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

log-Coulomb gases in the projective line of a pp-field

Joe Webster
Abstract

This article extends recent results on log-Coulomb gases in a pp-field KK (i.e., a nonarchimedean local field) to those in its projective line 1(K)\mathbb{P}^{1}(K), where the latter is endowed with the PGL2PGL_{2}-invariant Borel probability measure and spherical metric. Our first main result is an explicit combinatorial formula for the canonical partition function of log-Coulomb gases in 1(K)\mathbb{P}^{1}(K) with arbitrary charge values. Our second main result is called the “(q+1)(q+1)th Power Law”, which relates the grand canonical partition functions for one-component gases in 1(K)\mathbb{P}^{1}(K) (where all particles have charge 1) to those in the open and closed unit balls of KK in a simple way. The final result is a quadratic recurrence for the canonical partition functions for one-component gases in both unit balls of KK and in 1(K)\mathbb{P}^{1}(K). In addition to efficient computation of the canonical partition functions, the recurrence provides their “q1q\to 1” limits and “qq1q\mapsto q^{-1}” functional equations.

footnotetext: Keywords: nonarchimedean local field, projective line, log-Coulomb gas, canonical partition function
Mathematics subject classification 2020: 05A18, 11S40, 12J25, 32A99, 82D05

1 Introduction

1.1 Canonical partition functions for log-Coulomb gases

Let XX be a topological space endowed with a metric dd and a finite positive Borel measure μ\mu satisfying μN{(x1,,xN)XN:xi=xj for some ij}=0\mu^{N}\{(x_{1},\dots,x_{N})\in X^{N}:x_{i}=x_{j}\text{ for some $i\neq j$}\}=0 for every N1N\geq 1. A log-Coulomb gas in XX is a statistical model described as follows: Consider N1N\geq 1 particles with fixed charge values 𝔮1,,𝔮N\mathfrak{q}_{1},\dots,\mathfrak{q}_{N}\in\mathbb{R} and corresponding variable locations x1,,xNXx_{1},\dots,x_{N}\in X. Whether the charge values are distinct or not, we assume particles are distinguished by the labels 1,2,,N1,2,\dots,N, so that unique configurations of the system correspond to unique tuples (x1,,xN)XN(x_{1},\dots,x_{N})\in X^{N}. Each tuple is called a microstate and has an energy defined by

E(x1,,xN):={1i<jN𝔮i𝔮𝔧logd(xi,xj)if xixj for all i<j,otherwise.E(x_{1},\dots,x_{N}):=\begin{cases}-\sum_{1\leq i<j\leq N}\mathfrak{q}_{i}\mathfrak{q_{j}}\log d(x_{i},x_{j})&\text{if $x_{i}\neq x_{j}$ for all $i<j$},\\ \infty&\text{otherwise}.\end{cases}

Note that E1()E^{-1}(\infty) has measure 0 in XNX^{N} by our choice of μ\mu, and that EE is identically zero if N=1N=1. We assume the system is in thermal equilibrium with a heat reservoir at some inverse temperature β>0\beta>0, so that its microstates form a canonical ensemble distributed according to the density eβE(x1,,xN)e^{-\beta E(x_{1},\dots,x_{N})}. The canonical partition function βZN(X,β)\beta\mapsto Z_{N}(X,\beta) is defined as the total mass of this density, namely

ZN(X,β):=XNeβE(x1,,xN)𝑑μN=XN1i<jNd(xi,xj)𝔮i𝔮jβdμN.Z_{N}(X,\beta):=\int_{X^{N}}e^{-\beta E(x_{1},\dots,x_{N})}\,d\mu^{N}=\int_{X^{N}}\prod_{1\leq i<j\leq N}d(x_{i},x_{j})^{\mathfrak{q}_{i}\mathfrak{q}_{j}\beta}\,d\mu^{N}~{}. (1.1.1)

Given (X,d,μ)(X,d,\mu) and 𝔮1,,𝔮N\mathfrak{q}_{1},\dots,\mathfrak{q}_{N}, the explicit formula for ZN(X,β)Z_{N}(X,\beta) is of primary interest, as it yields fundamental relationships between the observable parameters of the system and its temperature. For instance, the system’s dimensionless free energy, mean energy, and energy fluctuation (variance) are respectively given by logZN(X,β)-\log Z_{N}(X,\beta), /βlogZN(X,β)-\partial/\partial\beta\log Z_{N}(X,\beta), and 2/β2logZN(X,β)\partial^{2}/\partial\beta^{2}\log Z_{N}(X,\beta), all of which are functions of β\beta (and hence of temperature). Below are two examples in which the formula for ZN(X,β)Z_{N}(X,\beta) is known.

Example 1.1.

If X=X=\mathbb{R} with the standard metric dd, the standard Gaussian measure μ\mu, and charge values 𝔮1==𝔮N=1\mathfrak{q}_{1}=\dots=\mathfrak{q}_{N}=1, then Mehta’s integral formula [FW08] states that ZN(,β)Z_{N}(\mathbb{R},\beta) converges absolutely for all complex β\beta with Re(β)>2/N\operatorname{Re}(\beta)>-2/N, and in this case

ZN(,β)=j=1NΓ(1+jβ/2)Γ(1+β/2).Z_{N}(\mathbb{R},\beta)=\prod_{j=1}^{N}\frac{\Gamma(1+j\beta/2)}{\Gamma(1+\beta/2)}~{}.

At the special values β{1,2,4}\beta\in\{1,2,4\}, the probability density 1ZN(,β)eβE(x1,,xN)\frac{1}{Z_{N}(\mathbb{R},\beta)}e^{-\beta E(x_{1},\dots,x_{N})} coincides with the joint density of the NN eigenvalues x1,,xNx_{1},\dots,x_{N} (counted with multiplicity) of the N×NN\times N Gaussian orthogonal (β=1)(\beta=1), unitary (β=2)(\beta=2), and symplectic (β=4)(\beta=4) matrix ensembles.

Example 1.2.

If X=pX=\mathbb{Z}_{p} with d(x,y)=|xy|pd(x,y)=|x-y|_{p}, the Haar probability measure μ\mu, and N=3N=3 charges with 𝔮1=1\mathfrak{q}_{1}=1, 𝔮2=2\mathfrak{q}_{2}=2, and 𝔮3=3\mathfrak{q}_{3}=3, then a general theorem in [Web21b] implies that Z3(p,β)Z_{3}(\mathbb{Z}_{p},\beta) converges absolutely for all complex β\beta with Re(β)>1/6\operatorname{Re}(\beta)>-1/6, and in this case

Z3(p,β)=p11βp2+11β1((p1)(p2)+(p1)2[1p1+2β1+1p1+3β1+1p1+6β1]).Z_{3}(\mathbb{Z}_{p},\beta)=\frac{p^{11\beta}}{p^{2+11\beta}-1}\cdot\left((p-1)(p-2)+(p-1)^{2}\left[\frac{1}{p^{1+2\beta}-1}+\frac{1}{p^{1+3\beta}-1}+\frac{1}{p^{1+6\beta}-1}\right]\right)~{}.

Note that βZN(X,β)\beta\mapsto Z_{N}(X,\beta) always extends to a complex domain containing the line Re(β)=0\operatorname{Re}(\beta)=0. To simultaneously treat all possible choices of 𝔮i\mathfrak{q}_{i}, we extend further to a subset of N(N1)/2\mathbb{C}^{N(N-1)/2} as follows:

Definition 1.3.

