots,r_{k},s_{1},\ldots,s_{j}}\right)q_{1}^{r_{1}}\cdots q_{k}^{r_{k}}u_{1}^{s_{1}}\cdots u_{j}^{s_{j}},

and the multigraded Frobenius series by

Frob\displaystyle\operatorname{Frob} (Rn(k,j);q1,,qk;u1,,uj)\displaystyle(R_{n}^{(k,j)};q_{1},\ldots,q_{k};u_{1},\ldots,u_{j})
(5) :=r1,,rk,s1,,sj0FChar((Rn(k,j))r1,,rk,s1,,sj)q1r1qkrku1s1ujsj,\displaystyle:=\sum_{r_{1},\ldots,r_{k},s_{1},\ldots,s_{j}\geq 0}F\operatorname{Char}\left((R_{n}^{(k,j)})_{r_{1},\ldots,r_{k},s_{1},\ldots,s_{j}}\right)q_{1}^{r_{1}}\cdots q_{k}^{r_{k}}u_{1}^{s_{1}}\cdots u_{j}^{s_{j}},

where FF denotes the Frobenius characteristic map and Char\operatorname{Char} denotes the character. For simplicity, if k2k\leq 2, we will use q,tq,t for q1,q2q_{1},q_{2} and if j4j\leq 4, we will use u,v,w,xu,v,w,x for u1,u2,u3,u4u_{1},u_{2},u_{3},u_{4}. Recall that Frob(Rn(k,j);q1,,qk;u1,,uj),h1n=Hilb(Rn(k,j);q1,,qk;u1,,uj)\langle\operatorname{Frob}(R_{n}^{(k,j)};q_{1},\ldots,q_{k};u_{1},\ldots,u_{j}),h_{1}^{n}\rangle=\operatorname{Hilb}(R_{n}^{(k,j)};q_{1},\ldots,q_{k};u_{1},\ldots,u_{j}). Furthermore, Rn(k,j)R_{n}^{(k,j)} is a GLk×GLj×𝔖n\operatorname{GL}_{k}\times\operatorname{GL}_{j}\times\mathfrak{S}_{n}-module (see [2, Section 2]), so its multigraded Frobenius character is a sum of products of three Schur functions, which are irreducible characters of polynomial representations of GLk\operatorname{GL}_{k} and GLj\operatorname{GL}_{j}, along with a Frobenius character. The main focus of this paper is on the case (k,j)=(0,3)(k,j)=(0,3). We now describe the setting in more detail, from the perspective of both coinvariants and, isomorphically, harmonics.

Now specialize to Rn(0,3):=[𝜽n,𝝃n,𝝆n]/[𝜽n,𝝃n,𝝆n]+𝔖nR_{n}^{(0,3)}:=\mathbb{Q}[\bm{\theta}_{n},\bm{\xi}_{n},\bm{\rho}_{n}]/\langle\mathbb{Q}[\bm{\theta}_{n},\bm{\xi}_{n},\bm{\rho}_{n}]_{+}^{\mathfrak{S}_{n}}\rangle, which we call the triagonal fermionic coinvariant ring.111Note that [𝜽n,𝝃n,𝝆n]\mathbb{Q}[\bm{\theta}_{n},\bm{\xi}_{n},\bm{\rho}_{n}] is an exterior algebra over the variables 𝜽n,𝝃n,𝝆n\bm{\theta}_{n},\bm{\xi}_{n},\bm{\rho}_{n}. For more on this perspective, see [11]. Its defining ideal [𝜽n,𝝃n,𝝆n]+𝔖n\langle\mathbb{Q}[\bm{\theta}_{n},\bm{\xi}_{n},\bm{\rho}_{n}]_{+}^{\mathfrak{S}_{n}}\rangle is the ideal generated by polynomials in [𝜽n,𝝃n,𝝆n]\mathbb{Q}[\bm{\theta}_{n},\bm{\xi}_{n},\bm{\rho}_{n}], without constant term, invariant under the diagonal action of 𝔖n\mathfrak{S}_{n}. A generating set for the ideal [𝜽n,𝝃n,𝝆n]+𝔖n\langle\mathbb{Q}[\bm{\theta}_{n},\bm{\xi}_{n},\bm{\rho}_{n}]_{+}^{\mathfrak{S}_{n}}\rangle is given by all the nonzero monomial symmetric functions in any of the three sets of variables:

{θ1++θn,ξ1++ξn,ρ1++ρn,θ1ξ1++θnξn,\displaystyle\{\theta_{1}+\cdots+\theta_{n},\xi_{1}+\cdots+\xi_{n},\rho_{1}+\cdots+\rho_{n},\theta_{1}\xi_{1}+\cdots+\theta_{n}\xi_{n},
(6) θ1ρ1++θnρn,ξ1ρ1++ξnρn,θ1ξ1ρ1++θnξnρn}.\displaystyle\theta_{1}\rho_{1}+\cdots+\theta_{n}\rho_{n},\xi_{1}\rho_{1}+\cdots+\xi_{n}\rho_{n},\theta_{1}\xi_{1}\rho_{1}+\cdots+\theta_{n}\xi_{n}\rho_{n}\}.

Consider the ring [𝜽n,𝝃n,𝝆n]\mathbb{Q}[\bm{\theta}_{n},\bm{\xi}_{n},\bm{\rho}_{n}] in three sets of nn anticommuting variables, which we arrange into the 3×n3\times n matrix

(7) M:=[θ1θ2θnξ1ξ2ξnρ1ρ2ρn].M:=\begin{bmatrix}\theta_{1}&\theta_{2}&\cdots&\theta_{n}\\ \xi_{1}&\xi_{2}&\cdots&\xi_{n}\\ \rho_{1}&\rho_{2}&\cdots&\rho_{n}\end{bmatrix}.

For each 3×33\times 3 invertible matrix AA in GL3\operatorname{GL}_{3}, multiply MM on the left by AA to obtain the product AMA\cdot M. Then, in any polynomial f[𝜽n,𝝃n,𝝆n]f\in\mathbb{Q}[\bm{\theta}_{n},\bm{\xi}_{n},\bm{\rho}_{n}], replace each variable θi,ξi,ρi\theta_{i},\xi_{i},\rho_{i} by the corresponding entry of the matrix AMA\cdot M. This defines a left GL3\operatorname{GL}_{3}-action on [𝜽n,𝝃n,𝝆n]\mathbb{Q}[\bm{\theta}_{n},\bm{\xi}_{n},\bm{\rho}_{n}].

The ring [𝜽n,𝝃n,𝝆n]\mathbb{Q}[\bm{\theta}_{n},\bm{\xi}_{n},\bm{\rho}_{n}] is naturally trigraded. Denote by ([𝜽n,𝝃n,𝝆n])a,b,c(\mathbb{Q}[\bm{\theta}_{n},\bm{\xi}_{n},\bm{\rho}_{n}])_{a,b,c} the homogeneous component spanned by all monomials that contain exactly aa variables in 𝜽n\bm{\theta}_{n}, bb variables in 𝝃n\bm{\xi}_{n}, and cc variables in 𝝆n\bm{\rho}_{n}. Since GL3\operatorname{GL}_{3} acts by linear combinations of the three rows, the GL3\operatorname{GL}_{3}-action preserves the total degree d=a+b+cd=a+b+c. Hence each direct sum a+b+c=d([𝜽n,𝝃n,𝝆n])a,b,c\bigoplus_{a+b+c=d}(\mathbb{Q}[\bm{\theta}_{n},\bm{\xi}_{n},\bm{\rho}_{n}])_{a,b,c} is GL3\operatorname{GL}_{3}-invariant.

Each permutation σ𝔖n\sigma\in\mathfrak{S}_{n} corresponds to a permutation matrix PσP_{\sigma}, defined so that (Pσ)i,j=1(P_{\sigma})_{i,j}=1 if i=σ(j)i=\sigma(j) and 0 otherwise. Then the right multiplication MPσM\cdot P_{\sigma} has the effect on MM of letting σ\sigma act on the indices of all variables: θiθσ(i)\theta_{i}\mapsto\theta_{\sigma(i)}, ξiξσ(i)\xi_{i}\mapsto\xi_{\sigma(i)}, and ρiρσ(i)\rho_{i}\mapsto\rho_{\sigma(i)}. Then, in any polynomial f[𝜽n,𝝃n,𝝆n]f\in\mathbb{Q}[\bm{\theta}_{n},\bm{\xi}_{n},\bm{\rho}_{n}], replace each variable θi,ξi,ρi\theta_{i},\xi_{i},\rho_{i} by the corresponding entry of the matrix MPσM\cdot P_{\sigma}. This defines a right 𝔖n\mathfrak{S}_{n}-action on [𝜽n,𝝃n,𝝆n]\mathbb{Q}[\bm{\theta}_{n},\bm{\xi}_{n},\bm{\rho}_{n}].

Because left multiplication by AA and right multiplication by PσP_{\sigma} commute, these GL3\operatorname{GL}_{3} and 𝔖n\mathfrak{S}_{n}-actions commute, so [𝜽n,𝝃n,𝝆n]\mathbb{Q}[\bm{\theta}_{n},\bm{\xi}_{n},\bm{\rho}_{n}] is a GL3×𝔖n\operatorname{GL}_{3}\times\mathfrak{S}_{n}-module.

Recall the definition of derivative of anticommuting variables (where here each θi\theta_{i} could be from any of the three sets of anticommuting variables) is (see for example [15, Section 1.5]):

(8) θjθi1θik={(1)1θi1θ^iθik if j=i,0 otherwise.\partial_{\theta_{j}}\theta_{i_{1}}\cdots\theta_{i_{k}}=\begin{cases}(-1)^{\ell-1}\theta_{i_{1}}\cdots\hat{\theta}_{i_{\ell}}\cdots\theta_{i_{k}}&\text{ if }j=i_{\ell},\\ 0&\text{ otherwise.}\end{cases}

Define the space of triagonal fermionic harmonics by

(9) Tn:={f[𝜽n,𝝃n,𝝆n]|i=1nθihξikρif=0 for all h+k+>0}.T_{n}:=\left\{f\in\mathbb{Q}[\bm{\theta}_{n},\bm{\xi}_{n},\bm{\rho}_{n}]\,\Bigg{|}\,\sum_{i=1}^{n}\partial_{\theta_{i}}^{h}\partial_{\xi_{i}}^{k}\partial_{\rho_{i}}^{\ell}f=0\text{ for all }h+k+\ell>0\right\}.

We also note that since any θi2=0\theta_{i}^{2}=0, we need not check any second derivatives, hence

(10) Tn={f[𝜽n,𝝃n,𝝆n]|i=1nθihξikρif=0 for all h+k+>0 and h,k,{0,1}}.T_{n}=\left\{f\in\mathbb{Q}[\bm{\theta}_{n},\bm{\xi}_{n},\bm{\rho}_{n}]\,\Bigg{|}\,\sum_{i=1}^{n}\partial_{\theta_{i}}^{h}\partial_{\xi_{i}}^{k}\partial_{\rho_{i}}^{\ell}f=0\text{ for all }h+k+\ell>0\text{ and }h,k,\ell\in\{0,1\}\right\}.

We wish to study TnT_{n} as a GL3\operatorname{GL}_{3}-module, so we look at the infinitesimal action of the Lie algebra 𝔰𝔩3\mathfrak{sl}_{3} on TnT_{n}. Similarly to [9, Section 3.1], we define the operators

(11) Fθξ:=i=1nξiθi, and Fξρ:=i=1nρiξi,F^{\theta\rightarrow\xi}:=\sum_{i=1}^{n}\xi_{i}\partial_{\theta_{i}},\text{ and }F^{\xi\rightarrow\rho}:=\sum_{i=1}^{n}\rho_{i}\partial_{\xi_{i}},

and

(12) Eθξ:=i=1nθiξi and Eξρ:=i=1nξiρi.E^{\theta\leftarrow\xi}:=\sum_{i=1}^{n}\theta_{i}\partial_{\xi_{i}}\text{ and }E^{\xi\leftarrow\rho}:=\sum_{i=1}^{n}\xi_{i}\partial_{\rho_{i}}.

Along with two HH operators given by taking the commutators of the EE and FF operators, these generate the Lie algebra 𝔰𝔩3\mathfrak{sl}_{3}.

Furthermore TnT_{n} is a GL3×𝔖n\operatorname{GL}_{3}\times\mathfrak{S}_{n}-module. Write a trigraded component of TnT_{n} as (Tn)a,b,c(T_{n})_{a,b,c} and Rn(0,3)R_{n}^{(0,3)} as (Rn(0,3))a,b,c(R_{n}^{(0,3)})_{a,b,c}. Then we have that TnRn(0,3)T_{n}\cong R_{n}^{(0,3)} as trigraded 𝔖n\mathfrak{S}_{n}-modules. Working with harmonics was advocated by Garsia: a benefit of working with harmonics over coinvariants is that one works with polynomials instead of equivalence classes.

Observe that the Schur function s(k2)(u,v,w)s_{(k-2)}(u,v,w) is a u,v,wu,v,w-analogue of the binomial coefficient (k2)\binom{k}{2}; denote it by

(13) (k2)u,v,w:=s(k2)(u,v,w).\binom{k}{2}_{u,v,w}:=s_{(k-2)}(u,v,w).

