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

The characters of symmetric groups that depend only on length

Alexander Rossi Miller alexander.r.miller@univie.ac.at
Abstract.

The characters χ(π)\chi(\pi) of SnS_{n} that depend only on the number of cycles of π\pi are completely determined.

1. Introduction

Let (π)\ell(\pi) denote the number of cycles of a permutation πSn\pi\in S_{n}. Let ϕ0,ϕ1,,ϕn1\phi_{0},\phi_{1},\ldots,\phi_{n-1} be the Foulkes characters of SnS_{n}, so ϕi\phi_{i} is afforded by the sum of Specht modules VβV_{\beta} with β\beta of border shape with exactly nn boxes and i+1i+1 rows. For history, properties, and a number of recent developments, including the Diaconis–Fulman connection with adding random numbers, see [2, 3, 6, 7, 8, 9, 10, 11].

The Foulkes characters have long been studied for their remarkable properties:

  1. a) 

    They depend only on length in the sense that

    ϕi(π)=ϕi(σ)whenever(π)=(σ).\phi_{i}(\pi)=\phi_{i}(\sigma)\ \text{whenever}\ \ell(\pi)=\ell(\sigma).
  2. b) 

    They form a basis for the space CF(Sn)\mathrm{CF}_{\ell}(S_{n}) of class functions of SnS_{n} that depend only on \ell, with each θCF(Sn)\theta\in\mathrm{CF}_{\ell}(S_{n}) decomposing uniquely as

    θ=i=0n1θ,εiεi(1)ϕi,\theta=\sum_{i=0}^{n-1}\frac{\langle\theta,\varepsilon_{i}\rangle}{\varepsilon_{i}(1)}\phi_{i},

    where εi\varepsilon_{i} is the irreducible character χ(ni,1,,1)\chi_{(n-i,1,\ldots,1)}.

  3. c) 

    They decompose the character ρ\rho of the regular representation:

    ϕ0+ϕ1++ϕn1=ρ.\phi_{0}+\phi_{1}+\ldots+\phi_{n-1}=\rho.
  4. d) 

    Their degrees are the Eulerian numbers:

    ϕi(1)=#{πSndes(π)=i},des(π)=#{iπ(i)>π(i+1)}.\phi_{i}(1)=\#\{\pi\in S_{n}\mid\mathrm{des}(\pi)=i\},\quad\mathrm{des}(\pi)=\#\{i\mid\pi(i)>\pi(i+1)\}.
  5. e) 

    They branch according to

    ϕi|Sn1=(ni)ϕi1+(i+1)ϕi,where ϕ1=0.\phi_{i}|_{S_{n-1}}=(n-i)\phi_{i-1}+(i+1)\phi_{i},\ \ \text{where\ \ $\phi_{-1}=0$}.
  6. f) 

    And they even admit an explicit expression:

    ϕi(π)=j=0n1(1)ij(n+1ij)(j+1)(π),\phi_{i}(\pi)=\sum_{j=0}^{n-1}(-1)^{i-j}\binom{n+1}{i-j}(j+1)^{\ell(\pi)},

    with the usual convention, used throughout this paper, that (uv)=0\binom{u}{v}=0 if vv does not satisfy uv0u\geq v\geq 0.

Two missing properties were recently added to the list in [11], the first being a solution to the problem of decomposing products:

  1. g) 

    For any two Foulkes characters ϕi\phi_{i} and ϕj\phi_{j} of SnS_{n},

    ϕiϕj=k=0n1cijkϕk,\phi_{i}\phi_{j}=\sum_{k=0}^{n-1}c_{ijk}\phi_{k},

    where

    cijk=#{(x,y)Sn×Sndes(x)=i,des(y)=j,xy=z}c_{ijk}=\#\{(x,y)\in S_{n}\times S_{n}\mid\mathrm{des}(x)=i,\ \mathrm{des}(y)=j,\ xy=z\}

    for any fixed zSnz\in S_{n} with des(z)=k\mathrm{des}(z)=k, or more explicitly,

    cijk=0ui0vj(1)iu+jv(n+1iu)(n+1jv)(uv+u+v+nkn).c_{ijk}=\sum_{{0\leq u\leq i}\atop{0\leq v\leq j}}(-1)^{i-u+j-v}\binom{n+1}{i-u}\binom{n+1}{j-v}\binom{uv+u+v+n-k}{n}.

In addition to the combinatorial solution and the explicit solution presented here, a recursive solution was also found by specializing results of Delsarte that predate the introduction of Foulkes characters and were discovered in a completely different context, see [11]. The combinatorial solution follows from an earlier result [7, Thm. 9] that connects Foulkes characters with Eulerian idempotents from cyclic homology, namely that

𝒟i=j=1nϕi(Cj)j1,\mathscr{D}_{i}=\sum_{j=1}^{n}\phi_{i}(C_{j})\mathscr{E}_{j-1},

where ϕi(Ck)\phi_{i}(C_{k}) denotes the value ϕi(σ)\phi_{i}(\sigma) at any σCk={πSn(π)=k}\sigma\in C_{k}=\{\pi\in S_{n}\mid\ell(\pi)=k\},

𝒟i=πSndes(π)=iπ,\mathscr{D}_{i}=\sum_{{\pi\in S_{n}}\atop{\mathrm{des}(\pi)=i}}\pi,

and the i\mathscr{E}_{i}’s are the Eulerian idempotents, which are certain orthogonal idempotents in [Sn]\mathbb{Q}[S_{n}], see [11, (2.5)].

The second property added to the list in [11] solves the problem of finding a natural description of the unique inner product on CF(Sn)\mathrm{CF}_{\ell}(S_{n}) with respect to which the ϕi\phi_{i}’s form an orthonormal basis:

  1. h) 

    The Foulkes characters ϕ0,ϕ1,,ϕn1\phi_{0},\phi_{1},\ldots,\phi_{n-1} of SnS_{n} form an orthonormal basis for the Hilbert space CF(Sn)\mathrm{CF}_{\ell}(S_{n}) with inner product [-,-][\,\text{-}\,,\,\text{-}\,] defined by

    [θ,ψ]=1|Sn|i,j=1nθ(Ci)ψ(Cj)¯𝐄|σCiτCj|,[\theta,\psi]=\frac{1}{|S_{n}|}\sum_{i,j=1}^{n}\theta(C_{i})\overline{\psi(C_{j})}\mathbf{E}|\sigma C_{i}\cap\tau C_{j}|,

    where σ\sigma and τ\tau are nn-cycles chosen uniformly at random from C1C_{1}, so 𝐄|σCiτCj|\mathbf{E}|\sigma C_{i}\cap\tau C_{j}| equals the expected number of ways that the product στ\sigma\tau can be written as a product αβ\alpha\beta with αCi\alpha\in C_{i} and βCj\beta\in C_{j}.

This inner product [-,-][\,\text{-}\,,\,\text{-}\,] gives a new and natural construction of Foulkes characters by applying the Gram–Schmidt process to a natural choice of basis for CF(Sn)\mathrm{CF}_{\ell}(S_{n}) that is composed of characters, analogous to how the irreducible characters themselves can be constructed by applying the Gram–Schmidt process to the characters (1Sλ)Sn(1_{S_{\lambda}})^{S_{n}}, see [11].

