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

Three Examples of Quasisymmetric Compatible 𝔖n\mathfrak{S}_{n}-modules

Angela Hicks, Samantha Miller-Brown
(Date: September 30, 2025)
Abstract.

The Schur functions, a basis for the symmetric polynomials (Sym), encode the irreducible representations of the symmetric group 𝔖n\mathfrak{S}_{n} via the Frobenius characteristic map. In 1996, Krob and Thibon defined a quasisymmetric Frobenius map on the representations of β„‹n​(0)\mathcal{H}_{n}(0), mapping them to the quasisymmetric functions (QSym). Despite the obvious inclusion of Sym in QSym and the close relationship between 𝔖n\mathfrak{S}_{n} and β„‹n​(0)\mathcal{H}_{n}(0), there is no known direct link between these two Frobenius characteristic maps and the related representations. We explore three specific situations in which a deformation of an 𝔖n\mathfrak{S}_{n} action results in a valid β„‹n​(0)\mathcal{H}_{n}(0) action and gives a quasisymmetric Frobenius characteristic that is equal to the symmetric Frobenius characteristic. We introduce the concept of quasisymmetric compatibility, which formalizes a link between the two maps, and we show it applies to all 𝔖n\mathfrak{S}_{n}-modules.

1. Introduction

The Frobenius characteristic map Fc​h​a​rF_{char} a well-studied map on the representations of 𝔖n\mathfrak{S}_{n}, encodes modules as symmetric functions, allowing tools from linear algebra to be used to determine representation theoretic information about these 𝔖n\mathfrak{S}_{n}-modules. Here, we refer to the image of an 𝔖n\mathfrak{S}_{n}-module under this map as the symmetric Frobenius characteristic. In particular, if SΞ»S^{\lambda} is the irreducible 𝔖n\mathfrak{S}_{n}-module associated to the partition Ξ»\lambda of nn, then

Fc​h​a​r​(SΞ»)=sΞ»,F_{char}(S^{\lambda})=s_{\lambda},

where sΞ»s_{\lambda} is the Schur function associated to Ξ»\lambda.

In 1996, Krob and Thibon [1] defined two similar maps, namely the quasisymmetric Frobenius characteristic Fc​h​a​rQF_{char}^{Q} and the noncommutative Frobenius characteristic Fc​h​a​rNF_{char}^{N}, that encode modules of β„‹n​(0)\mathcal{H}_{n}(0), a deformation of the group algebra ℂ​[𝔖n]\mathbb{C}[\mathfrak{S}_{n}], as quasisymmetric functions and noncommutative symmetric functions, respectively. Since β„‹n​(0)\mathcal{H}_{n}(0) is not semisimple, indecomposable modules are not always irreducible. Fc​h​a​rNF_{char}^{N}, in particular, is defined only on the projective modules of β„‹n​(0)\mathcal{H}_{n}(0), sending the indecomposable modules PΞ±P_{\alpha}, where Ξ±\alpha is a (strong) composition of nn, to the noncommutative ribbon Schur functions:

Fc​h​a​rN​(PΞ±)=RΞ±F_{char}^{N}(P_{\alpha})=R_{\alpha}

Fc​h​a​rQF_{char}^{Q} is defined more broadly on all (equivalence classes of) β„‹n​(0)\mathcal{H}_{n}(0)-modules, and if CΞ±C_{\alpha} is the irreducible β„‹n​(0)\mathcal{H}_{n}(0)-module associated to Ξ±\alpha, then

Fc​h​a​rQ​(CΞ±)=FΞ±,F_{char}^{Q}(C_{\alpha})=F_{\alpha},

where FΞ±F_{\alpha} is Gessel’s fundamental quasisymmetric function associated to composition Ξ±\alpha.

When MM is a projective module,

Fc​h​a​rQ​(M)=χ​(Fc​h​a​rN​(PΞ±)),F_{char}^{Q}(M)=\chi(F_{char}^{N}(P_{\alpha})),

where Ο‡\chi is the forgetful map on noncommutative symmetric functions which allows the variables to commute: thus if MM is a projective β„‹n​(0)\mathcal{H}_{n}(0)-module, it must have a symmetric image under Fc​h​a​rQ​(M)F_{char}^{Q}(M). The converse is not the case, and in particular Fc​h​a​rQF_{char}^{Q} is not injective.

The simple and projective indecomposible modules of β„‹n​(0)\mathcal{H}_{n}(0) were first classified by NortonΒ [2]. In 2016, Huang [3] introduced a tableau-based definition of the indecomposable module PΞ±P_{\alpha}. See Figure 1 for one such example. A reader familiar with Specht modules may quickly notice the similarities to the Specht module associated to the ribbon Schur function sΞ±s_{\alpha}, which we denote by SΞ±S^{\alpha}. In particular, the underlying vector space of SΞ±S^{\alpha} is isomorphic to PΞ±P_{\alpha}. While the symmetric action on the polytabloids of ribbon shape Ξ±\alpha is not identical to the β„‹n​(0)\mathcal{H}_{n}(0)-action on the ribbon tableaux, there are clear similarities.

\ytableausetup𝐏𝟏𝟐𝟏\mathbf{P_{121}}\ytableaushort​2,14,\none​3{\ytableaushort{2,14,\none 3}}\ytableaushort​3,14,\none​2{\ytableaushort{3,14,\none 2}}\ytableaushort​4,13,\none​2{\ytableaushort{4,13,\none 2}}\ytableaushort​3,24,\none​1{\ytableaushort{3,24,\none 1}}\ytableaushort​4,23,\none​1{\ytableaushort{4,23,\none 1}}0π¯2\overline{\pi}_{2}π¯3\overline{\pi}_{3}π¯1\overline{\pi}_{1}π¯3\overline{\pi}_{3}π¯1\overline{\pi}_{1}π¯2\overline{\pi}_{2}π¯1=βˆ’1\overline{\pi}_{1}=-1π¯3=βˆ’1\overline{\pi}_{3}=-1π¯2=βˆ’1\overline{\pi}_{2}=-1π¯1,π¯3=βˆ’1\overline{\pi}_{1},\overline{\pi}_{3}=-1π¯1,π¯3=βˆ’1\overline{\pi}_{1},\overline{\pi}_{3}=-1
Figure 1. A Visual Representation of P121P_{121}

Since Fc​h​a​rN​(PΞ±)=RΞ±F_{char}^{N}(P_{\alpha})=R_{\alpha} and χ​(RΞ±)=sΞ±\chi(R_{\alpha})=s_{\alpha},

Fc​h​a​r​(SΞ±)=sΞ±=χ​(Fc​h​a​rN​(PΞ±))=Fc​h​a​rQ​(PΞ±).F_{char}(S^{\alpha})=s_{\alpha}=\chi(F_{char}^{N}(P_{\alpha}))=F_{char}^{Q}(P_{\alpha}).

Similar coincidences have been observed elsewhere. For example, Huang and Rhoades in [4] expanded on the quasisymmetric Frobenius image of a β„‹n​(0)\mathcal{H}_{n}(0)-deformation of the generalized ring of coinvariants, showing that the result is the same symmetric polynomial as the symmetric Frobenius characteristic of the 𝔖n\mathfrak{S}_{n}-module. While their construction of the β„‹n​(0)\mathcal{H}_{n}(0)-module is specific to a quotient space of ℂ​[x1,…,xn]\mathbb{C}[x_{1},\dots,x_{n}] and not generalizable to all β„‹n​(0)\mathcal{H}_{n}(0)-modules, these coincidences suggest one should think more generally about deformations of 𝔖n\mathfrak{S}_{n}-modules which result in the same quasisymmetric Frobenius image as the classical Frobenius image.

In particular, since the space of symmetric functions Sym is a subspace of the space of quasisymmetric functions QSym and β„‹n​(0)\mathcal{H}_{n}(0) is a deformation of 𝔖n\mathfrak{S}_{n}, a natural question is:

Main Question 1.

What map ff causes the following diagram to commute, where ΞΉ\iota denotes inclusion?

Rep’ns of 𝔖n\mathfrak{S}_{n}222Formally, ff should be a map from the Grothendieck group of ℂ​[𝔖n]\mathbb{C}[\mathfrak{S}_{n}] to the Grothendieck group of β„‹n​(0)\mathcal{H}_{n}(0), as explained in Section 2.6.SymRep’ns of β„‹n​(0)\mathcal{H}_{n}(0)QSymFc​h​a​r\scriptstyle{F_{char}}f\scriptstyle{f}ΞΉ\scriptstyle{\iota}Fc​h​a​rQ\scriptstyle{F_{char}^{Q}}

This paper provides an answer to this question. Our work suggests and is inspired by a broader open question, which we do not attempt to answer, but which motivates our approach:

Motivation.

Can one determine the symmetric Frobenius image of an 𝔖n\mathfrak{S}_{n}-module by computing the quasisymmetric Frobenius image of a related β„‹n​(0)\mathcal{H}_{n}(0)-module?

In Section 2, we give necessary background on the representations of 𝔖n\mathfrak{S}_{n} and the representations of β„‹n​(0)\mathcal{H}_{n}(0). Next in Sections 3 and 4, we give two 𝔖n\mathfrak{S}_{n}-modules and their corresponding β„‹n​(0)\mathcal{H}_{n}(0) modules. These two examples suggest a broader implication in Section 5, where every 𝔖n\mathfrak{S}_{n}-module may be deformed in such a way. However, in Section 6, we explore our final example, where the implication and subsequent restrictions must be weakened. Finally, we conclude this paper in Section 7 by comparing this paper to previous work, leading to two open questions that may be explored in future work.

2. Background

2.1. The Symmetric Group 𝔖n\mathfrak{S}_{n} and its Representations

While we give a brief review in this section of relevant facts and notation regarding 𝔖n\mathfrak{S}_{n} and its representations, we assume the reader has some familiarity with these 1423ics. The less familiar reader who is motivated to learn more is recommended to consult [5] and [6].

Recall that the symmetric group 𝔖n\mathfrak{S}_{n} is generated by the simple transpositions sis_{i}, where

si2\displaystyle s_{i}^{2} =1​ for ​i≀nβˆ’1\displaystyle=1\text{ for }i\leq n-1
si​sj\displaystyle s_{i}s_{j} =sj​si​ for ​|iβˆ’j|β‰₯2\displaystyle=s_{j}s_{i}\text{ for }|i-j|\geq 2
si​si+1​si\displaystyle s_{i}s_{i+1}s_{i} =si+1​si​si+1.\displaystyle=s_{i+1}s_{i}s_{i+1}.

Given a permutation w=w1​w2​⋯​wnβˆˆπ”–nw=w_{1}w_{2}\cdots w_{n}\in\mathfrak{S}_{n}, we define the following statistics: let the descent set of ww be

des​(w)={i∣wi>wi+1},\text{des}(w)=\{i\mid w_{i}>w_{i+1}\},

the major index of ww to be

maj​(w)=βˆ‘i∈des​(w)i,\text{maj}(w)=\sum\limits_{i\in\text{des}(w)}i,

and the i-descent set of ww to be

ides​(w)={i∣i​ occurs after ​i+1​ in ​w}.\text{ides}(w)=\{i\mid i\text{ occurs after }i+1\text{ in }w\}.

Now, for w=w1​w2​⋯​wnw=w_{1}w_{2}\cdots w_{n}, let wβˆ’1w^{-1} represent the inverse of the permutation ww and rev​(w)\text{rev}(w) be the reverse of ww such that rev​(w)=wn​wnβˆ’1​⋯​w1\text{rev}(w)=w_{n}w_{n-1}\cdots w_{1}. With this, we offer an alternative definition of the i-descent set:

ides={i∣wiβˆ’1>wi+1βˆ’1}\text{ides}=\{i\mid w^{-1}_{i}>w^{-1}_{i+1}\}

Next, we introduce several basic notions in the representation theory of a finite group GG. Recall that a well-defined action of GG on a vector space MM is a group homomorphism from GG to G​L​(M)GL(M). This group homomorphism is called a representation of GG, and we say that MM is a GG-module. Given a GG-module MM, we define the character of an element g∈Gg\in G, written χ​(g)\chi(g), by the trace of the linear transformation Mβ†’MM\rightarrow M defined by v↦g​vv\mapsto gv. The character Ο‡\chi is constant on conjugacy classes; in the case of G=𝔖nG=\mathfrak{S}_{n}, the conjugacy classes are determined by the cycle types of permutations, which is always a partition of nn. Thus, χμ=χ​(Οƒ)\chi_{\mu}=\chi(\sigma), where Οƒ\sigma is any permutation of cycle type ΞΌ\mu.

A submodule AβŠ†MA\subseteq M is a subspace of MM that is closed under the action of GG. An irreducible submodule is one where it has no non-trivial submodules. In [5], it is shown that any 𝔖n\mathfrak{S}_{n}-module can be written as a direct sum of its irreducible submodules, that is, the group algebra, ℂ​[𝔖n]\mathbb{C}[\mathfrak{S}_{n}], is semisimple. The irreducible submodules of 𝔖n\mathfrak{S}_{n} are usually referred to as the Specht modules SΞ»S^{\lambda}.

2.2. Combinatorial Objects

In the representation theory of the symmetric group 𝔖n\mathfrak{S}_{n}, combinatorial objects like tableaux play an important role.

Recall that a partition ΞΌ=(ΞΌ1,ΞΌ2,…,ΞΌk)\mu=(\mu_{1},\mu_{2},\dots,\mu_{k}) of nn, written μ⊒n\mu\vdash n, is a weakly decreasing sequence of positive integers that sum to nn and that l​(ΞΌ)=kl(\mu)=k is the length of ΞΌ\mu. Similarly, recall a weak composition Ξ±=(Ξ±1,Ξ±2,…,Ξ±k)\alpha=(\alpha_{1},\alpha_{2},\dots,\alpha_{k}) of nn, written α⊨wn\alpha\vDash_{w}n, is a sequence of nonnegative integers that sum to nn. In the case where all integers are strictly positive, a strong composition is denoted by α⊨n\alpha\vDash n. Partitions give rise to Young tableaux, whereby in French notation, we stack ΞΌ1\mu_{1} boxes in the bottom row, ΞΌ2\mu_{2} boxes in the second, etc. Given another partition Ξ»\lambda, we say a Young tableau has content Ξ»\lambda if we fill the tableaux with Ξ»1\lambda_{1} 1s, Ξ»2\lambda_{2} 2s, etc.

We say that a Young tableau of shape ΞΌ\mu and content Ξ»\lambda is semistandard if the filling is such that the rows are weakly increasing left to right and the columns are strictly increasing bottom to 1423. We denote the space of semistandard Young tableaux of shape ΞΌ\mu and content Ξ»\lambda by SSYT(ΞΌ,Ξ»\mu,\lambda), where the number of semistandard Young tableaux of shape ΞΌ\mu and content Ξ»\lambda is counted by the Kostka numbers Kμ​λK_{\mu\lambda}. On the other hand, a Young tableau of shape ΞΌ\mu and content 1n1^{n} is standard if both the rows and the columns are strictly increasing. The space of standard Young tableaux of shape ΞΌ\mu is denoted by SYT(ΞΌ\mu), where fΞ»=|SYT​(ΞΌ)|f^{\lambda}=|\text{SYT}(\mu)|.

Example 2.1.

The following tableau is a semistandard Young tableau of shape (3,1)(3,1) and content (2,1,1)(2,1,1): \ytableausetupsmalltableaux

p=\ytableaushort​2​\none,113p={{\ytableaushort{2\none,113}}}

On the other hand, the following tableau is a standard Young tableau with the same shape:

S=\ytableaushort​4​\none,123S={{\ytableaushort{4\none,123}}}

A useful fact, which we will use later, restricts the relative positions of ii and i+1i+1:

Fact 2.2.

Given a standard tableau TT, ii must be directly below, to the left of, strictly southeast of, or strictly northwest of i+1i+1.

One important statistic on Young tableaux is the i-descent set; given a tableau tt, we say

ides​(t)={i∣i​ is south of ​i+1​ in ​t}.\text{ides}(t)=\{i\mid i\text{ is south of }i+1\text{ in }t\}.

For the readers familiar with reading words, note that the definition of the i-descent set of a tableau is precisely the i-descent set of the reading word of tt. Using this definition of the i-descent set, we define the major index of tt:

maj​(t)=βˆ‘i∈ides​(t)i.\text{maj}(t)=\sum\limits_{i\in\text{ides}(t)}i.

Furthermore, we can define a partial ordering on Young tableaux with content 1n1^{n}. In fact, it can be shown that this ordering is a total ordering on standard Young tableaux.

Definition 2.3.

Given two standard tableaux SS and TT, we say that S≺TS\prec T if and only if the smallest letter that lies in a different position is further north in TT than SS.

We observe that in standard tableaux S≺TS\prec T, the smallest letter must be strictly further north and weakly further west in TT than in SS. For this reason, we use northwest to emphasize that we are comparing standard tableaux.

Example 2.4.

Given the following two tableaux of shape (3,2,1)(3,2,1),

\ytableausetup​s​m​a​l​l​t​a​b​l​e​a​u​x​S=\ytableaushort​4,35,126​ and ​T=\ytableaushort​6,24,135,\ytableausetup{smalltableaux}S={{\ytableaushort{4,35,126}}}\text{ and }T={{\ytableaushort{6,24,135}}},

we have S≺T,S\prec T, since 2 is further to the northwest in TT than in SS.

We define a natural action of 𝔖n\mathfrak{S}_{n} on tableaux by si​t=ss_{i}t=s, where ss is the tableau that results from swapping the letters ii and i+1i+1 in the filling of tt. Note that the sis_{i} action on a standard tableau TT does not necessarily guarantee that SS is a standard tableau.

Lemma 2.5.

If TT and si​Ts_{i}T are standard tableaux, then Tβ‰Ίsi​TT\prec s_{i}T if and only if i+1i+1 is strictly northwest of ii in TT.

Proof.

First, given standard tableau TT, we note that si​Ts_{i}T is standard only if ii is strictly northwest or southeast of i+1i+1 by Fact 2.2.

Now, if ii is strictly southeast of i+1i+1 in TT, then ii is strictly northwest of i+1i+1 in si​Ts_{i}T. So, by definition, we have Tβ‰Ίsi​TT\prec s_{i}T.