Recall the notation [n]u,v:=un1+un2v++uvn2+vn1[n]_{u,v}:=u^{n-1}+u^{n-2}v+\cdots+uv^{n-2}+v^{n-1}. It follows that as a polynomial in ww with coefficients in uu and vv we have

(14) (k2)u,v,w\displaystyle\binom{k}{2}_{u,v,w} =[k1]u,v+[k2]u,vw+[k3]u,vw2++[1]u,vwk2,\displaystyle=[k-1]_{u,v}+[k-2]_{u,v}w+[k-3]_{u,v}w^{2}+\cdots+[1]_{u,v}w^{k-2},

and there are two similar expressions for (k2)u,v,w\binom{k}{2}_{u,v,w} as a polynomial in uu or in vv.

A polynomial p[𝜽n,𝝃n,𝝆n]p\in\mathbb{Q}[\bm{\theta}_{n},\bm{\xi}_{n},\bm{\rho}_{n}] is called antisymmetric if σ(p)=sgn(σ)p\sigma(p)=\operatorname{sgn}(\sigma)p for all σ𝔖n\sigma\in\mathfrak{S}_{n}, where σ\sigma acts diagonally by permuting the variables. Let (Tn)ϵ(T_{n})^{\epsilon} denote the antisymmetric subspace222This is sometimes also referred to as the alternating subspace. of TnT_{n}, which consists of all elements pTnp\in T_{n} which are antisymmetric. This corresponds to the u,v,wu,v,w-graded multiplicity of the sign character in Rn(0,3)R_{n}^{(0,3)}, that is, Hilb((Tn)ϵ;u,v,w)=Frob(Rn(0,3);u,v,w),s(1n)\operatorname{Hilb}((T_{n})^{\epsilon};u,v,w)=\langle\operatorname{Frob}(R_{n}^{(0,3)};u,v,w),s_{(1^{n})}\rangle. Our main result is the following.

Theorem 1.1.
Frob(Rn(0,3);u,v,w),s(1n)\displaystyle\langle\operatorname{Frob}(R_{n}^{(0,3)};u,v,w),s_{(1^{n})}\rangle =s(n1)(u,v,w)+s(n2,1,1)(u,v,w)\displaystyle=s_{(n-1)}(u,v,w)+s_{(n-2,1,1)}(u,v,w)
=(n+12)u,v,w+uvw(n12)u,v,w.\displaystyle=\binom{n+1}{2}_{u,v,w}+uvw\binom{n-1}{2}_{u,v,w}.

The theorem immediately implies the following corollary, which was conjectured by Bergeron [2, Table 3].

Corollary 1.2.
Frob(Rn(0,3);1,1,1),s(1n)=n2n+1.\langle\operatorname{Frob}(R_{n}^{(0,3)};1,1,1),s_{(1^{n})}\rangle=n^{2}-n+1.
Proof.

Recall that (k2)u,v,w|u=v=w=1=(k2)\binom{k}{2}_{u,v,w}|_{u=v=w=1}=\binom{k}{2}. By evaluating Theorem 1.1 at u=v=w=1u=v=w=1, the dimension is (n+12)+(n12)=n2n+1\binom{n+1}{2}+\binom{n-1}{2}=n^{2}-n+1. ∎

The organization of the paper is as follows. In Section 2, we recall some results of Haglund–Sergel [8] and Kim–Rhoades [11], along with proving some preliminary results. In Section 3, building on the work of Haglund–Sergel and Kim–Rhoades, we give an upper bound on the multiplicity of the sign character of Rn(0,3)R_{n}^{(0,3)} (Corollary 3.2). In Section 4, we construct two elements in the ring of triagonal fermionic harmonics TnT_{n} (Proposition 4.4), and study the GL3\operatorname{GL}_{3}-representations that they generate (Proposition 4.6), which constructs enough elements in TnT_{n} to show that the upper bound on the multiplicity of the sign character is achieved with equality, proving the main theorem.

In Section 5, we derive a formula for double hook characters of Rn(0,2)R_{n}^{(0,2)} (Theorem 5.2). In Section 6, we discuss methods to analyze the sign character of Rn(0,4)R_{n}^{(0,4)}. In Section 7, we provide a q,u,v,wq,u,v,w-refinement of a conjecture of Bergeron [2] on the sign character of Rn(1,3)R_{n}^{(1,3)} (Conjecture 7.6).

2. Background and preliminary results

Haglund and Sergel gave a formula for the graded Frobenius series of the fermionic coinvariants Rn(0,1):=[𝜽n]/[𝜽n]+𝔖nR_{n}^{(0,1)}:=\mathbb{Q}[\bm{\theta}_{n}]/\langle\mathbb{Q}[\bm{\theta}_{n}]_{+}^{\mathfrak{S}_{n}}\rangle.

Lemma 2.1 ([8, Lemma 4.10]).

For n1n\geq 1,

Frob(Rn(0,1);w)=k=0n1wks(nk,1k).\operatorname{Frob}(R_{n}^{(0,1)};w)=\sum_{k=0}^{n-1}w^{k}s_{(n-k,1^{k})}.

Kim and Rhoades gave a formula for the bigraded Frobenius series of the diagonal fermionic coinvariants Rn(0,2):=[𝜽n,𝝃n]/[𝜽n,𝝃n]+𝔖nR_{n}^{(0,2)}:=\mathbb{Q}[\bm{\theta}_{n},\bm{\xi}_{n}]/\langle\mathbb{Q}[\bm{\theta}_{n},\bm{\xi}_{n}]_{+}^{\mathfrak{S}_{n}}\rangle.

Theorem 2.2 ([11, Theorem 6.1]).

For n1n\geq 1,

Frob(Rn(0,2);u,v)=0i+j<nuivj(s(ni,1i)s(nj,1j)s(ni+1,1i1)s(nj+1,1j1)),\operatorname{Frob}(R_{n}^{(0,2)};u,v)=\sum_{0\leq i+j<n}u^{i}v^{j}\left(s_{(n-i,1^{i})}*s_{(n-j,1^{j})}-s_{(n-i+1,1^{i-1})}*s_{(n-j+1,1^{j-1})}\right),

where * denotes the Kronecker product and s(n+1,11)s_{(n+1,1^{-1})} is interpreted as 0.

In particular, they found bigraded multiplicities for the trivial, sign, and hook characters.

Proposition 2.3 ([11, Proposition 6.2]).

In Rn(0,2)R_{n}^{(0,2)}, the multiplicity of the trivial character is

Frob(Rn(0,2);u,v),s(n)=1,\langle\operatorname{Frob}(R_{n}^{(0,2)};u,v),s_{(n)}\rangle=1,

the bigraded multiplicity of the sign character is

Frob(Rn(0,2);u,v),s(1n)=[n]u,v,\langle\operatorname{Frob}(R_{n}^{(0,2)};u,v),s_{(1^{n})}\rangle=[n]_{u,v},

and for 0<k<n10<k<n-1, the bigraded multiplicity of a hook character is

Frob(Rn(0,2);u,v),s(nk,1k)=[k+1]u,v+uv[k]u,v.\langle\operatorname{Frob}(R_{n}^{(0,2)};u,v),s_{(n-k,1^{k})}\rangle=[k+1]_{u,v}+uv[k]_{u,v}.

As a consequence of their result on the sign character, we conclude the following.

Corollary 2.4.
  1. (1)

    If a nonzero harmonic polynomial in Tn[𝜽n]T_{n}\cap\mathbb{Q}[\bm{\theta}_{n}] is antisymmetric, then it must be of degree exactly n1n-1.

  2. (2)

    If a nonzero harmonic polynomial in Tn[𝜽n,𝝃n]T_{n}\cap\mathbb{Q}[\bm{\theta}_{n},\bm{\xi}_{n}] is antisymmetric, then it must be of degree exactly n1n-1.

Proof.

The first claim is implied by the second. To show the second, notice that Proposition 2.3 shows that the dimension of each component of the sign character is n1n-1. ∎

We also will use the following result, which establishes that an antisymmetric polynomial must use enough distinct indices of variables to be nonzero.

Lemma 2.5.

If a polynomial in [𝛉n,𝛏n,𝛒n]\mathbb{Q}[\bm{\theta}_{n},\bm{\xi}_{n},\bm{\rho}_{n}] is antisymmetric, and for the variables that appear, at most n2n-2 unique indices are used, then the polynomial is 0.

Proof.

Without loss of generality, say that the indices used are {1,2,,n2}\{1,2,\ldots,n-2\}. Introduce the notation α1,β2,,ωn2\alpha_{1},\beta_{2},\ldots,\omega_{n-2} where each letter, when indexed by ii, is the (ordered) product of any nonempty subset of {θi,ξi,ρi}\{\theta_{i},\xi_{i},\rho_{i}\}. We use unique letters, so it is known which subset of {θi,ξi,ρi}\{\theta_{i},\xi_{i},\rho_{i}\} is being referred to after applying permutations.

Consider the symmetric group 𝔖n\mathfrak{S}_{n} as a set, and partition it 𝔖n=𝔖n𝔖n′′\mathfrak{S}_{n}=\mathfrak{S}_{n}^{\prime}\sqcup\mathfrak{S}_{n}^{\prime\prime} into two equal-sized subsets such that for each σ𝔖n\sigma^{\prime}\in\mathfrak{S}_{n}^{\prime}, we have that σsn1=σ′′\sigma^{\prime}\cdot s_{n-1}=\sigma^{\prime\prime}, for some σ′′𝔖n′′\sigma^{\prime\prime}\in\mathfrak{S}_{n}^{\prime\prime}, where sn1s_{n-1} denotes the simple transposition which interchanges n1n-1 and nn, and σsn1\sigma^{\prime}\cdot s_{n-1} will act on polynomials by first applying sn1s_{n-1} and then σ\sigma^{\prime}.

Any antisymmetric polynomial p[𝜽n,𝝃n,𝝆n]p\in\mathbb{Q}[\bm{\theta}_{n},\bm{\xi}_{n},\bm{\rho}_{n}] can be obtained by applying an antisymmetrizing operator σ𝔖nsgn(σ)σ\sum_{\sigma\in\mathfrak{S}_{n}}\operatorname{sgn}(\sigma)\sigma to some polynomial q[𝜽n,𝝃n,𝝆n]q\in\mathbb{Q}[\bm{\theta}_{n},\bm{\xi}_{n},\bm{\rho}_{n}]. Consider each monomial in qq, say α1β2ωn2=:r\alpha_{1}\beta_{2}\cdots\omega_{n-2}=:r. When σ\sigma^{\prime} and σ′′\sigma^{\prime\prime} are related by σsn1=σ′′\sigma^{\prime}\cdot s_{n-1}=\sigma^{\prime\prime}, we have that σ′′(r)=σsn1(r)=σ(r)\sigma^{\prime\prime}(r)=\sigma^{\prime}\cdot s_{n-1}(r)=\sigma^{\prime}(r), since rr does not contain any variables indexed by n1n-1 or nn. Hence, by antisymmetrizing the monomial rr, we get that

p\displaystyle p =σ𝔖nsgn(σ)σ(α1β2ωn2)\displaystyle=\sum_{\sigma\in\mathfrak{S}_{n}}\operatorname{sgn}(\sigma)\sigma(\alpha_{1}\beta_{2}\cdots\omega_{n-2})
=σ𝔖nsgn(σ)σ(α1β2ωn2)+σ′′𝔖n′′sgn(σ′′)σ′′(α1β2ωn2)\displaystyle=\sum_{\sigma^{\prime}\in\mathfrak{S}_{n}^{\prime}}\operatorname{sgn}(\sigma^{\prime})\sigma^{\prime}(\alpha_{1}\beta_{2}\cdots\omega_{n-2})+\sum_{\sigma^{\prime\prime}\in\mathfrak{S}_{n}^{\prime\prime}}\operatorname{sgn}(\sigma^{\prime\prime})\sigma^{\prime\prime}(\alpha_{1}\beta_{2}\cdots\omega_{n-2})
=σ𝔖nsgn(σ)σ(α1β2ωn2)+σ𝔖nsgn(σsn1)σ(α1β2ωn2)\displaystyle=\sum_{\sigma^{\prime}\in\mathfrak{S}_{n}^{\prime}}\operatorname{sgn}(\sigma^{\prime})\sigma^{\prime}(\alpha_{1}\beta_{2}\cdots\omega_{n-2})+\sum_{\sigma^{\prime}\in\mathfrak{S}_{n}^{\prime}}\operatorname{sgn}(\sigma^{\prime}\cdot s_{n-1})\sigma^{\prime}(\alpha_{1}\beta_{2}\cdots\omega_{n-2})
=σ𝔖n(sgn(σ)+sgn(σsn1))σ(α1β2ωn2)\displaystyle=\sum_{\sigma^{\prime}\in\mathfrak{S}_{n}^{\prime}}(\operatorname{sgn}(\sigma^{\prime})+\operatorname{sgn}(\sigma^{\prime}\cdot s_{n-1}))\sigma^{\prime}(\alpha_{1}\beta_{2}\cdots\omega_{n-2})
=0,\displaystyle=0,

where the last line follows since sn1s_{n-1} is a simple transposition, so σ\sigma^{\prime} and σsn1\sigma^{\prime}\cdot s_{n-1} have opposite signs. ∎

3. An upper bound on degree

The following result is inspired by a similar result of Haglund–Sergel (on a different ring, Rn(2,1)R_{n}^{(2,1)}) [8, Theorem 4.11].

Proposition 3.1.

For n1n\geq 1,

Frob(Rn(0,2)Rn(0,1);u,v,w),s(1n)=(n+12)u,v,w+uvw(n12)u,v,w+uv[nk1]u,v.\langle\operatorname{Frob}(R_{n}^{(0,2)}\otimes R_{n}^{(0,1)};u,v,w),s_{(1^{n})}\rangle=\binom{n+1}{2}_{u,v,w}+uvw\binom{n-1}{2}_{u,v,w}+uv[n-k-1]_{u,v}.
Proof.