The author gave a different construction of Foulkes characters in [7] using sums of modules afforded by reduced homology groups of subcomplexes of certain equivariant strong deformation retracts of wedges of spheres called Milnor fibers coming from invariant theory, which works not only for SnG(1,1,n){S_{n}\cong G(1,1,n)} but for all of the full monomial groups G(r,1,n)G(r,1,n), including the hyperoctahedral group G(2,1,n)G(2,1,n). In this more general setting, the role of \ell is played by n𝔩n-\mathfrak{l}, where 𝔩\mathfrak{l} is the most natural choice of “length”,

𝔩(x)=min{k0x=y1y2ykfor some reflections yiG(r,1,n)}.\mathfrak{l}(x)=\min\{k\geq 0\mid x=y_{1}y_{2}\ldots y_{k}\ \text{for some reflections }y_{i}\in G(r,1,n)\}.

In addition to new properties and proofs in the classical case, analogues of all the properties that we have described so far have been established for G(r,1,n)G(r,1,n). These generalized Foulkes characters also have connections with certain Markov chains, just as in the case of SnS_{n}. Most notably, Diaconis and Fulman [3] connected the hyperoctahedral Foulkes characters with a Markov chain for adding random numbers in balanced ternary, a number system that reduces carries and that Donald Knuth famously described as “perhaps the prettiest number system of all.”

But what is missing from the list of properties presented so far is an analogue of the fundamental characterization of characters, by which we mean if χ0,χ1,,χk1\chi_{0},\chi_{1},\ldots,\chi_{k-1} are the irreducible characters of a finite group GG, then the characters of GG are the non-negative integer linear combinations

𝔞0χ0+𝔞1χ1++𝔞k1χk1,𝔞=(𝔞0,𝔞1,,𝔞k1)tk,\mathfrak{a}_{0}\chi_{0}+\mathfrak{a}_{1}\chi_{1}+\ldots+\mathfrak{a}_{k-1}\chi_{k-1},\quad\mathfrak{a}=(\mathfrak{a}_{0},\mathfrak{a}_{1},\ldots,\mathfrak{a}_{k-1})^{t}\in\mathbb{N}^{k},

where by \mathbb{N} we mean the set of non-negative integers. Interestingly, for the hyperoctahedral groups and all the other full monomial groups with r2r\geq 2, there is an analogue of this characterization [8]:

  1. i 

    If r2r\geq 2, then the characters of G(r,1,n)G(r,1,n) that depend only on length are precisely the non-negative integer linear combinations of the Foulkes characters of the group.

So for the monomial groups G(r,1,n)G(r,1,n) with r2r\geq 2, the picture is complete.

For the symmetric group, however, the problem of determining the characters that depend only on \ell is more complicated and has remained open for a few decades. Already for n=3n=3, one finds that not all characters in CF(Sn)\mathrm{CF}_{\ell}(S_{n}) are integer linear combinations of Foulkes characters, for example χ(2,1)\chi_{(2,1)}, and it was recently shown in [8], using symmetric functions and Chebyshev’s theorem about primes, that the same is true for each n3n\geq 3, but little else is known about the characters in CF(Sn)\mathrm{CF}_{\ell}(S_{n}).

Given this gap in the story for SnS_{n}, a natural question to ask is, instead of the Foulkes characters, does there exist a different sequence of nn characters that is better suited for the role of irreducibles in CF(Sn)\mathrm{CF}_{\ell}(S_{n}) in that their non-negative integer linear combinations are the characters that depend only on length? Our first theorem answers in the negative.

Theorem 1.1.

If n>3n>3, then there does not exist a collection of nn characters χ0,χ1,,χn1CF(Sn)\chi_{0},\chi_{1},\ldots,\chi_{n-1}\in\mathrm{CF}_{\ell}(S_{n}) such that every character of SnS_{n} that depends only on \ell can be written as a non-negative integer linear combination of the χi\chi_{i}’s.

In our main theorem, we completely resolve the problem of determining the characters of SnS_{n} that depend only on length. Our solution is explicit and complete with a parametrization in terms of non-negative integers, just as in the case of ordinary characters. Let {r}=rr\{r\}=r-\lfloor r\rfloor denote the fractional part of any given rational number rr.

Theorem 1.2.

The characters of SnS_{n} that depend only on \ell are the linear combinations

θ𝔞=𝔞~0ϕ0+𝔞~1ϕ1++𝔞~n1ϕn1,𝔞n,\theta_{\mathfrak{a}}=\tilde{\mathfrak{a}}_{0}\phi_{0}+\tilde{\mathfrak{a}}_{1}\phi_{1}+\ldots+\tilde{\mathfrak{a}}_{n-1}\phi_{n-1},\quad\mathfrak{a}\in\mathbb{N}^{n},
𝔞~k=𝔞kdk+1+{j=0n1(nk1jk)𝔞jdj+1},dk=kgcd(n,k),\tilde{\mathfrak{a}}_{k}=\left\lfloor\frac{\mathfrak{a}_{k}}{d_{k+1}}\right\rfloor+\left\{\sum_{j=0}^{n-1}\binom{n-k-1}{j-k}\frac{\mathfrak{a}_{j}}{d_{j+1}}\right\},\quad d_{k}=\frac{k}{\gcd(n,k)},

and moreover,

θ𝔞=θ𝔟if and only if𝔞=𝔟.\theta_{\mathfrak{a}}=\theta_{\mathfrak{b}}\quad\text{if and only if}\quad\mathfrak{a}=\mathfrak{b}.

As a consequence, we find that the non-negative integer linear combinations of Foulkes characters of SnS_{n} have exponentially decaying density among the characters of SnS_{n} that depend only on length. We also obtain an upper bound for this density and information about the denominators of the rational coefficients that occur in Theorem 1.2.

Theorem 1.3.

The number of characters of SnS_{n} that depend only on \ell and lie in the fundamental parallelepiped {tiϕiti[0,1)}\{\sum t_{i}\phi_{i}\mid t_{i}\in[0,1)\} equals

n!gcd(1,n)gcd(2,n)gcd(n,n).\frac{n!}{\gcd(1,n)\gcd(2,n)\ldots\gcd(n,n)}.
Theorem 1.4.

If σn\sigma_{n} denotes the smallest positive integer such that, for each character χ\chi of SnS_{n} that depends only on \ell, σnχ\sigma_{n}\chi is a non-negative integer linear combination of Foulkes characters, then

σn=lcm(1,2,,n)n=ef(n)n,\sigma_{n}=\frac{\mathrm{lcm}(1,2,\ldots,n)}{n}=\frac{e^{f(n)}}{n},

where ff is the second Chebyshev function.

Asymptotics of ff, and in turn σn\sigma_{n}, have important number-theoretic significance, with the prime number theorem being equivalent to

f(n)nf(n)\sim n

and more precise statements being equivalent to the Riemann Hypothesis.

2. Number theoretic preliminaries

Important for our analysis is a certain result which compliments and improves an old result of Cauchy.