On the other hand, if Tβ‰Ίsi​TT\prec s_{i}T and si​Ts_{i}T is standard, we can infer from the definition of dominance ordering that ii must be strictly northwest of i+1i+1 in si​Ts_{i}T. Therefore, i+1i+1 must be strictly northwest of ii in TT, as desired. ∎

We can further define an equivalence relation on Young tableaux with content 1n1^{n} by saying that t∼rst\sim_{r}s iff each row of tt contains the same letters as the corresponding row of ss. The resulting equivalence classes are tabloids of shape μ\mu, written {t}\{t\}. Each equivalence class has a tableau that is strictly increasing in all rows, and we refer to that tableau as the tabloid representative.

Example 2.6.

If t=\ytableausetup​s​m​a​l​l​t​a​b​l​e​a​u​x​{ytableau}​3​\none​\none​124t=\ytableausetup{smalltableaux}{{\ytableau 3&\none&\none\\ 1&2&4}}, then the tabloid {t}=\ytableausetup​b​o​x​s​i​z​e=n​o​r​m​a​l,t​a​b​l​o​i​d​s,s​m​a​l​l​t​a​b​l​e​a​u​x​{ytableau}​3​\none​\none​124\{t\}=\ytableausetup{boxsize=normal,tabloids,smalltableaux}{{\ytableau 3&\none&\none\\ 1&2&4}} is the equivalence class: \ytableausetupnotabloids,smalltableaux

{t}={{ytableau}​3​\none​\none​124,{ytableau}​3​\none​\none​214,{ytableau}​3​\none​\none​142,{ytableau}​3​\none​\none​241,{ytableau}​3​\none​\none​412,{ytableau}​3​\none​\none​421}\{t\}=\left\{{{\ytableau 3&\none&\none\\ 1&2&4}},{{\ytableau 3&\none&\none\\ 2&1&4}},{{\ytableau 3&\none&\none\\ 1&4&2}},{{\ytableau 3&\none&\none\\ 2&4&1}},{{\ytableau 3&\none&\none\\ 4&1&2}},{{\ytableau 3&\none&\none\\ 4&2&1}}\right\}

Going forward, we introduce two well-studied 𝔖n\mathfrak{S}_{n}-modules, which are each based on tabloids.

Definition 2.7.

Given a partition Ξ»\lambda, let MΞ»M^{\lambda} be the vector space generated by tabloids of shape Ξ»\lambda.

We will make use of the following bijection on this basis of MΞ»M^{\lambda}:

Fact 2.8.

Tabloids of shape Ξ»\lambda are in bijection with the set of words of type rev​(Ξ»)\text{rev}(\lambda) W​(rev​(Ξ»))W(\text{rev}(\lambda)). The straightforward bijection is given as {t}↦w\{t\}\mapsto w, where w=w1​w2​⋯​wnw=w_{1}w_{2}\cdots w_{n} and wi=l​(Ξ»)βˆ’ri+1w_{i}=\text{l}(\lambda)-r_{i}+1, where rir_{i} is the row of {t}\{t\} in which the letter ii sits.

Example 2.9.

Consider the tabloid {t}=\ytableausetup​b​o​x​s​i​z​e=n​o​r​m​a​l,t​a​b​l​o​i​d​s,s​m​a​l​l​t​a​b​l​e​a​u​x​{ytableau}​7​\none​\none​1452368\{t\}=\ytableausetup{boxsize=normal,tabloids,smalltableaux}\ytableau 7&\none&\none\\ 1&4&5\\ 2&3&6&8, then wt=23322313w_{t}=23322313.

Now, we give some statistics on tabloids, as well as some related results. Given a tabloid {t}∈Mλ\{t\}\in M^{\lambda}, let wtw_{t} be the word of content λ\lambda defined by the bijection in Fact 2.8. We give an analogous definition of i-descent sets on tabloids.

Definition 2.10.

For a tabloid {t}∈Mλ\{t\}\in M^{\lambda}, define the i-descent set of {t}\{t\} as

ides​({t})={i∣i​ is lower than ​i+1​ in ​t}\text{ides}(\{t\})=\{i\mid i\text{ is lower than }i+1\text{ in }t\}

Our bijection on tabloids was chosen to have the following property:

Lemma 2.11.

By construction, for {t}∈MΞ»\{t\}\in M^{\lambda} and corresponding wt∈W​(rev​(Ξ»))w_{t}\in W(\text{rev}(\lambda)), we have

ides​({t})=des​(wt).\text{ides}(\{t\})=\text{des}(w_{t}).

We also define an analogous sis_{i} action on tabloids, where we swap the letters ii and i+1i+1. Under this action, MΞ»M^{\lambda} is an 𝔖n\mathfrak{S}_{n}-module.

Finally, we make explicit a partial ordering on MΞ»M^{\lambda}, by extending β‰Ί\prec in Definition 2.3.

Definition 2.12.

Given tabloids {s}\{s\} and {t}\{t\} of shape Ξ»\lambda, we say {s}β‰Ί{t}\{s\}\prec\{t\} if the smallest letter in a different position is further north in {t}\{t\} than in {s}\{s\}.

For convenience, we often want a total ordering, rather than a partial ordering. Thus, extend this partial ordering to a total ordering β‰Ίc\prec_{c} on MΞ»M^{\lambda} arbitrarily.

Next, we consider a submodule of MΞ»M^{\lambda}. In particular, the Specht module SΞ»S^{\lambda} is defined using a signed sum of tabloids. Given a tabloid representative tt, we define a signed sum of tabloids based on the column space of tt, col​(t)\text{col}(t). The column space of tt is the cross product of sets of permutations that permute the letters within each column of tt. Define a polytabloid, ete_{t}, by a signed sum of tabloids:

et=βˆ‘Οƒβˆˆcol​(t)sgn​(Οƒ)​{σ​t}e_{t}=\sum\limits_{\sigma\in\text{col}(t)}\text{sgn}(\sigma)\{\sigma t\}

SΞ»S^{\lambda}, the vector space spanned by polytabloids of a given shape Ξ»\lambda, has basis {eT}T∈SYT​(Ξ»)\{e_{T}\}_{T\in\text{SYT}(\lambda)}.

We can define an sis_{i} action on polytabloids to ensure that we have an 𝔖n\mathfrak{S}_{n}-module. If we define this action on the standard polytabloids, we would say that

si​eT=esi​T.s_{i}e_{T}=e_{s_{i}T}.

However, recall that si​Ts_{i}T is not necessarily a standard tableau. Thus, we employ a straightening algorithm through the use of Garnir elements (see [5] for a full description) to write the sis_{i} action in terms of the basis of SΞ»S^{\lambda}. In particular, we see that si​(eT)s_{i}(e_{T}) may be written as a linear combination of larger standard polytabloids with coefficients cT,Siβˆˆβ„‚c_{T,S}^{i}\in\mathbb{C}:

si​(eT)={βˆ’eTi,i+1​in the same columneTΒ±βˆ‘Tβ‰ΊScT,Si​eSi,i+1​in the same rowesi​Ti,i+1​ in different rows/columnss_{i}(e_{T})=\begin{cases}-e_{T}&i,i+1\hskip 5.69054pt\text{in the same column}\\ e_{T}\pm\sum\limits_{T\prec S}c_{T,S}^{i}e_{S}&i,i+1\hskip 5.69054pt\text{in the same row}\\ e_{s_{i}T}&i,i+1\text{ in different rows/columns}\end{cases}
Definition 2.13 (Specht module).

For λ⊒n\lambda\vdash n, the Specht module SΞ»S^{\lambda} is the submodule of MΞ»M^{\lambda} with basis {eT∣T∈SYT​(Ξ»)}\{e_{T}\mid T\in\text{SYT}(\lambda)\}.

These Specht modules form the irreducible representations of 𝔖n\mathfrak{S}_{n}. Thus, any 𝔖n\mathfrak{S}_{n}-module MM is isomorphic to a direct sum of Specht modules:

Mβ‰…β¨Ξ»βŠ’ncλ​SΞ»M\cong\bigoplus\limits_{\lambda\vdash n}c_{\lambda}S^{\lambda}

On this basis, we can once again define a total ordering that is analogous to the total ordering on standard Young tableaux.

Definition 2.14.

Given two polytabloids, we say eS≺eTe_{S}\prec e_{T} if and only if S≺TS\prec T.

2.3. Symmetric Functions

In this paper, we consider two subspaces of ℂ​[x1,…,xn]\mathbb{C}[x_{1},\dots,x_{n}]. 𝔖n\mathfrak{S}_{n} acts on ℂ​[x1,…,xn]\mathbb{C}[x_{1},\dots,x_{n}] by swapping indices such that, for 1<i≀n1<i\leq n,

si​f​(x1,…,xi,xi+1,…,xn)=f​(x1,…,xi+1,xi,…,xn).s_{i}f(x_{1},\dots,x_{i},x_{i+1},\dots,x_{n})=f(x_{1},\dots,x_{i+1},x_{i},\dots,x_{n}).

The first subspace is the space of symmetric functions Symn\text{Sym}_{n}. A polynomial fβˆˆβ„‚β€‹[x1,…,xn]f\in\mathbb{C}[x_{1},\dots,x_{n}] is considered symmetric if, for all 1≀i<n1\leq i<n,

si​f​(x1,…,xn)=f​(x1,…,xn).s_{i}f(x_{1},\dots,x_{n})=f(x_{1},\dots,x_{n}).

The symmetric functions form a graded subspace of ℂ​[x1,…,xn]\mathbb{C}[x_{1},\dots,x_{n}], and we list some of the relevant bases here.

Definition 2.15 (Bases of symmetric functions).

The power sum, the complete homogeneous, and the elementary symmetric functions are defined multiplicatively, so that pΞΌ=pΞΌ1​pΞΌ2​⋯​pΞΌlp_{\mu}=p_{\mu_{1}}p_{\mu_{2}}\cdots p_{\mu_{l}}, hΞΌ=hΞΌ1​hΞΌ2​⋯​hΞΌlh_{\mu}=h_{\mu_{1}}h_{\mu_{2}}\cdots h_{\mu_{l}}, and eΞΌ=eΞΌ1​eΞΌ2​⋯​eΞΌle_{\mu}=e_{\mu_{1}}e_{\mu_{2}}\cdots e_{\mu_{l}}, where

pk​(x1,…,xn)=βˆ‘i=1nxik,p_{k}(x_{1},\dots,x_{n})=\sum\limits_{i=1}^{n}x_{i}^{k},
hk​(x1,…,xn)=βˆ‘1≀i1≀⋯≀ik≀nxi1​⋯​xik,h_{k}(x_{1},\dots,x_{n})=\sum\limits_{1\leq i_{1}\leq\cdots\leq i_{k}\leq n}x_{i_{1}}\cdots x_{i_{k}},

and

ek​(x1,…,xn)=βˆ‘1<i1<β‹―<ik≀nxi1​⋯​xik.e_{k}(x_{1},\dots,x_{n})=\sum\limits_{1<i_{1}<\cdots<i_{k}\leq n}x_{i_{1}}\cdots x_{i_{k}}.
Definition 2.16 (Schur functions).

The Schur functions are defined using semi-standard Young tableaux:

sΞ»=βˆ‘T∈SSYT​(Ξ»)xTs_{\lambda}=\sum\limits_{T\in\text{SSYT}(\lambda)}x^{T}

We will need the relationships between some of these bases in this paper. For relevant proofs, see [6].

Lemma 2.17.
hΞ»=βˆ‘ΞΌβŠ’nKμ​λ​sΞΌ.h_{\lambda}=\sum\limits_{\mu\vdash n}K_{\mu\lambda}s_{\mu}.

Given any nn, it turns out that Symn\text{Sym}_{n} is isomorphic to the vector space of nn-dimensional representations of 𝔖n\mathfrak{S}_{n}. To see this, we introduce the Frobenius characteristic map, which employs the use of the character of a permutation. Going forward, we refer to this as the symmetric Frobenius characteristic image.

Definition 2.18 (Frobenius characteristic map).

The Frobenius characteristic of an 𝔖n\mathfrak{S}_{n}-module MM is defined using the character function Ο‡:𝔖nβ†’β„‚\chi:\mathfrak{S}_{n}\rightarrow\mathbb{C}, where χμ\chi_{\mu} is the image of a permutation with cycle type ΞΌ\mu under Ο‡\chi:

Fc​h​a​r​(M)=βˆ‘ΞΌβŠ’nχμ​pΞΌzΞΌF_{char}(M)=\sum\limits_{\mu\vdash n}\chi_{\mu}\frac{p_{\mu}}{z_{\mu}}

We note that, if MM is graded module, then one may encode the degree of the submodules using a graded version of the Frobenius characteristic, Fc​h​a​r​(M;q)F_{char}(M;q).

Using the above definition, we obtain the following results about the modules defined in the previous section (see [5] for relevant proofs):

Fc​h​a​r​(SΞ»)=sΞ»F_{char}(S^{\lambda})=s_{\lambda}
Fc​h​a​r​(MΞ»)=hΞ»F_{char}(M^{\lambda})=h_{\lambda}

2.4. Quasisymmetric Functions

The second subspace of formal power series that we will use often is the the quasisymmetric functions QSymn\text{QSym}_{n}. Like Sym, QSym is a sub-Hopf algebra of ℂ​[x1,…,xn]\mathbb{C}[x_{1},\dots,x_{n}]. Quasisymmetric functions are shift invariant, so the coefficient of x1Ξ±1​⋯​xkΞ±kx_{1}^{\alpha_{1}}\cdots x_{k}^{\alpha_{k}} is equal to the coefficient of xi1Ξ±1​⋯​xikΞ±kx_{i_{1}}^{\alpha_{1}}\cdots x_{i_{k}}^{\alpha_{k}}, where i1<i2<β‹―<iki_{1}<i_{2}<\cdots<i_{k}.

Definition 2.19 (Gessel’s fundamental quasisymmetric functions).

Define the Gessel’s fundamental quasisymmetric function FKF_{K} as the following sum:

FK=βˆ‘1≀i1≀i2≀⋯≀ikj∈Kβ‡’ij<ij+1xi1​xi2​⋯​xikF_{K}=\sum\limits_{\begin{subarray}{c}1\leq i_{1}\leq i_{2}\leq\cdots\leq i_{k}\\ j\in K\Rightarrow i_{j}<i_{j+1}\end{subarray}}x_{i_{1}}x_{i_{2}}\cdots x_{i_{k}}

.

Example 2.20.
F{1}​(x1,x2,x3)=x1​x22+x1​x2​x3+x1​x32+x2​x32F_{\{1\}}(x_{1},x_{2},x_{3})=x_{1}x_{2}^{2}+x_{1}x_{2}x_{3}+x_{1}x_{3}^{2}+x_{2}x_{3}^{2}

Every symmetric function is shift invariant, so SymnβŠ‚QSymn\text{Sym}_{n}\subset\text{QSym}_{n}. As a result, we are able to write any symmetric functions in terms of the fundamental quasisymmetric functions.

Fact 2.21.

We can write the Schur functions as the following sum:

sΞ»=βˆ‘T∈SYT​(Ξ»)Fides​(T)s_{\lambda}=\sum\limits_{T\in\text{SYT}(\lambda)}F_{\text{ides}(T)}

2.5. The 0-Hecke Algebra β„‹n​(0)\mathcal{H}_{n}(0) and its Representations

The 0-Hecke algebra is a deformation of 𝔖n\mathfrak{S}_{n} with generators π¯i\overline{\pi}_{i} that satisfy the following relations.

π¯i2\displaystyle\overline{\pi}_{i}^{2} =βˆ’Ο€Β―i​ for ​i≀nβˆ’1\displaystyle=-\overline{\pi}_{i}\text{ for }i\leq n-1
π¯i​π¯j\displaystyle\overline{\pi}_{i}\overline{\pi}_{j} =π¯j​π¯i​ for ​|iβˆ’j|β‰₯2\displaystyle=\overline{\pi}_{j}\overline{\pi}_{i}\text{ for }|i-j|\geq 2
π¯i​π¯i+1​π¯i\displaystyle\overline{\pi}_{i}\overline{\pi}_{i+1}\overline{\pi}_{i} =π¯i+1​π¯i​π¯i+1\displaystyle=\overline{\pi}_{i+1}\overline{\pi}_{i}\overline{\pi}_{i+1}

Given wβˆˆπ”–nw\in\mathfrak{S}_{n} written as a product of simple transpositions, we can define π¯w\overline{\pi}_{w} multiplicatively.

The irreducible representations of β„‹n​(0)\mathcal{H}_{n}(0) are one-dimensional submodules naturally indexed by subsets of [nβˆ’1][n-1]. We can define these one-dimensional irreducibles using the actions of the generators π¯i\overline{\pi}_{i}.

Definition 2.22 (Irreducible β„‹n​(0)\mathcal{H}_{n}(0)-modules).

For IβŠ†[nβˆ’1]I\subseteq[n-1], let CIC_{I} be the one-dimensional β„‹n​(0)\mathcal{H}_{n}(0) representation defined by

CI​(π¯i)={0iβˆ‰Iβˆ’1i∈IC_{I}(\overline{\pi}_{i})=\begin{cases}0&i\notin I\\ -1&i\in I\end{cases}

In [2], it is shown that {CI∣IβŠ†[nβˆ’1]}\{C_{I}\mid I\subseteq[n-1]\} are all of the 2nβˆ’12^{n-1} irreducible representations of β„‹n​(0)\mathcal{H}_{n}(0).

Using this characterization of the irreducible submodules, Krob and Thibon [1] define the following characteristic map:

Definition 2.23 (Quasisymmetric Frobenius characteristic).

Given an irreducible β„‹n​(0)\mathcal{H}_{n}(0)-submodule CIC_{I}, define the quasisymmetric Frobenius characteristic map using fundamental quasisymmetric functions:

Fc​h​a​rQ​(CI)=FIF_{char}^{Q}(C_{I})=F_{I}

Now, given an arbitrary β„‹n​(0)\mathcal{H}_{n}(0)-module MM, let M=M1βŠƒM2βŠƒβ‹―β€‹Mk=βˆ…M=M_{1}\supset M_{2}\supset\cdots M_{k}=\emptyset be a finite composition series for MM. Then, since each \faktor​Mj​Mj+1\faktor{M_{j}}{M_{j+1}} is irreducible, define

Fc​h​a​rQ​(M)=βˆ‘j=1kβˆ’1Fc​h​a​rQ​(\faktor​Mj​Mj+1).F_{char}^{Q}(M)=\sum\limits_{j=1}^{k-1}F_{char}^{Q}\left(\faktor{M_{j}}{M_{j+1}}\right).