For any N0N\geq 0 and (X,d,μ)(X,d,\mu) as above, we write 𝒔\bm{s} for a complex tuple (sij)1i<jN(s_{ij})_{1\leq i<j\leq N} (the empty tuple if N=0N=0 or N=1N=1) and define 𝒵0(X,𝒔):=1\mathcal{Z}_{0}(X,\bm{s}):=1 and

𝒵N(X,𝒔):=XN1i<jNd(xi,xj)sijdμNfor N1.\mathcal{Z}_{N}(X,\bm{s}):=\int_{X^{N}}\prod_{1\leq i<j\leq N}d(x_{i},x_{j})^{s_{ij}}\,d\mu^{N}\qquad\text{for $N\geq 1$.}

Once the formula and domain for 𝒔𝒵N(X,𝒔)\bm{s}\mapsto\mathcal{Z}_{N}(X,\bm{s}) are known, then for any choice of 𝔮1,,𝔮N\mathfrak{q}_{1},\dots,\mathfrak{q}_{N}\in\mathbb{R}, the formula and domain for βZN(X,β)\beta\mapsto Z_{N}(X,\beta) follow by specializing 𝒵N(X,𝒔)\mathcal{Z}_{N}(X,\bm{s}) to sij=𝔮i𝔮jβ{s_{ij}=\mathfrak{q}_{i}\mathfrak{q}_{j}\beta}. Thus, given (X,d,μ)(X,d,\mu), the main problem is to determine the formula and domain of 𝒔𝒵N(X,𝒔)\bm{s}\mapsto\mathcal{Z}_{N}(X,\bm{s}).

1.2 pp-fields, projective lines, and splitting chains

Of our two main goals in this paper, the first is to determine the explicit formula and domain of 𝒔𝒵N(1(K),𝒔)\bm{s}\mapsto\mathcal{Z}_{N}(\mathbb{P}^{1}(K),\bm{s}), where KK is a pp-field and its projective line 1(K)\mathbb{P}^{1}(K) is endowed with a natural metric and measure. To make this precise, we briefly recall well-known properties of pp-fields (see [Wei95], for instance) and establish some notation. Fix a pp-field KK, write |||\cdot| for its canonical absolute value, write dd for the associated metric (i.e., d(x,y):=|xy|d(x,y):=|x-y|), and define

R:={xK:|x|1}andP:={xK:|x|<1}.R:=\{x\in K:|x|\leq 1\}\qquad\text{and}\qquad P:=\{x\in K:|x|<1\}~{}.

The closed unit ball RR is the maximal compact subring of KK, the open unit ball PP is the unique maximal ideal in RR, and the group of units is R×=RP={xK:|x|=1}R^{\times}=R\setminus P=\{x\in K:|x|=1\}. The residue field R/PR/P is isomorphic to 𝔽q\mathbb{F}_{q} for some prime power qq, and there is a canonical isomorphism of the cyclic group (R/P)×(R/P)^{\times} onto the group of (q1)(q-1)th roots of unity in KK. Fixing a primitive such root ξK\xi\in K and sending P0P\mapsto 0 extends the isomorphism (R/P)×{1,ξ,,ξq2}(R/P)^{\times}\to\{1,\xi,\dots,\xi^{q-2}\} to a bijection R/P{0,1,ξ,,ξq2}R/P\to\{0,1,\xi,\dots,\xi^{q-2}\} with inverse yy+Py\mapsto y+P. Therefore {0,1,ξ,,ξq2}\{0,1,\xi,\dots,\xi^{q-2}\} is a complete set of representatives for the cosets of PRP\subset R, i.e.,

R=P(1+P)(ξ+P)(ξq2+P)R×.R=P\sqcup\underbrace{(1+P)\sqcup(\xi+P)\sqcup\dots\sqcup(\xi^{q-2}+P)}_{R^{\times}}~{}. (1.2.1)

Fix a uniformizer πK\pi\in K (any element satisfying P=πRP=\pi R) and let μ\mu be the unique additive Haar measure on KK satisfying μ(R)=1\mu(R)=1. The open balls in KK are precisely the sets of the form y+πvRy+\pi^{v}R with yKy\in K and vv\in\mathbb{Z}, and every such ball is compact with measure equal to its radius, i.e., μ(y+πvR)=|πv|=qv\mu(y+\pi^{v}R)=|\pi^{v}|=q^{-v}. In particular, each of the qq cosets of PP in (1.2.1) is a compact open ball with measure (and radius) q1q^{-1}, and two elements x,yRx,y\in R satisfy |xy|=1|x-y|=1 if and only if they belong to different cosets. Henceforth, we reserve the symbols KK, |||\cdot|, dd, RR, PP, qq, ξ\xi, π\pi, and μ\mu for the items above, and we distinguish |||\cdot| from the standard absolute value on \mathbb{C} by writing |||\cdot|_{\mathbb{C}} for the latter.

We now recall some useful facts from [FP15] in our present notation. The projective line of KK is the quotient space 1(K)=(K2{(0,0)})/\mathbb{P}^{1}(K)=(K^{2}\setminus\{(0,0)\})/\sim, where (x0,x1)(y0,y1)(x_{0},x_{1})\sim(y_{0},y_{1}) if and only if y0=λx0y_{0}=\lambda x_{0} and y1=λx1y_{1}=\lambda x_{1} for some λK×\lambda\in K^{\times}. Thus we may understand 1(K)\mathbb{P}^{1}(K) concretely as the set of symbols [x0:x1][x_{0}:x_{1}] with (x0,x1)K2{(0,0)}(x_{0},x_{1})\in K^{2}\setminus\{(0,0)\}, subject to the relation [λx0:λx1]=[x0:x1][\lambda x_{0}:\lambda x_{1}]=[x_{0}:x_{1}] for all λK×\lambda\in K^{\times} and endowed with the topology induced by the quotient (x0,x1)[x0:x1](x_{0},x_{1})\mapsto[x_{0}:x_{1}]. The projective line is compact and metrizable by the spherical metric δ:1(K)×1(K){0}{qv:v0}\delta:\mathbb{P}^{1}(K)\times\mathbb{P}^{1}(K)\to\{0\}\cup\{q^{-v}:v\in\mathbb{Z}_{\geq 0}\}, which is defined via

δ([x0:x1],[y0:y1]):=|x0y1x1y0|max{|x0|,|x1|}max{|y0|,|y1|}.\delta([x_{0}:x_{1}],[y_{0}:y_{1}]):=\frac{|x_{0}y_{1}-x_{1}y_{0}|}{\max\{|x_{0}|,|x_{1}|\}\cdot\max\{|y_{0}|,|y_{1}|\}}~{}. (1.2.2)

In particular, every open set in 1(K)\mathbb{P}^{1}(K) is a union of balls of the form

Bv[x0:x1]:={[y0:y1]1(K):δ([x0:x1],[y0:y1])qv}B_{v}[x_{0}:x_{1}]:=\{[y_{0}:y_{1}]\in\mathbb{P}^{1}(K):\delta([x_{0}:x_{1}],[y_{0}:y_{1}])\leq q^{-v}\} (1.2.3)

with [x0:x1]1(K)[x_{0}:x_{1}]\in\mathbb{P}^{1}(K) and v0v\in\mathbb{Z}_{\geq 0}, and every such ball is open and compact. The projective linear group PGL2(R)PGL_{2}(R) is the quotient of GL2(R)={AM2(R):det(A)R×}GL_{2}(R)=\{A\in M_{2}(R):\det(A)\in R^{\times}\} by its center, namely Z={(λ00λ):λR×}R×Z=\{\left(\begin{smallmatrix}\lambda&0\\ 0&\lambda\end{smallmatrix}\right):\lambda\in R^{\times}\}\cong R^{\times}. It is straightforward to verify that the rule

ϕ[x0:x1]:=[ax0+bx1:cx0+dx1],\phi[x_{0}:x_{1}]:=[ax_{0}+bx_{1}:cx_{0}+dx_{1}]~{},

where ϕPGL2(R)\phi\in PGL_{2}(R) and (abcd)GL2(R)\left(\begin{smallmatrix}a&b\\ c&d\end{smallmatrix}\right)\in GL_{2}(R) is any representative of ϕ\phi, gives a well-defined transitive action of PGL2(R)PGL_{2}(R) on 1(K)\mathbb{P}^{1}(K).

Lemma 1.4 (PGL2(R)PGL_{2}(R)-invariance [FP15]).