Given a partition λ\lambda of nn, abbreviated λn\lambda\vdash n, we shall write mk(λ)m_{k}(\lambda) for the number of parts of λ\lambda that equal kk,

(λ)=k=1nmk(λ),\ell(\lambda)=\sum_{k=1}^{n}m_{k}(\lambda),

and

𝖬(λ)=((λ)m1(λ),m2(λ),,mn(λ))=(λ)!m1(λ)!m2(λ)!mn(λ)!.{\sf M}(\lambda)=\binom{\ell(\lambda)}{m_{1}(\lambda),m_{2}(\lambda),\ldots,m_{n}(\lambda)}=\frac{\ell(\lambda)!}{m_{1}(\lambda)!m_{2}(\lambda)!\ldots m_{n}(\lambda)!}.

Schönemann proved that

gcd(m1(λ),m2(λ),,mn(λ))(λ)𝖬(λ)\frac{\gcd(m_{1}(\lambda),m_{2}(\lambda),\ldots,m_{n}(\lambda))}{\ell(\lambda)}{\sf M}(\lambda)\in\mathbb{Z}

and Cauchy independently proved that

n(λ)𝖬(λ),\frac{n}{\ell(\lambda)}{\sf M}(\lambda)\in\mathbb{Z},

see Dickson’s book [4, Chap. IX, p. 265]. We find the exact gcd of the 𝖬(λ){\sf M}(\lambda)’s where λ\lambda partitions nn into exactly kk non-zero parts.

Proposition 2.1.

For any positive integers nk1n\geq k\geq 1,

(2.1) gcd{𝖬(λ)λn,(λ)=k}=kgcd(n,k).\gcd\{{\sf M}(\lambda)\mid\lambda\vdash n,\ \ell(\lambda)=k\}=\frac{k}{\gcd(n,k)}.
Proof.

Fixing positive integers nn and kk with nkn\geq k, let

g=gcd{𝖬(λ)λn,(λ)=k}.g=\gcd\{{\sf M}(\lambda)\mid\lambda\vdash n,\ \ell(\lambda)=k\}.

Let λ\lambda be a partition of nn with (λ)=k\ell(\lambda)=k. For 1jn1\leq j\leq n,

mj(λ)k(kmj(λ))=(k1mj(λ)1),\frac{m_{j}(\lambda)}{k}\binom{k}{m_{j}(\lambda)}=\binom{k-1}{m_{j}(\lambda)-1},

so

mj(λ)k𝖬(λ).\frac{m_{j}(\lambda)}{k}{\sf M}(\lambda)\in\mathbb{Z}.

Multiplying the left-hand side by jj and then summing over jj gives Cauchy’s result

nk𝖬(λ).\frac{n}{k}{\sf M}(\lambda)\in\mathbb{Z}.

Writing gcd(n,k)=un+vk\gcd(n,k)=un+vk with u,vu,v\in\mathbb{Z}, we therefore have

gcd(n,k)k𝖬(λ)=unk𝖬(λ)+vkk𝖬(λ).\frac{\gcd(n,k)}{k}{\sf M}(\lambda)=u\frac{n}{k}{\sf M}(\lambda)+v\frac{k}{k}{\sf M}(\lambda)\in\mathbb{Z}.

Hence kgcd(n,k)\frac{k}{\gcd(n,k)} divides gg.

To show that gg divides kgcd(n,k)\frac{k}{\gcd(n,k)}, we show that for each prime pp,

vp(g)vp(kgcd(n,k)),v_{p}(g)\leq v_{p}\left(\frac{k}{\gcd(n,k)}\right),

where vp(m)v_{p}(m) denotes the pp-adic valuation of mm. Fix pp and let e=vp(k)e=v_{p}(k), so pekp^{e}\mid k and pe+1kp^{e+1}\nmid k. Write n=ak+rn=ak+r for non-negative integers aa and rr with 0r<k0\leq r<k. Similarly write r=bpe+sr=bp^{e}+s with 0s<pe0\leq s<p^{e}. Writing u=pvp(s)u=p^{v_{p}(s)}, let μ\mu be the partition with uu parts of size a+rua+\frac{r}{u} and kuk-u parts of size aa. Then μ\mu is a partition of nn, (μ)=k\ell(\mu)=k, and

𝖬(μ)=(ku).{\sf M}(\mu)=\binom{k}{u}.

Using Kummer’s theorem, then

vp(𝖬(μ))=vp(k)vp(u)=vp(kgcd(n,k)).v_{p}({\sf M}(\mu))=v_{p}(k)-v_{p}(u)=v_{p}\left(\frac{k}{\gcd(n,k)}\right).

Hence vp(g)vp(kgcd(n,k))v_{p}(g)\leq v_{p}(\frac{k}{\gcd(n,k)}). ∎

3. Proof of Theorem 1.1

Lemma 3.1.

The Frobenius characteristic of (n1)1CF(Sn)(n-1)^{\ell-1}\in\mathrm{CF}_{\ell}(S_{n}) is a non-negative integer linear combination of the symmetric functions hλh_{\lambda}. In particular, the class function (n1)1(n-1)^{\ell-1} is a character of SnS_{n}.

Proof.

The case n=1n=1 is trivial, so assume n>1n>1. By [8, Cor. 8], the Frobenius characteristic ch(θ)\mathrm{ch}(\theta) of the class function θ(π)=(n1)(π)1\theta(\pi)=(n-1)^{\ell(\pi)-1} satisfies

ch(θ)=1n1λncλhλ,\mathrm{ch}(\theta)=\frac{1}{n-1}\sum_{\lambda\vdash n}c_{\lambda}h_{\lambda},
cλ=(n1(λ))𝖬(λ)=(n1m1(λ),m2(λ),,mn(λ),n(λ)1).c_{\lambda}=\binom{n-1}{\ell(\lambda)}{\sf M}(\lambda)=\binom{n-1}{m_{1}(\lambda),m_{2}(\lambda),\ldots,m_{n}(\lambda),n-\ell(\lambda)-1}.

For any non-negative integers k1,k2,,ksk_{1},k_{2},\ldots,k_{s} with sum kk, we have

kik(kk1,k2,,ks)=(k1k1,k2,,ki1,ki1,ki+1,,ks).\frac{k_{i}}{k}\binom{k}{k_{1},k_{2},\ldots,k_{s}}=\binom{k-1}{k_{1},k_{2},\ldots,k_{i-1},k_{i}-1,k_{i+1},\ldots,k_{s}}\in\mathbb{Z}.

So for λn\lambda\vdash n and 1kn1\leq k\leq n,

mk(λ)n1cλ,\frac{m_{k}(\lambda)}{n-1}c_{\lambda}\in\mathbb{Z},

and in turn

nn1cλ=k=1nkmk(λ)n1cλ.\frac{n}{n-1}c_{\lambda}=\sum_{k=1}^{n}k\frac{m_{k}(\lambda)}{n-1}c_{\lambda}\in\mathbb{Z}.

Therefore, for each partition λ\lambda of nn,

1n1cλ=nn1cλcλ.\frac{1}{n-1}c_{\lambda}=\frac{n}{n-1}c_{\lambda}-c_{\lambda}\in\mathbb{Z}.