By the Jordan-HΓΆlder Theorem, the collection of isomorphism classes of these irreducible quotients is uniquely determined. Moreover, since each quotient is irreducible, it is isomorphic to CIC_{I} for some II. Moving forward, we use [β‹…]j[\cdot]_{j} to denote equivalence classes in \faktor​Mj​Mj+1\faktor{M_{j}}{M_{j+1}}. Then, if [v]j[v]_{j} is the generator of \faktor​Mj​Mj+1\faktor{M_{j}}{M_{j+1}},

Fc​h​a​rQ​(\faktor​Mj​Mj+1)=F{iβˆ£Ο€Β―i​v=βˆ’v}.F_{char}^{Q}\left(\faktor{M_{j}}{M_{j+1}}\right)=F_{\{i\mid\overline{\pi}_{i}v=-v\}}.

2.6. Grothendieck Groups

Through this point, we have two Frobenius characteristic maps: the symmetric Frobenius map on the representations of 𝔖n\mathfrak{S}_{n}, Fc​h​a​rF_{char}, and the quasisymmetric Frobenius map on the representations of β„‹n​(0)\mathcal{H}_{n}(0), Fc​h​a​rQF_{char}^{Q}. In this section, we offer a more formal, and uniform, definition of these two maps.

Definition 2.24 (Grothendieck group).

For an algebra AA, consider the β„€\mathbb{Z}-vector space of isomorphism classes of finite-dimensional AA-modules ℛ​(A)\mathcal{R}(A). Let [M][M] denote the isomorphism class of MM. Then, the Grothendieck group 𝒒0​(A)\mathcal{G}_{0}(A) is the quotient space of ℛ​(A)\mathcal{R}(A) obtained by the relation [M]=[M1]+[M2][M]=[M_{1}]+[M_{2}] when we have the short exact sequence 0β†’M1β†’Mβ†’M2β†’00\rightarrow M_{1}\rightarrow M\rightarrow M_{2}\rightarrow 0.

This simplifies considerably if the algebra is semisimple, as in the case of ℂ​[𝔖n]\mathbb{C}[\mathfrak{S}_{n}], the group algebra of 𝔖n\mathfrak{S}_{n}. For any algebra, when 0β†’M1→𝑓M→𝑔M2β†’00\rightarrow M_{1}\xrightarrow{f}M\xrightarrow{g}M_{2}\rightarrow 0 is a short exact sequence, we have M2β‰…\faktor​M​im​(f)M_{2}\cong\faktor{M}{\text{im}(f)} with im​(f)β‰…M1\text{im}(f)\cong M_{1} a submodule of MM. If the algebra is semisimple, however, then this short exact sequence always splits so that Mβ‰…M1βŠ•M2M\cong M_{1}\oplus M_{2}. Thus, if the algebra AA is semisimple, the relation defining 𝒒0​(A)\mathcal{G}_{0}(A) is simply [M]=[M1]+[M2][M]=[M_{1}]+[M_{2}], when Mβ‰…M1βŠ•M2M\cong M_{1}\oplus M_{2}, without the added complexity of short exact sequences.

Since β„‹n​(0)\mathcal{H}_{n}(0) is not semisimple, a finitely-generated module MM cannot necessarily be written as a direct sum of its irreducible submodules. The irreducible representations CIC_{I} form the basis of the Grothendieck group of β„‹n​(0)\mathcal{H}_{n}(0) 𝒒0​(β„‹n​(0))\mathcal{G}_{0}(\mathcal{H}_{n}(0)). Using this and the Specht modules SΞ»S^{\lambda} as the basis for 𝒒0​(ℂ​[𝔖n])\mathcal{G}_{0}(\mathbb{C}[\mathfrak{S}_{n}]), one can formalize the definitions of the symmetric Frobenius characteristic map and the quasisymmetric Frobenius characteristic map. In particular, these linear transformations are defined by

Fc​h​a​r:𝒒0​(ℂ​[𝔖n])\displaystyle F_{char}:\mathcal{G}_{0}\left(\mathbb{C}[\mathfrak{S}_{n}]\right) β†’Symn\displaystyle\rightarrow\text{Sym}^{n}
[SΞ»]\displaystyle[S^{\lambda}] ↦sΞ»\displaystyle\mapsto s_{\lambda}

and

Fc​h​a​rQ:𝒒0​(β„‹n​(0))\displaystyle F_{char}^{Q}:\mathcal{G}_{0}\left(\mathcal{H}_{n}(0)\right) β†’QSymn\displaystyle\rightarrow\text{QSym}^{n}
[CI]\displaystyle[C_{I}] ↦FI\displaystyle\mapsto F_{I}

Readers wishing for a first concrete example of applying this map are encouraged to consult [7], where Tewari and van Willigenburg give the nice example of a β„‹n​(0)\mathcal{H}_{n}(0)-module whose quasisymmetric Frobenius image is the quasischur basis for the quasisymmetric functions.

3. An β„‹n​(0)\mathcal{H}_{n}(0)-Module, MΞ»^\widehat{M^{\lambda}}

Our goal for this section is to deform the 𝔖n\mathfrak{S}_{n}-action on the permutation module MΞ»M^{\lambda} to create a new β„‹n​(0)\mathcal{H}_{n}(0)-module MΞ»^\widehat{M^{\lambda}} such that

Fc​h​a​rQ​(MΞ»^)=Fc​h​a​r​(MΞ»).F_{char}^{Q}(\widehat{M^{\lambda}})=F_{char}(M^{\lambda}).

To define a β„‹n​(0)\mathcal{H}_{n}(0)-action on MΞ»^\widehat{M^{\lambda}}, we use the vector space M22M^{22} as a guiding example. Recall the 𝔖4\mathfrak{S}_{4} action on M22M^{22} on the left of Figure 2, drawn such that if {s}\{s\} lies below {t}\{t\}, then {s}β‰Ί{t}\{s\}\prec\{t\}. Now for a valid β„‹n​(0)\mathcal{H}_{n}(0)-action, we must have π¯i2=βˆ’Ο€Β―i\overline{\pi}_{i}^{2}=-\overline{\pi}_{i}, and thus all edges must be directed. To this end, we choose one direction in the diagram and add loops when necessary. To ensure that the quasisymmetric Frobenius characteristic is the same as the symmetric Frobenius characteristic, we will direct the π¯i\overline{\pi}_{i} edges so that a non-trivial π¯i\overline{\pi}_{i} action on a tabloid creates a new i-descent. In particular, consider the right of Figure 2, where we direct the edges as desired, while adding β€œnegative” loops to satisfy the first commuting relation. Furthermore, note that all previous sis_{i} actions resulting in si​v=vs_{i}v=v cannot be analogously replaced by π¯i​v=v\overline{\pi}_{i}v=v and must be adjusted.

\ytableausetup

boxsize=normal,tabloids \ytableaushort​14,23{\ytableaushort{14,23}}\ytableaushort​24,13{\ytableaushort{24,13}}\ytableaushort​13,24{\ytableaushort{13,24}}\ytableaushort​12,34{\ytableaushort{12,34}}\ytableaushort​34,12{\ytableaushort{34,12}}\ytableaushort​23,14{\ytableaushort{23,14}}s1s_{1}s3s_{3}s2s_{2}s3s_{3}s2s_{2}s1s_{1}s2=1s_{2}=1s1=s3=1s_{1}=s_{3}=1s1=s3=1s_{1}=s_{3}=1s2=1s_{2}=1 \ytableausetup\ytableaushort​14,23{\ytableaushort{14,23}}\ytableaushort​24,13{\ytableaushort{24,13}}\ytableaushort​13,24{\ytableaushort{13,24}}\ytableaushort​12,34{\ytableaushort{12,34}}\ytableaushort​34,12{\ytableaushort{34,12}}\ytableaushort​23,14{\ytableaushort{23,14}}000π¯1\overline{\pi}_{1}π¯3\overline{\pi}_{3}π¯2\overline{\pi}_{2}π¯3\overline{\pi}_{3}π¯2\overline{\pi}_{2}π¯1\overline{\pi}_{1}π¯2\overline{\pi}_{2}π¯2\overline{\pi}_{2}π¯1,π¯3\overline{\pi}_{1},\overline{\pi}_{3}π¯1,π¯3\overline{\pi}_{1},\overline{\pi}_{3}π¯1=π¯3=βˆ’1\overline{\pi}_{1}=\overline{\pi}_{3}=-1π¯2=βˆ’1\overline{\pi}_{2}=-1π¯2=βˆ’1\overline{\pi}_{2}=-1π¯1=βˆ’1\overline{\pi}_{1}=-1π¯3=βˆ’1\overline{\pi}_{3}=-1

Figure 2. An 𝔖4\mathfrak{S}_{4}-action on M22M^{22} and an β„‹4​(0)\mathcal{H}_{4}(0)-action on M22^\widehat{M^{22}}

With this figure in mind, we define the following β„‹n​(0)\mathcal{H}_{n}(0)-action on the tabloids of MΞ»^\widehat{M^{\lambda}}:

Definition 3.1.
π¯i​{t}={0i,i+1​ in the same rowβˆ’{t}i​ south of ​i+1si​{t}i​ north of ​i+1\overline{\pi}_{i}\{t\}=\begin{cases}0&i,i+1\text{ in the same row}\\ -\{t\}&i\text{ south of }i+1\\ s_{i}\{t\}&i\text{ north of }i+1\end{cases}

The next three propositions will show that this action is a valid β„‹n​(0)\mathcal{H}_{n}(0)-action.

Proposition 3.2.

The action defined on MΞ»^\widehat{M^{\lambda}} satisfies π¯i2=βˆ’Ο€Β―i\overline{\pi}_{i}^{2}=-\overline{\pi}_{i}.

Proof.

We have three cases to consider, based on the relative positions of ii and i+1i+1 in an arbitrary tabloid {t}\{t\}.

For the first and shortest case, suppose ii and i+1i+1 are in the same row in {t}\{t\}. Then, si​{t}={t}s_{i}\{t\}=\{t\}. Thus, π¯i​{t}=0\overline{\pi}_{i}\{t\}=0. This is easy to see that π¯i2=βˆ’Ο€Β―i\overline{\pi}_{i}^{2}=-\overline{\pi}_{i} trivially.

For the second case, suppose ii is north of i+1i+1 in {t}\{t\}. Then, π¯i​{t}=si​{t}\overline{\pi}_{i}\{t\}=s_{i}\{t\}, by definition of the action. In si​{t}s_{i}\{t\}, ii is now south of i+1i+1, so π¯i2​{t}=π¯i​(si​{t})=βˆ’si​{t}=βˆ’Ο€Β―i​{t}\overline{\pi}_{i}^{2}\{t\}=\overline{\pi}_{i}(s_{i}\{t\})=-s_{i}\{t\}=-\overline{\pi}_{i}\{t\}.

For the third and final case, suppose ii is south of i+1i+1. Then, in si​{t}s_{i}\{t\}, ii is north of i+1i+1. Thus, by definition, π¯i​{t}=βˆ’{t}\overline{\pi}_{i}\{t\}=-\{t\}, and π¯i​(βˆ’{t})={t}=βˆ’Ο€Β―i​{t}\overline{\pi}_{i}(-\{t\})=\{t\}=-\overline{\pi}_{i}\{t\}. ∎

Proposition 3.3.

The action defined on MΞ»^\widehat{M^{\lambda}} satisfies π¯i​π¯j=π¯j​π¯i\overline{\pi}_{i}\overline{\pi}_{j}=\overline{\pi}_{j}\overline{\pi}_{i} when |iβˆ’j|β‰₯2|i-j|\geq 2.

Proof.

It is straightforward to see that the relative positions of ii and i+1i+1 and the relative positions of jj and j+1j+1 have no effect on π¯j\overline{\pi}_{j} and π¯i\overline{\pi}_{i}, respectively. ∎

Proposition 3.4.

The action defined on MΞ»^\widehat{M^{\lambda}} satisfies π¯i​π¯i+1​π¯i=π¯i+1​π¯i​π¯i+1\overline{\pi}_{i}\overline{\pi}_{i+1}\overline{\pi}_{i}=\overline{\pi}_{i+1}\overline{\pi}_{i}\overline{\pi}_{i+1}.

Proof.

While there are 12 individual cases to check for an arbitrary tabloid {t}\{t\}, across three families, we proceed by showing one illustrative case from each of the three families. The remaining cases are similar to these illustrative cases and are straightforward computations.

Case 1.

For the first representative case, we assume that, in a tabloid {t}\{t\}, i+2i+2 is above ii, which is above i+1i+1. Then, we have

π¯i​π¯i+1​π¯i​{t}\displaystyle\overline{\pi}_{i}\overline{\pi}_{i+1}\overline{\pi}_{i}\{t\} =π¯i​π¯i+1​(si​{t})\displaystyle=\overline{\pi}_{i}\overline{\pi}_{i+1}(s_{i}\{t\})
=π¯i​(βˆ’si​{t})\displaystyle=\overline{\pi}_{i}(-s_{i}\{t\})
=si​{t}\displaystyle=s_{i}\{t\}
π¯i+1​π¯i​π¯i+1​{t}\displaystyle\overline{\pi}_{i+1}\overline{\pi}_{i}\overline{\pi}_{i+1}\{t\} =π¯i+1​π¯i​(βˆ’{t})\displaystyle=\overline{\pi}_{i+1}\overline{\pi}_{i}(-\{t\})
=π¯i+1​(βˆ’si​{t})\displaystyle=\overline{\pi}_{i+1}(-s_{i}\{t\})
=si​{t}\displaystyle=s_{i}\{t\}
Case 2.

Next, suppose ii and i+2i+2 are in the same row, while i+1i+1 is in a row above. Then,

π¯i​π¯i+1​π¯i​{t}\displaystyle\overline{\pi}_{i}\overline{\pi}_{i+1}\overline{\pi}_{i}\{t\} =π¯i​π¯i+1​(βˆ’{t})\displaystyle=\overline{\pi}_{i}\overline{\pi}_{i+1}(-\{t\})
=π¯i​(βˆ’si+1​{t})\displaystyle=\overline{\pi}_{i}(-s_{i+1}\{t\})
=0\displaystyle=0
π¯i+1​π¯i​π¯i+1​{t}\displaystyle\overline{\pi}_{i+1}\overline{\pi}_{i}\overline{\pi}_{i+1}\{t\} =π¯i+1​π¯i​(si+1​{t})\displaystyle=\overline{\pi}_{i+1}\overline{\pi}_{i}(s_{i+1}\{t\})
=0\displaystyle=0
Case 3.

Finally, suppose ii and i+2i+2 are in the same row, while i+1i+1 is in a row below. Then,

π¯i​π¯i+1​π¯i​{t}\displaystyle\overline{\pi}_{i}\overline{\pi}_{i+1}\overline{\pi}_{i}\{t\} =π¯i​π¯i+1​(si​{t})\displaystyle=\overline{\pi}_{i}\overline{\pi}_{i+1}(s_{i}\{t\})
=0\displaystyle=0
π¯i+1​π¯i​π¯i+1​{t}\displaystyle\overline{\pi}_{i+1}\overline{\pi}_{i}\overline{\pi}_{i+1}\{t\} =π¯i+1​π¯i​(βˆ’{t})\displaystyle=\overline{\pi}_{i+1}\overline{\pi}_{i}(-\{t\})
=π¯i+1​(βˆ’si​{t})\displaystyle=\overline{\pi}_{i+1}(-s_{i}\{t\})
=0\displaystyle=0

∎

Now that we have shown that this action is valid, we can start creating the composition series that will lead us to the quasisymmetric Frobenius characteristic.

𝐌𝟐𝟐^=𝐌𝟏\widehat{\mathbf{M^{22}}}=\mathbf{M_{1}}𝐌𝟐\mathbf{M_{2}}πŒπŸ‘\mathbf{M_{3}}\ytableaushort​14,23{\ytableaushort{14,23}}\ytableaushort​24,13{\ytableaushort{24,13}}\ytableaushort​13,24{\ytableaushort{13,24}}\ytableaushort​12,34{\ytableaushort{12,34}}\ytableaushort​34,12{\ytableaushort{34,12}}\ytableaushort​23,14{\ytableaushort{23,14}}000π¯1\overline{\pi}_{1}π¯3\overline{\pi}_{3}π¯2\overline{\pi}_{2}π¯3\overline{\pi}_{3}π¯2\overline{\pi}_{2}π¯1\overline{\pi}_{1}π¯2\overline{\pi}_{2}π¯2\overline{\pi}_{2}π¯1,π¯3\overline{\pi}_{1},\overline{\pi}_{3}π¯1,π¯3\overline{\pi}_{1},\overline{\pi}_{3}π¯1=π¯3=βˆ’1\overline{\pi}_{1}=\overline{\pi}_{3}=-1π¯2=βˆ’1\overline{\pi}_{2}=-1π¯2=βˆ’1\overline{\pi}_{2}=-1π¯1=βˆ’1\overline{\pi}_{1}=-1π¯3=βˆ’1\overline{\pi}_{3}=-1
Figure 3. Part of the Composition Series of M22^\widehat{M^{22}}

In Figure 3, we draw the first three submodules in the composition series MΞ»^=M1βŠƒM2βŠƒβ‹―β€‹Mk=βˆ…\widehat{M^{\lambda}}=M_{1}\supset M_{2}\supset\cdots M_{k}=\emptyset. We see that \faktor​Mi​Mi+1\faktor{M_{i}}{M_{i+1}} is indeed one-dimensional, generated by a single tabloid from the basis. If we consider the figure further, we can observe that we remove tabloids from largest to smallest according to β‰Ίc\prec_{c}. In particular, we can also define the following action that we claim is equivalent to Definition 3.1.

Definition 3.5.
π¯i​{t}={0si​{t}={t}βˆ’{t}{t}β‰Ίsi​{t}si​{t}si​{t}β‰Ί{t}\overline{\pi}_{i}\{t\}=\begin{cases}0&s_{i}\{t\}=\{t\}\\ -\{t\}&\{t\}\prec s_{i}\{t\}\\ s_{i}\{t\}&s_{i}\{t\}\prec\{t\}\end{cases}

It is straightforward to see that Definition 3.1 and Definition 3.5 are equivalent. Thus, we can assume that the commuting relations hold for this action, as well. In particular, this action will make it easier to define the composition series, so that each MiM_{i} is determined by removing tabloids that are greater according to β‰Ίc\prec_{c}. Again, referring to Figure 3, we can work through the quasisymmetric Frobenius characteristic map.

Example 3.6 (Frobenius Characteristic of M22^\widehat{M^{22}}).