The spherical metric satisfies

δ(ϕ[x0:x1],ϕ[y0:y1])=δ([x0:x1],[y0:y1])\delta(\phi[x_{0}:x_{1}],\phi[y_{0}:y_{1}])=\delta([x_{0}:x_{1}],[y_{0}:y_{1}])

for all ϕPGL2(R)\phi\in PGL_{2}(R) and all [x0:x1],[y0:y1]1(K)[x_{0}:x_{1}],[y_{0}:y_{1}]\in\mathbb{P}^{1}(K). There is also a unique Borel probability measure ν\nu on 1(K)\mathbb{P}^{1}(K) satisfying

ν(ϕ(M))=ν(M)\nu(\phi(M))=\nu(M)

for all ϕPGL2(R)\phi\in PGL_{2}(R) and all Borel subsets M1(K)M\subset\mathbb{P}^{1}(K). In particular, for each v0v\in\mathbb{Z}_{\geq 0} the relation ϕ(Bv[x0:x1])=Bv(ϕ[x0:x1])\phi(B_{v}[x_{0}:x_{1}])=B_{v}(\phi[x_{0}:x_{1}]) defines a transitive PGL2(R)PGL_{2}(R) action on the set of balls of radius qvq^{-v}, and thus ν(Bv[x0:x1])\nu(B_{v}[x_{0}:x_{1}]) depends only on vv.

It is routine to verify that μN{(x1,,xN)(y+πvR)N:xi=xj for some ij}=0\mu^{N}\{(x_{1},\dots,x_{N})\in(y+\pi^{v}R)^{N}:x_{i}=x_{j}\text{ for some $i\neq j$}\}=0 and

νN{([x1,0:x1,1],,[xN,0:xN,1])(1(K))N:[xi,0:xi,1]=[xj,0:xj,1] for some ij}=0\nu^{N}\{([x_{1,0}:x_{1,1}],\dots,[x_{N,0}:x_{N,1}])\in(\mathbb{P}^{1}(K))^{N}:[x_{i,0}:x_{i,1}]=[x_{j,0}:x_{j,1}]\text{ for some $i\neq j$}\}=0

for all N1N\geq 1, so we have suitable metrics and measures to define log-Coulomb gases in X=y+πvRX=y+\pi^{v}R and X=1(K)X=\mathbb{P}^{1}(K). Definition 1.3 specializes to these XX as follows:

Definition 1.5.

With N0N\geq 0 and 𝒔\bm{s} as before, we have 𝒵0(y+πvR,𝒔)=𝒵0(1(K),𝒔)=1\mathcal{Z}_{0}(y+\pi^{v}R,\bm{s})=\mathcal{Z}_{0}(\mathbb{P}^{1}(K),\bm{s})=1, and

𝒵N(y+πvR,𝒔)\displaystyle\mathcal{Z}_{N}(y+\pi^{v}R,\bm{s}) =(y+πvR)N1i<jN|xixj|sijdμNand\displaystyle=\int_{(y+\pi^{v}R)^{N}}\prod_{1\leq i<j\leq N}|x_{i}-x_{j}|^{s_{ij}}\,d\mu^{N}\qquad\text{and}
𝒵N(1(K),𝒔)\displaystyle\mathcal{Z}_{N}(\mathbb{P}^{1}(K),\bm{s}) =(1(K))N1i<jNδ([xi,0:xi,1],[xj,0:xj,1])sijdνN\displaystyle=\int_{(\mathbb{P}^{1}(K))^{N}}\prod_{1\leq i<j\leq N}\delta([x_{i,0}:x_{i,1}],[x_{j,0}:x_{j,1}])^{s_{ij}}\,d\nu^{N}

for N1N\geq 1. The first integral is independent of yy, and equal to 𝒵N(R,𝒔)\mathcal{Z}_{N}(R,\bm{s}) if v=0v=0 or 𝒵N(P,𝒔)\mathcal{Z}_{N}(P,\bm{s}) if v=1v=1.

The formulas and domains of absolute convergence for the integrals above can be stated neatly in terms of the following items from [Web21b]:

Definition 1.6 (Splitting chains).

A splitting chain of order N2N\geq 2 and length L1L\geq 1 is a tuple =(0,,L){\bm{\pitchfork}}=({\pitchfork}_{0},\dots,{\pitchfork}_{L}) of partitions of [N]={1,,N}[N]=\{1,\dots,N\} satisfying

{[N]}=0>1>2>>L={{1},,{N}}.\{[N]\}={\pitchfork}_{0}>{\pitchfork}_{1}>{\pitchfork}_{2}>\dots>{\pitchfork}_{L}=\{\{1\},\dots,\{N\}\}~{}.
  • (a)

    Each non-singleton part λ01L\lambda\in{\pitchfork}_{0}\cup{\pitchfork}_{1}\cup\dots\cup{\pitchfork}_{L} is called a branch of {\bm{\pitchfork}}. We write ()\mathcal{B}({\bm{\pitchfork}}) for the set of all branches of {\bm{\pitchfork}}, i.e.,

    ():=(0L1)L.\mathcal{B}({\bm{\pitchfork}}):=({\pitchfork}_{0}\cup\dots\cup{\pitchfork}_{L-1})\setminus{\pitchfork}_{L}~{}.
  • (b)

    Each λ()\lambda\in\mathcal{B}({\bm{\pitchfork}}) must appear in a final partition {\pitchfork}_{\ell} before refining into two or more parts in +1{\pitchfork}_{\ell+1}, so we define its depth (λ){0,1,,L1}\ell_{\bm{\pitchfork}}(\lambda)\in\{0,1,\dots,L-1\} and degree deg(λ){2,3,,N}\deg_{\bm{\pitchfork}}(\lambda)\in\{2,3,\dots,N\} by

    (λ):=max{:λ}anddeg(λ):=#{λ(λ)+1:λλ}.\ell_{\bm{\pitchfork}}(\lambda):=\max\{\ell:\lambda\in{\pitchfork}_{\ell}\}\qquad\text{and}\qquad\deg_{\bm{\pitchfork}}(\lambda):=\#\{\lambda^{\prime}\in{\pitchfork}_{\ell_{\bm{\pitchfork}}(\lambda)+1}:\lambda^{\prime}\subset\lambda\}~{}.
  • (c)

    We say that {\bm{\pitchfork}} is reduced if each λ()\lambda\in\mathcal{B}({\bm{\pitchfork}}) satisfies λ=(λ)\lambda\in{\pitchfork}_{\ell}\iff\ell=\ell_{\bm{\pitchfork}}(\lambda).

Write N\mathcal{R}_{N} for the set of reduced splitting chains of order NN and define

ΩN:=λ[N]#λ>1{𝒔:Re(eλ(𝒔))>0}whereeλ(𝒔):=#λ1+i<ji,jλsij.\Omega_{N}:=\bigcap_{\begin{subarray}{c}\lambda\subset[N]\\ \#\lambda>1\end{subarray}}\left\{\bm{s}:\operatorname{Re}(e_{\lambda}(\bm{s}))>0\right\}\qquad\text{where}\qquad e_{\lambda}(\bm{s}):=\#\lambda-1+\sum_{\begin{subarray}{c}i<j\\ i,j\in\lambda\end{subarray}}s_{ij}~{}.

Proposition 3.15 and Theorem 2.6(c) in [Web21b] imply the following proposition, which shall be generalized slightly in order to prove the main results of this paper:

Proposition 1.7.

For N2N\geq 2, the integral 𝒵N(R,𝒔)\mathcal{Z}_{N}(R,\bm{s}) converges absolutely if and only if 𝒔ΩN\bm{s}\in\Omega_{N}, and in this case it can be written as the finite sum

𝒵N(R,𝒔)=qi<jsijNλ()(q1)deg(λ)1qeλ(𝒔)1.\mathcal{Z}_{N}(R,\bm{s})=q^{\sum_{i<j}s_{ij}}\sum_{{\bm{\pitchfork}}\in\mathcal{R}_{N}}\prod_{\lambda\in\mathcal{B}({\bm{\pitchfork}})}\frac{(q-1)_{\deg_{{\bm{\pitchfork}}}(\lambda)-1}}{q^{e_{\lambda}(\bm{s})}-1}~{}.