So ch(θ)\mathrm{ch}(\theta) is a non-negative integer linear combination of the hλh_{\lambda}’s. ∎

Proof of Theorem 1.1.

Suppose that n>3n>3 and χ0,χ1,,χn1\chi_{0},\chi_{1},\ldots,\chi_{n-1} are characters in CF(Sn)\mathrm{CF}_{\ell}(S_{n}) with the property that every character in CF(Sn)\mathrm{CF}_{\ell}(S_{n}) is a non-negative integer linear combination of the χi\chi_{i}’s. Then the change of basis matrices A=(Aij)0i,jn1A=(A_{ij})_{0\leq i,j\leq n-1} and A1A^{-1}, where χj=iAijϕi\chi_{j}=\sum_{i}A_{ij}\phi_{i}, both have all entries non-negative. If at least two entries in a row of AA were non-zero, say in columns ss and tt, then from the non-negativity and the relation AA1=IAA^{-1}=I, the ss and tt rows of A1A^{-1} would be linearly dependent. So AA must be a positive definite diagonal matrix times a permutation matrix, meaning that for some positive scalars s0,s1,,sn1s_{0},s_{1},\ldots,s_{n-1} and some permutation σ\sigma of the indices, χj=sjϕσ(j)\chi_{j}=s_{j}\phi_{\sigma(j)} for 0jn10\leq j\leq n-1.

Let θ\theta be the class function of SnS_{n} given by

θ(π)=(n1)(π)1.\theta(\pi)=(n-1)^{\ell(\pi)-1}.

By Lemma 3.1, θ\theta is a character of SnS_{n}. We claim that θ\theta can not be written as a non-negative integer linear combination of characters η0,η1,,ηn1\eta_{0},\eta_{1},\ldots,\eta_{n-1} with ηi{sϕis0}\eta_{i}\in\{s\phi_{i}\mid s\geq 0\} for 0in10\leq i\leq n-1. Equivalently, if θ\theta is written as

(3.1) θ=c0ϕ0+c1ϕ1++cn1ϕn1,\theta=c_{0}\phi_{0}+c_{1}\phi_{1}+\ldots+c_{n-1}\phi_{n-1},

then at least one of the summands ciϕic_{i}\phi_{i} is not a character. We will show that in fact cn2ϕn2c_{n-2}\phi_{n-2} is not a character.

Let γ0,γ1,,γn1\gamma_{0},\gamma_{1},\ldots,\gamma_{n-1} be the characters of SnS_{n} given by

γk(π)=(k+1)(π),\gamma_{k}(\pi)=(k+1)^{\ell(\pi)},

so γk\gamma_{k} is afforded by (k+1)n(\mathbb{C}^{k+1})^{\otimes n} with

π.(u1u2un)=uπ1(1)uπ1(2)uπ1(n).\pi.(u_{1}\otimes u_{2}\otimes\ldots\otimes u_{n})=u_{\pi^{-1}(1)}\otimes u_{\pi^{-1}(2)}\otimes\ldots\otimes u_{\pi^{-1}(n)}.

Then

θ=γn2n1\theta=\frac{\gamma_{n-2}}{n-1}

and

(3.2) ϕi=j=0n1(1)ij(n+1ij)γj,0in1.\phi_{i}=\sum_{j=0}^{n-1}(-1)^{i-j}\binom{n+1}{i-j}\gamma_{j},\quad 0\leq i\leq n-1.

The inverse of the matrix

[(1)ij(n+1ij)]0i,jn1\left[(-1)^{i-j}\binom{n+1}{i-j}\right]_{0\leq i,j\leq n-1}

equals

[(n+ijn)]0i,jn1,\left[\binom{n+i-j}{n}\right]_{0\leq i,j\leq n-1},

so

(3.3) γi=j=0n1(n+ijn)ϕj,0in1.\gamma_{i}=\sum_{j=0}^{n-1}\binom{n+i-j}{n}\phi_{j},\quad 0\leq i\leq n-1.

Therefore, the coefficient of ϕj\phi_{j} in θ\theta is

(3.4) cj=1n1(2n2jn),0jn1.c_{j}=\frac{1}{n-1}\binom{2n-2-j}{n},\quad 0\leq j\leq n-1.

Let ν\nu be the partition of nn with m2(ν)=2m_{2}(\nu)=2 and m1(ν)=n4m_{1}(\nu)=n-4. Using that for any kk,

k,χν=bk+c(b)h(b),\langle k^{\ell},\chi_{\nu}\rangle=\prod_{b}\frac{k+c(b)}{h(b)},

where bb runs over the boxes in the diagram of ν\nu and where c(b)c(b) and h(b)h(b) denote the content and hook length of bb, we have

(3.5) γj,χν={0if jn4,(n2)(n3)2if j=n3,n(n1)(n3)2if j=n2,n2(n+1)(n3)4if j=n1.\langle\gamma_{j},\chi_{\nu}\rangle=\begin{cases}0&\text{if $j\leq n-4$},\\ \frac{(n-2)(n-3)}{2}&\text{if $j=n-3$},\\ \frac{n(n-1)(n-3)}{2}&\text{if $j=n-2$},\\ \frac{n^{2}(n+1)(n-3)}{4}&\text{if $j=n-1$}.\\ \end{cases}

Combining (3.2), (3.4), and (3.5), we have

cn2ϕn2,χν\displaystyle\langle c_{n-2}\phi_{n-2},\chi_{\nu}\rangle =1n1[(n+11)(n2)(n3)2+n(n1)(n3)2]\displaystyle=\frac{1}{n-1}\left[-\binom{n+1}{1}\frac{(n-2)(n-3)}{2}+\frac{n(n-1)(n-3)}{2}\right]
=n3n2,\displaystyle=\frac{n-3}{n-2},

which is not an integer for n>3n>3. So cn2ϕn2c_{n-2}\phi_{n-2} is not a character. ∎

4. Proofs of Theorems 1.21.4

Fixing nn, we have a chain of lattices

𝒳𝒴𝒵,\mathcal{X}\subset\mathcal{Y}\subset\mathcal{Z},

where 𝒵\mathcal{Z} is the lattice of virtual characters of SnS_{n},

𝒵={χIrr(Sn)aχχ|aχ},\mathcal{Z}=\left\{\textstyle{\sum_{\chi\in\mathrm{Irr}(S_{n})}}a_{\chi}\chi\ \,\big{|}\,\ a_{\chi}\in\mathbb{Z}\right\},

𝒴\mathcal{Y} is the lattice of virtual characters of SnS_{n} that depend only on \ell,

𝒴={χ𝒵χdepends only on },\mathcal{Y}=\{\chi\in\mathcal{Z}\mid\chi\ \text{depends only on $\ell$}\},

and 𝒳\mathcal{X} is the sublattice of 𝒴\mathcal{Y} spanned by the Foulkes characters of SnS_{n},

𝒳={a0ϕ0+a1ϕ1++an1ϕn1ai}.\mathcal{X}=\left\{a_{0}\phi_{0}+a_{1}\phi_{1}+\ldots+a_{n-1}\phi_{n-1}\mid a_{i}\in\mathbb{Z}\right\}.