Given \faktor​Mj​Mj+1β‰…CI\faktor{M_{j}}{M_{j+1}}\cong C_{I} with generator [v]j[v]_{j}, we have I={iβˆ£Ο€Β―i​v=βˆ’v}I=\{i\mid\overline{\pi}_{i}v=-v\}. Thus,

Fc​h​a​rQ​(\faktor​Mj​Mj+1)=FI.F_{char}^{Q}(\faktor{M_{j}}{M_{j+1}})=F_{I}.

Again, from Figure 3, we can see that each successive quotient module \faktor​Mj​Mj+1\faktor{M_{j}}{M_{j+1}} is generated by a single equivalence class [{t}j]j[\{t\}_{j}]_{j}. So, for each tabloid, we have I={iβˆ£Ο€Β―i​{t}j=βˆ’{t}j}I=\{i\mid\overline{\pi}_{i}\{t\}_{j}=-\{t\}_{j}\}. Thus, for each tabloid, the β€œnegative” loops that appear will index the fundamental quasisymmetric function that appears in the quasisymmetric Frobenius characteristic image of M22^\widehat{M^{22}}. However, we know that a β€œnegative” loop occurs when we would remove an i-descent, so I=ides​({t})I=\text{ides}(\{t\}).

So, given the figure above, we have

Fc​h​a​rQ​(M22^)=Fβˆ…+F{1}+F{3}+2​F{2}+F{1,3}F_{char}^{Q}(\widehat{M^{22}})=F_{\emptyset}+F_{\{1\}}+F_{\{3\}}+2F_{\{2\}}+F_{\{1,3\}}

However, this sum of quasisymmetric functions is symmetric and can be written in terms of a basis of Sym. In particular, we have

Fc​h​a​rQ​(M22^)=h22F_{char}^{Q}(\widehat{M^{22}})=h_{22}

For the upcoming result, we will need a result of the RSK correspondence, which we state here for use later.

Fact 3.7.

Given a tabloid {t}∈Mλ\{t\}\in M^{\lambda}, let wtw_{t} be the word that corresponds to the tabloid.

Under RSK, wtw_{t} maps to the pair (P,Q)(P,Q) with PP, a semistandard Young tableau, and QQ, a standard Young tableau, where

ides​(Q)=des​(wt)=ides​({t})\text{ides}(Q)=\text{des}(w_{t})=\text{ides}(\{t\})

As anticipated, we have the following result.

Theorem 3.8.

There exists a β„‹n​(0)\mathcal{H}_{n}(0)-action on the permutation modules MΞ»^\widehat{M^{\lambda}} such that

Fc​h​a​rQ​(MΞ»^)=hΞ»=Fc​h​a​r​(MΞ»).F_{char}^{Q}(\widehat{M^{\lambda}})=h_{\lambda}=F_{char}(M^{\lambda}).
Proof.

Let {t}1,…,{t}k\{t\}_{1},\dots,\{t\}_{k} be the list of tabloids that form the basis of MΞ»^\widehat{M^{\lambda}} given in the total ordering consistent with dominance ordering, β‰Ίc\prec_{c}, where {t}1\{t\}_{1} is the largest. Define submodules Mj=span​({t}j,{t}j+1,…,{t}k)M_{j}=\text{span}(\{t\}_{j},\{t\}_{j+1},\dots,\{t\}_{k}). Thus, we have a composition series MΞ»^=M1βŠƒM2βŠƒβ‹―βŠƒMk=βˆ…\widehat{M^{\lambda}}=M_{1}\supset M_{2}\supset\cdots\supset M_{k}=\emptyset.

We note a couple of facts about this composition series. First, we observe that \faktor​Mj​Mj+1\faktor{M_{j}}{M_{j+1}} is generated by [{t}j]j[\{t\}_{j}]_{j}. Furthermore, this composition series is, in fact, a composition series of submodules as we have MjβŠƒMj+1M_{j}\supset M_{j+1}, by construction of our action.

Then, \faktor​Mj​Mj+1β‰…CI\faktor{M_{j}}{M_{j+1}}\cong C_{I}, where CI={iβˆ£Ο€Β―i​{t}j=βˆ’{t}j}C_{I}=\{i\mid\overline{\pi}_{i}\{t\}_{j}=-\{t\}_{j}\}. So, by the definition of the quasisymmetric Frobenius characteristic map, we have

Fc​h​a​rQ​(MΞ»^)=βˆ‘{t}∈MΞ»^F{iβˆ£Ο€Β―i​{t}=βˆ’{t}}F_{char}^{Q}(\widehat{M^{\lambda}})=\sum\limits_{\{t\}\in\widehat{M^{\lambda}}}F_{\{i\mid\overline{\pi}_{i}\{t\}=-\{t\}\}}

However, since π¯i​{t}=βˆ’{t}\overline{\pi}_{i}\{t\}=-\{t\} when ii is south of i+1i+1, this means that I={i∣i​ south of ​i+1​ in ​t}I=\{i\mid i\text{ south of }i+1\text{ in }t\}. By definition of the i-descent set of a tabloid, we can conclude that I=ides​({t})I=\text{ides}(\{t\}). Thus, overall, we have the following formula for the quasisymmetric Frobenius characteristic of MΞ»^\widehat{M^{\lambda}}:

Fc​h​a​rQ​(MΞ»^)\displaystyle F_{char}^{Q}(\widehat{M^{\lambda}}) =βˆ‘{t}∈MΞ»^Fides​({t})\displaystyle=\sum\limits_{\{t\}\in\widehat{M^{\lambda}}}F_{\text{ides}(\{t\})}
=βˆ‘w∈W​(rev​(Ξ»))Fdes​(w)\displaystyle=\sum\limits_{w\in W(\text{rev}(\lambda))}F_{\text{des}(w)}

where the second sum comes from the bijection on tabloids given in Fact 2.8 and Lemma 2.11.

Splitting the sum, we have

Fc​h​a​rQ​(MΞ»^)=βˆ‘BβŠ†[nβˆ’1]FBβ€‹βˆ‘w∈W​(rev​(Ξ»))πŸ™des​(w)=B,F_{char}^{Q}(\widehat{M^{\lambda}})=\sum\limits_{B\subseteq[n-1]}F_{B}\sum\limits_{w\in W(\text{rev}(\lambda))}\mathbbm{1}_{\text{des}(w)=B},

and by a result of the RSK correspondence given in Fact 3.7, we have

Fc​h​a​rQ​(MΞ»^)=βˆ‘BβŠ†[nβˆ’1]FBβ€‹βˆ‘ΞΌβŠ’nβˆ‘P∈SSYT​(ΞΌ,rev​(Ξ»))βˆ‘Q∈SYT​(ΞΌ)πŸ™ides​(Q)=B.F_{char}^{Q}(\widehat{M^{\lambda}})=\sum\limits_{B\subseteq[n-1]}F_{B}\sum\limits_{\mu\vdash n}\sum\limits_{P\in\text{SSYT}(\mu,\text{rev}(\lambda))}\sum\limits_{Q\in\text{SYT}(\mu)}\mathbbm{1}_{\text{ides}(Q)=B}.

Straightforward manipulations will then show that

Fc​h​a​rQ​(MΞ»^)\displaystyle F_{char}^{Q}(\widehat{M^{\lambda}}) =βˆ‘BβŠ†[nβˆ’1]βˆ‘ΞΌβŠ’nβˆ‘Q∈SYT​(ΞΌ)Kμ​rev​(Ξ»)β€‹πŸ™ides​(Q)=B​FB\displaystyle=\sum\limits_{B\subseteq[n-1]}\sum\limits_{\mu\vdash n}\sum\limits_{Q\in\text{SYT}(\mu)}K_{\mu\text{rev}(\lambda)}\mathbbm{1}_{\text{ides}(Q)=B}F_{B}
=βˆ‘ΞΌβŠ’nβˆ‘Q∈SYT​(ΞΌ)Kμ​rev​(Ξ»)​Fides​(Q)\displaystyle=\sum\limits_{\mu\vdash n}\sum\limits_{Q\in\text{SYT}(\mu)}K_{\mu\text{rev}(\lambda)}F_{\text{ides}(Q)}
=βˆ‘ΞΌβŠ’nKμ​rev​(Ξ»)β€‹βˆ‘Q∈SYT​(ΞΌ)Fides​(Q)\displaystyle=\sum\limits_{\mu\vdash n}K_{\mu\text{rev}(\lambda)}\sum\limits_{Q\in\text{SYT}(\mu)}F_{\text{ides}(Q)}
=βˆ‘ΞΌβŠ’nKμ​rev​(Ξ»)​sΞΌ\displaystyle=\sum\limits_{\mu\vdash n}K_{\mu\text{rev}(\lambda)}s_{\mu}

In his textbook [5], Sagan takes a representation theoretic approach to showing that Kμ​λ~=Kμ​λK_{\mu\tilde{\lambda}}=K_{\mu\lambda} for any rearrangement Ξ»~\tilde{\lambda} of Ξ»\lambda. Thus,

Fc​h​a​rQ​(MΞ»^)\displaystyle F_{char}^{Q}(\widehat{M^{\lambda}}) =βˆ‘ΞΌβŠ’nKμ​rev​(Ξ»)​sΞΌ\displaystyle=\sum\limits_{\mu\vdash n}K_{\mu\text{rev}(\lambda)}s_{\mu}
=βˆ‘ΞΌβŠ’nKμ​λ​sΞΌ\displaystyle=\sum\limits_{\mu\vdash n}K_{\mu\lambda}s_{\mu}
=hΞ»\displaystyle=h_{\lambda}

∎

4. An β„‹n​(0)\mathcal{H}_{n}(0)-module, SΞ»^\widehat{S^{\lambda}}

Given a valid β„‹n​(0)\mathcal{H}_{n}(0)-action on tabloids, we can extend to a valid action on polytabloids. In particular, recall that SΞ»S^{\lambda} is the Specht module, generated by standard polytabloids. We defined a dominance ordering on tableaux that was then extended to an ordering on polytabloids in Definition 2.14.

Similarly to the previous section, our main goals will be to define a related β„‹n​(0)\mathcal{H}_{n}(0)-action on SΞ»S^{\lambda} and a compatible composition series of submodules closed under that action:

SΞ»=M1βŠƒM2βŠƒβ‹―βŠƒMfΞ»βŠƒβˆ….S^{\lambda}=M_{1}\supset M_{2}\supset\cdots\supset M_{f^{\lambda}}\supset\emptyset.

As above, it turns out that selecting a total ordering on the basis {eT∣T∈SYT​(Ξ»)}\{e_{T}\mid T\in\text{SYT}(\lambda)\} can easily give rise to both. Using Definition 2.14 above, consider the ordering eT1β‰ΊeT2β‰Ίβ‹―β‰ΊeTfΞ»e_{T_{1}}\prec e_{T_{2}}\prec\cdots\prec e_{T_{f^{\lambda}}}. Then we can easily define the composition series:

span​(eT1,…,eTfΞ»)βŠƒspan​(eT2,…,eTfΞ»)βŠƒβ‹―βŠƒspan​(eTfΞ»)βŠƒβˆ….\text{span}(e_{T_{1}},\dots,e_{T_{f^{\lambda}}})\supset\text{span}(e_{T_{2}},\dots,e_{T_{f^{\lambda}}})\supset\cdots\supset\text{span}(e_{T_{f^{\lambda}}})\supset\emptyset.

Next, to ensure a compatible β„‹n​(0)\mathcal{H}_{n}(0) action to the composition series, it must be the case that if

π¯i​(eTj)=βˆ‘k=1fΞ»ck​eTk,\overline{\pi}_{i}(e_{T_{j}})=\sum\limits_{k=1}^{f^{\lambda}}c_{k}e_{T_{k}},

then ck=0c_{k}=0 when k>jk>j. Moreover, cjc_{j} must be either 0 or -1, since π¯i2=βˆ’Ο€Β―i\overline{\pi}_{i}^{2}=-\overline{\pi}_{i}.

To formalize the action, we will need the concepts of the leading term and trailing term with respect to a given basis and ordering.

Definition 4.1.

Assume ℬ={v1,…,vl}\mathcal{B}=\{v_{1},\dots,v_{l}\} is a basis of a vector space, given in decreasing order according to ≀\leq defined on ℬ\mathcal{B}.

Let the leading term of a vector vv be lt​(v)=lt​(βˆ‘ci​vi)=cj​vj\text{lt}(v)=\text{lt}(\sum c_{i}v_{i})=c_{j}v_{j}, where jj is the smallest index such that cjβ‰ 0c_{j}\neq 0.

Let the trailing term of a vector vv be tt​(v)=tt​(βˆ‘ci​vi)=cj​vj\text{tt}(v)=\text{tt}\left(\sum c_{i}v_{i}\right)=c_{j}v_{j}, where jj is the largest index such that cjβ‰ 0c_{j}\neq 0.

Now, we recall the 𝔖n\mathfrak{S}_{n}-action on polytabloids, which we will deform to define the compatible β„‹n​(0)\mathcal{H}_{n}(0)-action.

Recall.
si​(eT)={βˆ’eTi​andΒ i+1Β in the same columneTΒ±βˆ‘Tβ‰ΊScT,Si​eSi​andΒ i+1Β in the same rowesi​Ti​andΒ i+1Β in a different row and column,s_{i}(e_{T})=\begin{cases}-e_{T}&i\hskip 5.69054pt\text{and $i+1$ in the same column}\\ e_{T}\pm\sum\limits_{T\prec S}c_{T,S}^{i}e_{S}&i\hskip 5.69054pt\text{and $i+1$ in the same row}\\ e_{s_{i}T}&i\hskip 5.69054pt\text{and $i+1$ in a different row and column}\end{cases},

where cT,Siβˆˆβ„‚c_{T,S}^{i}\in\mathbb{C}.

Definition 4.2.
π¯i​eT={0tt​(si​eT)=eTβˆ’eTeTβ‰Ίesi​T​ or ​si​eT=βˆ’eTesi​Tesi​Tβ‰ΊeT\overline{\pi}_{i}e_{T}=\begin{cases}0&\text{tt}(s_{i}e_{T})=e_{T}\\ -e_{T}&e_{T}\prec e_{s_{i}T}\text{ or }s_{i}e_{T}=-e_{T}\\ e_{s_{i}T}&e_{s_{i}T}\prec e_{T}\end{cases}

Note that this π¯i\overline{\pi}_{i} action is more involved than that on tabloids. This is because si​(eT)s_{i}(e_{T}) may be a linear combination of standard polytabloids, when written in terms of the basis. Thus, we must account for additional cases when comparing via β‰Ί\prec. However, it is straightforward to see that, much like tabloids, there is an easier, equivalent action for which it will be more straightforward to show that the commuting relations hold.

Definition 4.3.
π¯i​eT={0i,i+1​ in the same row of ​Tβˆ’eTi​ southeast of ​i+1​ in ​Tesi​Ti​ northwest of ​i+1​ in ​T\overline{\pi}_{i}e_{T}=\begin{cases}0&i,i+1\text{ in the same row of }T\\ -e_{T}&i\text{ southeast of }i+1\text{ in }T\\ e_{s_{i}T}&i\text{ northwest of }i+1\text{ in }T\end{cases}

As with the actions on MΞ»^\widehat{M^{\lambda}}, it is relatively straightforward to see that these two β„‹n​(0)\mathcal{H}_{n}(0)-actions on SΞ»^\widehat{S^{\lambda}} are equivalent.

Proposition 4.4.

SΞ»^\widehat{S^{\lambda}} is a β„‹n​(0)\mathcal{H}_{n}(0)-module under the π¯i\overline{\pi}_{i} action defined above.

Proof.

Since both definitions are equivalent, we can use the second action to show that the commuting relations hold, and thus SΞ»^\widehat{S^{\lambda}} is a β„‹n​(0)\mathcal{H}_{n}(0)-module. The proofs for each of the commuting relations are analogous to those of Propositions 3.2, 3.3, and 3.4. ∎

As with MΞ»^\widehat{M^{\lambda}}, we see that the ordering β‰Ί\prec gives a natural way to construct a composition series of SΞ»^\widehat{S^{\lambda}} such that each quotient space is a one-dimensional irreducible β„‹n​(0)\mathcal{H}_{n}(0)-module that is isomorphic to some CIC_{I}. Recall that I={iβˆ£Ο€Β―i​eT=βˆ’eT}I=\{i\mid\overline{\pi}_{i}e_{T}=-e_{T}\}, so we monitor where the π¯i\overline{\pi}_{i} action results in a negative loop.

Theorem 4.5.

Under the quasisymmetric Frobenius characteristic map, we have

Fc​h​a​rQ​(SΞ»^)=sΞ»=Fc​h​a​r​(SΞ»)F_{char}^{Q}(\widehat{S^{\lambda}})=s_{\lambda}=F_{char}(S^{\lambda})
Proof.

First, order the fΞ»f^{\lambda} standard polytabloids that form the basis of SΞ»^\widehat{S^{\lambda}} according to the total ordering β‰Ί\prec, from greatest to least. Call eTje_{T_{j}} the jthj^{\text{th}} polytabloid in the basis ordering. To construct the composition series, SΞ»^=M1βŠƒM2βŠƒβ‹―β€‹MfΞ»=βˆ…\widehat{S^{\lambda}}=M_{1}\supset M_{2}\supset\cdots M_{f^{\lambda}}=\emptyset, let Mj=span​(eTj,eTj+1,…,eTfΞ»)M_{j}=\text{span}{(e_{T_{j}},e_{T_{j+1}},\dots,e_{T_{f^{\lambda}}})}.

We note a couple of facts about this composition series. First, we observe that \faktor​Mj​Mj+1\faktor{M_{j}}{M_{j+1}} is one-dimensional, generated by [eTj]j[e_{T_{j}}]_{j}. Furthermore, this composition series is, in fact, a composition series of submodules, and we do have MjβŠƒMj+1M_{j}\supset M_{j+1}, as our action is based on β‰Ί\prec.

Now, \faktor​Mj​Mj+1β‰…CI\faktor{M_{j}}{M_{j+1}}\cong C_{I}, where I={iβˆ£Ο€Β―i​eTj=βˆ’eTj}I=\{i\mid\overline{\pi}_{i}e_{T_{j}}=-e_{T_{j}}\}. So, by the definition of the quasisymmetric Frobenius characteristic map, we have

Fc​h​a​rQ​(SΞ»^)=βˆ‘T∈SYT​(Ξ»)F{iβˆ£Ο€Β―i​eT=βˆ’eT}.F_{char}^{Q}(\widehat{S^{\lambda}})=\sum\limits_{T\in\text{SYT}(\lambda)}F_{\{i\mid\overline{\pi}_{i}e_{T}=-e_{T}\}}.