Here (q1)n(q-1)_{n} stands for the degree nn falling factorial (z)n=z(z1)(zn+1)[z](z)_{n}=z(z-1)\dots(z-n+1)\in\mathbb{Z}[z] evaluated at the integer z=q1z=q-1.

2 Statement of results

2.1 The projective analogue

Our first main result is the following analogue of Proposition 1.7:

Theorem 2.1.

For N2N\geq 2, the integral 𝒵N(1(K),𝒔)\mathcal{Z}_{N}(\mathbb{P}^{1}(K),\bm{s}) converges absolutely if and only if 𝒔ΩN\bm{s}\in\Omega_{N}, and in this case it can be written as the finite sum

𝒵N(1(K),𝒔)=1(q+1)N1NqN+i<jsij+1deg([N])q+1deg([N])λ()(q1)deg(λ)1qeλ(𝒔)1.\mathcal{Z}_{N}(\mathbb{P}^{1}(K),\bm{s})=\frac{1}{(q+1)^{N-1}}\sum_{{\bm{\pitchfork}}\in\mathcal{R}_{N}}\frac{q^{N+\sum_{i<j}s_{ij}}+1-\deg_{{\bm{\pitchfork}}}([N])}{q+1-\deg_{{\bm{\pitchfork}}}([N])}\prod_{\lambda\in\mathcal{B}({\bm{\pitchfork}})}\frac{(q-1)_{\deg_{{\bm{\pitchfork}}}(\lambda)-1}}{q^{e_{\lambda}(\bm{s})}-1}~{}.

The summand for each N{\bm{\pitchfork}}\in\mathcal{R}_{N} is defined for all prime powers qq, as the denominator q+1deg([N])q+1-\deg_{\bm{\pitchfork}}([N]) is cancelled by the factor (q1)deg([N])1(q-1)_{\deg_{\bm{\pitchfork}}([N])-1} inside the product over λ()\lambda\in\mathcal{B}({\bm{\pitchfork}}).

The evident similarities between Proposition 1.7 and Theorem 2.1 follow from explicit relationship between the metrics and measures on KK and those on 1(K)\mathbb{P}^{1}(K). These relationships also play a role in the upcoming results, so they are worth recalling now. Note that [x0:x1][1:0][x_{0}:x_{1}]\neq[1:0] if and only if x10x_{1}\neq 0, in which case x=x0/x1x=x_{0}/x_{1} is the unique element of KK satisfying [x:1]=[x0:x1][x:1]=[x_{0}:x_{1}]. Therefore the rule ι(x):=[x:1]\iota(x):=[x:1] defines a bijection ι:K1(K){[1:0]}\iota:K\to\mathbb{P}^{1}(K)\setminus\{[1:0]\} and relates the metric structures of KK and 1(K)\mathbb{P}^{1}(K) in a simple way: Given x,yKx,y\in K, (1.2.2) implies δ(ι(x),[1:0])=(max{1,|x|})1\delta(\iota(x),[1:0])=(\max\{1,|x|\})^{-1} and

δ(ι(x),ι(y))={|xy|if x,yR,1if xR and yR,|1/x1/y|if x,yR.\delta(\iota(x),\iota(y))=\begin{cases}|x-y|&\text{if $x,y\in R$},\\ 1&\text{if $x\in R$ and $y\notin R$},\\ |1/x-1/y|&\text{if }x,y\notin R.\end{cases} (2.1.1)

Using the definitions (1.2.3) and vK(x):=logq|x|v_{K}(x):=-\log_{q}|x| for xK×x\in K^{\times}, along with (2.1.1) and the strong triangle equality (i.e., |x+y|=max{|x|,|y|}|x+y|=\max\{|x|,|y|\} whenever |x||y||x|\neq|y|), one easily verifies that

ι(y+πvR)={Bv(ι(y))if yR,Bv2vK(y)(ι(y))if yR,\iota(y+\pi^{v}R)=\begin{cases}B_{v}(\iota(y))&\text{if }y\in R,\\ B_{v-2v_{K}(y)}(\iota(y))&\text{if }y\notin R,\end{cases} (2.1.2)

whenever yKy\in K and v>0v\in\mathbb{Z}_{>0}. That is, ι\iota sends the open ball of radius r(0,1)r\in(0,1) centered at yKy\in K onto the open ball of radius r/max{1,|y|2}r/\max\{1,|y|^{2}\} centered at ι(y)1(K){[1:0]}\iota(y)\in\mathbb{P}^{1}(K)\setminus\{[1:0]\}, so ι:K1(K){[1:0]}\iota:K\to\mathbb{P}^{1}(K)\setminus\{[1:0]\} is a homeomorphism that restricts to an isometry on RR and a contraction on KRK\setminus R.

The map ι\iota also relates the measures on KK and 1(K)\mathbb{P}^{1}(K) in a simple way: Given v>0v>0 and a complete set of representatives y1,,yqvRy_{1},\dots,y_{q^{v}}\in R for the cosets of πvRR\pi^{v}R\subset R, applying (2.1.2) to the partition R=(y1+πvR)(yqv+πvR)R=(y_{1}+\pi^{v}R)\sqcup\dots\sqcup(y_{q^{v}}+\pi^{v}R) yields

ι(R)=Bv[y1:1]Bv[yqv:1].\iota(R)=B_{v}[y_{1}:1]\sqcup\cdots\sqcup B_{v}[y_{q^{v}}:1]~{}.

Therefore PGL2(R)PGL_{2}(R)-invariance of ν\nu (Lemma 1.4) implies ν(ι(R))=qvν(Bv[0:1])\nu(\iota(R))=q^{v}\cdot\nu(B_{v}[0:1]). On the other hand,

ι(KR)=ι({x:|x|q})={ι(x):δ(ι(x),[1:0])q1}=B1[1:0]{[1:0]}\iota(K\setminus R)=\iota(\{x:|x|\geq q\})=\{\iota(x):\delta(\iota(x),[1:0])\leq q^{-1}\}=B_{1}[1:0]\setminus\{[1:0]\}

implies ι(R)B1[1:0]=ι(R)ι(KR){[1:0]}=1(K)\iota(R)\sqcup B_{1}[1:0]=\iota(R)\sqcup\iota(K\setminus R)\sqcup\{[1:0]\}=\mathbb{P}^{1}(K), which has measure 1. But ν(ι(R))=qν(B1[1:0])\nu(\iota(R))=q\cdot\nu(B_{1}[1:0]), so the measure of ι(R)\iota(R) must be q/(q+1)q/(q+1), and therefore every ball Bv[x0:x1]1(K)B_{v}[x_{0}:x_{1}]\subset\mathbb{P}^{1}(K) with v>0v>0 has measure qvq/(q+1)q^{-v}\cdot q/(q+1). Combining this with (2.1.2), one concludes that the measure ν\nu on 1(K){[1:0]}\mathbb{P}^{1}(K)\setminus\{[1:0]\} pulls back along ι\iota to an explicit measure on KK:

ν(ι(M))=qq+1M(max{1,|x|2})1𝑑μfor any Borel subset MK.\nu(\iota(M))=\frac{q}{q+1}\int_{M}\left(\max\{1,|x|^{2}\}\right)^{-1}\,d\mu\qquad\text{for any Borel subset }M\subset K~{}. (2.1.3)

Finally, (1.2.1) and (2.1.2) give a nice refinement of 1(K)=ι(R)B1[1:0]\mathbb{P}^{1}(K)=\iota(R)\sqcup B_{1}[1:0] in terms of the (q1)(q-1)th roots of unity in KK, which should be understood as the projective analogue of (1.2.1):

1(K)=B1[0:1]B1[1:1]B1[ξ:1]B1[ξq2:1]ι(R×)B1[1:0].\mathbb{P}^{1}(K)=B_{1}[0:1]\sqcup\underbrace{B_{1}[1:1]\sqcup B_{1}[\xi:1]\sqcup\dots\sqcup B_{1}[\xi^{q-2}:1]}_{\iota(R^{\times})}\sqcup B_{1}[1:0]~{}. (2.1.4)

Indeed, all q+1q+1 of the parts in the partition are balls with measure 1/(q+1)1/(q+1) and radius q1q^{-1}, and two points [x0:x1],[y0:y1]1(K)[x_{0}:x_{1}],[y_{0}:y_{1}]\in\mathbb{P}^{1}(K) satisfy δ([y0:y1],[y0:y1])=1\delta([y_{0}:y_{1}],[y_{0}:y_{1}])=1 if and only if [x0:x1][x_{0}:x_{1}] and [y0:y1][y_{0}:y_{1}] belong to different parts. Note that ι\iota sends R×R^{\times} onto the “equator” ι(R×)\iota(R^{\times}), i.e., the set of points in 1(K)\mathbb{P}^{1}(K) with δ\delta-distance 1 from both the “south pole” [0:1][0:1] and the “north pole” [1:0][1:0].