That the lattice of virtual characters that depend only on length can also be characterized as the set of virtual characters in the rational span of Foulkes characters is fundamental to our approach.

Proposition 4.1.
𝒴=𝒵{r0ϕ0+r1ϕ1++rn1ϕn1ri}.\mathcal{Y}=\mathcal{Z}\cap\{r_{0}\phi_{0}+r_{1}\phi_{1}+\ldots+r_{n-1}\phi_{n-1}\mid r_{i}\in\mathbb{Q}\}.
Proof.

If a rational linear combination of Foulkes characters is a virtual character, then it is a virtual character that depends only on \ell. For the other inclusion, if θ𝒴\theta\in\mathcal{Y}, then θ𝒵\theta\in\mathcal{Z} and

θ=i=0n1θ,εiεi(1)ϕi,εi=χ(ni,1,,1),\theta=\sum_{i=0}^{n-1}\frac{\langle\theta,\varepsilon_{i}\rangle}{\varepsilon_{i}(1)}\phi_{i},\quad\varepsilon_{i}=\chi_{(n-i,1,\ldots,1)},

so θ\theta is a virtual character in the rational span of Foulkes characters. ∎

Let 𝒫\mathcal{P} be the fundamental parallelepiped

𝒫={t0ϕ0+t1ϕ1++tn1ϕn1ti[0,1)}.\mathcal{P}=\{t_{0}\phi_{0}+t_{1}\phi_{1}+\ldots+t_{n-1}\phi_{n-1}\mid t_{i}\in[0,1)\}.

Then each θ𝒴\theta\in\mathcal{Y} decomposes uniquely as

θ=θ𝒳+θ𝒫,θ𝒳𝒳,θ𝒫𝒴𝒫.\theta=\theta_{\mathcal{X}}+\theta_{\mathcal{P}},\qquad\theta_{\mathcal{X}}\in\mathcal{X},\quad\theta_{\mathcal{P}}\in\mathcal{Y}\cap\mathcal{P}.

4.1.

We are interested in the genuine characters in our lattices. Let

={characters of Sn that depend only on }\mathcal{L}=\{\text{characters of $S_{n}$ that depend only on $\ell$}\}

and

={non-negative integer linear combinations of ϕi’s}.\mathcal{F}=\{\text{non-negative integer linear combinations of $\phi_{i}$'s}\}.

So \mathcal{F}\subset\mathcal{L} and both \mathcal{F} and \mathcal{L} are closed under addition.

Lemma 4.2.
𝒵𝒫=𝒴𝒫=𝒫.\mathcal{Z}\cap\mathcal{P}=\mathcal{Y}\cap\mathcal{P}=\mathcal{L}\cap\mathcal{P}\subset\mathcal{L}.
Proof.

Let θ𝒵𝒫\theta\in\mathcal{Z}\cap\mathcal{P}. Then θ\theta is a linear combination of the Foulkes characters with non-negative coefficients, so θ\theta depends only on length and satisfies θ,χ0\langle\theta,\chi\rangle\geq 0 for each χIrr(Sn)\chi\in\mathrm{Irr}(S_{n}). Since θ,χ\langle\theta,\chi\rangle is an integer for each χIrr(Sn)\chi\in\mathrm{Irr}(S_{n}), we conclude that θ\theta is a character of SnS_{n} that depends only on length. ∎

Proposition 4.3.

The characters θ\theta of SnS_{n} that depend only on length are precisely the sums

θ=θ+θ𝒫,θ,θ𝒫𝒵𝒫,\theta=\theta_{\mathcal{F}}+\theta_{\mathcal{P}},\qquad\theta_{\mathcal{F}}\in\mathcal{F},\quad\theta_{\mathcal{P}}\in\mathcal{Z}\cap\mathcal{P},

and the components θ\theta_{\mathcal{F}} and θ𝒫\theta_{\mathcal{P}} are uniquely determined by θ\theta. Equivalently, the addition map (x,y)x+y(x,y)\mapsto x+y takes ×(𝒵𝒫)\mathcal{F}\times(\mathcal{Z}\cap\mathcal{P}) bijectively onto \mathcal{L}. Moreover, if

θ=i=0n1riϕi,\theta=\sum_{i=0}^{n-1}r_{i}\phi_{i},

then

θ=i=0n1riϕiandθ𝒫=i=0n1{ri}ϕi.\theta_{\mathcal{F}}=\sum_{i=0}^{n-1}\lfloor r_{i}\rfloor\phi_{i}\quad\text{and}\quad\theta_{\mathcal{P}}=\sum_{i=0}^{n-1}\{r_{i}\}\phi_{i}.
Proof.

By Lemma 4.2, we have a mapping

α:×(𝒵𝒫)given byα(x,y)=x+y.\alpha:\mathcal{F}\times(\mathcal{Z}\cap\mathcal{P})\to\mathcal{L}\quad\text{given by}\quad\alpha(x,y)=x+y.

Each element of 𝒴\mathcal{Y} is uniquely expressible as an element from the sublattice 𝒳\mathcal{X} plus an element from the fundamental domain 𝒴𝒫=𝒵𝒫\mathcal{Y}\cap\mathcal{P}=\mathcal{Z}\cap\mathcal{P}, and we also have the inclusions 𝒳\mathcal{F}\subset\mathcal{X} and 𝒴\mathcal{L}\subset\mathcal{Y}, so α\alpha is injective.

To show that α\alpha is surjective, consider a character θ\theta that depends only on length. Then

θ=i=0n1riϕi,ri=θ,χ(ni,1,,1)(n1i)0.\theta=\sum_{i=0}^{n-1}r_{i}\phi_{i},\qquad r_{i}=\frac{\langle\theta,\chi_{(n-i,1,\ldots,1)}\rangle}{\binom{n-1}{i}}\geq 0.

Hence

θ=ϕ+ψ,\theta=\phi+\psi,

where

ϕ=i=0n1riϕi\phi=\sum_{i=0}^{n-1}\lfloor r_{i}\rfloor\phi_{i}\in\mathcal{F}

and

ψ=θϕ=i=0n1{ri}ϕi𝒵𝒫.\psi=\theta-\phi=\sum_{i=0}^{n-1}\{r_{i}\}\phi_{i}\in\mathcal{Z}\cap\mathcal{P}.

4.2.

We now proceed to describe a particularly good pair of bases for the lattice 𝒴\mathcal{Y} and the sublattice 𝒳\mathcal{X}, from which we will be able to better understand the fundamental domain 𝒵𝒫\mathcal{Z}\cap\mathcal{P} for 𝒳𝒴\mathcal{X}\subset\mathcal{Y}. Specifically, we find a basis for 𝒴\mathcal{Y} that extends to a basis of 𝒵\mathcal{Z} and has the property that certain multiples of the basis elements form a basis for the sublattice 𝒳\mathcal{X}.

4.2.1.

The new basis for 𝒳\mathcal{X} will be denoted by ψ0,ψ1,,ψn1\psi_{0},\psi_{1},\ldots,\psi_{n-1}.