As noted in [8, equation (4.9)], following from [4], for any 𝔖n\mathfrak{S}_{n}-modules AA and BB tensored under the diagonal action of 𝔖n\mathfrak{S}_{n}, we have that

Frob(AB)=νnsνλ,μnFrob(A),sλFrob(B),sμsλsμ,sν,\operatorname{Frob}(A\otimes B)=\sum_{\nu\vdash n}s_{\nu}\sum_{\lambda,\mu\vdash n}\langle\operatorname{Frob}(A),s_{\lambda}\rangle\langle\operatorname{Frob}(B),s_{\mu}\rangle\langle s_{\lambda}*s_{\mu},s_{\nu}\rangle,

where sλsμ,sν=g(λ,μ,ν)\langle s_{\lambda}*s_{\mu},s_{\nu}\rangle=g(\lambda,\mu,\nu) are the Kronecker coefficients. In our case, we are interested in

Frob(Rn(0,2)Rn(0,1);u,v,w),s(1n)\displaystyle\langle\operatorname{Frob}(R_{n}^{(0,2)}\otimes R_{n}^{(0,1)};u,v,w),s_{(1^{n})}\rangle =λ,μnFrob(Rn(0,2);u,v),sλFrob(Rn(0,1);w),sμg(λ,μ,1n)\displaystyle=\sum_{\lambda,\mu\vdash n}\langle\operatorname{Frob}(R_{n}^{(0,2)};u,v),s_{\lambda}\rangle\langle\operatorname{Frob}(R_{n}^{(0,1)};w),s_{\mu}\rangle g(\lambda,\mu,1^{n})
(15) =λnFrob(Rn(0,2);u,v),sλFrob(Rn(0,1);w),sλ,\displaystyle=\sum_{\lambda\vdash n}\langle\operatorname{Frob}(R_{n}^{(0,2)};u,v),s_{\lambda^{\prime}}\rangle\langle\operatorname{Frob}(R_{n}^{(0,1)};w),s_{\lambda}\rangle,

since g(λ,μ,1n)=δμ,λg(\lambda,\mu,1^{n})=\delta_{\mu,\lambda^{\prime}}. By Lemma 2.1, which states that only hook Schur functions appear in Frob(Rn(0,1);w)\operatorname{Frob}(R_{n}^{(0,1)};w), we reduce to

Frob(Rn(0,2)Rn(0,1);u,v,w),s(1n)\displaystyle\langle\operatorname{Frob}(R_{n}^{(0,2)}\otimes R_{n}^{(0,1)};u,v,w),s_{(1^{n})}\rangle =k=0n1wkFrob(Rn(0,2);u,v),s(k+1,1nk1)\displaystyle=\sum_{k=0}^{n-1}w^{k}\langle\operatorname{Frob}(R_{n}^{(0,2)};u,v),s_{(k+1,1^{n-k-1})}\rangle
(16) =k=0n1wk([nk]u,v+uv[nk1]u,v),\displaystyle=\sum_{k=0}^{n-1}w^{k}([n-k]_{u,v}+uv[n-k-1]_{u,v}),

where we applied Proposition 2.3. After a bit of algebra, we obtain the claimed formula. ∎

Now the following result gives us an upper bound on the occurrences of the sign character in Rn(0,3)R_{n}^{(0,3)}. For multivariate polynomials AA and BB, the notation ABA\leq B means that BAB-A is a sum of monomials with only nonnegative coefficients.

Corollary 3.2.

For n1n\geq 1,

Frob(Rn(0,3);u,v,w),s(1n)(n+12)u,v,w+uvw(n12)u,v,w.\langle\operatorname{Frob}(R_{n}^{(0,3)};u,v,w),s_{(1^{n})}\rangle\leq\binom{n+1}{2}_{u,v,w}+uvw\binom{n-1}{2}_{u,v,w}.
Proof.

First note that any occurrence of the sign character in Rn(0,3)R_{n}^{(0,3)} is predicated on it occurring in Rn(0,2)Rn(0,1)R_{n}^{(0,2)}\otimes R_{n}^{(0,1)}, since Rn(0,3)R_{n}^{(0,3)} is a quotient of Rn(0,2)Rn(0,1)R_{n}^{(0,2)}\otimes R_{n}^{(0,1)} under the diagonal action of 𝔖n\mathfrak{S}_{n} (see [8]).

By permuting which sets of variables out of 𝜽n,𝝃n,𝝆n\bm{\theta}_{n},\bm{\xi}_{n},\bm{\rho}_{n} are assigned to Rn(0,2)R_{n}^{(0,2)} and to Rn(0,1)R_{n}^{(0,1)} in Proposition 3.1, we conclude that a sign character in Rn(0,3)R_{n}^{(0,3)} must have trigraded multiplicity bounded above by

(17) (n+12)u,v,w+uvw(n12)u,v,w+uv[nk1]u,v,\binom{n+1}{2}_{u,v,w}+uvw\binom{n-1}{2}_{u,v,w}+uv[n-k-1]_{u,v},
(18) (n+12)u,v,w+uvw(n12)u,v,w+uw[nk1]u,w,\binom{n+1}{2}_{u,v,w}+uvw\binom{n-1}{2}_{u,v,w}+uw[n-k-1]_{u,w},

and

(19) (n+12)u,v,w+uvw(n12)u,v,w+vw[nk1]v,w.\binom{n+1}{2}_{u,v,w}+uvw\binom{n-1}{2}_{u,v,w}+vw[n-k-1]_{v,w}.

Note that if a polynomial AA satisfies Auv[nk1]u,vA\leq uv[n-k-1]_{u,v}, Auw[nk1]u,wA\leq uw[n-k-1]_{u,w}, and Avw[nk1]v,wA\leq vw[n-k-1]_{v,w}, then A=0A=0. Hence by taking the bounds given by equations (17-19) together, we conclude that the sign character in Rn(0,3)R_{n}^{(0,3)} must have trigraded multiplicity bounded above by

(20) (n+12)u,v,w+uvw(n12)u,v,w.\binom{n+1}{2}_{u,v,w}+uvw\binom{n-1}{2}_{u,v,w}.

4. Construction of basis elements

Now we will work towards the proof of the main theorem, by constructing two explicit elements in TnT_{n}, which are highest weight vectors for certain GL3\operatorname{GL}_{3}-representations. We ultimately show that the upper bound given in Corollary 3.2 is obtained with equality.

Definition 4.1.

For n1n\geq 1, define the primary theta-seed by

Δ1(𝜽n):=σ𝔖nsgn(σ)σ(θ1θ2θn1).\Delta_{1}(\bm{\theta}_{n}):=\sum_{\sigma\in\mathfrak{S}_{n}}\operatorname{sgn}(\sigma)\sigma(\theta_{1}\theta_{2}\cdots\theta_{n-1}).
Definition 4.2.

For n3n\geq 3, define the secondary theta-seed by

Δ2(𝜽n):=σ𝔖nsgn(σ)σ((θ1ξ2ρ2+θ2ξ1ρ2+θ2ξ2ρ1)θ3θ4θn1).\Delta_{2}(\bm{\theta}_{n}):=\sum_{\sigma\in\mathfrak{S}_{n}}\operatorname{sgn}(\sigma)\sigma((\theta_{1}\xi_{2}\rho_{2}+\theta_{2}\xi_{1}\rho_{2}+\theta_{2}\xi_{2}\rho_{1})\theta_{3}\theta_{4}\cdots\theta_{n-1}).

When n=1n=1 or 33 in the definitions of the primary and secondary theta-seed, respectively, the empty product of θi\theta_{i}’s is interpreted as 11.

We now prove a technical lemma.

Lemma 4.3.

The antisymmetrization operator σ𝔖nsgn(σ)σ\sum_{\sigma\in\mathfrak{S}_{n}}\operatorname{sgn}(\sigma)\sigma commutes with the differential operator i=1nθihξikρi\sum_{i=1}^{n}\partial_{\theta_{i}}^{h}\partial_{\xi_{i}}^{k}\partial_{\rho_{i}}^{\ell}.

Proof.

Let f[𝜽n,𝝃n,𝝆n]f\in\mathbb{Q}[\bm{\theta}_{n},\bm{\xi}_{n},\bm{\rho}_{n}]. Observe that for any σ𝔖n\sigma\in\mathfrak{S}_{n},

(21) i=1nθihξikρi=i=1nσ(θi)hσ(ξi)kσ(ρi),\sum_{i=1}^{n}\partial_{\theta_{i}}^{h}\partial_{\xi_{i}}^{k}\partial_{\rho_{i}}^{\ell}=\sum_{i=1}^{n}\partial_{\sigma(\theta_{i})}^{h}\partial_{\sigma(\xi_{i})}^{k}\partial_{\sigma(\rho_{i})}^{\ell},

since addition is commutative. Thus,

i=1nθihξikρi(σ𝔖nsgn(σ)σ(f))\displaystyle\sum_{i=1}^{n}\partial_{\theta_{i}}^{h}\partial_{\xi_{i}}^{k}\partial_{\rho_{i}}^{\ell}\left(\sum_{\sigma\in\mathfrak{S}_{n}}\operatorname{sgn}(\sigma)\sigma(f)\right) =σ𝔖nsgn(σ)i=1nθihξikρi(σ(f))\displaystyle=\sum_{\sigma\in\mathfrak{S}_{n}}\operatorname{sgn}(\sigma)\sum_{i=1}^{n}\partial_{\theta_{i}}^{h}\partial_{\xi_{i}}^{k}\partial_{\rho_{i}}^{\ell}(\sigma(f))
=σ𝔖nsgn(σ)i=1nσ(θi)hσ(ξi)kσ(ρi)(σ(f))\displaystyle=\sum_{\sigma\in\mathfrak{S}_{n}}\operatorname{sgn}(\sigma)\sum_{i=1}^{n}\partial_{\sigma(\theta_{i})}^{h}\partial_{\sigma(\xi_{i})}^{k}\partial_{\sigma(\rho_{i})}^{\ell}(\sigma(f))
(22) =σ𝔖nsgn(σ)σ(i=1nθihξikρif),\displaystyle=\sum_{\sigma\in\mathfrak{S}_{n}}\operatorname{sgn}(\sigma)\sigma\left(\sum_{i=1}^{n}\partial_{\theta_{i}}^{h}\partial_{\xi_{i}}^{k}\partial_{\rho_{i}}^{\ell}f\right),

so the operators commute as claimed. ∎

Now we show that Δ1(𝜽n)\Delta_{1}(\bm{\theta}_{n}) and Δ2(𝜽n)\Delta_{2}(\bm{\theta}_{n}) are harmonic.

Proposition 4.4.
  1. (1)

    The primary theta-seed Δ1(𝜽n)\Delta_{1}(\bm{\theta}_{n}) is in TnT_{n}.

  2. (2)

    The secondary theta-seed Δ2(𝜽n)\Delta_{2}(\bm{\theta}_{n}) is in TnT_{n}.

Proof.

We show (1). Consider the primary theta-seed

Δ1(𝜽n)=σ𝔖nsgn(σ)σ(θ1θ2θn1).\Delta_{1}(\bm{\theta}_{n})=\sum_{\sigma\in\mathfrak{S}_{n}}\operatorname{sgn}(\sigma)\sigma(\theta_{1}\theta_{2}\cdots\theta_{n-1}).

As discussed in the introduction, we only need to show that i=1nθihξikρiΔ1(𝜽n)=0\sum_{i=1}^{n}\partial_{\theta_{i}}^{h}\partial_{\xi_{i}}^{k}\partial_{\rho_{i}}^{\ell}\Delta_{1}(\bm{\theta}_{n})=0 for (h,k,){(1,0,0),(0,1,0),(0,0,1),(1,1,0),(1,0,1),(0,1,1),(1,1,1)}.(h,k,\ell)\in\{(1,0,0),(0,1,0),(0,0,1),(1,1,0),(1,0,1),(0,1,1),(1,1,1)\}. Since only θi\theta_{i} variables appear in Δ1(𝜽n)\Delta_{1}(\bm{\theta}_{n}), it only remains to check (1,0,0)(1,0,0). Consider θ1θn1\theta_{1}\cdots\theta_{n-1}, which is what we will antisymmetrize to get the primary theta-seed. By Lemma 4.3, the differential operator commutes with the antisymmetrization operator. We write

(23) i=1nθiθ1θn1=i=1n(1)i1θ1θ^iθn1,\sum_{i=1}^{n}\partial_{\theta_{i}}\theta_{1}\cdots\theta_{n-1}=\sum_{i=1}^{n}(-1)^{i-1}\theta_{1}\cdots\hat{\theta}_{i}\cdots\theta_{n-1},

where each monomial on the right hand side is in only n2n-2 variables. Upon antisymmetrization, by Lemma 2.5, this becomes 0.

We show (2). For n3n\geq 3, consider the secondary theta-seed

Δ2(𝜽n)=σ𝔖nsgn(σ)σ((θ1ξ2ρ2+θ2ξ1ρ2+θ2ξ2ρ1)θ3θ4θn1).\Delta_{2}(\bm{\theta}_{n})=\sum_{\sigma\in\mathfrak{S}_{n}}\operatorname{sgn}(\sigma)\sigma((\theta_{1}\xi_{2}\rho_{2}+\theta_{2}\xi_{1}\rho_{2}+\theta_{2}\xi_{2}\rho_{1})\theta_{3}\theta_{4}\cdots\theta_{n-1}).