2.2 Relationships between grand canonical partition functions

So far we have only considered log-Coulomb gases with NN labeled (and hence distinguishable) particles. Our second main result concerns the situation in which all particles are identical with charge 𝔮i=1\mathfrak{q}_{i}=1 for all ii, in which case a microstate (x1,,xN)XN(x_{1},\dots,x_{N})\in X^{N} is regarded as unique only up to permutations of its entries. Since the energy E(x1,,xN)E(x_{1},\dots,x_{N}) and measure on XNX^{N} are invariant under such permutations, each unlabeled microstate makes the contribution eβE(x1,,xN)dμNe^{-\beta E(x_{1},\dots,x_{N})}d\mu^{N} to the integral ZN(X,β)Z_{N}(X,\beta) in (1.1.1) precisely N!N! times. Therefore the canonical partition function for the unlabeled microstates is given by ZN(X,β)/N!Z_{N}(X,\beta)/N!. We further assume that the system exchanges particles with the heat reservoir with chemical potential η\eta and define the fugacity parameter f=eηβf=e^{\eta\beta}. In this situation the particle number N0N\geq 0 is treated as a random variable and the canonical partition function is replaced by the grand canonical partition function

Z(f,X,β):=N=0ZN(X,β)fNN!Z(f,X,\beta):=\sum_{N=0}^{\infty}Z_{N}(X,\beta)\frac{f^{N}}{N!} (2.2.1)

with the familiar convention Z0(X,β)=1Z_{0}(X,\beta)=1. Many properties of the system can be deduced from the grand canonical partition function. For instance, if β>0\beta>0 is fixed and ZN(X,β)Z_{N}(X,\beta) is sub-exponential in NN, then Z(f,X,β)Z(f,X,\beta) is analytic in ff and the expected number of particles in the system is given by fflog(Z(f,X,β))f\frac{\partial}{\partial f}\log(Z(f,X,\beta)). The canonical partition function for each N0N\geq 0 can also be recovered by evaluating the NNth derivative of Z(f,X,β)Z(f,X,\beta) with respect to ff at f=0f=0.

We are interested in the examples Z(f,R,β)Z(f,R,\beta), Z(f,P,β)Z(f,P,\beta), and Z(f,1(K),β)Z(f,\mathbb{P}^{1}(K),\beta), which turn out to share several common properties and simple relationships. By setting sij=βs_{ij}=\beta in Definition 1.5, one sees that |ZN(R,β)||Z_{N}(R,\beta)|_{\mathbb{C}}, |ZN(P,β)||Z_{N}(P,\beta)|_{\mathbb{C}}, and |ZN(1(K),β)||Z_{N}(\mathbb{P}^{1}(K),\beta)|_{\mathbb{C}} are bounded above by 1 for all N0N\geq 0 and all β>0\beta>0, and hence Z(f,R,β)Z(f,R,\beta), Z(f,P,β)Z(f,P,\beta), and Z(f,1(K),β)Z(f,\mathbb{P}^{1}(K),\beta) are analytic in ff when β>0\beta>0. Sinclair recently found an elegant relationship between the first two, which is closely related to the partition of RR in (1.2.1):

Proposition 2.2 (The qqth Power Law [Sin20]).

For β>0\beta>0 we have

Z(f,R,β)=(Z(f,P,β))q.Z(f,R,\beta)=(Z(f,P,\beta))^{q}~{}.

Roughly speaking, the qqth Power Law states that a log-Coulomb gas in RR exchanging energy and particles with a heat reservoir “factors” into qq identical sub-gases (one in each coset of PP) that exchange energy and particles with the reservoir. For β>0\beta>0, note that the series equation Z(f,R,β)=(Z(f,P,β))qZ(f,R,\beta)=(Z(f,P,\beta))^{q} is equivalent to the coefficient identities

ZN(R,β)N!=N0++Nq1=NN0,,Nq10k=0q1ZNk(P,β)Nk!for all N0.\frac{Z_{N}(R,\beta)}{N!}=\sum_{\begin{subarray}{c}N_{0}+\dots+N_{q-1}=N\\ N_{0},\dots,N_{q-1}\geq 0\end{subarray}}\prod_{k=0}^{q-1}\frac{Z_{N_{k}}(P,\beta)}{N_{k}!}\qquad\text{for all $N\geq 0$.} (2.2.2)

The β=1\beta=1 case of (2.2.2) is given in [BGMR06], in which the positive number ZN(R,1)/N!Z_{N}(R,1)/N! is recognized as the probability that a random monic polynomial in R[x]R[x] splits completely in RR. The more general β>0\beta>0 case given in [Sin20] makes explicit use of the partition of RR into cosets of PP (as in (1.2.1)). In Section 3, we will use the analogous partition of 1(K)\mathbb{P}^{1}(K) into q+1q+1 balls (recall (2.1.4)) to show that

ZN(1(K),β)N!=N0++Nq=NN0,,Nq0k=0q(qq+1)NkZNk(P,β)Nk!for all β>0 and N0,\frac{Z_{N}(\mathbb{P}^{1}(K),\beta)}{N!}=\sum_{\begin{subarray}{c}N_{0}+\dots+N_{q}=N\\ N_{0},\dots,N_{q}\geq 0\end{subarray}}\prod_{k=0}^{q}\left(\frac{q}{q+1}\right)^{N_{k}}\frac{Z_{N_{k}}(P,\beta)}{N_{k}!}\qquad\text{for all $\beta>0$ and $N\geq 0$,} (2.2.3)

which immediately implies our second main result:

Theorem 2.3 (The (q+1)(q+1)th Power Law).

For all β>0\beta>0 we have

Z(f,1(K),β)=(Z(qfq+1,P,β))q+1.Z(f,\mathbb{P}^{1}(K),\beta)=(Z(\tfrac{qf}{q+1},P,\beta))^{q+1}~{}.

Like the qqth Power Law, the (q+1)(q+1)th Power Law roughly states that a log-Coulomb gas in 1(K)\mathbb{P}^{1}(K) exchanging energy and particles with a heat reservoir “factors” into q+1q+1 identical sub-gases in the balls B1[0:1]B_{1}[0:1], B[1:1]B[1:1], B[ξ:1]B[\xi:1], \dots, B[ξq2:1]B[\xi^{q-2}:1], B1[1:0]B_{1}[1:0] (each isometrically homeomorphic to PP), with fugacity qfq+1\frac{qf}{q+1}. The qqth Power Law allows the (q+1)(q+1)th Power Law to be written more crudely as

Z(f,1(K),β)=Z(qfq+1,R,β)Z(qfq+1,P,β),Z(f,\mathbb{P}^{1}(K),\beta)=Z(\tfrac{qf}{q+1},R,\beta)\cdot Z(\tfrac{qf}{q+1},P,\beta)~{}, (2.2.4)

which is to say that the gas in 1(K)\mathbb{P}^{1}(K) “factors” into two sub-gases: one in ι(R)\iota(R) and one in B[1:0]B[1:0] (which are respectively isometrically homeomorphic to RR and PP), both with fugacity qfq+1\frac{qf}{q+1}.