Definition 4.4.

Define ψ0,ψ1,,ψn1CF(Sn)\psi_{0},\psi_{1},\ldots,\psi_{n-1}\in\mathrm{CF}_{\ell}(S_{n}) by

(4.1) ψi=j=0n1(1)ij(i+1ij)(j+1).\psi_{i}=\sum_{j=0}^{n-1}(-1)^{i-j}\binom{i+1}{i-j}(j+1)^{\ell}.

Denoting by γk\gamma_{k} the character (k+1)(k+1)^{\ell} in CF(Sn)\mathrm{CF}_{\ell}(S_{n}), we have the following relations between the ψi\psi_{i}’s, the ϕi\phi_{i}’s, and the γi\gamma_{i}’s.

Proposition 4.5.
(4.2) ϕi\displaystyle\phi_{i} =j=0n1(1)ij(n+1ij)γj,\displaystyle=\sum_{j=0}^{n-1}(-1)^{i-j}\binom{n+1}{i-j}\gamma_{j}, γi\displaystyle\qquad\gamma_{i} =j=0n1(n+ijij)ϕj,\displaystyle=\sum_{j=0}^{n-1}\binom{n+i-j}{i-j}\phi_{j},
(4.3) ψi\displaystyle\psi_{i} =j=0n1(1)ij(i+1ij)γj,\displaystyle=\sum_{j=0}^{n-1}(-1)^{i-j}\binom{i+1}{i-j}\gamma_{j}, γi\displaystyle\qquad\gamma_{i} =j=0n1(i+1ij)ψj,\displaystyle=\sum_{j=0}^{n-1}\binom{i+1}{i-j}\psi_{j},
(4.4) ϕi\displaystyle\phi_{i} =j=0n1(1)ij(nj1ij)ψj,\displaystyle=\sum_{j=0}^{n-1}(-1)^{i-j}\binom{n-j-1}{i-j}\psi_{j}, ψi\displaystyle\qquad\psi_{i} =j=0n1(nj1ij)ϕj.\displaystyle=\sum_{j=0}^{n-1}\binom{n-j-1}{i-j}\phi_{j}.
Proof.

The first equality in (4.2) is the usual well-known description of ϕi\phi_{i}. The second equality in (4.2) is (3.3), which can alternatively be obtained by first writing

γi=j=0n1γi,εj(n1j)ϕj,\gamma_{i}=\sum_{j=0}^{n-1}\frac{\langle\gamma_{i},\varepsilon_{j}\rangle}{\binom{n-1}{j}}\phi_{j},

where εj=χ(nj,1,,1)\varepsilon_{j}=\chi_{(n-j,1,\ldots,1)}, and then using that for any partition λ\lambda of nn,

(4.5) γi,χλ=bi+1+c(b)h(b),\langle\gamma_{i},\chi_{\lambda}\rangle=\prod_{b}\frac{i+1+c(b)}{h(b)},

where the product is taken over all boxes bb in the diagram of λ\lambda and where c(b)c(b) and h(b)h(b) denote the content and hook length of bb, which gives

(4.6) γi,εj(n1j)=(n+ijij).\frac{\langle\gamma_{i},\varepsilon_{j}\rangle}{\binom{n-1}{j}}=\binom{n+i-j}{i-j}.

The first equality in (4.3) is the definition of ψi\psi_{i}, and the second equality in (4.3) follows from the first by using the fact that the inverse of the matrix [(ji)]1i,jn[\binom{j}{i}]_{1\leq i,j\leq n} equals [(1)ij(ji)]1i,jn[(-1)^{i-j}\binom{j}{i}]_{1\leq i,j\leq n}.

For the second equality in (4.4), by combining the first relation in (4.3) and the second relation in (4.2), we have

(4.7) ψi=j=0n1u=0n1(1)iu(n+ujuj)(i+1iu)ϕj.\psi_{i}=\sum_{j=0}^{n-1}\sum_{u=0}^{n-1}(-1)^{i-u}\binom{n+u-j}{u-j}\binom{i+1}{i-u}\phi_{j}.

For 1s,tn1\leq s,t\leq n,

u=1n(nsnu)(tu)=(ns+tn),\sum_{u=1}^{n}\binom{n-s}{n-u}\binom{t}{u}=\binom{n-s+t}{n},

so, for 1s,tn1\leq s,t\leq n,

(4.8) u=1n(1)ut(ns+un)(tu)=(nsnt).\sum_{u=1}^{n}(-1)^{u-t}\binom{n-s+u}{n}\binom{t}{u}=\binom{n-s}{n-t}.

Using (4.8) in (4.7), we conclude that

ψi=j=0n1(nj1ij)ϕj.\psi_{i}=\sum_{j=0}^{n-1}\binom{n-j-1}{i-j}\phi_{j}.

The first equality in (4.4) follows from the second equality in (4.4) by using both that the inverse of [(ji)]0i,jn1[\binom{j}{i}]_{0\leq i,j\leq n-1} equals [(1)ij(ji)]0i,jn1[(-1)^{i-j}\binom{j}{i}]_{0\leq i,j\leq n-1} and that matrix inversion commutes with the operation of transposing across the anti-diagonal, i.e. AJAtJA\mapsto JA^{t}J with J=[δi+j,n1]0i,jn1J=[\delta_{i+j,n-1}]_{0\leq i,j\leq n-1}. ∎

Theorem 4.6.

The sequences {ϕi}i=0n1,{γi}i=0n1,{ψi}i=0n1\{\phi_{i}\}_{i=0}^{n-1},\ \{\gamma_{i}\}_{i=0}^{n-1},\ \{\psi_{i}\}_{i=0}^{n-1} are bases for 𝒳\mathcal{X}.

Proof.

The sequence of ϕi\phi_{i}’s is a basis for 𝒳\mathcal{X}, so by Proposition 4.5, the sequence of γi\gamma_{i}’s and the sequence of ψi\psi_{i}’s are also bases for 𝒳\mathcal{X}. ∎

Proposition 4.7.

The Frobenius characteristic ch(ψk)\mathrm{ch}(\psi_{k}) of the character ψk\psi_{k} of SnS_{n} satisfies

(4.9) ch(ψk)=λn(λ)=k+1𝖬(λ)hλ.\mathrm{ch}(\psi_{k})=\sum_{{\lambda\vdash n}\atop\ell(\lambda)=k+1}{\sf M}(\lambda)h_{\lambda}.
Proof.

By [8, Cor. 8],

(4.10) ch(γj)=λn(j+1(λ))M(λ)hλ,\mathrm{ch}(\gamma_{j})=\sum_{\lambda\vdash n}\binom{j+1}{\ell(\lambda)}M(\lambda)h_{\lambda},

so

ch(ψk)\displaystyle\mathrm{ch}(\psi_{k}) =j=0n1(1)kj(k+1j+1)λn(j+1(λ))𝖬(λ)hλ\displaystyle=\sum_{j=0}^{n-1}(-1)^{k-j}\binom{k+1}{j+1}\sum_{\lambda\vdash n}\binom{j+1}{\ell(\lambda)}{\sf M}(\lambda)h_{\lambda}
=λn𝖬(λ)hλj=1n(1)k+1j(k+1j)(j(λ))\displaystyle=\sum_{\lambda\vdash n}{\sf M}(\lambda)h_{\lambda}\sum_{j=1}^{n}(-1)^{k+1-j}\binom{k+1}{j}\binom{j}{\ell(\lambda)}
=λn(λ)=k+1𝖬(λ)hλ.\displaystyle=\sum_{{\lambda\vdash n}\atop{\ell(\lambda)=k+1}}{\sf M}(\lambda)h_{\lambda}.