Using the definition of the action, we have

Fc​h​a​rQ​(SΞ»^)\displaystyle F_{char}^{Q}(\widehat{S^{\lambda}}) =βˆ‘T∈SYT​(Ξ»)F{iβˆ£Ο€Β―i​eT=βˆ’eT}\displaystyle=\sum\limits_{T\in\text{SYT}(\lambda)}F_{\{i\mid\overline{\pi}_{i}e_{T}=-e_{T}\}}
=βˆ‘T∈SYT​(Ξ»)F{i∣i​ southeast of ​i+1​ in ​T}\displaystyle=\sum\limits_{T\in\text{SYT}(\lambda)}F_{\{i\mid i\text{ southeast of }i+1\text{ in }T\}}
=βˆ‘T∈SYT​(Ξ»)Fides​(T)\displaystyle=\sum\limits_{T\in\text{SYT}(\lambda)}F_{\text{ides}(T)}
=sΞ»\displaystyle=s_{\lambda}

∎

5. Implications of the previous sections

The proofs of Theorems 3.8 and 4.5 follow the same general approach. In particular, we created an action and a composition series according to the following steps:

  1. (1)

    Define an ordering ≀\leq on a basis ℬ\mathcal{B} of the 𝔖n\mathfrak{S}_{n}-module MM.

  2. (2)

    Using ℬ\mathcal{B} and ≀\leq, create a maximal decreasing sequence of nested subspaces:

    M=M1βŠƒM2βŠƒβ‹―βŠƒMk=βˆ….M=M_{1}\supset M_{2}\supset\cdots\supset M_{k}=\emptyset.
  3. (3)

    Create a compatible β„‹n​(0)\mathcal{H}_{n}(0)-action based on ≀\leq that causes the subspaces MiM_{i} to be closed under that action, so that the sequence

    M=M1βŠƒM2βŠƒβ‹―βŠƒMk=βˆ…M=M_{1}\supset M_{2}\supset\cdots\supset M_{k}=\emptyset

    forms a composition series of MM.

This approach suggests the following definition:

Definition 5.1.

Let MM be a 𝔖n\mathfrak{S}_{n}-module with basis ℬ={v1,…,vk}\mathcal{B}=\{v_{1},\dots,v_{k}\} and a total ordering ≀\leq on ℬ\mathcal{B} such that v1≀v2≀⋯≀vkv_{1}\leq v_{2}\leq\cdots\leq v_{k}. On the elements of ℬ\mathcal{B}, let

π¯i​(vt)={0tt​(si​(vt))=vtβˆ’vtsi​(vt)=vj>vt​ or ​si​(vt)=βˆ’vtsi​(vt)si​(vt)=vj<vt.\overline{\pi}_{i}(v_{t})=\begin{cases}0&\text{tt}(s_{i}(v_{t}))=v_{t}\\ -v_{t}&s_{i}(v_{t})=v_{j}>v_{t}\text{ or }s_{i}(v_{t})=-v_{t}\\ s_{i}(v_{t})&s_{i}(v_{t})=v_{j}<v_{t}\end{cases}.

In particular, assume only these three cases appear on the basis ℬ\mathcal{B}. When π¯i\overline{\pi}_{i} is a valid β„‹n​(0)\mathcal{H}_{n}(0)-action on the vector space of MM, call the resulting module M^\widehat{M}. Then we must have a resulting composition series

span​(v1,β‹―,vk)βŠƒspan​(v2,β‹―,vk)βŠƒβ‹―βŠƒspan​(vk)βŠƒβˆ…\text{span}(v_{1},\cdots,v_{k})\supset\text{span}(v_{2},\cdots,v_{k})\supset\cdots\supset\text{span}(v_{k})\supset\emptyset

as in particular each of the vector spaces are closed under the action.

We say (M,ℬ,≀)(M,\mathcal{B},\leq) is strongly quasisymmetric characteristic compatible (SQCC) if

Fc​h​a​rQ​(M^)=Fc​h​a​r​(M).F_{char}^{Q}(\widehat{M})=F_{char}(M).
Theorem 5.2.

(MΞ»,{{t}∣t​ is a row-strict tableau of shape ​λ},β‰Ί)(M^{\lambda},\{\{t\}\mid t\text{ is a row-strict tableau of shape }\lambda\},\prec) is strongly quasisymmetric characteristic compatible.

Proof.

This follows from Theorem 3.8 and Definition 5.1. ∎

Theorem 5.3.

(SΞ»,{eT∣T∈SYT​(Ξ»)},β‰Ί)(S^{\lambda},\{e_{T}\mid T\in\text{SYT}(\lambda)\},\prec) is strongly quasisymmetric characteristic compatible.

Proof.

This follows from Theorem 4.5 and Definition 5.1. ∎

Proposition 5.4.

For every 𝔖n\mathfrak{S}_{n}-module MM, there exists a basis ℬM\mathcal{B}_{M} and a total ordering ≀M\leq_{M} such that (M,ℬM,≀M)(M,\mathcal{B}_{M},\leq_{M}) is strongly quasisymmetric characteristic compatible.

Proof.

If MM is an 𝔖n\mathfrak{S}_{n}-module, then we can write Mβ‰…βŠ•cλ​SΞ»M\cong\oplus c_{\lambda}S^{\lambda}. Consider the ordering β‰Ί\prec on the basis of SΞ»S^{\lambda} defined in Definition 2.14. Using the isomorphism, create a compatible basis ℬM\mathcal{B}_{M} for MM from the bases {eT|T∈SYT​(Ξ»)}\{e_{T}|T\in\text{SYT}(\lambda)\} within an isomorphic copy of SΞ»S^{\lambda} such that within each submodule, the relative order of the basis elements in Definition 2.14 are preserved. Call this new order on the basis elements of ≀M\leq_{M}. When M^\widehat{M} is a β„‹n​(0)\mathcal{H}_{n}(0)-module, we also have M^β‰…βŠ•cλ​SΞ»^\widehat{M}\cong\oplus c_{\lambda}\widehat{S^{\lambda}} as vector spaces. Then, we have the following:

Fc​h​a​rQ​(M^)\displaystyle F_{char}^{Q}(\widehat{M}) =Fc​h​a​rQ​(βŠ•cλ​SΞ»^)\displaystyle=F_{char}^{Q}\left(\oplus c_{\lambda}\widehat{S^{\lambda}}\right)
=βˆ‘cλ​Fc​h​a​rQ​(SΞ»^)\displaystyle=\sum c_{\lambda}F_{char}^{Q}\left(\widehat{S^{\lambda}}\right)
=βˆ‘cλ​Fc​h​a​r​(SΞ»)\displaystyle=\sum c_{\lambda}F_{char}\left(S^{\lambda}\right)
=βˆ‘cλ​sΞ»\displaystyle=\sum c_{\lambda}s_{\lambda}
=Fc​h​a​r​(βŠ•cλ​SΞ»)\displaystyle=F_{char}\left(\oplus c_{\lambda}S^{\lambda}\right)
=Fc​h​a​r​(M)\displaystyle=F_{char}(M)

∎

Corollary 5.5.

Given an 𝔖n\mathfrak{S}_{n}-module MM, let ℬM\mathcal{B}_{M} and total ordering ≀M\leq_{M} be as constructed in Proposition 5.4. Let ff be a map defined on the representations of 𝔖n\mathfrak{S}_{n} such that f​(M)=M^f(M)=\widehat{M}, and ΞΉ:Symβ†’QSym\iota:\text{Sym}\rightarrow\text{QSym} denote the inclusion map. Then, the following diagram commutes:

𝒒0​(𝔖n){\mathcal{G}_{0}(\mathfrak{S}_{n})}Sym𝒒0​(β„‹n​(0)){\mathcal{G}_{0}(\mathcal{H}_{n}(0))}QSymFc​h​a​r\scriptstyle{F_{char}}f\scriptstyle{f}ΞΉ\scriptstyle{\iota}Fc​h​a​rQ\scriptstyle{F_{char}^{Q}}

This proposition and subsequent corollary imply that for every action on an 𝔖n\mathfrak{S}_{n}-module there exists a naturual deformed to a β„‹n​(0)\mathcal{H}_{n}(0)-action with the same Frobenius image. However, the specific deformation mentioned above may not always be easy to compute, due to the complexities of defining a 1-1 map between SΞ»S^{\lambda} and MM.

6. The Coinvariant Algebra as a β„‹n​(0)\mathcal{H}_{n}(0)-module

In this section, we are inspired by a broader motivation:

Motivation.

Study the symmetric Frobenius image of MM by computing the quasisymmetric Frobenius image of M^\widehat{M}.

Corollary 5.5, alas, does not address this motivation directly, since ℬM\mathcal{B}_{M} and ≀M\leq_{M} may not be natural on a generic module MM. In fact, as the next example suggests, the action presented in Definition 5.1 may be too limiting if this is our goal, but a weaker condition may be more realistic.

For our final example, we transition to defining a compatible β„‹n​(0)\mathcal{H}_{n}(0)-action on a well-studied quotient space: the coinvariant algebra β„›n\mathcal{R}_{n}. Before giving the formal definition, we briefly give some necessary background on GrΓΆbner bases in Section 6.1 before returning to the example at hand in section 6.2.

Let ℂ​[x1,…,xn]\mathbb{C}[x_{1},\dots,x_{n}] be the space of all functions in nn variables. Given a weak composition α⊨wd\alpha\vDash_{w}d, xΞ±βˆˆβ„‚β€‹[x1,…,xn]x^{\alpha}\in\mathbb{C}[x_{1},\dots,x_{n}] is the monomial x1Ξ±1​x2Ξ±2​⋯​xnΞ±nx_{1}^{\alpha_{1}}x_{2}^{\alpha_{2}}\cdots x_{n}^{\alpha_{n}}, and we say that xΞ±x^{\alpha} has degree dd.

6.1. GrΓΆbner Bases

This subsection serves as a brief overview of GrΓΆbner bases and related results. We adopt the notation of Dummit and FooteΒ [8], which gives a good introduction to the 1423ic for the unfamiliar reader. Following their notation, when the underlying ideal II is obvious, for any pβˆˆβ„‚β€‹[x1,…,xn]p\in\mathbb{C}[x_{1},\dots,x_{n}], we use pΒ―\overline{p} for the coset in \faktor​ℂ​[x1,…,xn]​I\faktor{\mathbb{C}[x_{1},\dots,x_{n}]}{I} corresponding to pp.

One difficulty with computations in quotient spaces is that elements of quotient spaces have multiple coset representatives and thus it may be difficult to determine if pΒ―\overline{p} equals qΒ―\overline{q} for two distinct elements p,qβˆˆβ„‚β€‹[x1,…,xn]p,q\in\mathbb{C}[x_{1},\dots,x_{n}]. In the case that II is a nonzero ideal of ℂ​[x1,…,xn]\mathbb{C}[x_{1},\dots,x_{n}], GrΓΆbner bases are employed as a tool to overcome this obstacle when doing computations in Q=\faktor​ℂ​[x1,…,xn]​IQ=\faktor{\mathbb{C}[x_{1},\dots,x_{n}]}{I}. In such a setting, we must first select a monomial ordering on ℂ​[x1,…,xn]\mathbb{C}[x_{1},\dots,x_{n}]:

Definition 6.1.

A monomial ordering on ℂ​[x1,…,xn]\mathbb{C}[x_{1},\dots,x_{n}] is a total ordering such that, if xα≀xΞ²x^{\alpha}\leq x^{\beta} and xΞ³x^{\gamma} is any other monomial, then xα​xγ≀xβ​xΞ³x^{\alpha}x^{\gamma}\leq x^{\beta}x^{\gamma}.

The lexicographic monomial ordering, explicitly defined below, is a widely used monomial ordering on ℂ​[x1,…,xn]\mathbb{C}[x_{1},\dots,x_{n}].

Definition 6.2.

Given two monomials xΞ±x^{\alpha} and xΞ²x^{\beta}, we say xΞ±<xΞ²x^{\alpha}<x^{\beta} in lexicographic monomial ordering if the first nonzero entry in the sequence Ξ²βˆ’Ξ±\beta-\alpha is strictly positive.

With respect to the monomial ordering, we then have a β€œfirst” monomial that occurs with nonzero coefficient in every polynomial, leading to two related definitions:

Definition 6.3.

Given a polynomial pβˆˆβ„‚β€‹[x1,…,xn]p\in\mathbb{C}[x_{1},\dots,x_{n}], let

p=c1​xΞ±1+c2​xΞ±2+β‹―+cm​xΞ±mp=c_{1}x^{\alpha_{1}}+c_{2}x^{\alpha_{2}}+\cdots+c_{m}x^{\alpha_{m}}

with xΞ±1>xΞ±2>β‹―>xΞ±mx^{\alpha_{1}}>x^{\alpha_{2}}>\cdots>x^{\alpha_{m}} and ciβ‰ 0c_{i}\neq 0 for all ii. Then, the leading term of pp, denoted lt​(p)\text{lt}(p), is c1​xΞ±1c_{1}x^{\alpha_{1}} and the leading monomial of pp, denoted lm​(p)\text{lm}(p), is xΞ±1x^{\alpha_{1}}.

Note that as defined monomial lt​(p)\text{lt}(p) may have any nonzero coefficient, while the monomial lm​(p)\text{lm}(p) will have a coefficient of 1.

GrΓΆbner bases, misleadingly named as they are not actually bases of a vector space, are defined using ideals generated by leading terms of polynomials.

Definition 6.4.

Given a nonzero ideal II in ℂ​[x1,…,xn]\mathbb{C}[x_{1},\dots,x_{n}], let lt​(I)\text{lt}(I) be the ideal generated by the leading terms of polynomials in II:

lt​(I)=⟨lt​(p)∣p∈I⟩\text{lt}(I)=\langle\text{lt}(p)\mid p\in I\rangle
Definition 6.5.

A GrΓΆbner basis G={g1,…,gm}G=\{g_{1},\dots,g_{m}\} of a nonzero ideal II in the ring ℂ​[x1,…,xn]\mathbb{C}[x_{1},\dots,x_{n}] is a finite generating set of II whose leading terms generate lt​(I)\text{lt}(I). Thus,

I=⟨g1,…,gmβŸ©β€‹Β andΒ lt​(I)=⟨lt​(g1),…,lt​(gm)⟩.I=\langle g_{1},\dots,g_{m}\rangle\text{ and }\text{lt}(I)=\langle\text{lt}(g_{1}),\dots,\text{lt}(g_{m})\rangle.
Definition 6.6.

A reduced GrΓΆbner basis G={g1,…,gm}G=\{g_{1},\dots,g_{m}\} of nonzero ideal II in the ring ℂ​[x1,…,xn]\mathbb{C}[x_{1},\dots,x_{n}] is a GrΓΆbner basis where

  • β€’

    lt​(gi)\text{lt}(g_{i}) has coefficient 1 for each ii

  • β€’

    no monomial appearing in gjg_{j} is divisible by lt​(gi)\text{lt}(g_{i}) when iβ‰ ji\neq j.

Lemma 6.7 (Theorem 27 in [8]).

Given a monomial ordering on ℂ​[x1,…,xn]\mathbb{C}[x_{1},\dots,x_{n}], there is a unique reduced GrΓΆbner basis for every nonzero ideal II in ℂ​[x1,…,xn]\mathbb{C}[x_{1},\dots,x_{n}].

By using a reduced GrΓΆbner basis G={g1,…,gm}G=\{g_{1},\dots,g_{m}\}, we are able to define a division algorithm on ℂ​[x1,…,xn]\mathbb{C}[x_{1},\dots,x_{n}] that results in a unique remainder. Reminiscent of the Euclidean algorithm, if pβˆˆβ„‚β€‹[x1,…,xn]p\in\mathbb{C}[x_{1},\dots,x_{n}], then:

  • β€’

    if lt​(p)\text{lt}(p) is divisible by lt​(gi)\text{lt}(g_{i}) such that lt​(p)=ai​lt​(gi)\text{lt}(p)=a_{i}\text{lt}(g_{i}), add aia_{i} to the β€œquotient” qiq_{i}, and replace pp by pβˆ’ai​gip-a_{i}g_{i}. Repeat this process.

  • β€’

    if lt​(p)\text{lt}(p) is not divisible by lt​(gi)\text{lt}(g_{i}) for any ii, add the leading term of pp to the β€œremainder” rr, and replace pp by pβˆ’lt​(p)p-\text{lt}(p). Repeat this process.

This algorithm does terminate, resulting in

p=q1​g1+q2​g2+β‹―+qm​gm+r.p=q_{1}g_{1}+q_{2}g_{2}+\cdots+q_{m}g_{m}+r.

Moreover, when working only with the reduced GrΓΆbner basis GG, the remainder of any element in ℂ​[x1,…,xn]\mathbb{C}[x_{1},\dots,x_{n}] depends only on II and the choice of ≀\leq. Otherwise, the remainder also depends on the choice of GrΓΆbner basis GG.

Lemma 6.8 (Theorem 23 in [8]).

Let G={g1,…,gm}G=\{g_{1},\dots,g_{m}\} be a GrΓΆbner basis for a nonzero ideal II in ℂ​[x1,…,xn]\mathbb{C}[x_{1},\dots,x_{n}]. Then, every polynomial pβˆˆβ„‚β€‹[x1,…,xn]p\in\mathbb{C}[x_{1},\dots,x_{n}] can be written uniquely as

p=fI+r,p=f_{I}+r,

where fI∈If_{I}\in I and no nonzero monomial term in rr is divisible by the leading terms of the polynomials in GG. Moreover, the remainder rr is a unique representative in the coset of pp in \faktor​ℂ​[x1,…,xn]​I\faktor{\mathbb{C}[x_{1},\dots,x_{n}]}{I}.

Definition 6.9.

Let GG be a GrΓΆbner basis of nonzero ideal II and pβˆˆβ„‚β€‹[x1,…,xn]p\in\mathbb{C}[x_{1},\dots,x_{n}], and write p=fI+rp=f_{I}+r as given in Lemma 6.8. If fIf_{I} is a nonzero polynomial, then we say that pp is reducible by GG. If fI=0f_{I}=0, then pp is reduced with respect to GG.

Corollary 6.10.

Let II be a nonzero ideal in ℂ​[x1,…,xn]\mathbb{C}[x_{1},\dots,x_{n}] with GrΓΆbner basis GG. Then, every coset p¯∈\faktor​ℂ​[x1,…,xn]​I\overline{p}\in\faktor{\mathbb{C}[x_{1},\dots,x_{n}]}{I} has a unique coset representative p~βˆˆβ„‚β€‹[x1,…,xn]\widetilde{p}\in\mathbb{C}[x_{1},\dots,x_{n}] that is reduced with respect to GG.