We must check that Δ2(𝜽n)\Delta_{2}(\bm{\theta}_{n}) is harmonic. We only need to show that i=1nθihξikρiΔ2(𝜽n)=0\sum_{i=1}^{n}\partial_{\theta_{i}}^{h}\partial_{\xi_{i}}^{k}\partial_{\rho_{i}}^{\ell}\Delta_{2}(\bm{\theta}_{n})=0 for the following seven choices of (h,k,)(h,k,\ell):

{(1,0,0),(0,1,0),(0,0,1),(1,1,0),(1,0,1),(0,1,1),(1,1,1)}.\{(1,0,0),(0,1,0),(0,0,1),(1,1,0),(1,0,1),(0,1,1),(1,1,1)\}.

Case 1: (h,k,)=(1,0,0)(h,k,\ell)=(1,0,0). Let f=(θ1ξ2ρ2+θ2ξ1ρ2+θ2ξ2ρ1)θ3θ4θn1f=(\theta_{1}\xi_{2}\rho_{2}+\theta_{2}\xi_{1}\rho_{2}+\theta_{2}\xi_{2}\rho_{1})\theta_{3}\theta_{4}\cdots\theta_{n-1}, which is what we will antisymmetrize to get the secondary theta-seed. For concision, write i=3n1θi\prod_{i=3}^{n-1}\theta_{i} for the ordered product θ3θ4θn1\theta_{3}\theta_{4}\cdots\theta_{n-1}. We write that

i=1nθif\displaystyle\sum_{i=1}^{n}\partial_{\theta_{i}}f =θ1f+θ2f+i=3n1θif\displaystyle=\partial_{\theta_{1}}f+\partial_{\theta_{2}}f+\sum_{i=3}^{n-1}\partial_{\theta_{i}}f
(24) =ξ2ρ2i=3n1θi+(ξ1ρ2+ξ2ρ1)i=3n1θi+i=3n1(1)i(θ1ξ2ρ2+θ2ξ1ρ2+θ2ξ2ρ1)θ3θ^iθn1.\displaystyle=\xi_{2}\rho_{2}\prod_{i=3}^{n-1}\theta_{i}+(\xi_{1}\rho_{2}+\xi_{2}\rho_{1})\prod_{i=3}^{n-1}\theta_{i}+\sum_{i=3}^{n-1}(-1)^{i}(\theta_{1}\xi_{2}\rho_{2}+\theta_{2}\xi_{1}\rho_{2}+\theta_{2}\xi_{2}\rho_{1})\theta_{3}\cdots\hat{\theta}_{i}\cdots\theta_{n-1}.

Then upon antisymmetrization, both of the terms ξ2ρ2i=3n1θi\xi_{2}\rho_{2}\prod_{i=3}^{n-1}\theta_{i} and i=3n1(1)i(θ1ξ2ρ2+θ2ξ1ρ2+θ2ξ2ρ1)θ3θ^iθn1\sum_{i=3}^{n-1}(-1)^{i}(\theta_{1}\xi_{2}\rho_{2}+\theta_{2}\xi_{1}\rho_{2}+\theta_{2}\xi_{2}\rho_{1})\theta_{3}\cdots\hat{\theta}_{i}\cdots\theta_{n-1} will be 0 due to Lemma 2.5, since they only contain n2n-2 indices for variables.

For the term (ξ1ρ2+ξ2ρ1)i=3n1θi(\xi_{1}\rho_{2}+\xi_{2}\rho_{1})\prod_{i=3}^{n-1}\theta_{i}, there are n1n-1 indices for variables, so we cannot apply Lemma 2.5 directly. However, we can apply a similar argument as in the proof of Lemma 2.5. Partition 𝔖n=𝔖n𝔖n′′\mathfrak{S}_{n}=\mathfrak{S}_{n}^{\prime}\sqcup\mathfrak{S}_{n}^{\prime\prime} into two equal-sized subsets such that for each σ𝔖n\sigma^{\prime}\in\mathfrak{S}_{n}^{\prime}, we have that σs1=σ′′\sigma^{\prime}\cdot s_{1}=\sigma^{\prime\prime}, where s1s_{1} denotes the simple transposition which interchanges 1 and 2. Upon symmetrization, we write

σ𝔖n\displaystyle\sum_{\sigma\in\mathfrak{S}_{n}} sgn(σ)σ((ξ1ρ2+ξ2ρ1)θ3θn1)\displaystyle\operatorname{sgn}(\sigma)\sigma((\xi_{1}\rho_{2}+\xi_{2}\rho_{1})\theta_{3}\cdots\theta_{n-1})
=σ𝔖nsgn(σ)σ((ξ1ρ2+ξ2ρ1)θ3θn1)+σ′′𝔖n′′sgn(σ′′)σ′′((ξ1ρ2+ξ2ρ1)θ3θn1)\displaystyle=\sum_{\sigma^{\prime}\in\mathfrak{S}_{n}^{\prime}}\operatorname{sgn}(\sigma^{\prime})\sigma^{\prime}((\xi_{1}\rho_{2}+\xi_{2}\rho_{1})\theta_{3}\cdots\theta_{n-1})+\sum_{\sigma^{\prime\prime}\in\mathfrak{S}_{n}^{\prime\prime}}\operatorname{sgn}(\sigma^{\prime\prime})\sigma^{\prime\prime}((\xi_{1}\rho_{2}+\xi_{2}\rho_{1})\theta_{3}\cdots\theta_{n-1})
=σ𝔖nsgn(σ)σ((ξ1ρ2+ξ2ρ1)θ3θn1)+σ𝔖nsgn(σs1)(σs1)((ξ1ρ2+ξ2ρ1)θ3θn1)\displaystyle=\sum_{\sigma^{\prime}\in\mathfrak{S}_{n}^{\prime}}\operatorname{sgn}(\sigma^{\prime})\sigma^{\prime}((\xi_{1}\rho_{2}+\xi_{2}\rho_{1})\theta_{3}\cdots\theta_{n-1})+\sum_{\sigma^{\prime}\in\mathfrak{S}_{n}^{\prime}}\operatorname{sgn}(\sigma^{\prime}\cdot s_{1})(\sigma^{\prime}\cdot s_{1})((\xi_{1}\rho_{2}+\xi_{2}\rho_{1})\theta_{3}\cdots\theta_{n-1})
=σ𝔖nsgn(σ)σ((ξ1ρ2+ξ2ρ1)θ3θn1)+σ𝔖nsgn(σ)σ((ξ2ρ1+ξ1ρ2)θ3θn1)\displaystyle=\sum_{\sigma^{\prime}\in\mathfrak{S}_{n}^{\prime}}\operatorname{sgn}(\sigma^{\prime})\sigma^{\prime}((\xi_{1}\rho_{2}+\xi_{2}\rho_{1})\theta_{3}\cdots\theta_{n-1})+\sum_{\sigma^{\prime}\in\mathfrak{S}_{n}^{\prime}}-\operatorname{sgn}(\sigma^{\prime})\sigma^{\prime}((\xi_{2}\rho_{1}+\xi_{1}\rho_{2})\theta_{3}\cdots\theta_{n-1})
(25) =0.\displaystyle=0.

Case 2: (h,k,)=(0,1,0)(h,k,\ell)=(0,1,0). Before antisymmetrization, we write that

i=1nξif\displaystyle\sum_{i=1}^{n}\partial_{\xi_{i}}f =ξ1f+ξ2f\displaystyle=\partial_{\xi_{1}}f+\partial_{\xi_{2}}f
=θ2ρ2i=3n1θi(θ1ρ2+θ2ρ1)i=3n1θi.\displaystyle=-\theta_{2}\rho_{2}\prod_{i=3}^{n-1}\theta_{i}-(\theta_{1}\rho_{2}+\theta_{2}\rho_{1})\prod_{i=3}^{n-1}\theta_{i}.

Observe that θ2ρ2i=3n1θi\theta_{2}\rho_{2}\prod_{i=3}^{n-1}\theta_{i} becomes 0 upon antisymmetrization, by Lemma 2.5. Additionally, (θ1ρ2+θ2ρ1)i=3n1θi(\theta_{1}\rho_{2}+\theta_{2}\rho_{1})\prod_{i=3}^{n-1}\theta_{i} becomes 0 upon antisymmetrization by the same argument used in Case 1.

Case 3: (h,k,)=(0,0,1)(h,k,\ell)=(0,0,1). The same argument as in Case 2 applies.

Case 4: (h,k,)=(1,1,0)(h,k,\ell)=(1,1,0). Before antisymmetrization, we write that

i=1nθiξif\displaystyle\sum_{i=1}^{n}\partial_{\theta_{i}}\partial_{\xi_{i}}f =θ1ξ1f+θ2ξ2f+i=3n1θiξif\displaystyle=\partial_{\theta_{1}}\partial_{\xi_{1}}f+\partial_{\theta_{2}}\partial_{\xi_{2}}f+\sum_{i=3}^{n-1}\partial_{\theta_{i}}\partial_{\xi_{i}}f
=0ρ1i=3n1θi+0,\displaystyle=0-\rho_{1}\prod_{i=3}^{n-1}\theta_{i}+0,

which becomes 0 upon antisymmetrization by Lemma 2.5.

Case 5: (h,k,)=(1,0,1)(h,k,\ell)=(1,0,1). The same argument as in Case 4 applies.

Case 6: (h,k,)=(0,1,1)(h,k,\ell)=(0,1,1). The same argument as in Case 4 applies.

Case 7: (h,k,)=(1,1,1)(h,k,\ell)=(1,1,1). Before antisymmetrization, we write that

i=1nθiξiρif\displaystyle\sum_{i=1}^{n}\partial_{\theta_{i}}\partial_{\xi_{i}}\partial_{\rho_{i}}f =θ1ξ1ρ1f+θ2ξ2ρ2f+i=3n1θiξiρif\displaystyle=\partial_{\theta_{1}}\partial_{\xi_{1}}\partial_{\rho_{1}}f+\partial_{\theta_{2}}\partial_{\xi_{2}}\partial_{\rho_{2}}f+\sum_{i=3}^{n-1}\partial_{\theta_{i}}\partial_{\xi_{i}}\partial_{\rho_{i}}f
=0,\displaystyle=0,

which remains 0 upon antisymmetrization.

This completes the proof that the secondary theta-seed is harmonic. ∎

A priori, a harmonic polynomial in TnT_{n} could equal zero if certain term cancellations occur. However, the following result demonstrates that this does not happen for Δ1(𝜽n)\Delta_{1}(\bm{\theta}_{n}) and Δ2(𝜽n)\Delta_{2}(\bm{\theta}_{n}).

Proposition 4.5.
  1. (1)

    The primary theta-seed Δ1(𝜽n)\Delta_{1}(\bm{\theta}_{n}) is nonzero in TnT_{n}.

  2. (2)

    The secondary theta-seed Δ2(𝜽n)\Delta_{2}(\bm{\theta}_{n}) is nonzero in TnT_{n}.

Proof.

We show (1). Consider in TnT_{n},

Δ1(𝜽n)=σ𝔖nsgn(σ)σ(θ1θ2θn1).\Delta_{1}(\bm{\theta}_{n})=\sum_{\sigma\in\mathfrak{S}_{n}}\operatorname{sgn}(\sigma)\sigma(\theta_{1}\theta_{2}\cdots\theta_{n-1}).

When the antisymmetrization operator σ𝔖nsgn(σ)σ\sum_{\sigma\in\mathfrak{S}_{n}}\operatorname{sgn}(\sigma)\sigma is applied to θ1θ2θn1\theta_{1}\theta_{2}\cdots\theta_{n-1}, consider which σ\sigma will output the monomial θ1θ2θn1\theta_{1}\theta_{2}\cdots\theta_{n-1}. Such a σ\sigma must fix nn, but 1,,n11,\ldots,n-1 can be permuted. For any σ\sigma in the symmetric group on letters 1,,n11,\ldots,n-1, we have that sgn(σ)σ(θ1θ2θn1)=θ1θ2θn1\operatorname{sgn}(\sigma)\sigma(\theta_{1}\theta_{2}\cdots\theta_{n-1})=\theta_{1}\theta_{2}\cdots\theta_{n-1} (see for example [5, Chapter III, Section 7.3, Proposition 5]). Hence antisymmetrization produces (n1)!(n-1)! copies of θ1θ2θn1\theta_{1}\theta_{2}\cdots\theta_{n-1}, so Δ1(𝜽n)\Delta_{1}(\bm{\theta}_{n}) is nonzero in TnT_{n}, as desired.

Next we show (2). Consider in TnT_{n},

Δ2(𝜽n)=σ𝔖nsgn(σ)σ((θ1ξ2ρ2+θ2ξ1ρ2+θ2ξ2ρ1)θ3θ4θn1).\Delta_{2}(\bm{\theta}_{n})=\sum_{\sigma\in\mathfrak{S}_{n}}\operatorname{sgn}(\sigma)\sigma((\theta_{1}\xi_{2}\rho_{2}+\theta_{2}\xi_{1}\rho_{2}+\theta_{2}\xi_{2}\rho_{1})\theta_{3}\theta_{4}\cdots\theta_{n-1}).