4.2.2.

Next, we show that certain scalar multiples of the ψi\psi_{i} basis for 𝒳\mathcal{X} give a basis for 𝒴\mathcal{Y} and that this basis for 𝒴\mathcal{Y} can be extended to a basis for 𝒵\mathcal{Z}.

Definition 4.8.

Define ω0,ω1,,ωn1CF(Sn)\omega_{0},\omega_{1},\ldots,\omega_{n-1}\in\mathrm{CF}_{\ell}(S_{n}) by

(4.11) ωk=1dk+1ψk,dk=kgcd(n,k).\omega_{k}=\frac{1}{d_{k+1}}\psi_{k},\quad d_{k}=\frac{k}{\gcd(n,k)}.
Proposition 4.9.

ω0,ω1,,ωn1\omega_{0},\omega_{1},\ldots,\omega_{n-1} are linearly independent characters in \mathcal{L}.

Proof.

The ωi\omega_{i}’s are linearly independent because the ψi\psi_{i}’s are linearly independent, and they are characters by Proposition 2.1 and Proposition 4.7. ∎

Theorem 4.10.

ω0,ω1,,ωn1\omega_{0},\omega_{1},\ldots,\omega_{n-1} can be extended to a basis of 𝒵\mathcal{Z}.

Proof.

For 0kn10\leq k\leq n-1, let

Bk={ch1(hλ)λn,(λ)=k+1},B_{k}=\{\mathrm{ch}^{-1}(h_{\lambda})\mid\lambda\vdash n,\ \ell(\lambda)=k+1\},

so B0,B1,,Bn1B_{0},B_{1},\ldots,B_{n-1} are pairwise disjoint and B=k=0n1BkB=\bigcup_{k=0}^{n-1}B_{k} is a basis for 𝒵\mathcal{Z}.

By Proposition 2.1 and Proposition 4.7, for 0kn10\leq k\leq n-1,

(4.12) ωk=ξBkaξξ\omega_{k}=\sum_{\xi\in B_{k}}a_{\xi}\xi

with coefficients aξa_{\xi} that are positive integers satisfying

(4.13) gcd{aξξBk}=1.\gcd\{a_{\xi}\mid\xi\in B_{k}\}=1.

Now, if SS is any non-empty subset of BB, and if χ\chi is any character of shape

χ=ξSaξξ\chi=\sum_{\xi\in S}a_{\xi}\xi

with positive integer coefficients aξa_{\xi} that satisfy gcd{aξξS}=1,\gcd\{a_{\xi}\mid\xi\in S\}=1, then

(4.14) χ\chi can be extended to a basis of ξSξ\textstyle{\bigoplus}_{\xi\in S}\mathbb{Z}\xi,

which we establish by induction on |S||S| as follows. The |S|=1|S|=1 case is trivial. Assuming the statement for all subsets of BB with exactly kk elements, suppose χ\chi is of the described shape as a sum over a subset SBS\subset B with |S|=k+1|S|=k+1. Let T=S{η}T=S\smallsetminus\{\eta\} for some ηS\eta\in S and write

χ=aη+bθ,\chi=a\eta+b\theta,

where aa is the coefficient of η\eta in χ\chi and b=gcd{coeff. of ξ in χξT}b=\gcd\{\text{coeff.\ of $\xi$ in $\chi$}\mid\xi\in T\}, so gcd(a,b)=1\gcd(a,b)=1. Writing au+bv=1au+bv=1 for some u,vu,v\in\mathbb{Z}, let ψ=uθvη\psi=u\theta-v\eta. Then

η=uχbψandθ=vχ+aψ,\eta=u\chi-b\psi\quad\text{and}\quad\theta=v\chi+a\psi,

so the pair χ,ψ\chi,\psi is a basis for ηθ\mathbb{Z}\eta\oplus\mathbb{Z}\theta. By hypothesis, θ\theta can be extended to a basis of ξTξ\bigoplus_{\xi\in T}\mathbb{Z}\xi. Therefore, χ\chi can be extended to a basis of ξSξ\bigoplus_{\xi\in S}\mathbb{Z}\xi.

By (4.12), (4.13), and (4.14), for 0kn10\leq k\leq n-1, ωk\omega_{k} can be extended to a basis of ξBkξ\bigoplus_{\xi\in B_{k}}\mathbb{Z}\xi. So ω0,ω1,,ωn1\omega_{0},\omega_{1},\ldots,\omega_{n-1} can be extended to a basis of 𝒵\mathcal{Z}. ∎

Theorem 4.11.

ω0,ω1,,ωn1\omega_{0},\omega_{1},\ldots,\omega_{n-1} is a basis for 𝒴\mathcal{Y}.

Proof.

By Proposition 4.9, the ωi\omega_{i}’s are linearly independent, so it remains to show that they span 𝒴\mathcal{Y}. By Theorem 4.6, the ϕi\phi_{i}’s and the ψi\psi_{i}’s are bases of 𝒳\mathcal{X}, so they have the same rational span, which, by Definition 4.8, must be the rational span of the ωi\omega_{i}’s. So by Proposition 4.1,

(4.15) 𝒴=𝒵{r0ω0+r1ω1++rn1ωn1ri}.\mathcal{Y}=\mathcal{Z}\cap\{r_{0}\omega_{0}+r_{1}\omega_{1}+\ldots+r_{n-1}\omega_{n-1}\mid r_{i}\in\mathbb{Q}\}.

By Theorem 4.10, the right-hand side of (4.15) is the set of integer linear combinations of ω0,ω1,,ωn1\omega_{0},\omega_{1},\ldots,\omega_{n-1}. So ω0,ω1,,ωn1\omega_{0},\omega_{1},\ldots,\omega_{n-1} is a basis of 𝒴\mathcal{Y}. ∎

4.3. Proofs of Theorems 1.21.4

From Theorem 4.6 and Theorem 4.11 we will obtain an explicit description of the characters of SnS_{n} that lie in the fundamental parallelepiped 𝒫\mathcal{P}. We start by reading off the number of these characters.

Theorem 4.12.
𝒴/𝒳/d1×/d2××/dn,dk=kgcd(n,k).\mathcal{Y}/\mathcal{X}\cong\mathbb{Z}/d_{1}\mathbb{Z}\times\mathbb{Z}/d_{2}\mathbb{Z}\times\ldots\times\mathbb{Z}/d_{n}\mathbb{Z},\quad d_{k}=\frac{k}{\gcd(n,k)}.