2.3 Functional equations and a quadratic recurrence

Although Proposition 1.7 and Theorem 2.1 provide explicit formulas for ZN(R,β)Z_{N}(R,\beta) and ZN(1(K),β)Z_{N}(\mathbb{P}^{1}(K),\beta), they are not efficient for computation because they require a complete list of reduced splitting chains of order NN. For a practical alternative, we take advantage of both Power Laws and the following ideas from [BGMR06] and [Sin20]: Apply Z(f,P,β)fZ(f,P,\beta)\cdot\frac{\partial}{\partial f} to the equation Z(f,R,β)=(Z(f,P,β))qZ(f,R,\beta)=(Z(f,P,\beta))^{q} to get

Z(f,P,β)fZ(f,R,β)=qZ(f,R,β)fZ(f,P,β),Z(f,P,\beta)\cdot\frac{\partial}{\partial f}Z(f,R,\beta)=q\cdot Z(f,R,\beta)\cdot\frac{\partial}{\partial f}Z(f,P,\beta)~{},

then expand both sides as power series in ff to obtain the coefficient equations

k=1NZNk(P,β)(Nk)!Zk(R,β)(k1)!=qk=1NZNk(R,β)(Nk)!Zk(P,β)(k1)!for all N1.\sum_{k=1}^{N}\frac{Z_{N-k}(P,\beta)}{(N-k)!}\frac{Z_{k}(R,\beta)}{(k-1)!}=q\cdot\sum_{k=1}^{N}\frac{Z_{N-k}(R,\beta)}{(N-k)!}\frac{Z_{k}(P,\beta)}{(k-1)!}\qquad\text{for all }N\geq 1. (2.3.1)

The identities Zj(P,β)=qj(j2)βZj(R,β)Z_{j}(P,\beta)=q^{-j-\binom{j}{2}\beta}Z_{j}(R,\beta) follow easily from Definition 1.5 and eliminate all instances of Zj(P,β)Z_{j}(P,\beta) in (2.3.1) while introducing powers of the form qj(j2)βq^{-j-\binom{j}{2}\beta}. For N2N\geq 2, a careful rearrangement of these powers, the factorials, and the terms in (2.3.1) yields the explicit recurrence

ZN(R,β)N!q12(N2)β=k=1N1kNsinh(log(q)2[(N+(N2)β)(12kN)+1])sinh(log(q)2[(N+(N2)β)1])ZNk(R,β)(Nk)!q12(Nk2)βZk(R,β)k!q12(k2)β.\frac{Z_{N}(R,\beta)}{N!q^{\frac{1}{2}\binom{N}{2}\beta}}=\sum_{k=1}^{N-1}\frac{k}{N}\cdot\frac{\sinh\left(\frac{\log(q)}{2}\left[\left(N+\binom{N}{2}\beta\right)\left(1-\frac{2k}{N}\right)+1\right]\right)}{\sinh\left(\frac{\log(q)}{2}\left[\left(N+\binom{N}{2}\beta\right)-1\right]\right)}\cdot\frac{Z_{N-k}(R,\beta)}{(N-k)!q^{\frac{1}{2}\binom{N-k}{2}\beta}}\cdot\frac{Z_{k}(R,\beta)}{k!q^{\frac{1}{2}\binom{k}{2}\beta}}~{}.

The expression at left is identically 1 if N=0N=0 or N=1N=1, so induction confirms that it is polynomial in ratios of hyperbolic sines for all N0N\geq 0. In particular, its dependence on qq is carried only by the factor log(q)\log(q) appearing inside the hyperbolic sines, which motivates the following lemma:

Lemma 2.4 (The Quadratic Recurrence).

Set F0(t,β)=F1(t,β)=1F_{0}(t,\beta)=F_{1}(t,\beta)=1 for all β\beta\in\mathbb{C} and all tt\in\mathbb{R}. For N2N\geq 2, Re(β)>2/N\operatorname{Re}(\beta)>-2/N, and tt\in\mathbb{R}, define FN(t,β)F_{N}(t,\beta) by the recurrence

FN(t,β):={k=1N1kNsinh(t2[(N+(N2)β)(12kN)+1])sinh(t2[(N+(N2)β)1])FNk(t,β)Fk(t,β)if t0,k=1N1kN(N+(N2)β)(12kN)+1(N+(N2)β)1FNk(0,β)Fk(0,β)if t=0.F_{N}(t,\beta):=\begin{cases}\displaystyle{\sum_{k=1}^{N-1}\frac{k}{N}\cdot\frac{\sinh\left(\frac{t}{2}\left[\left(N+\binom{N}{2}\beta\right)\left(1-\frac{2k}{N}\right)+1\right]\right)}{\sinh\left(\frac{t}{2}\left[\left(N+\binom{N}{2}\beta\right)-1\right]\right)}\cdot F_{N-k}(t,\beta)\cdot F_{k}(t,\beta)}&\text{if $t\neq 0$,}\\ \\ \displaystyle{\sum_{k=1}^{N-1}\frac{k}{N}\cdot\frac{\left(N+\binom{N}{2}\beta\right)\left(1-\frac{2k}{N}\right)+1}{\left(N+\binom{N}{2}\beta\right)-1}\cdot F_{N-k}(0,\beta)\cdot F_{k}(0,\beta)}&\text{if $t=0$.}\end{cases}
  1. (a)

    For fixed N2N\geq 2 and fixed tt, the function βFN(t,β)\beta\mapsto F_{N}(t,\beta) is holomorphic for Re(β)>2/N\operatorname{Re}(\beta)>-2/N.

  2. (b)

    For fixed N2N\geq 2 and fixed β\beta, the function tFN(t,β)t\mapsto F_{N}(t,\beta) is defined, smooth, and even on \mathbb{R}.

Both parts of the Quadratic Recurrence are straightforward to verify by induction. An interesting and immediate consequence of The Quadratic Recurrence and the preceding discussion is the formula

ZN(R,β)=N!q12(N2)βFN(log(q),β).Z_{N}(R,\beta)=N!q^{\frac{1}{2}\binom{N}{2}\beta}F_{N}(\log(q),\beta)~{}.

It offers a computationally efficient alternative to the “univariate case” of Proposition 1.7 (when sij=βs_{ij}=\beta for all i<ji<j) and extends ZN(R,β)Z_{N}(R,\beta) to a smooth function of q(0,)q\in(0,\infty). Moreover, it transforms nicely under the involution qq1q\mapsto q^{-1}:

ZN(R,β)|qq1=N!q12(N2)βFN(log(q1),β)=N!q12(N2)βFN(log(q),β)=q(N2)βZN(R,β).Z_{N}(R,\beta)\big{|}_{q\mapsto q^{-1}}=N!q^{-\frac{1}{2}\binom{N}{2}\beta}F_{N}\left(\log(q^{-1}),\beta\right)=N!q^{-\frac{1}{2}\binom{N}{2}\beta}F_{N}\left(\log(q),\beta\right)=q^{-\binom{N}{2}\beta}Z_{N}(R,\beta)~{}.

The Quadratic Recurrence serves the projective analogue as well. Expanding (2.2.4) into powers of ff yields the coefficient equations

ZN(1(K),β)N!=k=0N(qq+1)NZNk(R,β)(Nk)!Zk(P,β)k!for all N0,\frac{Z_{N}(\mathbb{P}^{1}(K),\beta)}{N!}=\sum_{k=0}^{N}\left(\frac{q}{q+1}\right)^{N}\frac{Z_{N-k}(R,\beta)}{(N-k)!}\frac{Z_{k}(P,\beta)}{k!}\qquad\text{for all }N\geq 0, (2.3.2)

and the identities Zj(P,β)=qj(j2)βZj(R,β)Z_{j}(P,\beta)=q^{-j-\binom{j}{2}\beta}Z_{j}(R,\beta) and Zj(R,β)=j!q12(j2)βFj(log(q),β)Z_{j}(R,\beta)=j!q^{\frac{1}{2}\binom{j}{2}\beta}F_{j}(\log(q),\beta) allow the kkth summand to be rewritten as

(qq+1)NZNk(R,β)(Nk)!Zk(P,β)k!=q12(N+(N2)β)(12kN)(2cosh(log(q)2))NFNk(log(q),β)Fk(log(q),β)for N1.\left(\frac{q}{q+1}\right)^{N}\frac{Z_{N-k}(R,\beta)}{(N-k)!}\frac{Z_{k}(P,\beta)}{k!}=\frac{q^{\frac{1}{2}\left(N+\binom{N}{2}\beta\right)\left(1-\frac{2k}{N}\right)}}{\left(2\cosh\left(\frac{\log(q)}{2}\right)\right)^{N}}\cdot F_{N-k}(\log(q),\beta)\cdot F_{k}(\log(q),\beta)\qquad\text{for $N\geq 1$}.