When the operator σ𝔖nsgn(σ)σ\sum_{\sigma\in\mathfrak{S}_{n}}\operatorname{sgn}(\sigma)\sigma is applied to (θ1ξ2ρ2+θ2ξ1ρ2+θ2ξ2ρ1)θ3θ4θn1(\theta_{1}\xi_{2}\rho_{2}+\theta_{2}\xi_{1}\rho_{2}+\theta_{2}\xi_{2}\rho_{1})\theta_{3}\theta_{4}\cdots\theta_{n-1}, consider which σ\sigma will include the monomial θ1ξ2ρ2θ3θ4θn1\theta_{1}\xi_{2}\rho_{2}\theta_{3}\theta_{4}\cdots\theta_{n-1} in its output. Since the ξi\xi_{i} and ρi\rho_{i} have the same index, it must come from an antisymmetrization of θ1ξ2ρ2θ3θ4θn1\theta_{1}\xi_{2}\rho_{2}\theta_{3}\theta_{4}\cdots\theta_{n-1} (and not of θ2ξ1ρ2θ3θ4θn1\theta_{2}\xi_{1}\rho_{2}\theta_{3}\theta_{4}\cdots\theta_{n-1} nor θ2ξ2ρ1θ3θ4θn1\theta_{2}\xi_{2}\rho_{1}\theta_{3}\theta_{4}\cdots\theta_{n-1}). Thus such a σ\sigma must fix 22 and nn, but 1,3,4,,n11,3,4,\ldots,n-1 can be permuted. For any σ\sigma in the symmetric group on letters 1,3,4,,n11,3,4,\ldots,n-1, it follows that sgn(σ)σ(θ1θ3θ4θn1)=θ1θ3θ4θn1\operatorname{sgn}(\sigma)\sigma(\theta_{1}\theta_{3}\theta_{4}\cdots\theta_{n-1})=\theta_{1}\theta_{3}\theta_{4}\cdots\theta_{n-1}. This implies that sgn(σ)σ(θ1ξ2ρ2θ3θ4θn1)=θ1ξ2ρ2θ3θ4θn1\operatorname{sgn}(\sigma)\sigma(\theta_{1}\xi_{2}\rho_{2}\theta_{3}\theta_{4}\cdots\theta_{n-1})=\theta_{1}\xi_{2}\rho_{2}\theta_{3}\theta_{4}\cdots\theta_{n-1}. Hence antisymmetrization produces (n2)!(n-2)! copies of θ1ξ2ρ2θ3θ4θn1\theta_{1}\xi_{2}\rho_{2}\theta_{3}\theta_{4}\cdots\theta_{n-1}, so Δ2(𝜽n)\Delta_{2}(\bm{\theta}_{n}) is nonzero in TnT_{n}, as desired. ∎

Starting with Δ1(𝜽n)\Delta_{1}(\bm{\theta}_{n}) and Δ2(𝜽n)\Delta_{2}(\bm{\theta}_{n}), the operators defined in equation (11) FθξF^{\theta\rightarrow\xi} and FξρF^{\xi\rightarrow\rho} create GL3\operatorname{GL}_{3}-representations.

Proposition 4.6.
  1. (1)

    The primary theta-seed Δ1(𝜽n)\Delta_{1}(\bm{\theta}_{n}) is the highest weight vector for the GL3\operatorname{GL}_{3}-representation with character s(n1)(u,v,w)s_{(n-1)}(u,v,w).

  2. (2)

    The secondary theta-seed Δ2(𝜽n)\Delta_{2}(\bm{\theta}_{n}) is the highest weight vector for the GL3\operatorname{GL}_{3}-representation with character s(n2,1,1)(u,v,w)s_{(n-2,1,1)}(u,v,w).

Proof.

We show (1). Start with the primary theta-seed

Δ1(𝜽n)=σ𝔖nsgn(σ)σ(θ1θ2θn1).\Delta_{1}(\bm{\theta}_{n})=\sum_{\sigma\in\mathfrak{S}_{n}}\operatorname{sgn}(\sigma)\sigma(\theta_{1}\theta_{2}\cdots\theta_{n-1}).

Since we constructed it using an antisymmetrizing operator, it is antisymmetric. By Proposition 4.4, it is in TnT_{n}, and by Proposition 4.5, it is nonzero. Observe that EθξE^{\theta\leftarrow\xi} and EξρE^{\xi\leftarrow\rho} both kill Δ1(𝜽n)\Delta_{1}(\bm{\theta}_{n}). Thus Δ1(𝜽n)\Delta_{1}(\bm{\theta}_{n}) is a highest weight vector for the irreducible GL3\operatorname{GL}_{3}-representation with highest weight (n1,0,0)(n-1,0,0), where the weight of any pTnp\in T_{n} is given by (degθp,degξp,degρp)(\deg_{\theta}p,\deg_{\xi}p,\deg_{\rho}p). Thus the GL3\operatorname{GL}_{3}-character of this representation is s(n1)(u,v,w)s_{(n-1)}(u,v,w).

We show (2). Consider the secondary theta-seed

Δ2(𝜽n)=σ𝔖nsgn(σ)σ((θ1ξ2ρ2+θ2ξ1ρ2+θ2ξ2ρ1)θ3θ4θn1).\Delta_{2}(\bm{\theta}_{n})=\sum_{\sigma\in\mathfrak{S}_{n}}\operatorname{sgn}(\sigma)\sigma((\theta_{1}\xi_{2}\rho_{2}+\theta_{2}\xi_{1}\rho_{2}+\theta_{2}\xi_{2}\rho_{1})\theta_{3}\theta_{4}\cdots\theta_{n-1}).

Note that this is antisymmetric, as it is constructed using an antisymmetrization operator. By Corollary 2.4, there are no antisymmetric elements of degree nn solely in any one or two sets of variables. This implies that EθξE^{\theta\leftarrow\xi} and EξρE^{\xi\leftarrow\rho} both kill Δ2(𝜽n)\Delta_{2}(\bm{\theta}_{n}). By Proposition 4.4, it is in TnT_{n}, and by Proposition 4.5, it is nonzero. Thus Δ2(𝜽n)\Delta_{2}(\bm{\theta}_{n}) is a highest weight vector for the irreducible GL3\operatorname{GL}_{3}-representation with highest weight (n2,1,1)(n-2,1,1). Thus the GL3\operatorname{GL}_{3}-character of this representation is s(n2,1,1)(u,v,w)s_{(n-2,1,1)}(u,v,w). ∎

Now we can prove the main theorem.

Theorem 1.1.
Frob(Rn(0,3);u,v,w),s(1n)\displaystyle\langle\operatorname{Frob}(R_{n}^{(0,3)};u,v,w),s_{(1^{n})}\rangle =s(n1)(u,v,w)+s(n2,1,1)(u,v,w)\displaystyle=s_{(n-1)}(u,v,w)+s_{(n-2,1,1)}(u,v,w)
=(n+12)u,v,w+uvw(n12)u,v,w.\displaystyle=\binom{n+1}{2}_{u,v,w}+uvw\binom{n-1}{2}_{u,v,w}.
Proof.

The representations constructed in Proposition 4.6 are sufficient to force the bound in Corollary 3.2 to be achieved with equality. A bit of algebra verifies that the formulation in terms of Schur functions is equivalent to the formulation in terms of u,v,wu,v,w-binomial coefficients. ∎

Remark 4.7.

While it is not necessarily true in general that the Lie algebra operators FF and EE will correspond to crystal operators, in the present case, it is true since every weight space for the representations studied in Proposition 4.6 is one-dimensional. The representation with highest weight (n1,0,0)(n-1,0,0) has crystal structure isomorphic to a crystal of tableaux (n1)\mathcal{B}_{(n-1)} and the representation with highest weight (n2,1,1)(n-2,1,1) has crystal structure isomorphic to a crystal of tableaux (n2,1,1)\mathcal{B}_{(n-2,1,1)} (see for example [6, Chapter 3]).

5. Double hook characters of the diagonal fermionic coinvariant ring

With Proposition 2.3, Kim and Rhoades gave explicit formulas for the trivial, sign, and hook characters of Rn(0,2)R_{n}^{(0,2)}. In this section, we extend the analysis to give an explicit formula for double hook characters of Rn(0,2)R_{n}^{(0,2)}. All characters of Rn(0,2)R_{n}^{(0,2)} indexed by shapes not contained in a double hook are 0.

Rosas gave a combinatorial formula for the Kronecker coefficients g(λ,(ne,1e),(nf,1f))g(\lambda,(n-e,1^{e}),(n-f,1^{f})), for any shape λ\lambda. These Kronecker coefficients are only nonzero if λ\lambda is contained in a double hook shape. Here, we recall her formulas for double hook shapes, rows, and hook shapes.

Theorem 5.1 ([14, Theorem 3]).

Let (ne,1e)(n-e,1^{e}) and (nf,1f)(n-f,1^{f}) be hook shapes.

  1. (1)

    If λ\lambda is not contained in a double hook, then g(λ,(ne,1e),(nf,1f))=0g(\lambda,(n-e,1^{e}),(n-f,1^{f}))=0.

  2. (2)

    Let λ=(λ1,λ2,2,1n2λ1λ2)\lambda=(\lambda_{1},\lambda_{2},2^{\ell},1^{n-2\ell-\lambda_{1}-\lambda_{2}}) where λ1λ22\lambda_{1}\geq\lambda_{2}\geq 2 be a double hook. Then

    g(\displaystyle g( λ,(ne,1e),(nf,1f))\displaystyle\lambda,(n-e,1^{e}),(n-f,1^{f}))
    =χ(λ21e+fn+λ1+λ22λ1)χ(|fe|n2λ1λ2)\displaystyle=\chi(\lambda_{2}-1\leq\frac{e+f-n+\lambda_{1}+\lambda_{2}}{2}\leq\lambda_{1})\cdot\chi(|f-e|\leq n-2\ell-\lambda_{1}-\lambda_{2})
    +χ(λ2e+fn+λ1+λ2+12λ1)χ(|fe|n2λ1λ2+1),\displaystyle+\chi(\lambda_{2}\leq\frac{e+f-n+\lambda_{1}+\lambda_{2}+1}{2}\leq\lambda_{1})\cdot\chi(|f-e|\leq n-2\ell-\lambda_{1}-\lambda_{2}+1),

    where χ(P)\chi(P) is 11 if the proposition PP is true and 0 if it is false.

  3. (3)

    If λ=(n)\lambda=(n), then g(λ,(ne,1e),(nf,1f))=χ(e=f)g(\lambda,(n-e,1^{e}),(n-f,1^{f}))=\chi(e=f).

  4. (4)

    If λ\lambda is a hook shape (nd,1d)(n-d,1^{d}), then

    g(λ,(ne,1e),(nf,1f))=χ(|ef|d)χ(de+f2nd2).g(\lambda,(n-e,1^{e}),(n-f,1^{f}))=\chi(|e-f|\leq d)\cdot\chi(d\leq e+f\leq 2n-d-2).

Note that this implies that these coefficients are in {0,1,2}\{0,1,2\} [14, Corollary 4]. We are ready to prove the following result, establishing a formula for double hook characters in Rn(0,2)R_{n}^{(0,2)}.

Theorem 5.2.

Let λn\lambda\vdash n be a partition contained in a double hook shape, which is not itself contained within a hook shape. That is, λ=(λ1,λ2,2,1n2λ1λ2)\lambda=(\lambda_{1},\lambda_{2},2^{\ell},1^{n-2\ell-\lambda_{1}-\lambda_{2}}) where λ1λ22\lambda_{1}\geq\lambda_{2}\geq 2. Let m:=n2λ1λ2+2m:=n-2\ell-\lambda_{1}-\lambda_{2}+2. Then if λ1=λ2\lambda_{1}=\lambda_{2}, we have that

(26) Frob(Rn(0,2);u,v),sλ=(uv)+λ21(uv[m2]u,v+[m]u,v+[m1]u,v).\langle\operatorname{Frob}(R_{n}^{(0,2)};u,v),s_{\lambda}\rangle=(uv)^{\ell+\lambda_{2}-1}(uv[m-2]_{u,v}+[m]_{u,v}+[m-1]_{u,v}).

If λ1>λ2\lambda_{1}>\lambda_{2}, we have that

(27) Frob(Rn(0,2);u,v),sλ=(uv)+λ21(uv[m1]u,v+uv[m2]u,v+[m]u,v+[m1]u,v).\langle\operatorname{Frob}(R_{n}^{(0,2)};u,v),s_{\lambda}\rangle=(uv)^{\ell+\lambda_{2}-1}(uv[m-1]_{u,v}+uv[m-2]_{u,v}+[m]_{u,v}+[m-1]_{u,v}).
Proof.

Recall from Theorem 2.2 that

Frob(Rn(0,2);u,v)=0i+j<nuivj(s(ni,1i)s(nj,1j)s(ni+1,1i1)s(nj+1,1j1)).\operatorname{Frob}(R_{n}^{(0,2)};u,v)=\sum_{0\leq i+j<n}u^{i}v^{j}\left(s_{(n-i,1^{i})}*s_{(n-j,1^{j})}-s_{(n-i+1,1^{i-1})}*s_{(n-j+1,1^{j-1})}\right).

Let λ=(λ1,λ2,2,1n2λ1λ2)\lambda=(\lambda_{1},\lambda_{2},2^{\ell},1^{n-2\ell-\lambda_{1}-\lambda_{2}}) where λ1λ22\lambda_{1}\geq\lambda_{2}\geq 2. Hence,

\displaystyle\langle Frob(Rn(0,2);u,v),sλ\displaystyle\operatorname{Frob}(R_{n}^{(0,2)};u,v),s_{\lambda}\rangle
(28) =0i+j<nuivj(g(λ,(ni,1i),(nj,1j))g(λ,(ni+1,1i1),(nj+1,1j1))).\displaystyle=\sum_{0\leq i+j<n}u^{i}v^{j}\left(g(\lambda,(n-i,1^{i}),(n-j,1^{j}))-g(\lambda,(n-i+1,1^{i-1}),(n-j+1,1^{j-1}))\right).