In particular,

[𝒴:𝒳]=|𝒵𝒫|=d1d2dn=n!gcd(1,n)gcd(2,n)gcd(n,n).[\mathcal{Y}:\mathcal{X}]=|\mathcal{Z}\cap\mathcal{P}|=d_{1}d_{2}\ldots d_{n}=\frac{n!}{\gcd(1,n)\gcd(2,n)\ldots\gcd(n,n)}.
Proof of Theorem 4.12 and Theorem 1.3.

By Theorem 4.6, Theorem 4.11, and the relation ψk=dk+1ωk\psi_{k}=d_{k+1}\omega_{k} for 0kn10\leq k\leq n-1. ∎

Proof of Theorem 1.4.

From Theorem 4.12, we have

σn=lcm(d1,d2,,dn),dk=kgcd(n,k).\sigma_{n}=\mathrm{lcm}(d_{1},d_{2},\ldots,d_{n}),\quad d_{k}=\frac{k}{\gcd(n,k)}.

So

σn\displaystyle\sigma_{n} =1nlcm(ngcd(1,n),2ngcd(2,n),,n2gcd(n,n))\displaystyle=\frac{1}{n}\mathrm{lcm}\left(\frac{n}{\gcd(1,n)},\frac{2n}{\gcd(2,n)},\ldots,\frac{n^{2}}{\gcd(n,n)}\right)
=1nlcm(lcm(1,n),lcm(2,n),,lcm(n,n))\displaystyle=\frac{1}{n}\mathrm{lcm}(\mathrm{lcm}(1,n),\mathrm{lcm}(2,n),\ldots,\mathrm{lcm}(n,n))
=lcm(1,2,,n)n.\displaystyle=\frac{\mathrm{lcm}(1,2,\ldots,n)}{n}.

Theorem 4.13.

The elements of 𝒵𝒫\mathcal{Z}\cap\mathcal{P} are

θ𝔞=𝔞~0ϕ0+𝔞~1ϕ1++𝔞~n1ϕn1,𝔞n,0𝔞k<dk+1,\theta_{\mathfrak{a}}=\tilde{\mathfrak{a}}_{0}\phi_{0}+\tilde{\mathfrak{a}}_{1}\phi_{1}+\ldots+\tilde{\mathfrak{a}}_{n-1}\phi_{n-1},\quad\mathfrak{a}\in\mathbb{N}^{n},\quad 0\leq\mathfrak{a}_{k}<d_{k+1},

where

𝔞~k={j=0n1(nk1jk)𝔞jdj+1},dk=kgcd(n,k).\tilde{\mathfrak{a}}_{k}=\left\{\sum_{j=0}^{n-1}\binom{n-k-1}{j-k}\frac{\mathfrak{a}_{j}}{d_{j+1}}\right\},\quad d_{k}=\frac{k}{\gcd(n,k)}.
Proof.

Let

𝒜={(𝔞0,𝔞1,,𝔞n1)tn0𝔞k<dk+1}.\mathcal{A}=\{(\mathfrak{a}_{0},\mathfrak{a}_{1},\ldots,\mathfrak{a}_{n-1})^{t}\in\mathbb{N}^{n}\mid 0\leq\mathfrak{a}_{k}<d_{k+1}\}.

Then by Theorem 4.6, Theorem 4.11, and the relation ψk=dk+1ωk\psi_{k}=d_{k+1}\omega_{k}, the characters

ξ𝔞=𝔞0ω0+𝔞1ω1++𝔞n1ωn1,𝔞𝒜,\xi_{\mathfrak{a}}=\mathfrak{a}_{0}\omega_{0}+\mathfrak{a}_{1}\omega_{1}+\ldots+\mathfrak{a}_{n-1}\omega_{n-1},\quad\mathfrak{a}\in\mathcal{A},

form a complete set of pairwise distinct representatives for the cosets in 𝒴/𝒳\mathcal{Y}/\mathcal{X}. Using the second equality in (4.4),

ξ𝔞=𝔞^0ϕ0+𝔞^1ϕ1++𝔞^n1ϕn1,𝔞𝒜,\xi_{\mathfrak{a}}=\hat{\mathfrak{a}}_{0}\phi_{0}+\hat{\mathfrak{a}}_{1}\phi_{1}+\ldots+\hat{\mathfrak{a}}_{n-1}\phi_{n-1},\quad\mathfrak{a}\in\mathcal{A},

where

𝔞^k=j=0n1(nk1jk)𝔞jdj+1.\hat{\mathfrak{a}}_{k}=\sum_{j=0}^{n-1}\binom{n-k-1}{j-k}\frac{\mathfrak{a}_{j}}{d_{j+1}}.

So

ξ𝔞+𝒳=θ𝔞+𝒳,𝔞𝒜,\xi_{\mathfrak{a}}+\mathcal{X}=\theta_{\mathfrak{a}}+\mathcal{X},\quad\mathfrak{a}\in\mathcal{A},

and hence the θ𝔞\theta_{\mathfrak{a}} with 𝔞𝒜\mathfrak{a}\in\mathcal{A} also form a complete set of pairwise distinct representatives for the cosets in 𝒴/𝒳\mathcal{Y}/\mathcal{X}. Since θ𝔞𝒫\theta_{\mathfrak{a}}\in\mathcal{P} for each 𝔞𝒜\mathfrak{a}\in\mathcal{A}, we conclude that the θ𝔞\theta_{\mathfrak{a}}’s are precisely the pairwise distinct elements of 𝒵𝒫{\mathcal{Z}\cap\mathcal{P}}. ∎

Proof of Theorem 1.2.

By Theorem 4.13 and Proposition 4.3. ∎

References

  • [1] P. Diaconis, Group Representations in Probability and Statistics, Lecture Notes–Monograph Series 11 (1988).
  • [2] P. Diaconis and J. Fulman, Foulkes characters, Eulerian idempotents, and an amazing matrix. J. Algebr. Comb. 36 (2012) 425–440.
  • [3] P. Diaconis and J. Fulman, Combinatorics of balanced carries. Adv. in Appl. Math. 59 (2014) 8–25.
  • [4] L. E. Dickson, History of the Theory of Numbers, vol. I: Divisibility and Primality, Carnegie Institution of Washington, Washington, 1919.
  • [5] A. Gnedin, V. Gorin, S. Kerov, Block characters of the symmetric groups. J. Algebr. Comb. 38 (2013) 79–101.
  • [6] A. Kerber and K.-J. Thürlings, Eulerian numbers, Foulkes characters and Lefschetz characters of SnS_{n}. Sém. Lothar. 8 (1984) 31–36.
  • [7] A. R. Miller, Foulkes characters for complex reflection groups. Proc. Amer. Math. Soc. 143 (2015) 3281–3293.
  • [8] A. R. Miller, Some characters that depend only on length. Math. Res. Lett. 24 (2017) 879–891.
  • [9] A. R. Miller, Walls in Milnor fiber complexes. Doc. Math. 23 (2018) 1247–1261.
  • [10] A. R. Miller, Milnor fiber complexes and some representations, in “Topology of Arrangements and Representation Stability”, pp. 43–123, Oberwolfach reports 15, issue 1, 2018.
  • [11] A. R. Miller, On Foulkes characters, Math. Ann. 381 (2021) 1589–1614.