Thus, adding two copies of the sum in (2.3.2) together, pairing the kkth term of the first copy with the (Nk)(N-k)th term of the second copy, and dividing by 22 gives

ZN(1(K),β)N!=k=0Ncosh(log(q)2(N+(N2)β)(12kN))(2cosh(log(q)2))NFNk(log(q),β)Fk(log(q),β)for N1,\frac{Z_{N}(\mathbb{P}^{1}(K),\beta)}{N!}=\sum_{k=0}^{N}\frac{\cosh\left(\frac{\log(q)}{2}\left(N+\binom{N}{2}\beta\right)\left(1-\frac{2k}{N}\right)\right)}{\left(2\cosh\left(\frac{\log(q)}{2}\right)\right)^{N}}\cdot F_{N-k}(\log(q),\beta)\cdot F_{k}(\log(q),\beta)\qquad\text{for $N\geq 1$},

which is valid for Re(β)>2/N\operatorname{Re}(\beta)>-2/N. Through this formula, ZN(1(K),β)Z_{N}(\mathbb{P}^{1}(K),\beta) clearly extends to a smooth function of q(0,)q\in(0,\infty) and is invariant under the involution qq1q\mapsto q^{-1}. We conclude this section by summarizing these observations:

Theorem 2.5 (Efficient Formulas and Functional Equations).

Suppose N2N\geq 2 and Re(β)>2/N\operatorname{Re}(\beta)>-2/N, and define (Fk(t,β))k=0N(F_{k}(t,\beta))_{k=0}^{N} as in Lemma 2.4. The NNth canonical partition functions are given by the formulas

ZN(R,β)\displaystyle Z_{N}(R,\beta) =N!q12(N2)βFN(log(q),β)and\displaystyle=N!q^{\frac{1}{2}\binom{N}{2}\beta}F_{N}(\log(q),\beta)\qquad\text{and}
ZN((K),β)\displaystyle Z_{N}(\mathbb{P}(K),\beta) =N!k=0Ncosh(log(q)2(N+(N2)β)(12kN))(2cosh(log(q)2))NFNk(log(q),β)Fk(log(q),β),\displaystyle=N!\sum_{k=0}^{N}\frac{\cosh\left(\frac{\log(q)}{2}\left(N+\binom{N}{2}\beta\right)\left(1-\frac{2k}{N}\right)\right)}{\left(2\cosh\left(\frac{\log(q)}{2}\right)\right)^{N}}\cdot F_{N-k}(\log(q),\beta)\cdot F_{k}(\log(q),\beta)~{},

which extend ZN(R,β)Z_{N}(R,\beta) and ZN(1(K),β)Z_{N}(\mathbb{P}^{1}(K),\beta) to smooth functions of q(0,)q\in(0,\infty) satisfying

ZN(R,β)|qq1=q(N2)βZN(R,β)andZN(1(K),β)|qq1=ZN(1(K),β).Z_{N}(R,\beta)\big{|}_{q\mapsto q^{-1}}=q^{-\binom{N}{2}\beta}Z_{N}(R,\beta)\qquad\text{and}\qquad Z_{N}(\mathbb{P}^{1}(K),\beta)\big{|}_{q\mapsto q^{-1}}=Z_{N}(\mathbb{P}^{1}(K),\beta)~{}.

It should be noted here that the qq1q\mapsto q^{-1} functional equation at left is a special case of the one proved in [DM91], and that both functional equations closely resemble the ones in [Vol10].

3 Proofs of the main results

This section will establish the proofs of Theorems 2.1 and 2.3. The common step in both is a decomposition of (1(K))N(\mathbb{P}^{1}(K))^{N} into (q+1)N(q+1)^{N} cells that are isometrically isomorphic to PNP^{N}, which combines with the metric and measure properties in Section 1.2 to create the key relationship between the canonical partition functions for X=1(K)X=\mathbb{P}^{1}(K) and X=PX=P. We will prove this relationship first, then conclude the proofs of Theorems 2.1 and 2.3 in their own subsections.

3.1 Decomposing the integral over (1(K))N(\mathbb{P}^{1}(K))^{N}

We begin with an integer N2N\geq 2 that shall remain fixed for the rest of this section, reserve the symbol 𝒔\bm{s} for a complex tuple (sij)1i<jN(s_{ij})_{1\leq i<j\leq N}, and fix the following notation to better organize the forthcoming arguments:

Notation 3.1.

Let II be a subset of [N]={1,,N}[N]=\{1,\dots,N\}.

  • For any set XX we write XIX^{I} for the product iIX={xI=(xi)iI:xiX}\prod_{i\in I}X=\{x_{I}=(x_{i})_{i\in I}:x_{i}\in X\} and assume XIX^{I} has the product topology if XX is a topological space.

  • We write μI\mu^{I} for the product Haar measure on KIK^{I} satisfying μI(RI)=1\mu^{I}(R^{I})=1, and we make this consistent for I=I=\varnothing by giving the singleton space K=R={0}K^{\varnothing}=R^{\varnothing}=\{0\} measure 1. We also write νI\nu^{I} for the product measure on (1(K))I(\mathbb{P}^{1}(K))^{I}, with the same convention for I=I=\varnothing.

  • For a measurable subset XKX\subset K we set 𝒵(X,𝒔):=1\mathcal{Z}_{\varnothing}(X,\bm{s}):=1 and

    𝒵I(X,𝒔):=XIi<ji,jI|xixj|sijdμIifI.\mathcal{Z}_{I}(X,\bm{s}):=\int_{X^{I}}\prod_{\begin{subarray}{c}i<j\\ i,j\in I\end{subarray}}|x_{i}-x_{j}|^{s_{ij}}\,d\mu^{I}\qquad\text{if}\quad I\neq\varnothing~{}.

    Note that 𝒵I(X,𝒔)\mathcal{Z}_{I}(X,\bm{s}) is constant with respect to the entry sijs_{ij} if i[N]Ii\in[N]\setminus I or j[N]Ij\in[N]\setminus I, and it is equal to 𝒵N(X,𝒔)\mathcal{Z}_{N}(X,\bm{s}) if I=[N]I=[N].

  • Using the constant q=#(R/P)q=\#(R/P), we write (I0,,Iq)[N](I_{0},\dots,I_{q})\vdash[N] for an ordered partition of [N][N] into at most q+1q+1 parts. That is, (I0,,Iq)[N](I_{0},\dots,I_{q})\vdash[N] means I0,,IqI_{0},\dots,I_{q} are q+1q+1 disjoint ordered subsets of [N][N] with union equal to [N][N], where some IkI_{k} may be empty.

In addition to the above, it will be useful to consider II-analogues of splitting chains:

Definition 3.2.

Suppose I[N]I\subset[N]. An II-splitting chain of length L0L\geq 0 is a tuple =(0,,L){\bm{\pitchfork}}=({\pitchfork}_{0},\dots,{\pitchfork}_{L}) of partitions of II satisfying

{I}=0>1>2>>L={{i}:iI}.\{I\}={\pitchfork}_{0}>{\pitchfork}_{1}>{\pitchfork}_{2}>\dots>{\pitchfork}_{L}=\{\{i\}:i\in I\}~{}.