By Theorem 5.1,

g(\displaystyle g( λ,(ni,1i),(nj,1j))g(λ,(ni+1,1i1),(nj+1,1j1))\displaystyle\lambda,(n-i,1^{i}),(n-j,1^{j}))-g(\lambda,(n-i+1,1^{i-1}),(n-j+1,1^{j-1}))
=(χ(λ21i+jn+λ1+λ22λ1)χ(λ21i+jn+λ1+λ222λ1))\displaystyle=\left(\chi(\lambda_{2}-1\leq\frac{i+j-n+\lambda_{1}+\lambda_{2}}{2}\leq\lambda_{1})-\chi(\lambda_{2}-1\leq\frac{i+j-n+\lambda_{1}+\lambda_{2}-2}{2}\leq\lambda_{1})\right)
χ(|ji|n2λ1λ2)\displaystyle\qquad\cdot\chi(|j-i|\leq n-2\ell-\lambda_{1}-\lambda_{2})
+(χ(λ2i+jn+λ1+λ2+12λ1)χ(λ2i+jn+λ1+λ212λ1))\displaystyle+\left(\chi(\lambda_{2}\leq\frac{i+j-n+\lambda_{1}+\lambda_{2}+1}{2}\leq\lambda_{1})-\chi(\lambda_{2}\leq\frac{i+j-n+\lambda_{1}+\lambda_{2}-1}{2}\leq\lambda_{1})\right)
(29) χ(|ji|n2λ1λ2+1).\displaystyle\qquad\cdot\chi(|j-i|\leq n-2\ell-\lambda_{1}-\lambda_{2}+1).

Define d:=λ1λ2d:=\lambda_{1}-\lambda_{2}. We analyze a component of equation (5):

(30) χ(λ21i+jn+λ1+λ22λ1)χ(λ21i+jn+λ1+λ222λ1),\chi(\lambda_{2}-1\leq\frac{i+j-n+\lambda_{1}+\lambda_{2}}{2}\leq\lambda_{1})-\chi(\lambda_{2}-1\leq\frac{i+j-n+\lambda_{1}+\lambda_{2}-2}{2}\leq\lambda_{1}),

which is equivalent to

(31) χ(|i+j(n1)|d+1)χ(|i+j(n+1)|d+1).\chi(|i+j-(n-1)|\leq d+1)-\chi(|i+j-(n+1)|\leq d+1).

Since the bigraded component (Rn(0,2))i,j=0(R_{n}^{(0,2)})_{i,j}=0 whenever i+jni+j\geq n, we exclude these cases from our analysis. Thus |i+j(n1)|d+1|i+j-(n-1)|\leq d+1 is satisfied if i+j=n1i+j=n-1 or n2n-2 (if d0d\geq 0), if i+j=n3i+j=n-3 (if d1d\geq 1), if i+j=n4i+j=n-4 (if d2d\geq 2), etc. On the other hand, |i+j(n+1)|d+1|i+j-(n+1)|\leq d+1 is satisfied if i+j=n1i+j=n-1 (if d1d\geq 1), if i+j=n2i+j=n-2 (if d2d\geq 2), etc. Putting these together, we get that equation (31) is 1 when i+j{n1d,n2d}i+j\in\{n-1-d,n-2-d\} and 0 otherwise.

We analyze when condition χ(|ji|n2λ1λ2)\chi(|j-i|\leq n-2\ell-\lambda_{1}-\lambda_{2}) is satisfied. At i+j=n1di+j=n-1-d, the condition is true exactly for integers jj which satisfy

(32) +λ2jndλ21.\ell+\lambda_{2}\leq j\leq n-d-\ell-\lambda_{2}-1.

Summing up un1djvju^{n-1-d-j}v^{j} over such jj gives

(33) (uv)+λ2[nd22λ2]u,v.(uv)^{\ell+\lambda_{2}}[n-d-2\ell-2\lambda_{2}]_{u,v}.

At i+j=n2di+j=n-2-d, the condition is true exactly for integers jj which satisfy

(34) +λ21jndλ21.\ell+\lambda_{2}-1\leq j\leq n-d-\ell-\lambda_{2}-1.

Summing up un2djvju^{n-2-d-j}v^{j} over such jj gives

(35) (uv)+λ21[nd22λ2+1]u,v.(uv)^{\ell+\lambda_{2}-1}[n-d-2\ell-2\lambda_{2}+1]_{u,v}.

Next we analyze another component of equation (5):

(36) χ(λ2i+jn+λ1+λ2+12λ1)χ(λ2i+jn+λ1+λ212λ1),\chi(\lambda_{2}\leq\frac{i+j-n+\lambda_{1}+\lambda_{2}+1}{2}\leq\lambda_{1})-\chi(\lambda_{2}\leq\frac{i+j-n+\lambda_{1}+\lambda_{2}-1}{2}\leq\lambda_{1}),

which is equivalent to

(37) χ(|i+j(n1)|d)χ(|i+j(n+1)|d).\chi(|i+j-(n-1)|\leq d)-\chi(|i+j-(n+1)|\leq d).

We have that |i+j(n1)|d|i+j-(n-1)|\leq d is satisfied if i+j=n1i+j=n-1 (if d0d\geq 0), if i+j=n2i+j=n-2 (if d1d\geq 1), if i+j=n3i+j=n-3 (if d2d\geq 2), etc. On the other hand, |i+j(n+1)|d|i+j-(n+1)|\leq d is satisfied if i+j=n1i+j=n-1 (if d2d\geq 2), if i+j=n2i+j=n-2 (if d3d\geq 3), etc. Putting these together, when d=0d=0, we get that equation (37) is 1 when i+j=n1i+j=n-1 and 0 otherwise. When d1d\geq 1, we get that equation (37) is 1 when i+j{nd,n1d}i+j\in\{n-d,n-1-d\} and 0 otherwise.

We analyze when condition χ(|ji|n2λ1λ2+1)\chi(|j-i|\leq n-2\ell-\lambda_{1}-\lambda_{2}+1) is satisfied. At i+j=n1di+j=n-1-d, the condition is true exactly for integers jj which satisfy

(38) +λ21jndλ2.\ell+\lambda_{2}-1\leq j\leq n-d-\ell-\lambda_{2}.

Summing up un1djvju^{n-1-d-j}v^{j} over such jj gives

(39) (uv)+λ21[nd22λ2+2]u,v.(uv)^{\ell+\lambda_{2}-1}[n-d-2\ell-2\lambda_{2}+2]_{u,v}.

When d1d\geq 1, at i+j=ndi+j=n-d, the condition is true exactly for integers jj which satisfy

(40) +λ2jndλ2.\ell+\lambda_{2}\leq j\leq n-d-\ell-\lambda_{2}.

Summing up undjvju^{n-d-j}v^{j} over such jj gives

(41) (uv)+λ2[nd22λ2+1]u,v.(uv)^{\ell+\lambda_{2}}[n-d-2\ell-2\lambda_{2}+1]_{u,v}.

For concision, we use m:=n2λ1λ2+2=nd22λ2m:=n-2\ell-\lambda_{1}-\lambda_{2}+2=n-d-2\ell-2\lambda_{2}. When d=0d=0, i.e., λ1=λ2\lambda_{1}=\lambda_{2}, summing equations (33), (35), and (39) proves equation (26). When d1d\geq 1, i.e., λ1>λ2\lambda_{1}>\lambda_{2}, summing equations (33), (35), (39), and (41) proves equation (27).

6. Four sets of fermions

A natural further question is to determine Frob(Rn(0,j);u1,,uj),s(1n)\langle\operatorname{Frob}(R_{n}^{(0,j)};u_{1},\ldots,u_{j}),s_{(1^{n})}\rangle for j4j\geq 4, of which the simplest next case is Frob(Rn(0,4);u,v,w,x),s(1n)\langle\operatorname{Frob}(R_{n}^{(0,4)};u,v,w,x),s_{(1^{n})}\rangle. If the hook characters in Rn(0,3)R_{n}^{(0,3)} are all known, then a similar process as in Proposition 3.1 using Frob(Rn(0,3)Rn(0,1);u,v,w,x)\operatorname{Frob}(R_{n}^{(0,3)}\otimes R_{n}^{(0,1)};u,v,w,x) can be attempted. Using Theorem 5.1, we can first bound Frob(Rn(0,3);u,v,w),s(nd,1d)\langle\operatorname{Frob}(R_{n}^{(0,3)};u,v,w),s_{(n-d,1^{d})}\rangle.

Example 6.1.

Let n=3n=3. Suppose we want to calculate Frob(R3(0,3);u,v,w),s(2,1)\langle\operatorname{Frob}(R_{3}^{(0,3)};u,v,w),s_{(2,1)}\rangle. Consider that any s(2,1)s_{(2,1)} in Frob(R3(0,3);u,v,w)\operatorname{Frob}(R_{3}^{(0,3)};u,v,w) must appear in Frob(R3(0,2)R3(0,1);u,v,w)\operatorname{Frob}(R_{3}^{(0,2)}\otimes R_{3}^{(0,1)};u,v,w). We compute that

Frob\displaystyle\langle\operatorname{Frob} (R3(0,2)R3(0,1);u,v,w),s(2,1)\displaystyle(R_{3}^{(0,2)}\otimes R_{3}^{(0,1)};u,v,w),s_{(2,1)}\rangle
(42) =λ3d=02Frob(R3(0,2);u,v),sλFrob(R3(0,1);w),s(3d,1d)g(λ,(3d,1d),(2,1)).\displaystyle=\sum_{\lambda\vdash 3}\sum_{d=0}^{2}\langle\operatorname{Frob}(R_{3}^{(0,2)};u,v),s_{\lambda}\rangle\langle\operatorname{Frob}(R_{3}^{(0,1)};w),s_{(3-d,1^{d})}\rangle g(\lambda,(3-d,1^{d}),(2,1)).

Using Rosas’ formula for Kronecker coefficients of two hooks, we determine that the only pairs of λ\lambda and (3d,1d)(3-d,1^{d}) which do not have multiplicity of 0 are (3),(2,1)(3),(2,1); (2,1),(3)(2,1),(3); (2,1),(2,1)(2,1),(2,1); (2,1),(13)(2,1),(1^{3}); and (13),(2,1)(1^{3}),(2,1), which all have multiplicity of 11. Using Proposition 2.3 and Lemma 2.1, we obtain

(43) w+([2]u,v+uv)+([2]u,v+uv)w+([2]u,v+uv)w2+[3]u,vw.w+([2]_{u,v}+uv)+([2]_{u,v}+uv)w+([2]_{u,v}+uv)w^{2}+[3]_{u,v}w.

By permuting which sets of variables are assigned to Rn(0,2)R_{n}^{(0,2)} and Rn(0,1)R_{n}^{(0,1)}, we can similarly obtain

(44) v+([2]u,w+uw)+([2]u,w+uw)v+([2]u,w+uw)v2+[3]u,wv,v+([2]_{u,w}+uw)+([2]_{u,w}+uw)v+([2]_{u,w}+uw)v^{2}+[3]_{u,w}v,
(45) u+([2]w,v+wv)+([2]w,v+wv)u+([2]w,v+wv)u2+[3]w,vu.u+([2]_{w,v}+wv)+([2]_{w,v}+wv)u+([2]_{w,v}+wv)u^{2}+[3]_{w,v}u.

Summing the monomials which appear in all three equations, we obtain

(46) u+v+w+uv+uw+vw+2uvw.u+v+w+uv+uw+vw+2uvw.

Now a computer calculation finds that

(47) Frob(R3(0,3);u,v,w),s(2,1)=u+v+w+uv+uw+vw,\langle\operatorname{Frob}(R_{3}^{(0,3)};u,v,w),s_{(2,1)}\rangle=u+v+w+uv+uw+vw,

which shows that the bound is not tight in this case due to the monomial 2uvw2uvw. It will require additional work to determine how to cut out the extra monomials in general.

Example 6.2.

Let n=3n=3. Suppose we know all of the hook characters in R3(0,3)R_{3}^{(0,3)}; in this case Frob(R3(0,3);u,v,w)=(uvw+u2+uv+v2+uw+vw+w2)s(13)+(uv+uw+vw+u+v+w)s(2,1)+s(3)\operatorname{Frob}(R_{3}^{(0,3)};u,v,w)=(uvw+u^{2}+uv+v^{2}+uw+vw+w^{2})s_{(1^{3})}+(uv+uw+vw+u+v+w)s_{(2,1)}+s_{(3)}. Consider that any sign character in R3(0,4)R_{3}^{(0,4)} must appear in R3(0,3)R3(0,1)R_{3}^{(0,3)}\otimes R_{3}^{(0,1)}. We compute that

Frob(R3(0,3)R3(0,1);u,v,w,x),s(13)\displaystyle\langle\operatorname{Frob}(R_{3}^{(0,3)}\otimes R_{3}^{(0,1)};u,v,w,x),s_{(1^{3})}\rangle =λ3Frob(R3(0,3);u,v,w),sλFrob(R3(0,1);x),sλ\displaystyle=\sum_{\lambda\vdash 3}\langle\operatorname{Frob}(R_{3}^{(0,3)};u,v,w),s_{\lambda^{\prime}}\rangle\langle\operatorname{Frob}(R_{3}^{(0,1)};x),s_{\lambda}\rangle
(48) =d=02Frob(R3(0,3);u,v,w),s(d+1,12d)xd,\displaystyle=\sum_{d=0}^{2}\langle\operatorname{Frob}(R_{3}^{(0,3)};u,v,w),s_{(d+1,1^{2-d})}\rangle x^{d},

using Lemma 2.1. This simplifies to

(49) u2+v2+w2+x2+uv+uw+ux+vw+vx+wx+uvw+uvx+uwx+vwx,u^{2}+v^{2}+w^{2}+x^{2}+uv+uw+ux+vw+vx+wx+uvw+uvx+uwx+vwx,

which is unchanged under changing which sets of variables are assigned to Rn(0,3)R_{n}^{(0,3)} and Rn(0,1)R_{n}^{(0,1)}. In this example, equation (49) is equal to Frob(R3(0,4);u,v,w,x),s(13)\langle\operatorname{Frob}(R_{3}^{(0,4)};u,v,w,x),s_{(1^{3})}\rangle determined by computer calculation.