Proof.

By Lemma 6.8, to find the unique representative, we need only apply the division algorithm to any coset representative. ∎

Going forward, when the underlying ideal II and ordering ≀\leq is obvious, we consider the unique reduced GrΓΆbner basis GG. Then, we use p~βˆˆβ„‚β€‹[x1,…,xn]\widetilde{p}\in\mathbb{C}[x_{1},\dots,x_{n}] to denote this unique coset representative of pΒ―\overline{p} that is reduced with respect to GG.

6.2. The Coinvariant Algebra β„›n\mathcal{R}_{n}

Next, we turn our attention to a well-studied action of 𝔖n\mathfrak{S}_{n} and a compatible quotient space. Define an 𝔖n\mathfrak{S}_{n}-action on ℂ​[x1,…,xn]\mathbb{C}[x_{1},\dots,x_{n}] by

si​(xΞ±)=xsi​(Ξ±),s_{i}(x^{\alpha})=x^{s_{i}(\alpha)},

where si​(Ξ±)=si​((Ξ±1,…,Ξ±i,Ξ±i+1,…,Ξ±l))=(Ξ±1,…,Ξ±i+1,Ξ±i,…,Ξ±l)s_{i}(\alpha)=s_{i}((\alpha_{1},\dots,\alpha_{i},\alpha_{i+1},\dots,\alpha_{l}))=(\alpha_{1},\dots,\alpha_{i+1},\alpha_{i},\dots,\alpha_{l}).

Definition 6.11.

Let SS be the ideal ⟨e1​(x1,…,xn),…,en​(x1,…,xn)⟩\langle e_{1}(x_{1},\dots,x_{n}),\dots,e_{n}(x_{1},\dots,x_{n})\rangle. Then the coinvariant algebra β„›n\mathcal{R}_{n} is the quotient space \faktor​ℂ​[x1​…,xn]​S\faktor{\mathbb{C}[x_{1}\dots,x_{n}]}{S}.

Now, the ideal SS is invariant under the aforementioned 𝔖n\mathfrak{S}_{n}-action and thus one can consider the action on the quotient β„›n\mathcal{R}_{n}. In particular,

si​(xΞ±Β―)=xsi​(Ξ±)Β―s_{i}(\overline{x^{\alpha}})=\overline{x^{s_{i}(\alpha)}}

is thus a well-defined action on β„›n\mathcal{R}_{n} that does not depend on the choice of coset representative.

For additional details on β„›n\mathcal{R}_{n} and other well-known quotient spaces in this area, we recommend the unfamiliar reader see the excellent text by Bergeron [9].

As with the previous β„‹n​(0)\mathcal{H}_{n}(0)-modules studied here, an ordering will be used throughout this section. Here, we consider the lexicographic ordering on monomials defined in the previous subsection.

Lemma 6.12 (See [9]).

With respect to lexicographic ordering ≀\leq, the ideal SS has reduced GrΓΆbner basis

G={gi=hi​(xi,…,xn):1≀i≀n}.G=\{g_{i}=h_{i}(x_{i},\dots,x_{n}):1\leq i\leq n\}.

Artin [10] showed that the set of β€œsub-staircase” monomials form a basis of β„›n\mathcal{R}_{n}.

Definition 6.13.

Let An={Ξ±=(Ξ±1,Ξ±2,…,Ξ±n)∣0≀αj<j}A_{n}=\{\alpha=(\alpha_{1},\alpha_{2},\dots,\alpha_{n})\mid 0\leq\alpha_{j}<j\}. Then, define the set of sub-staircase monomials as

π’œn={xα¯∣α∈An}.\mathcal{A}_{n}=\{\overline{x^{\alpha}}\mid\alpha\in A_{n}\}.
Lemma 6.14.

π’œn\mathcal{A}_{n} is a basis for β„›n\mathcal{R}_{n}, and dim​(β„›n)=n!\text{dim}(\mathcal{R}_{n})=n!.

We will use lexicographic ordering to define a β„‹n​(0)\mathcal{H}_{n}(0)-action on β„›n\mathcal{R}_{n}. To do so, we will first need to analyze β„›n\mathcal{R}_{n} as an 𝔖n\mathfrak{S}_{n}-module.

Proposition 6.15.

Assume α∈An\alpha\in A_{n}. If si​(Ξ±)βˆ‰Ans_{i}(\alpha)\notin A_{n}, then it must be that lm​(xsi​(Ξ±)~)β‰₯xΞ±\text{lm}(\widetilde{x^{s_{i}(\alpha)}})\geq x^{\alpha}. If equality holds, then lt​(xsi​(Ξ±)~)=βˆ’xΞ±\text{lt}(\widetilde{x^{s_{i}(\alpha)}})=-x^{\alpha}.

Proof.

We write

xΞ±=x1Ξ±1​⋯​xiΞ±i​xi+1Ξ±i+1​⋯​xnΞ±nx^{\alpha}=x_{1}^{\alpha_{1}}\cdots x_{i}^{\alpha_{i}}x_{i+1}^{\alpha_{i+1}}\cdots x_{n}^{\alpha_{n}}

and

xsi​(Ξ±)=x1Ξ±1​⋯​xiΞ±i+1​xi+1Ξ±i​⋯​xnΞ±n.x^{s_{i}(\alpha)}=x_{1}^{\alpha_{1}}\cdots x_{i}^{\alpha_{i+1}}x_{i+1}^{\alpha_{i}}\cdots x_{n}^{\alpha_{n}}.

Since we assume that xsi​(Ξ±)x^{s_{i}(\alpha)} is reducible by GG and α∈An\alpha\in A_{n}, it must be that Ξ±i+1=i\alpha_{i+1}=i as a result of Lemma 6.14. Thus, we can rewrite

xsi​(Ξ±)=x1Ξ±1​⋯​xii​xi+1Ξ±i​⋯​xnΞ±n.x^{s_{i}(\alpha)}=x_{1}^{\alpha_{1}}\cdots x_{i}^{i}x_{i+1}^{\alpha_{i}}\cdots x_{n}^{\alpha_{n}}.

Now, in this way, it is easy to see that xsi​(Ξ±)x^{s_{i}(\alpha)} is reducible by gig_{i} specifically. In particular, the leading term of gig_{i} is xiix_{i}^{i} in lexicographic order. As such, we may reduce xsi​(Ξ±)x^{s_{i}(\alpha)} by gig_{i}, resulting in a new polynomial p​(x1,…,xn)p(x_{1},\dots,x_{n}) where

lt​(p​(x1,…,xn))=m​(x1,…,xn)=βˆ’x1Ξ±1​⋯​xiβˆ’1Ξ±iβˆ’1​xiiβˆ’1​xi+1Ξ±i+1​xi+2Ξ±i+2​⋯​xnΞ±n.\text{lt}(p(x_{1},\dots,x_{n}))=m(x_{1},\dots,x_{n})=-x_{1}^{\alpha_{1}}\cdots x_{i-1}^{\alpha_{i-1}}x_{i}^{i-1}x_{i+1}^{\alpha_{i}+1}x_{i+2}^{\alpha_{i+2}}\cdots x_{n}^{\alpha_{n}}.

We aim to show that, when fully reduced, m​(x1,…,xn)m(x_{1},\dots,x_{n}) remains the leading term of xsi​(Ξ±)~\widetilde{x^{s_{i}(\alpha)}} after successive reductions.

For further reduction, we note there may be monomials of the form x1Ξ±1​⋯​xkΞ±k+r​⋯​xnΞ±nx_{1}^{\alpha_{1}}\cdots x_{k}^{\alpha_{k}+r}\cdots x_{n}^{\alpha_{n}}, k>ik>i and r>0r>0, which must be reduced by gkg_{k}. Since gk=hk​(xk,…,xn)g_{k}=h_{k}(x_{k},\dots,x_{n}), we note that with the reduction, any new terms continue to have the same prefix x1Ξ±1​⋯​xiβˆ’1Ξ±iβˆ’1​xiiβˆ’1x_{1}^{\alpha_{1}}\cdots x_{i-1}^{\alpha_{i-1}}x_{i}^{i-1}. Thus,

lt​(xsi​(Ξ±)~)=m​(x1,…,xn).\text{lt}(\widetilde{x^{s_{i}(\alpha)}})=m(x_{1},\dots,x_{n}).

If Ξ±i=iβˆ’1\alpha_{i}=i-1, then Ξ±i+1=i=Ξ±i+1\alpha_{i}+1=i=\alpha_{i+1}. Thus, m​(x1,…,xn)=βˆ’xΞ±m(x_{1},\dots,x_{n})=-x^{\alpha}. Otherwise, if Ξ±i<iβˆ’1\alpha_{i}<i-1, we have lm​(xsi​(Ξ±)~)>xΞ±\text{lm}(\widetilde{x^{s_{i}(\alpha)}})>x^{\alpha}. ∎

It is a result of Chevalley [11] that, with this 𝔖n\mathfrak{S}_{n}-action, we have the left regular representation, so that

Fc​h​a​r​(β„›n)=βˆ‘Ξ»βŠ’nfλ​sΞ».F_{char}(\mathcal{R}_{n})=\sum\limits_{\lambda\vdash n}f^{\lambda}s_{\lambda}.

If we view β„›n\mathcal{R}_{n} as a graded 𝔖n\mathfrak{S}_{n}-module, we have from Lusztig, formalized by Stanley in [12], that

Fc​h​a​r​(β„›n;q)=βˆ‘Ξ»βŠ’nβˆ‘T∈SYT​(Ξ»)qmaj​(T)​sΞ».F_{char}(\mathcal{R}_{n};q)=\sum\limits_{\lambda\vdash n}\sum\limits_{T\in\text{SYT}(\lambda)}q^{\text{maj}(T)}s_{\lambda}.

We will notice that, at q=1q=1, we obtain Fc​h​a​r​(β„›n)F_{char}(\mathcal{R}_{n}) as an ungraded 𝔖n\mathfrak{S}_{n}-module. As such, we will work with the graded quasisymmetric Frobenius characteristic of β„›n\mathcal{R}_{n}, where, for monomial rings, we have

Fc​h​a​rQ​(⟨xα⟩;q)=qdeg​(xΞ±)​F{iβˆ£Ο€Β―i​(xΞ±)=βˆ’xΞ±}.F_{char}^{Q}(\langle x^{\alpha}\rangle;q)=q^{\text{deg}(x^{\alpha})}F_{\{i\mid\overline{\pi}_{i}(x^{\alpha})=-x^{\alpha}\}}.

Next, in the style of how we defined the β„‹n​(0)\mathcal{H}_{n}(0)-action on the Specht modules, let β„›n^\widehat{\mathcal{R}_{n}} be the vector space β„›n\mathcal{R}_{n} with the goal of defining a valid β„‹n​(0)\mathcal{H}_{n}(0)-action. In particular, we define an β„‹n​(0)\mathcal{H}_{n}(0)-action on β„›n^\widehat{\mathcal{R}_{n}}. In particular, we assume α∈An\alpha\in A_{n} so that the action is defined on a basis for β„›n\mathcal{R}_{n} and can be extended linearly:

π¯i​(xΞ±Β―)={0if ​xsi​α=xΞ±βˆ’xΞ±Β―ifΒ lm​(xsi​α~)>xα​ orΒ lt​(xsi​α~)=βˆ’xΞ±xsi​(Ξ±)Β―if ​xsi​α<xΞ±\overline{\pi}_{i}(\overline{x^{\alpha}})=\begin{cases}0&\text{if }x^{s_{i}\alpha}=x^{\alpha}\\ -\overline{x^{\alpha}}&\text{if }\text{lm}(\widetilde{x^{s_{i}\alpha}})>x^{\alpha}\text{ or }\text{lt}(\widetilde{x^{s_{i}\alpha}})=-x^{\alpha}\\ \overline{x^{s_{i}(\alpha)}}&\text{if }x^{s_{i}\alpha}<x^{\alpha}\end{cases}

By Proposition 6.15, these are the only cases to consider on β„›n^\widehat{\mathcal{R}_{n}}. By this same proposition, we know that if xΞ±>lm​(xsi​(Ξ±)~)x^{\alpha}>\text{lm}(\widetilde{x^{s_{i}(\alpha)}}), then xsi​(Ξ±)x^{s_{i}(\alpha)} is not reducible by GG, which is why the leading term notation is omitted in the third case above. If xsi​(Ξ±)<xΞ±x^{s_{i}(\alpha)}<x^{\alpha}, then we must have Ξ±i>Ξ±i+1\alpha_{i}>\alpha_{i+1}. Similarly, if xsi​(Ξ±)=xΞ±x^{s_{i}(\alpha)}=x^{\alpha}, then Ξ±i=Ξ±i+1\alpha_{i}=\alpha_{i+1}. This tells us that we can rewrite our β„‹n​(0)\mathcal{H}_{n}(0)-action on β„›n^\widehat{\mathcal{R}_{n}} as follows, again assuming that α∈An\alpha\in A_{n}:

π¯i​(xΞ±Β―)={0if ​αi=Ξ±i+1βˆ’xΞ±Β―if ​αi<Ξ±i+1xsi​(Ξ±)Β―if ​αi>Ξ±i+1\overline{\pi}_{i}(\overline{x^{\alpha}})=\begin{cases}0&\text{if }\alpha_{i}=\alpha_{i+1}\\ -\overline{x^{\alpha}}&\text{if }\alpha_{i}<\alpha_{i+1}\\ \overline{x^{s_{i}(\alpha)}}&\text{if }\alpha_{i}>\alpha_{i+1}\end{cases}

Now, we’ll show that β„›n^\widehat{\mathcal{R}_{n}} is a valid β„‹n​(0)\mathcal{H}_{n}(0)-module using this easier action. The proof is by contradiction. To reduce the number of cases, we will assume that xΞ±Β―\overline{x^{\alpha}} is a counterexample to the length three commuting relations at positions j,j+1,j+2j,j+1,j+2, and show that in fact this counterexample must occur at positions 3, 4, 5 (or earlier).

Lemma 6.16.

Let j,xΞ±Β―j,\overline{x^{\alpha}} be a counterexample to the length three commuting relations. In particular, suppose

π¯j​π¯j+1​π¯j​(xΞ±Β―)≠π¯j+1​π¯j​π¯j+1​(xΞ±Β―).\overline{\pi}_{j}\overline{\pi}_{j+1}\overline{\pi}_{j}(\overline{x^{\alpha}})\neq\overline{\pi}_{j+1}\overline{\pi}_{j}\overline{\pi}_{j+1}(\overline{x^{\alpha}}).

Then, there exists a non-distinct list β3,β4,β5∈{0,1,2}\beta_{3},\beta_{4},\beta_{5}\in\{0,1,2\} such that

π¯3​π¯4​π¯3​(x3Ξ²3​x4Ξ²4​x5Ξ²5Β―)≠π¯4​π¯3​π¯4​(x3Ξ²3​x4Ξ²4​x5Ξ²5Β―).\overline{\pi}_{3}\overline{\pi}_{4}\overline{\pi}_{3}(\overline{x_{3}^{\beta_{3}}x_{4}^{\beta_{4}}x_{5}^{\beta_{5}}})\neq\overline{\pi}_{4}\overline{\pi}_{3}\overline{\pi}_{4}(\overline{x_{3}^{\beta_{3}}x_{4}^{\beta_{4}}x_{5}^{\beta_{5}}}).
Proof.

Note that the action of π¯i\overline{\pi}_{i} on a monomial is dependent only on the relative order of the exponents themselves, and not the values of the exponents. Thus, π¯j​π¯j+1​π¯j\overline{\pi}_{j}\overline{\pi}_{j+1}\overline{\pi}_{j} and π¯j+1​π¯j​π¯j+1\overline{\pi}_{j+1}\overline{\pi}_{j}\overline{\pi}_{j+1} act on xΞ±Β―\overline{x^{\alpha}}, giving either 0 or a permutation of the exponents in positions j,j+1j,j+1, and j+2j+2, which depends only on the relative sizes of Ξ±j,Ξ±j+1\alpha_{j},\alpha_{j+1}, and Ξ±j+2\alpha_{j+2}. Thus, if

π¯j​π¯j+1​π¯j​(xΞ±Β―)≠π¯j+1​π¯j​π¯j+1​(xΞ±Β―),\overline{\pi}_{j}\overline{\pi}_{j+1}\overline{\pi}_{j}(\overline{x^{\alpha}})\neq\overline{\pi}_{j+1}\overline{\pi}_{j}\overline{\pi}_{j+1}(\overline{x^{\alpha}}),

then it must be that if Ξ²3,Ξ²4\beta_{3},\beta_{4} and Ξ²5\beta_{5} are chosen to have the same relative order amongst themselves that Ξ±j\alpha_{j}, Ξ±j+1\alpha_{j+1}, and Ξ±j+2\alpha_{j+2} have, acting by π¯3​π¯4​π¯3\overline{\pi}_{3}\overline{\pi}_{4}\overline{\pi}_{3} and π¯4​π¯3​π¯4\overline{\pi}_{4}\overline{\pi}_{3}\overline{\pi}_{4} will result in either 0 or an analogous permutation of Ξ²3,Ξ²4\beta_{3},\beta_{4}, and Ξ²5\beta_{5} on the exponents of xΞ²Β―\overline{x^{\beta}}, so

π¯3​π¯4​π¯3​(x3Ξ²3​x4Ξ²4​x5Ξ²5Β―)≠π¯4​π¯3​π¯4​(x3Ξ²3​x4Ξ²4​x5Ξ²5Β―).\overline{\pi}_{3}\overline{\pi}_{4}\overline{\pi}_{3}(\overline{x_{3}^{\beta_{3}}x_{4}^{\beta_{4}}x_{5}^{\beta_{5}}})\neq\overline{\pi}_{4}\overline{\pi}_{3}\overline{\pi}_{4}(\overline{x_{3}^{\beta_{3}}x_{4}^{\beta_{4}}x_{5}^{\beta_{5}}}).

Thus, we need only ensure that all relative orders of a triple (αj,αj+1,αj+2)(\alpha_{j},\alpha_{j+1},\alpha_{j+2}) may be achieved in the 3rd, 4th, and 5th exponents. Since the possible exponents of x3,x4x_{3},x_{4}, and x5x_{5} include 0, 1, and 2, any relative order of three exponents can be achieved. ∎

Example 6.17.