Example 6.3.

The previous example also gives tight bounds for n=4n=4, with graded multiplicity s(3)(u,v,w,x)+s(2,1,1)(u,v,w,x)+s(2,1,1,1)(u,v,w,x)s_{(3)}(u,v,w,x)+s_{(2,1,1)}(u,v,w,x)+s_{(2,1,1,1)}(u,v,w,x), and for n=5n=5, with graded multiplicity s(4)(u,v,w,x)+s(3,1,1)(u,v,w,x)+s(3,1,1,1)(u,v,w,x)+s(2,2,1,1)(u,v,w,x)+s(2,2,2,1)(u,v,w,x)s_{(4)}(u,v,w,x)+s_{(3,1,1)}(u,v,w,x)+s_{(3,1,1,1)}(u,v,w,x)+s_{(2,2,1,1)}(u,v,w,x)+s_{(2,2,2,1)}(u,v,w,x).

Another approach is to use Frob(Rn(0,2)Rn(0,2);u,v,w,x)\operatorname{Frob}(R_{n}^{(0,2)}\otimes R_{n}^{(0,2)};u,v,w,x) to bound Frob(Rn(0,4);u,v,w,x),s(1n)\langle\operatorname{Frob}(R_{n}^{(0,4)};u,v,w,x),s_{(1^{n})}\rangle directly. As Kim and Rhoades note, only Schur functions of shapes contained within double hooks occur in Frob(Rn(0,2);u,v)\operatorname{Frob}(R_{n}^{(0,2)};u,v), so the shapes which index sign characters are limited.

Example 6.4.

Let n=3n=3. Consider that any sign character in R3(0,4)R_{3}^{(0,4)} must appear in R3(0,2)R3(0,2)R_{3}^{(0,2)}\otimes R_{3}^{(0,2)}. We compute that

(50) Frob(R3(0,2)R3(0,2);u,v,w,x),s(13)\displaystyle\langle\operatorname{Frob}(R_{3}^{(0,2)}\otimes R_{3}^{(0,2)};u,v,w,x),s_{(1^{3})}\rangle =λ3Frob(R3(0,2);u,v),sλFrob(R3(0,2);w,x),sλ.\displaystyle=\sum_{\lambda\vdash 3}\langle\operatorname{Frob}(R_{3}^{(0,2)};u,v),s_{\lambda^{\prime}}\rangle\langle\operatorname{Frob}(R_{3}^{(0,2)};w,x),s_{\lambda}\rangle.

In this case, there are only three partitions of 33: (13),(2,1),(3)(1^{3}),(2,1),(3). Using Proposition 2.3, we get

(51) Frob(R3(0,2);u,v),s(3)Frob(R3(0,2);w,x),s(13)=[3]w,x,\displaystyle\langle\operatorname{Frob}(R_{3}^{(0,2)};u,v),s_{(3)}\rangle\langle\operatorname{Frob}(R_{3}^{(0,2)};w,x),s_{(1^{3})}\rangle=[3]_{w,x},
(52) Frob(R3(0,2);u,v),s(2,1)Frob(R3(0,2);w,x),s(2,1)=([2]u,v+uv)([2]w,x+wx),\displaystyle\langle\operatorname{Frob}(R_{3}^{(0,2)};u,v),s_{(2,1)}\rangle\langle\operatorname{Frob}(R_{3}^{(0,2)};w,x),s_{(2,1)}\rangle=([2]_{u,v}+uv)([2]_{w,x}+wx),
(53) Frob(R3(0,2);u,v),s(13)Frob(R3(0,2);w,x),s(3)=[3]u,v.\displaystyle\langle\operatorname{Frob}(R_{3}^{(0,2)};u,v),s_{(1^{3})}\rangle\langle\operatorname{Frob}(R_{3}^{(0,2)};w,x),s_{(3)}\rangle=[3]_{u,v}.

Putting these together, we obtain

(54) [3]w,x+([2]u,v+uv)([2]w,x+wx)+[3]u,v.[3]_{w,x}+([2]_{u,v}+uv)([2]_{w,x}+wx)+[3]_{u,v}.

By changing which sets of variables are assigned to each Rn(0,2)R_{n}^{(0,2)}, we can similarly obtain

(55) [3]v,x+([2]u,w+uw)([2]v,x+vx)+[3]u,w,[3]_{v,x}+([2]_{u,w}+uw)([2]_{v,x}+vx)+[3]_{u,w},
(56) [3]u,x+([2]w,v+wv)([2]u,x+ux)+[3]w,v.[3]_{u,x}+([2]_{w,v}+wv)([2]_{u,x}+ux)+[3]_{w,v}.

In this case, all three are equal to each other, and expand to give

(57) uvwx+uvw+uvx+uwx+vwx+u2+uv+v2+uw+vw+w2+ux+vx+wx+x2.uvwx+uvw+uvx+uwx+vwx+u^{2}+uv+v^{2}+uw+vw+w^{2}+ux+vx+wx+x^{2}.

Now a computer calculation finds that

(58) Frob(R3(0,4);u,v,w,x),s(13)=uvw+uvx+uwx+vwx+u2+uv+v2+uw+vw+w2+ux+vx+wx+x2,\langle\operatorname{Frob}(R_{3}^{(0,4)};u,v,w,x),s_{(1^{3})}\rangle=uvw+uvx+uwx+vwx+u^{2}+uv+v^{2}+uw+vw+w^{2}+ux+vx+wx+x^{2},

which shows that the bound is not tight in this case due to the monomial uvwxuvwx. It will require additional work to determine how to cut out the extra monomials in general.

7. One set of bosons and three sets of fermions

In this section, we study the sign character in Rn(1,3)R_{n}^{(1,3)}, the (1,3)(1,3)-bosonic-fermionic coinvariant ring. We first recall the “Theta conjecture” of D’Adderio, Iraci, and Vanden Wyngaerd, which expresses the multigraded Frobenius series of Rn(2,2)R_{n}^{(2,2)} in terms of certain Theta operators and the nabla operator.

Conjecture 7.1 ([7, Conjecture 8.2]).

For all n1n\geq 1,

Frob(Rn(2,2);q,t;u,v)=k+<nukvΘekΘeenk.\operatorname{Frob}(R_{n}^{(2,2)};q,t;u,v)=\sum_{k+\ell<n}u^{k}v^{\ell}\Theta_{e_{k}}\Theta_{e_{\ell}}\nabla e_{n-k-\ell}.

The Theta conjecture specialized at t=0t=0 is the following.

Conjecture 7.2 (D’Adderio, Iraci, and Vanden Wyngaerd).

For all n1n\geq 1,

Frob(Rn(1,2);q;u,v)=k+<nukv(ΘekΘeenk)|t=0.\operatorname{Frob}(R_{n}^{(1,2)};q;u,v)=\sum_{k+\ell<n}u^{k}v^{\ell}\left(\Theta_{e_{k}}\Theta_{e_{\ell}}\nabla e_{n-k-\ell}\right)|_{t=0}.

We recall the following result on the conjectural hook characters in Rn(1,2)R_{n}^{(1,2)}.

Theorem 7.3 ([12, Theorem 8.4]).

If the Theta conjecture specialized at t=0t=0 (Conjecture 7.2) is true, then

Frob(Rn(1,2);q;u,v),s(d+1,1nd1)=k+<nukvq(ndk2)[n1d]q[n1kd]q[n1k]q.\langle\operatorname{Frob}(R_{n}^{(1,2)};q;u,v),s_{(d+1,1^{n-d-1})}\rangle=\sum_{k+\ell<n}u^{k}v^{\ell}q^{\binom{n-d-k-\ell}{2}}\genfrac{[}{]}{0.0pt}{}{n-1-d}{\ell}_{q}\genfrac{[}{]}{0.0pt}{}{n-1-k}{d}_{q}\genfrac{[}{]}{0.0pt}{}{n-1-\ell}{k}_{q}.

Now we compute an example.

Example 7.4.

Let n=3n=3. Consider that any sign character in R3(1,3)R_{3}^{(1,3)} must appear in R3(1,2)R3(0,1)R_{3}^{(1,2)}\otimes R_{3}^{(0,1)}. We compute that

Frob\displaystyle\langle\operatorname{Frob} (R3(1,2)R3(0,1);q;u,v,w),s(13)\displaystyle(R_{3}^{(1,2)}\otimes R_{3}^{(0,1)};q;u,v,w),s_{(1^{3})}\rangle
=λ3Frob(R3(1,2);q;u,v),sλFrob(R3(0,1);w),sλ\displaystyle=\sum_{\lambda\vdash 3}\langle\operatorname{Frob}(R_{3}^{(1,2)};q;u,v),s_{\lambda^{\prime}}\rangle\langle\operatorname{Frob}(R_{3}^{(0,1)};w),s_{\lambda}\rangle
=d=02Frob(R3(1,2);q;u,v),s(d+1,12d)wd\displaystyle=\sum_{d=0}^{2}\langle\operatorname{Frob}(R_{3}^{(1,2)};q;u,v),s_{(d+1,1^{2-d})}\rangle w^{d}
(59) =(q3+vq[2]q+uq[2]q+uv[2]q+v2+u2)+(uv+q[2]q+v[2]q+u[2]q)w+w2.\displaystyle=(q^{3}+vq[2]_{q}+uq[2]_{q}+uv[2]_{q}+v^{2}+u^{2})+(uv+q[2]_{q}+v[2]_{q}+u[2]_{q})w+w^{2}.

In this example, this is equal to Frob(R3(1,3);q;u,v,w),s(13)\langle\operatorname{Frob}(R_{3}^{(1,3)};q;u,v,w),s_{(1^{3})}\rangle determined by computer calculation.

Proposition 7.5.

If the Theta conjecture specialized at t=0t=0 (Conjecture 7.2) is true, then we have the following upper-bound:

Frob(Rn(1,3);q;u,v,w),s(1n)k,,d0ukvwdq(ndk2)[n1d]q[n1kd]q[n1k]q.\langle\operatorname{Frob}(R_{n}^{(1,3)};q;u,v,w),s_{(1^{n})}\rangle\leq\sum_{k,\ell,d\geq 0}u^{k}v^{\ell}w^{d}q^{\binom{n-d-k-\ell}{2}}\genfrac{[}{]}{0.0pt}{}{n-1-d}{\ell}_{q}\genfrac{[}{]}{0.0pt}{}{n-1-k}{d}_{q}\genfrac{[}{]}{0.0pt}{}{n-1-\ell}{k}_{q}.
Proof.

Any sign character in Rn(1,3)R_{n}^{(1,3)} must appear in Rn(1,2)Rn(0,1)R_{n}^{(1,2)}\otimes R_{n}^{(0,1)}. Assume the Theta conjecture specialized at t=0t=0 (Conjecture 7.2) is true. By Theorem 7.3, we compute that

Frob\displaystyle\langle\operatorname{Frob} (Rn(1,2)Rn(0,1);q;u,v,w),s(1n)\displaystyle(R_{n}^{(1,2)}\otimes R_{n}^{(0,1)};q;u,v,w),s_{(1^{n})}\rangle
=λnFrob(Rn(1,2);q;u,v),sλFrob(Rn(0,1);w),sλ\displaystyle=\sum_{\lambda\vdash n}\langle\operatorname{Frob}(R_{n}^{(1,2)};q;u,v),s_{\lambda^{\prime}}\rangle\langle\operatorname{Frob}(R_{n}^{(0,1)};w),s_{\lambda}\rangle
=d=0n1wdFrob(Rn(1,2);q;u,v),s(d+1,1nd1)\displaystyle=\sum_{d=0}^{n-1}w^{d}\langle\operatorname{Frob}(R_{n}^{(1,2)};q;u,v),s_{(d+1,1^{n-d-1})}\rangle
=d=0n1wdk+<nukvq(ndk2)[n1d]q[n1kd]q[n1k]q\displaystyle=\sum_{d=0}^{n-1}w^{d}\sum_{k+\ell<n}u^{k}v^{\ell}q^{\binom{n-d-k-\ell}{2}}\genfrac{[}{]}{0.0pt}{}{n-1-d}{\ell}_{q}\genfrac{[}{]}{0.0pt}{}{n-1-k}{d}_{q}\genfrac{[}{]}{0.0pt}{}{n-1-\ell}{k}_{q}
(60) =k,,d0ukvwdq(ndk2)[n1d]q[n1kd]q[n1k]q,\displaystyle=\sum_{k,\ell,d\geq 0}u^{k}v^{\ell}w^{d}q^{\binom{n-d-k-\ell}{2}}\genfrac{[}{]}{0.0pt}{}{n-1-d}{\ell}_{q}\genfrac{[}{]}{0.0pt}{}{n-1-k}{d}_{q}\genfrac{[}{]}{0.0pt}{}{n-1-\ell}{k}_{q},

where the last line follows by considering when the qq-binomial coefficients must be 0 (specifically, one could take k+<nk+\ell<n, +d<n\ell+d<n, and d+k<nd+k<n). ∎

Based on data for n5n\leq 5, we propose the following conjecture.

Conjecture 7.6.
Frob(R(1,3);q;u,v,w),s(1n)=k,,d0ukvwdq(ndk2)[n1d]q[n1kd]q[n1k]q.\langle\operatorname{Frob}(R^{(1,3)};q;u,v,w),s_{(1^{n})}\rangle=\sum_{k,\ell,d\geq 0}u^{k}v^{\ell}w^{d}q^{\binom{n-d-k-\ell}{2}}\genfrac{[}{]}{0.0pt}{}{n-1-d}{\ell}_{q}\genfrac{[}{]}{0.0pt}{}{n-1-k}{d}_{q}\genfrac{[}{]}{0.0pt}{}{n-1-\ell}{k}_{q}.

Notice that upon specializing q,u,v,wq,u,v,w all to 11, the conjecture becomes

(61) Frob(R(1,3);1;1,1,1),s(1n)=k,,d0(n1d)(n1kd)(n1k).\langle\operatorname{Frob}(R^{(1,3)};1;1,1,1),s_{(1^{n})}\rangle=\sum_{k,\ell,d\geq 0}\binom{n-1-d}{\ell}\binom{n-1-k}{d}\binom{n-1-\ell}{k}.

Let FnF_{n} be the nnth Fibonacci number, defined by the initial conditions F0=0F_{0}=0, F1=1F_{1}=1, and recurrence Fn=Fn1+Fn2F_{n}=F_{n-1}+F_{n-2} for all n2n\geq 2. Inspired by a conjecture of Zabrocki on Frob(Rn(2,1);q,t;u)\operatorname{Frob}(R_{n}^{(2,1)};q,t;u) [16], Bergeron made a general conjecture on how the Frobenius series of bosonic-fermionic coinvariant rings could be derived from the Frobenius series of purely bosonic coinvariant rings [2, Conjecture 1]. While collecting computational evidence, Bergeron conjectured the following.

Conjecture 7.7 ([2, Table 3]).

For all n1n\geq 1,

Frob(Rn(1,3);1;1,1,1),s(1n)=12F3n.\langle\operatorname{Frob}(R_{n}^{(1,3)};1;1,1,1),s_{(1^{n})}\rangle=\frac{1}{2}F_{3n}.

Beginning at n=1n=1, the sequence 12F3n\frac{1}{2}F_{3n} is 1,4,17,72,305,1292,5473,23184,1,4,17,72,305,1292,5473,23184,\ldots (see [13, Sequence A001076]). We show the following result connecting the two proposed enumerations.

Proposition 7.8.

For n1n\geq 1,

k,,d0(n1d)(n1kd)(n1k)=12F3n.\sum_{k,\ell,d\geq 0}\binom{n-1-d}{\ell}\binom{n-1-k}{d}\binom{n-1-\ell}{k}=\frac{1}{2}F_{3n}.
Proof.

Shifting indices, the claim is equivalent to

(62) k,,d0(nd)(nkd)(nk)=12F3(n+1),\sum_{k,\ell,d\geq 0}\binom{n-d}{\ell}\binom{n-k}{d}\binom{n-\ell}{k}=\frac{1}{2}F_{3(n+1)},

for all n0n\geq 0. From [13, Sequence A001076] we have that for all n1n\geq 1, 12F3n=an\frac{1}{2}F_{3n}=a_{n}, where ana_{n} is defined by a1=1a_{1}=1, a2=4a_{2}=4, and an=4an1+an2a_{n}=4a_{n-1}+a_{n-2} for n3n\geq 3. Thus the claim is equivalent to

(63) k,,d0(nd)(nkd)(nk)=an+1,\sum_{k,\ell,d\geq 0}\binom{n-d}{\ell}\binom{n-k}{d}\binom{n-\ell}{k}=a_{n+1},

for all n0n\geq 0.

First, we check that the left hand side of equation (63) satisfies the initial conditions a1=1a_{1}=1 and a2=4a_{2}=4. For n=0n=0, the sum is (00)(00)(00)=1\binom{0}{0}\binom{0}{0}\binom{0}{0}=1. For n=1n=1, the sum is (10)(10)(10)+(10)(00)(11)+(11)(10)(00)+(00)(11)(10)=4\binom{1}{0}\binom{1}{0}\binom{1}{0}+\binom{1}{0}\binom{0}{0}\binom{1}{1}+\binom{1}{1}\binom{1}{0}\binom{0}{0}+\binom{0}{0}\binom{1}{1}\binom{1}{0}=4.

Next, we verify that the left hand side of equation (63) satisfies the recurrence for ana_{n}. Using Pascal’s identity repeatedly, along with shifting indices when appropriate, we write

k,,d0\displaystyle\sum_{k,\ell,d\geq 0} (nd)(nkd)(nk)\displaystyle\binom{n-d}{\ell}\binom{n-k}{d}\binom{n-\ell}{k}
=k,,d0(n1d)(nkd)(nk)+k,,d0(n1d1)(nkd)(nk)\displaystyle=\sum_{k,\ell,d\geq 0}\binom{n-1-d}{\ell}\binom{n-k}{d}\binom{n-\ell}{k}+\sum_{k,\ell,d\geq 0}\binom{n-1-d}{\ell-1}\binom{n-k}{d}\binom{n-\ell}{k}
=k,,d0(n1d)(nkd)(nk)+k,,d0(n1d)(nkd)(n1k)\displaystyle=\sum_{k,\ell,d\geq 0}\binom{n-1-d}{\ell}\binom{n-k}{d}\binom{n-\ell}{k}+\sum_{k,\ell,d\geq 0}\binom{n-1-d}{\ell}\binom{n-k}{d}\binom{n-1-\ell}{k}
=2k,,d0(n1d)(nkd)(n1k)+k,,d0(n1d)(nkd)(n1k1)\displaystyle=2\sum_{k,\ell,d\geq 0}\binom{n-1-d}{\ell}\binom{n-k}{d}\binom{n-1-\ell}{k}+\sum_{k,\ell,d\geq 0}\binom{n-1-d}{\ell}\binom{n-k}{d}\binom{n-1-\ell}{k-1}
=2k,,d0(n1d)(nkd)(n1k)+an1\displaystyle=2\sum_{k,\ell,d\geq 0}\binom{n-1-d}{\ell}\binom{n-k}{d}\binom{n-1-\ell}{k}+a_{n-1}
=2k,,d0(n1d)(n1kd1)(n1k)+3an1\displaystyle=2\sum_{k,\ell,d\geq 0}\binom{n-1-d}{\ell}\binom{n-1-k}{d-1}\binom{n-1-\ell}{k}+3a_{n-1}
(64) =2k,,d0(n2d)(n1kd)(n1k)+3an1.\displaystyle=2\sum_{k,\ell,d\geq 0}\binom{n-2-d}{\ell}\binom{n-1-k}{d}\binom{n-1-\ell}{k}+3a_{n-1}.

Note that (n2d)=(n1d)(n2d1)\binom{n-2-d}{\ell}=\binom{n-1-d}{\ell}-\binom{n-2-d}{\ell-1}, so we continue with

2k,,d0\displaystyle 2\sum_{k,\ell,d\geq 0} (n2d)(n1kd)(n1k)\displaystyle\binom{n-2-d}{\ell}\binom{n-1-k}{d}\binom{n-1-\ell}{k}
=an1+k,,d0(n2d)(n1kd)(n1k)\displaystyle=a_{n-1}+\sum_{k,\ell,d\geq 0}\binom{n-2-d}{\ell}\binom{n-1-k}{d}\binom{n-1-\ell}{k}
k,,d0(n2d1)(n1kd)(n1k)\displaystyle\qquad-\sum_{k,\ell,d\geq 0}\binom{n-2-d}{\ell-1}\binom{n-1-k}{d}\binom{n-1-\ell}{k}
=an1+k,,d0(n2d)(n1kd)(n1k)\displaystyle=a_{n-1}+\sum_{k,\ell,d\geq 0}\binom{n-2-d}{\ell}\binom{n-1-k}{d}\binom{n-1-\ell}{k}
k,,d0(n2d)(n1kd)(n2k)\displaystyle\qquad-\sum_{k,\ell,d\geq 0}\binom{n-2-d}{\ell}\binom{n-1-k}{d}\binom{n-2-\ell}{k}
=an1+k,,d0(n2d)(n1kd)(n2k)\displaystyle=a_{n-1}+\sum_{k,\ell,d\geq 0}\binom{n-2-d}{\ell}\binom{n-1-k}{d}\binom{n-2-\ell}{k}
+k,,d0(n2d)(n1kd)(n2k1)\displaystyle\qquad+\sum_{k,\ell,d\geq 0}\binom{n-2-d}{\ell}\binom{n-1-k}{d}\binom{n-2-\ell}{k-1}
k,,d0(n2d)(n1kd)(n2k)\displaystyle\qquad-\sum_{k,\ell,d\geq 0}\binom{n-2-d}{\ell}\binom{n-1-k}{d}\binom{n-2-\ell}{k}
(65) =an1+an2,\displaystyle=a_{n-1}+a_{n-2},

which upon substituting into equation (7) completes the proof. ∎

Remark 7.9.

If the Theta conjecture specialized at t=0t=0 (Conjecture 7.2) is true, and the conjecture of Bergeron (Conjecture 7.7) that Frob(R(1,3);1;1,1,1),s(1n)=12F3n\langle\operatorname{Frob}(R^{(1,3)};1;1,1,1),s_{(1^{n})}\rangle=\frac{1}{2}F_{3n} is true, then by Proposition 7.8, the sign character has multiplicity large enough to force equality in the upper bound given in Proposition 7.5. This would prove Conjecture 7.6.

Acknowledgements

The author would like to thank François Bergeron, Sylvie Corteel, Nicolle González, and Mark Haiman for helpful conversations. The author was partially supported by the National Science Foundation Graduate Research Fellowship DGE-2146752.

References

  • [1] François Bergeron, Jim Haglund, Alessandro Iraci, and Marino Romero, Bosonic-fermionic diagonal coinvariants and theta operators, preprint (2023), https://www2.math.upenn.edu/~jhaglund/preprints/BF2.pdf.
  • [2] François Bergeron, The bosonic-fermionic diagonal coinvariant modules conjecture, preprint (2020), https://arxiv.org/abs/2005.00924.
  • [3] by same author, (GLk×Sn)({GL}_{k}\times{S}_{n})-modules of multivariate diagonal harmonics, Open Problems in Algebraic Combinatorics (Christine Berkesch, Benjamin Brubaker, Gregg Musiker, Pavlo Pylyavskyy, and Victor Reiner, eds.), Proc. Symp. Pure Math., vol. 110, Providence, RI: American Mathematical Society (AMS), 2024, pp. 1–22.
  • [4] Christine Bessenrodt, Tensor products of representations of the symmetric groups and related groups, RIMS Kokyuroku (Proceedings of Research Insititute for Mathematical Sciences) 1149 (2000), 1–15.
  • [5] Nicolas Bourbaki, Algebra I, Elements of mathematics, Springer-Verlag, Berlin, 1989 (English).
  • [6] Daniel Bump and Anne Schilling, Crystal bases. Representations and combinatorics, Hackensack, NJ: World Scientific, 2017 (English).
  • [7] Michele D’Adderio, Alessandro Iraci, and Anna Vanden Wyngaerd, Theta operators, refined delta conjectures, and coinvariants, Adv. Math. 376 (2021), 60, Id/No 107447.
  • [8] James Haglund and Emily Sergel, Schedules and the delta conjecture, Annals of Combinatorics 25 (2020), no. 1, 1?31.
  • [9] Mark Haiman, Conjectures on the quotient ring by diagonal invariants, J. Algebr. Comb. 3 (1994), no. 1, 17–76.
  • [10] by same author, Vanishing theorems and character formulas for the Hilbert scheme of points in the plane, Invent. Math. 149 (2002), no. 2, 371–407.
  • [11] Jongwon Kim and Brendon Rhoades, Lefschetz theory for exterior algebras and fermionic diagonal coinvariants, Int. Math. Res. Not. 2022 (2022), no. 4, 2906–2933.
  • [12] John Lentfer, A conjectural basis for the (1,2)(1,2)-bosonic-fermionic coinvariant ring, preprint (2024), https://arxiv.org/abs/2406.19715.
  • [13] OEIS Foundation Inc. (2024), The On-Line Encyclopedia of Integer Sequences, Published electronically at https://oeis.org.
  • [14] Mercedes H. Rosas, The Kronecker product of Schur functions indexed by two-row shapes or hook shapes, Journal of Algebraic Combinatorics 14 (2001), no. 2, 153–173.
  • [15] Joshua P. Swanson and Nolan R. Wallach, Harmonic differential forms for pseudo-reflection groups II. Bi-degree bounds, Combinatorial Theory 3 (2023), no. 3, Paper no. 17.
  • [16] Mike Zabrocki, A module for the Delta conjecture, preprint (2019), https://arxiv.org/abs/1902.08966.
  • [17] by same author, Coinvariants and harmonics, 2020, https://realopacblog.wordpress.com/2020/01/26/coinvariants-and-harmonics/.