For n=9n=9, let xΞ±=x75​x82​x96x^{\alpha}=x_{7}^{5}x_{8}^{2}x_{9}^{6}. Then, we choose xΞ²=x31​x40​x52x^{\beta}=x_{3}^{1}x_{4}^{0}x_{5}^{2}, since Ξ±8<Ξ±7<Ξ±9\alpha_{8}<\alpha_{7}<\alpha_{9} and Ξ²4<Ξ²3<Ξ²5\beta_{4}<\beta_{3}<\beta_{5}, where the indices are now shifted downwards by 4. Note that π¯7​π¯8​π¯7​(xΞ±Β―)=x72​x85​x96Β―=s7​(xΞ±Β―)\overline{\pi}_{7}\overline{\pi}_{8}\overline{\pi}_{7}(\overline{x^{\alpha}})=\overline{x_{7}^{2}x_{8}^{5}x_{9}^{6}}=s_{7}(\overline{x^{\alpha}}) and π¯3​π¯4​π¯3​(xΞ²Β―)=x30​x41​x52Β―=s(7βˆ’4)​(xΞ²Β―)\overline{\pi}_{3}\overline{\pi}_{4}\overline{\pi}_{3}(\overline{x^{\beta}})=\overline{x_{3}^{0}x_{4}^{1}x_{5}^{2}}=s_{(7-4)}(\overline{x^{\beta}}). Thus, if xΞ±Β―\overline{x^{\alpha}} were a counterexample to the commuting relations, then xΞ²Β―\overline{x^{\beta}} would be as well.

Lemma 6.18.

With the π¯i\overline{\pi}_{i} action defined on β„›n^\widehat{\mathcal{R}_{n}}, let α∈An\alpha\in A_{n}. Then, we have

π¯3​π¯4​π¯3​(x1Ξ±1​⋯​x3Ξ±3​x4Ξ±4​x5Ξ±5​⋯​xnΞ±nΒ―)=π¯4​π¯3​π¯4​(x1Ξ±1​⋯​x3Ξ±3​x4Ξ±4​x5Ξ±5​⋯​xnΞ±nΒ―).\overline{\pi}_{3}\overline{\pi}_{4}\overline{\pi}_{3}(\overline{x_{1}^{\alpha_{1}}\cdots x_{3}^{\alpha_{3}}x_{4}^{\alpha_{4}}x_{5}^{\alpha_{5}}\cdots x_{n}^{\alpha_{n}}})=\overline{\pi}_{4}\overline{\pi}_{3}\overline{\pi}_{4}(\overline{x_{1}^{\alpha_{1}}\cdots x_{3}^{\alpha_{3}}x_{4}^{\alpha_{4}}x_{5}^{\alpha_{5}}\cdots x_{n}^{\alpha_{n}}}).
Proof.

The numerous cases, where we consider all relative orders of Ξ±3,Ξ±4\alpha_{3},\alpha_{4}, and Ξ±5\alpha_{5}, are similar in difficulty, so we give two specific cases here.

Case 1.

Suppose Ξ±3<Ξ±5<Ξ±4\alpha_{3}<\alpha_{5}<\alpha_{4}.

π¯3​π¯4​π¯3​(x1Ξ±1​⋯​x3Ξ±3​x4Ξ±4​x5Ξ±5​⋯​xnΞ±nΒ―)\displaystyle\overline{\pi}_{3}\overline{\pi}_{4}\overline{\pi}_{3}(\overline{x_{1}^{\alpha_{1}}\cdots x_{3}^{\alpha_{3}}x_{4}^{\alpha_{4}}x_{5}^{\alpha_{5}}\cdots x_{n}^{\alpha_{n}}}) =βˆ’Ο€Β―3​π¯4​(x1Ξ±1​⋯​x3Ξ±3​x4Ξ±4​x5Ξ±5​⋯​xnΞ±nΒ―)\displaystyle=-\overline{\pi}_{3}\overline{\pi}_{4}(\overline{x_{1}^{\alpha_{1}}\cdots x_{3}^{\alpha_{3}}x_{4}^{\alpha_{4}}x_{5}^{\alpha_{5}}\cdots x_{n}^{\alpha_{n}}})
=βˆ’Ο€Β―3​(x1Ξ±1​⋯​x3Ξ±3​x4Ξ±5​x5Ξ±4​⋯​xnΞ±nΒ―)\displaystyle=-\overline{\pi}_{3}(\overline{x_{1}^{\alpha_{1}}\cdots x_{3}^{\alpha_{3}}x_{4}^{\alpha_{5}}x_{5}^{\alpha_{4}}\cdots x_{n}^{\alpha_{n}}})
=x1Ξ±1​⋯​x3Ξ±3​x4Ξ±5​x5Ξ±4​⋯​xnΞ±nΒ―\displaystyle=\overline{x_{1}^{\alpha_{1}}\cdots x_{3}^{\alpha_{3}}x_{4}^{\alpha_{5}}x_{5}^{\alpha_{4}}\cdots x_{n}^{\alpha_{n}}}
π¯4​π¯3​π¯4​(x1Ξ±1​⋯​x3Ξ±3​x4Ξ±4​x5Ξ±5​⋯​xnΞ±nΒ―)\displaystyle\overline{\pi}_{4}\overline{\pi}_{3}\overline{\pi}_{4}(\overline{x_{1}^{\alpha_{1}}\cdots x_{3}^{\alpha_{3}}x_{4}^{\alpha_{4}}x_{5}^{\alpha_{5}}\cdots x_{n}^{\alpha_{n}}}) =π¯4​π¯3​(x1Ξ±1​⋯​x3Ξ±3​x4Ξ±5​x5Ξ±4​⋯​xnΞ±nΒ―)\displaystyle=\overline{\pi}_{4}\overline{\pi}_{3}(\overline{x_{1}^{\alpha_{1}}\cdots x_{3}^{\alpha_{3}}x_{4}^{\alpha_{5}}x_{5}^{\alpha_{4}}\cdots x_{n}^{\alpha_{n}}})
=βˆ’Ο€Β―4​(x1Ξ±1​⋯​x3Ξ±3​x4Ξ±5​x5Ξ±4​⋯​xnΞ±nΒ―)\displaystyle=-\overline{\pi}_{4}(\overline{x_{1}^{\alpha_{1}}\cdots x_{3}^{\alpha_{3}}x_{4}^{\alpha_{5}}x_{5}^{\alpha_{4}}\cdots x_{n}^{\alpha_{n}}})
=x1Ξ±1​⋯​x3Ξ±3​x4Ξ±5​x5Ξ±4​⋯​xnΞ±nΒ―\displaystyle=\overline{x_{1}^{\alpha_{1}}\cdots x_{3}^{\alpha_{3}}x_{4}^{\alpha_{5}}x_{5}^{\alpha_{4}}\cdots x_{n}^{\alpha_{n}}}
Case 2.

Now, assume Ξ±4<Ξ±5=Ξ±3\alpha_{4}<\alpha_{5}=\alpha_{3}.

π¯3​π¯4​π¯3​(x1Ξ±1​⋯​x3Ξ±3​x4Ξ±4​x5Ξ±5​⋯​xnΞ±nΒ―)\displaystyle\overline{\pi}_{3}\overline{\pi}_{4}\overline{\pi}_{3}(\overline{x_{1}^{\alpha_{1}}\cdots x_{3}^{\alpha_{3}}x_{4}^{\alpha_{4}}x_{5}^{\alpha_{5}}\cdots x_{n}^{\alpha_{n}}}) =π¯3​π¯4​(x1Ξ±1​⋯​x3Ξ±4​x4Ξ±3​x5Ξ±5​⋯​xnΞ±nΒ―)\displaystyle=\overline{\pi}_{3}\overline{\pi}_{4}(\overline{x_{1}^{\alpha_{1}}\cdots x_{3}^{\alpha_{4}}x_{4}^{\alpha_{3}}x_{5}^{\alpha_{5}}\cdots x_{n}^{\alpha_{n}}})
=π¯3​(0)\displaystyle=\overline{\pi}_{3}(0)
=0\displaystyle=0
π¯4​π¯3​π¯4​(x1Ξ±1​⋯​x3Ξ±3​x4Ξ±4​x5Ξ±5​⋯​xnΞ±nΒ―)\displaystyle\overline{\pi}_{4}\overline{\pi}_{3}\overline{\pi}_{4}(\overline{x_{1}^{\alpha_{1}}\cdots x_{3}^{\alpha_{3}}x_{4}^{\alpha_{4}}x_{5}^{\alpha_{5}}\cdots x_{n}^{\alpha_{n}}}) =βˆ’Ο€Β―4​π¯3​(x1Ξ±1​⋯​x3Ξ±3​x4Ξ±4​x5Ξ±5​⋯​xnΞ±nΒ―)\displaystyle=-\overline{\pi}_{4}\overline{\pi}_{3}(\overline{x_{1}^{\alpha_{1}}\cdots x_{3}^{\alpha_{3}}x_{4}^{\alpha_{4}}x_{5}^{\alpha_{5}}\cdots x_{n}^{\alpha_{n}}})
=βˆ’Ο€Β―4​(x1Ξ±1​⋯​x3Ξ±4​x4Ξ±3​x5Ξ±5​⋯​xnΞ±nΒ―)\displaystyle=-\overline{\pi}_{4}(\overline{x_{1}^{\alpha_{1}}\cdots x_{3}^{\alpha_{4}}x_{4}^{\alpha_{3}}x_{5}^{\alpha_{5}}\cdots x_{n}^{\alpha_{n}}})
=0\displaystyle=0

∎

Proposition 6.19.

β„›n^\widehat{\mathcal{R}_{n}} is a β„‹n​(0)\mathcal{H}_{n}(0)-module under the previously defined π¯i\overline{\pi}_{i} action.

Proof.

To show that β„›n^\widehat{\mathcal{R}_{n}} is a β„‹n​(0)\mathcal{H}_{n}(0)-action, we need only show that the commuting relations hold.

First, consider π¯i2​(x1Ξ±1​⋯​xiΞ±i​xi+1Ξ±i+1​⋯​xnΞ±nΒ―)\overline{\pi}_{i}^{2}(\overline{x_{1}^{\alpha_{1}}\cdots x_{i}^{\alpha_{i}}x_{i+1}^{\alpha_{i+1}}\cdots x_{n}^{\alpha_{n}}}). There are three cases, and we work through each.

Case 1.

Assume that Ξ±i>Ξ±i+1\alpha_{i}>\alpha_{i+1}. Then,

π¯i2​(x1Ξ±1​⋯​xiΞ±i​xi+1Ξ±i+1​⋯​xnΞ±nΒ―)\displaystyle\overline{\pi}_{i}^{2}(\overline{x_{1}^{\alpha_{1}}\cdots x_{i}^{\alpha_{i}}x_{i+1}^{\alpha_{i+1}}\cdots x_{n}^{\alpha_{n}}}) =π¯i​(x1Ξ±1​⋯​xiΞ±i+1​xi+1Ξ±i​⋯​xnΞ±nΒ―)\displaystyle=\overline{\pi}_{i}(\overline{x_{1}^{\alpha_{1}}\cdots x_{i}^{\alpha_{i+1}}x_{i+1}^{\alpha_{i}}\cdots x_{n}^{\alpha_{n}}})
=βˆ’x1Ξ±1​⋯​xiΞ±i+1​xi+1Ξ±i​⋯​xnΞ±nΒ―)\displaystyle=-\overline{x_{1}^{\alpha_{1}}\cdots x_{i}^{\alpha_{i+1}}x_{i+1}^{\alpha_{i}}\cdots x_{n}^{\alpha_{n}}})
=βˆ’Ο€Β―i​(x1Ξ±1​⋯​xiΞ±i​xi+1Ξ±i+1​⋯​xnΞ±nΒ―)\displaystyle=-\overline{\pi}_{i}(\overline{x_{1}^{\alpha_{1}}\cdots x_{i}^{\alpha_{i}}x_{i+1}^{\alpha_{i+1}}\cdots x_{n}^{\alpha_{n}}})
Case 2.

Assume that Ξ±i<Ξ±i+1\alpha_{i}<\alpha_{i+1}. Then,

π¯i2​(x1Ξ±1​⋯​xiΞ±i​xi+1Ξ±i+1​⋯​xnΞ±nΒ―)\displaystyle\overline{\pi}_{i}^{2}(\overline{x_{1}^{\alpha_{1}}\cdots x_{i}^{\alpha_{i}}x_{i+1}^{\alpha_{i+1}}\cdots x_{n}^{\alpha_{n}}}) =π¯i​(βˆ’x1Ξ±1​⋯​xiΞ±i​xi+1Ξ±i+1​⋯​xnΞ±nΒ―)\displaystyle=\overline{\pi}_{i}(-\overline{x_{1}^{\alpha_{1}}\cdots x_{i}^{\alpha_{i}}x_{i+1}^{\alpha_{i+1}}\cdots x_{n}^{\alpha_{n}}})
=x1Ξ±1​⋯​xiΞ±i+1​xi+1Ξ±i​⋯​xnΞ±nΒ―)\displaystyle=\overline{x_{1}^{\alpha_{1}}\cdots x_{i}^{\alpha_{i+1}}x_{i+1}^{\alpha_{i}}\cdots x_{n}^{\alpha_{n}}})
=βˆ’Ο€Β―i​(x1Ξ±1​⋯​xiΞ±i​xi+1Ξ±i+1​⋯​xnΞ±nΒ―)\displaystyle=-\overline{\pi}_{i}(\overline{x_{1}^{\alpha_{1}}\cdots x_{i}^{\alpha_{i}}x_{i+1}^{\alpha_{i+1}}\cdots x_{n}^{\alpha_{n}}})
Case 3.

Assume that Ξ±i=Ξ±i+1\alpha_{i}=\alpha_{i+1}. Then,

π¯i2​(x1Ξ±1​⋯​xiΞ±i​xi+1Ξ±i+1​⋯​xnΞ±nΒ―)\displaystyle\overline{\pi}_{i}^{2}(\overline{x_{1}^{\alpha_{1}}\cdots x_{i}^{\alpha_{i}}x_{i+1}^{\alpha_{i+1}}\cdots x_{n}^{\alpha_{n}}}) =π¯i​(0)\displaystyle=\overline{\pi}_{i}(0)
=0\displaystyle=0
=βˆ’Ο€Β―i​(x1Ξ±1​⋯​xiΞ±i​xi+1Ξ±i+1​⋯​xnΞ±nΒ―)\displaystyle=-\overline{\pi}_{i}(\overline{x_{1}^{\alpha_{1}}\cdots x_{i}^{\alpha_{i}}x_{i+1}^{\alpha_{i+1}}\cdots x_{n}^{\alpha_{n}}})

Now, we consider π¯i​π¯j​(x1Ξ±1​⋯​xnΞ±nΒ―)\overline{\pi}_{i}\overline{\pi}_{j}(\overline{x_{1}^{\alpha_{1}}\cdots x_{n}^{\alpha_{n}}}), where |iβˆ’j|β‰₯2|i-j|\geq 2. However, since |iβˆ’j|β‰₯2|i-j|\geq 2, Ξ±i,Ξ±i+1\alpha_{i},\alpha_{i+1} and Ξ±j,Ξ±j+1\alpha_{j},\alpha_{j+1} have no direct impact on the actions of π¯j\overline{\pi}_{j} and π¯i\overline{\pi}_{i}, respectively.

Finally, to show that π¯i​π¯i+1​π¯i​(x1Ξ±1​⋯​xnΞ±nΒ―)=π¯i+1​π¯i​π¯i+1​(x1Ξ±1​⋯​xnΞ±nΒ―)\overline{\pi}_{i}\overline{\pi}_{i+1}\overline{\pi}_{i}(\overline{x_{1}^{\alpha_{1}}\cdots x_{n}^{\alpha_{n}}})=\overline{\pi}_{i+1}\overline{\pi}_{i}\overline{\pi}_{i+1}(\overline{x_{1}^{\alpha_{1}}\cdots x_{n}^{\alpha_{n}}}), we note that any counterexample to this commuting relation must also be a counterexample if i=3i=3 by Lemma 6.16. Thus, we need only consider the case when i=3i=3, but then we are done by Lemma 6.18.

Therefore, β„›n^\widehat{\mathcal{R}_{n}} is a β„‹n​(0)\mathcal{H}_{n}(0)-module under the previously defined π¯i\overline{\pi}_{i} action. ∎

Now, we are ready to calculate the quasisymmetric Frobenius characteristic of β„›n^\widehat{\mathcal{R}_{n}}. Before we begin the proof, we need a definition and some facts about RSK.

Definition 6.20.

Define the inversion code of a permutation wβˆˆπ”–nw\in\mathfrak{S}_{n} to be the sequence c​(w)=(c1,c2,…,cn)c(w)=(c_{1},c_{2},\dots,c_{n}) for wβˆˆπ”–nw\in\mathfrak{S}_{n} such that ci=|{j​<i∣wjβˆ’1>​wiβˆ’1}|c_{i}=|\{j<i\mid w^{-1}_{j}>w^{-1}_{i}\}| is the number of letters to the right of ii in the one-line notation of ww that are smaller than it.

This sequence is such that 0≀ci<i0\leq c_{i}<i for all i≀ni\leq n. In particular, the inversion codes c​(w)c(w) are a variation of the well known Lehmer codes L​(w)L(w), with

c​(w)=rev​(L​(flip​(rev​(wβˆ’1))))c(w)=\text{rev}(L(\text{flip}(\text{rev}(w^{-1}))))

where flip​(w)\text{flip}(w) is the permutation defined by flip​(w)i=n+1βˆ’wi\text{flip}(w)_{i}=n+1-w_{i}. Thus, the inversion codes (and thus the permutations of 𝔖n\mathfrak{S}_{n}) are in bijection with the sub-staircase monomials π’œn\mathcal{A}_{n}.

Furthermore, we will also utilize the RSK correspondence and the following fact:

Fact 6.21.

(See [5].) Given a permutation wβˆˆπ”–nw\in\mathfrak{S}_{n}, RSK assigns a pair of standard Young tableaux (P,Q)(P,Q), where

ides​(P)=ides​(w)\text{ides}(P)=\text{ides}(w)

and

ides​(Q)=des​(w).\text{ides}(Q)=\text{des}(w).

With this, we have our main result of this section.

Theorem 6.22.
Fc​h​a​rQ​(β„›n^;q)=βˆ‘Ξ»βŠ’nβˆ‘Q∈SYT​(Ξ»)qmaj​(Q)​sΞ»=Fc​h​a​r​(β„›n;q)F_{char}^{Q}(\widehat{\mathcal{R}_{n}};q)=\sum\limits_{\lambda\vdash n}\sum\limits_{Q\in\text{SYT}(\lambda)}q^{\text{maj}(Q)}s_{\lambda}=F_{char}(\mathcal{R}_{n};q)
Proof.

Assume that Ξ²j∈An\beta^{j}\in A_{n} for 1≀j≀n!1\leq j\leq n!. Then, the monomials xΞ²jx^{\beta^{j}} can be ordered lexicographically, from greatest to least such that xΞ²1x^{\beta^{1}} is the largest monomial in the list. Then, let Ξ²kj\beta^{j}_{k} be the kthk^{\text{th}} part of Ξ²j\beta^{j}.

We can define a composition series of our module β„›n^\widehat{\mathcal{R}_{n}} such that, if Mj=span,(xΞ²jΒ―,xΞ²j+1Β―,…,xΞ²n!Β―)M_{j}=\text{span},(\overline{x^{\beta^{j}}},\overline{x^{\beta^{j+1}}},\dots,\overline{x^{\beta^{n!}}}), then

β„›n^=M1βŠƒM2βŠƒβ‹―β€‹Mn!=βˆ….\widehat{\mathcal{R}_{n}}=M_{1}\supset M_{2}\supset\cdots M_{n!}=\emptyset.

We note a couple of facts about this composition series. First, using [β‹…]j[\cdot]_{j} to denote the equivalence classes in \faktor​Mj​Mj+1\faktor{M_{j}}{M_{j+1}}, we observe that \faktor​Mj​Mj+1\faktor{M_{j}}{M_{j+1}} is one-dimensional, generated by [xΞ²jΒ―]j[\overline{x^{\beta^{j}}}]_{j}. Furthermore, this composition series is, in fact, a composition series of submodules since we have MjβŠƒMj+1M_{j}\supset M_{j+1}, as our action is based on ≀\leq.

Since \faktor​Mj​Mj+1=⟨[xΞ²jΒ―]jβŸ©β‰…CI\faktor{M_{j}}{M_{j+1}}=\langle[\overline{x^{\beta^{j}}}]_{j}\rangle\cong C_{I}, where I={iβˆ£Ο€Β―i​(xΞ²jΒ―)=βˆ’xΞ²jΒ―}I=\{i\mid\overline{\pi}_{i}(\overline{x^{\beta^{j}}})=-\overline{x^{\beta^{j}}}\}. However, from the definition of the π¯i\overline{\pi}_{i} action, this means I={i∣βij<Ξ²i+1j}I=\{i\mid\beta^{j}_{i}<\beta^{j}_{i+1}\}. Therefore,

Fc​h​a​rQ​(β„›n^;q)=βˆ‘k=1n!qdeg​(xΞ²k)​F{i∣βik<Ξ²i+1k}F_{char}^{Q}(\widehat{\mathcal{R}_{n}};q)=\sum\limits_{k=1}^{n!}q^{\text{deg}(x^{\beta^{k}})}F_{\{i\mid\beta^{k}_{i}<\beta^{k}_{i+1}\}}

Now, we know that the exponents Ξ²k\beta^{k} are in bijection with permutations wβˆˆπ”–nw\in\mathfrak{S}_{n} with corresponding inversion code c​(w)c(w), where c​(w)c(w) counts the number of inversions as defined above. So, if Ξ²ik<Ξ²i+1k\beta^{k}_{i}<\beta^{k}_{i+1} for some ii, then ci<ci+1c_{i}<c_{i+1} in c​(w)c(w). By definition of cic_{i}, this means there are more letters to the right of i+1i+1 that are smaller than it in ww than are to the right of ii that are smaller. In particular, this must mean that ii is to the right of i+1i+1 in ww. Conversely, if i∈ides​(w)i\in\text{ides}(w), i+1i+1 is to the left of ii and every element smaller than ii and to the right of it is also smaller than i+1i+1 and to it’s right. Since ii is also smaller than i+1i+1 and to the right of i+1i+1, we have ci+1>cic_{i+1}>c_{i}. Thus, for a given Ξ²k\beta^{k}, we have {i∣βik<Ξ²i+1k}=ides​(w)\{i\mid\beta^{k}_{i}<\beta^{k}_{i+1}\}=\text{ides}(w) for the unique wβˆˆπ”–nw\in\mathfrak{S}_{n} with c​(w)=Ξ²kc(w)=\beta^{k}.

Thus, we have the following:

Fc​h​a​rQ​(β„›n^;q)\displaystyle F_{char}^{Q}(\widehat{\mathcal{R}_{n}};q) =βˆ‘k=1n!qdeg​(xΞ²k)​F{i∣βik<Ξ²i+1k}\displaystyle=\sum\limits_{k=1}^{n!}q^{\text{deg}(x^{\beta^{k}})}F_{\{i\mid\beta^{k}_{i}<\beta^{k}_{i+1}\}}
=βˆ‘wβˆˆπ”–nqinv​(w)​Fides​(w)\displaystyle=\sum\limits_{w\in\mathfrak{S}_{n}}q^{\text{inv}(w)}F_{\text{ides}(w)}

Next, we use the Foata bijection [13] to exploit the relationship between the inversion number and major index. In particular, we also note that the Foata bijection preserves the i-descent sets of the permutation. Thus, we have

Fc​h​a​rQ​(β„›n^;q)=βˆ‘wβˆˆπ”–nqmaj​(w)​Fides​(w)F_{char}^{Q}(\widehat{\mathcal{R}_{n}};q)=\sum\limits_{w\in\mathfrak{S}_{n}}q^{\text{maj}(w)}F_{\text{ides}(w)}

By the RSK correspondence on permutations in 𝔖n\mathfrak{S}_{n}, we have a bijection to pairs of standard Young tableaux, giving us

Fc​h​a​rQ​(β„›n^;q)\displaystyle F_{char}^{Q}(\widehat{\mathcal{R}_{n}};q) =βˆ‘wβˆˆπ”–nqmaj​(w)​Fides​(w)\displaystyle=\sum\limits_{w\in\mathfrak{S}_{n}}q^{\text{maj}(w)}F_{\text{ides}(w)}
=βˆ‘Ξ»βŠ’nβˆ‘Q∈SYT​(Ξ»)qmaj​(Q)β€‹βˆ‘P∈SYT​(Ξ»)Fides​(P)\displaystyle=\sum\limits_{\lambda\vdash n}\sum\limits_{Q\in\text{SYT}(\lambda)}q^{\text{maj}(Q)}\sum\limits_{P\in\text{SYT}(\lambda)}F_{\text{ides}(P)}
=βˆ‘Ξ»βŠ’nβˆ‘Q∈SYT​(Ξ»)qmaj​(Q)​sΞ»\displaystyle=\sum\limits_{\lambda\vdash n}\sum\limits_{Q\in\text{SYT}(\lambda)}q^{\text{maj}(Q)}s_{\lambda}

∎

Corollary 6.23.
Fc​h​a​rQ​(β„›n^)=βˆ‘Ξ»βŠ’nfλ​sΞ»=Fc​h​a​r​(β„›n)F_{char}^{Q}(\widehat{\mathcal{R}_{n}})=\sum\limits_{\lambda\vdash n}f^{\lambda}s_{\lambda}=F_{char}(\mathcal{R}_{n})
Proof.

Given Theorem 6.22, set q=1q=1. The result follows, recalling that fλf^{\lambda} is the number of standard Young tableaux of shape λ\lambda. ∎

Remark 6.24.

Throughout this section, we recall that we used lexicographic ordering to define an ordering on β„›n\mathcal{R}_{n}. However, we could choose any monomial ordering for which we have xic>xi+1cx_{i}^{c}>x_{i+1}^{c} for all ii and cc and get the same resulting quasisymmetric Frobenius image. For example, this result holds for ≀\leq defined based on degree reverse lexicographic ordering, negative degree lexicographic ordering, and similar orderings.

(β„›n,π’œn,≀r)(\mathcal{R}_{n},\mathcal{A}_{n},\leq_{r}) does not satisfy the definition of SQCC given in Definition 5.1. In particular, our β„‹n​(0)\mathcal{H}_{n}(0)-action on β„›n^\widehat{\mathcal{R}_{n}} required comparisons of leading terms due to the more complicated nature of the 𝔖n\mathfrak{S}_{n}-action on β„›n\mathcal{R}_{n}. As a result, we now introduce a weaker compatibility, which is satisfied by all three examples presented.

Definition 6.25.

Let MM be a 𝔖n\mathfrak{S}_{n}-module with basis ℬ={v1,…,vn}\mathcal{B}=\{v_{1},\dots,v_{n}\} and a total ordering ≀\leq on ℬ\mathcal{B}. On the elements of ℬ\mathcal{B}, let

π¯i​(vk)={0tt​(si​(vk))=vkβˆ’vklt​(si​(vk))=vj>vk​ orΒ lt​(si​(vk))=βˆ’vksi​(vk)lt​(si​(vk))=vj<vk.\overline{\pi}_{i}(v_{k})=\begin{cases}0&\text{tt}(s_{i}(v_{k}))=v_{k}\\ -v_{k}&\text{lt}(s_{i}(v_{k}))=v_{j}>v_{k}\text{ or }\text{lt}(s_{i}(v_{k}))=-v_{k}\\ s_{i}(v_{k})&\text{lt}(s_{i}(v_{k}))=v_{j}<v_{k}\end{cases}.

In particular, assume all elements of ℬ\mathcal{B} fall into one of these three cases. When π¯i\overline{\pi}_{i} is a β„‹n​(0)\mathcal{H}_{n}(0)-action on the vector space of MM, call the resulting module M^\widehat{M}. We say (M,ℬ,≀)(M,\mathcal{B},\leq) is weakly quasisymmetric characteristic compatible (WQCC) if

Fc​h​a​rQ​(M^)=Fc​h​a​r​(M).F_{char}^{Q}(\widehat{M})=F_{char}(M).

As the name suggests, every triple (M,ℬ,≀)(M,\mathcal{B},\leq) that is SQCC is also WQCC. Additionally, we have:

Theorem 6.26.

(β„›n,π’œn,≀r)(\mathcal{R}_{n},\mathcal{A}_{n},\leq_{r}) is WQCC.

Proof.

This follows from Theorem 6.22 and Definition 6.25. ∎

Extensive computer simulation suggests the following additional example of a WQCC triple:

Conjecture.

Let β„›ΞΌ\mathcal{R}_{\mu} denote the Garsia-Procesi modules studied in [14] with basis ℬμ\mathcal{B}_{\mu}. Then, under the standard lexicographic monomial ordering ≀r\leq_{r}, (β„›ΞΌ,ℬμ,≀r)(\mathcal{R}_{\mu},\mathcal{B}_{\mu},\leq_{r}) is WQCC.

7. Open Questions

There are two natural motivations for deforming a 𝔖n\mathfrak{S}_{n}-module to create a β„‹n​(0)\mathcal{H}_{n}(0)-module:

  • β€’

    To better understand the 𝔖n\mathfrak{S}_{n}-module and

  • β€’

    To create natural β„‹n​(0)\mathcal{H}_{n}(0)-modules of interest.

Focusing on the first motivation, many of the best known open conjectures in this area of algebraic combinatorics give the graded Frobenius characteristic of an 𝔖n\mathfrak{S}_{n}-module as a conjectured sum expressed in terms of Gessel’s fundamental basis and only conjectured to be Schur positive. Such conjectures suggest an important open question:

Open Question 1.

Are there natural conditions one can place on a triple (M,ℬ,≀)(M,\mathcal{B},\leq) such that the π¯i\overline{\pi}_{i} action in Definition 6.25 is always valid and

Fc​h​a​rQ​(M^)=Fc​h​a​r​(M)​?F_{char}^{Q}(\widehat{M})=F_{char}(M)?

While this paper explores a uniform deformation of 𝔖n\mathfrak{S}_{n}-modules for the purposes of studying their Frobenius image, our approach has the drawback of mapping quotient spaces to less natural modules from a purely algebraic perspective. If a primary goal is to learn more about the underlying 𝔖n\mathfrak{S}_{n}-module, this is perhaps a reasonable trade off, but a different sensible goal is to try to preserve quotient spaces as quotient spaces. A limitation to this approach is that deformations of an 𝔖n\mathfrak{S}_{n}-action as a β„‹n​(0)\mathcal{H}_{n}(0)-action may cause an underlying ideal to no longer be closed under the newly defined action. Thus a a related underlying question is:

Open Question 2.

Let II be an ideal of 𝔖n\mathfrak{S}_{n}-module MM which is closed under the 𝔖n\mathfrak{S}_{n}-action. Is there a natural way to define a related ideal I~\widetilde{I} and β„‹n​(0)\mathcal{H}_{n}(0)-module M~\widetilde{M} such that I~\widetilde{I} is closed under the β„‹n​(0)\mathcal{H}_{n}(0)-action and such that

Fc​h​a​rQ​(\faktor​M~​I~)=Fc​h​a​r​(\faktor​M​I)​?F_{char}^{Q}\left(\faktor{\widetilde{M}}{\widetilde{I}}\right)=F_{char}\left(\faktor{M}{I}\right)?

Huang and Rhoades in [4] explored the answer to this open question in the case of the generalized coinvariant algebra defined by Haglund, Rhoades, and Shimozono in [15]. The following definition reduces to β„›n\mathcal{R}_{n}, the space we choose to explore earlier in this paper, when restricted to k=nk=n.

Definition 7.1 (Huang, Rhoades, Shimozono [15]).

Consider the ideal

In,k=⟨x1k,…,xnk,en​(x1,…,xn),enβˆ’1​(x1,…,xn),…,enβˆ’k+1​(x1,…,xn)⟩.I_{n,k}=\langle x_{1}^{k},\dots,x_{n}^{k},e_{n}(x_{1},\dots,x_{n}),e_{n-1}(x_{1},\dots,x_{n}),\dots,e_{n-k+1}(x_{1},\dots,x_{n})\rangle.

Define the generalized coinvariant algebra as the quotient

β„›n,k=\faktor​ℂ​[x1,…,xn]​In,k.\mathcal{R}_{n,k}=\faktor{\mathbb{C}[x_{1},\dots,x_{n}]}{I_{n,k}}.

In particular, Huang and Rhoades studied a different deformation of the natural 𝔖n\mathfrak{S}_{n}-action, where the β„‹n​(0)\mathcal{H}_{n}(0)-action is defined by the Demazure operators:

Ο€i​(f)=xi​fβˆ’xi+1​(si​(f))xiβˆ’xi+1.\pi_{i}(f)=\frac{x_{i}f-x_{i+1}(s_{i}(f))}{x_{i}-x_{i+1}}.

However, they noted the ideal In,kI_{n,k} is not closed under this Demazure action. In [4], they define an analogous ideal Jn,kJ_{n,k} that is closed under the Demazure action.

Definition 7.2 (Huang, Rhoades [4]).

Define the ideal

Jn,k=⟨en​(x1,…,xn),enβˆ’1​(x1,…,xn),…,enβˆ’k+1​(x1,…,xn),hk​(x1),hk​(x1,x2),…,hk​(x1,…,xn)⟩.J_{n,k}=\langle e_{n}(x_{1},\dots,x_{n}),e_{n-1}(x_{1},\dots,x_{n}),\dots,e_{n-k+1}(x_{1},\dots,x_{n}),h_{k}(x_{1}),h_{k}(x_{1},x_{2}),\dots,h_{k}(x_{1},\dots,x_{n})\rangle.

Then,

Sn,k=\faktor​ℂ​[x1,…,xn]​Jn,k.S_{n,k}=\faktor{\mathbb{C}[x_{1},\dots,x_{n}]}{J_{n,k}}.

In particular, their main result is as follows:

Theorem 7.3 (Huang, Rhoades [4]).
Fc​h​a​rQ​(Sn,k)=Fc​h​a​r​(Rn,k)F_{char}^{Q}(S_{n,k})=F_{char}(R_{n,k})

It would be interesting to see if a more general procedure could be developed on all quotient spaces of ℂ​[x1,…​xn]\mathbb{C}[x_{1},\dots x_{n}] that similarly leave the Frobenius image unchanged and have this construction of Rhoades and Huang as a special case.

Acknowledgement.

We’d like to thank Rob McCloskey and Jonathan Bloom for their thoughtful discussions throughout the development of this paper. The first author wishes to gratefully acknowledge the Institut Mittag-Leffler for graciously hosting her during a portion of this work.

References

  • [1] D.Β Krob and J-Y Thibon. Non-commutative symmetric functions iv: Quantum linear groups and hecke algebras at q=0. Journal of Algebraic Combinatorics, pages 339–376, 1997.
  • [2] P.N. Norton. 0-hecke algebras. Journal of the Australian Mathematical Society, Series A(27):337–357, 1979.
  • [3] J.Β Huang. A tableau approach to the representation theory of 0-hecke algebras. arXiv, Submitted 2016.
  • [4] Jia Huang and Brendon Rhoades. Ordered set partitions and the 0-hecke algebra. Algebraic Combinatorics, 1(1):47–80, 2018.
  • [5] B.Β Sagan. The Symmetric Group: Representations, Combinatorial Algorithms, and Symmetric Functions. Springer, 1991.
  • [6] R.Β P. Stanley. Enumerative Combinatorics, Volume 2. Cambridge University Press, 1999.
  • [7] V.Β Tewari and S.Β van Willigenburg. Modules of the 0-hecke algebra and quasisymmetric schur functions. Advances in Mathematics, 285:1025–1065, 2015.
  • [8] DavidΒ S. Dummit and RichardΒ M. Foote. Abstract Algebra Third Edition. John Wiley & Sons, 2004.
  • [9] FranΓ§ois Bergeron. Algebraic combinatorics and coinvariant spaces. CRC Press, 2009.
  • [10] E.Β Artin. Galois Theory. Notre Dame Math Lectures, no. 2, 1944.
  • [11] C.Β Chevalley. Invariants of finite groups generated by reflections. American Journal of Mathematics, 77:778–782, 1955.
  • [12] R.P. Stanley. Invariants of finite groups and their applications to combinatorics. Bulletin of the American Mathematical Society, 1:475–511, 1979.
  • [13] D.Β Foata and M.P. SchΓΌtzenburger. Major index and inversion number of permutations. Mathematische Nachrichten, 83:143–159, 1978.
  • [14] A.M. Garsia and C.Β Procesi. On certain graded Sn{S}_{n}-modules and the qq-kostka polynomials. Advances in Mathematics, 94:82–138, 1992.
  • [15] J.Β Haglund, B.Β Rhoades, and M.Β Shimozono. Ordered set partitions, generalized coinvairant algebras, and the delta conjecture. Advances in Mathematics, 329:851–915, 2018.