If #I2\#I\geq 2, we define ()\mathcal{B}({\bm{\pitchfork}}), (λ)\ell_{\bm{\pitchfork}}(\lambda), and deg(λ){2,3,,#I}\deg_{\bm{\pitchfork}}(\lambda)\in\{2,3,\dots,\#I\} just as in Definition 1.6. Otherwise ()\mathcal{B}({\bm{\pitchfork}}) will be treated as the empty set and there is no need to define \ell_{\bm{\pitchfork}} or deg\deg_{\bm{\pitchfork}}. Finally, we call an II-splitting chain {\bm{\pitchfork}} reduced if each λ()\lambda\in\mathcal{B}({\bm{\pitchfork}}) satisfies λ=(λ)\lambda\in{\pitchfork}_{\ell}\iff\ell=\ell_{\bm{\pitchfork}}(\lambda), write I\mathcal{R}_{I} for the set of reduced II-splitting chains, and define

ΩI:=λI#λ>1{𝒔:Re(eλ(𝒔))>0}whereeλ(𝒔):=#λ1+i<ji,jλsij.\Omega_{I}:=\bigcap_{\begin{subarray}{c}\lambda\subset I\\ \#\lambda>1\end{subarray}}\left\{\bm{s}:\operatorname{Re}(e_{\lambda}(\bm{s}))>0\right\}\qquad\text{where}\qquad e_{\lambda}(\bm{s}):=\#\lambda-1+\sum_{\begin{subarray}{c}i<j\\ i,j\in\lambda\end{subarray}}s_{ij}~{}.

Note that =\mathcal{R}_{\varnothing}=\varnothing because I=I=\varnothing has no partitions, Ω=N(N1)/2\Omega_{\varnothing}=\mathbb{C}^{N(N-1)/2} because Ω\Omega_{\varnothing} is an intersection of subsets of N(N1)/2\mathbb{C}^{N(N-1)/2} taken over an empty index set, and e(𝒔)=1e_{\varnothing}(\bm{s})=-1 for a similar reason. For each singleton {i}\{i\}, the set {i}\mathcal{R}_{\{i\}} is comprised of a single splitting chain of length zero, Ω{i}=N(N1)/2\Omega_{\{i\}}=\mathbb{C}^{N(N-1)/2} for the same reason as the I=I=\varnothing case, and similarly e{i}(𝒔)=0e_{\{i\}}(\bm{s})=0. At the other extreme, taking I=[N]I=[N] in Definition 3.2 recovers Definition 1.6 and gives ΩI=ΩN\Omega_{I}=\Omega_{N}.

Proposition 3.3.

For any vv\in\mathbb{Z} and any nonempty subset I[N]I\subset[N], the integral 𝒵I(πvR,𝒔)\mathcal{Z}_{I}(\pi^{v}R,\bm{s}) converges absolutely if and only if 𝒔ΩI\bm{s}\in\Omega_{I}, and in this case

𝒵I(πvR,𝒔)=1q(v1)(eI(𝒔)+1)+#IIλ()(q1)deg(λ)1qeλ(𝒔)1.\mathcal{Z}_{I}(\pi^{v}R,\bm{s})=\frac{1}{q^{(v-1)(e_{I}(\bm{s})+1)+\#I}}\sum_{{\bm{\pitchfork}}\in\mathcal{R}_{I}}\prod_{\lambda\in\mathcal{B}({\bm{\pitchfork}})}\frac{(q-1)_{\deg_{{\bm{\pitchfork}}}(\lambda)-1}}{q^{e_{\lambda}(\bm{s})}-1}~{}.

In particular, we recover Proposition 1.7 by taking I=[N]I=[N] and v=0v=0.

Proof.

First suppose II is a singleton, so that the product inside the integral ZI(πvR,𝒔)Z_{I}(\pi^{v}R,\bm{s}) is empty and hence

ZI(πvR,𝒔)=(πvR)I𝑑μI=πvR𝑑μ=qv.Z_{I}(\pi^{v}R,\bm{s})=\int_{(\pi^{v}R)^{I}}\,d\mu^{I}=\int_{\pi^{v}R}\,d\mu=q^{-v}~{}.

This integral is constant, and hence absolutely convergent, for all 𝒔N(N1)/2=ΩI\bm{s}\in\mathbb{C}^{N(N-1)/2}=\Omega_{I}. On the other hand, I\mathcal{R}_{I} consists of a single II-splitting chain, namely the one-tuple =({I}){\bm{\pitchfork}}=(\{I\}). Then ()=\mathcal{B}({\bm{\pitchfork}})=\varnothing and eI(𝒔)=0e_{I}(\bm{s})=0 imply

1q(v1)(eI(𝒔)+1)+#IIλ()(q1)deg(λ)1qeλ(𝒔)1=1q(v1)1+1λ(q1)deg(λ)1qeλ(𝒔)1=qv\frac{1}{q^{(v-1)(e_{I}(\bm{s})+1)+\#I}}\sum_{{\bm{\pitchfork}}\in\mathcal{R}_{I}}\prod_{\lambda\in\mathcal{B}({\bm{\pitchfork}})}\frac{(q-1)_{\deg_{{\bm{\pitchfork}}}(\lambda)-1}}{q^{e_{\lambda}(\bm{s})}-1}=\frac{1}{q^{(v-1)\cdot 1+1}}\prod_{\lambda\in\varnothing}\frac{(q-1)_{\deg_{{\bm{\pitchfork}}}(\lambda)-1}}{q^{e_{\lambda}(\bm{s})}-1}=q^{-v}

as well, so the claim holds for any singleton subset I[N]I\subset[N]. Now suppose II is not a singleton. By relabeling II we may assume I=[n]I=[n] where 2nN2\leq n\leq N. By Proposition 3.15 and Lemma 3.16(c) in [Web21b], the integral

𝒵I(R,𝒔)=𝒵n(R,𝒔)=Rn1i<jn|xixj|sijdμN\mathcal{Z}_{I}(R,\bm{s})=\mathcal{Z}_{n}(R,\bm{s})=\int_{R^{n}}\prod_{1\leq i<j\leq n}|x_{i}-x_{j}|^{s_{ij}}\,d\mu^{N}

converges absolutely if and only if 𝒔\bm{s} belongs to the intersection

nλ(){𝒔:Re(eλ(𝒔))>0},\bigcap_{{\bm{\pitchfork}}\in\mathcal{R}_{n}}\bigcap_{\lambda\in\mathcal{B}({\bm{\pitchfork}})}\left\{\bm{s}:\operatorname{Re}(e_{\lambda}(\bm{s}))>0\right\}~{}, (3.1.1)

and for such 𝒔\bm{s} we have

𝒵n(R,𝒔)=qe[n](𝒔)+1nnλ()(q1)deg(λ)1qeλ(𝒔)1.\mathcal{Z}_{n}(R,\bm{s})=q^{e_{[n]}(\bm{s})+1-n}\sum_{{\bm{\pitchfork}}\in\mathcal{R}_{n}}\prod_{\lambda\in\mathcal{B}({\bm{\pitchfork}})}\frac{(q-1)_{\deg_{{\bm{\pitchfork}}}(\lambda)-1}}{q^{e_{\lambda}(\bm{s})}-1}~{}.

Changing variables in the integral 𝒵n(R,𝒔)\mathcal{Z}_{n}(R,\bm{s}) by the homothety Rn(πvR)nR^{n}\to(\pi^{v}R)^{n} gives

𝒵N(πvR,𝒔)=qv(e[N](𝒔)+1)𝒵n(R,𝒔)=1q(v1)(e[n](𝒔)+1)+nnλ()(q1)deg(λ)1qeλ(𝒔)1,\mathcal{Z}_{N}(\pi^{v}R,\bm{s})=q^{-v(e_{[N]}(\bm{s})+1)}\cdot\mathcal{Z}_{n}(R,\bm{s})=\frac{1}{q^{(v-1)(e_{[n]}(\bm{s})+1)+n}}\sum_{{\bm{\pitchfork}}\in\mathcal{R}_{n}}\prod_{\lambda\in\mathcal{B}({\bm{\pitchfork}})}\frac{(q-1)_{\deg_{{\bm{\pitchfork}}}(\lambda)-1}}{q^{e_{\lambda}(\bm{s})}-1}~{},

and the first equality implies that the domain of absolute convergence for 𝒵n(πvR,𝒔)\mathcal{Z}_{n}(\pi^{v}R,\bm{s}) is also the intersection appearing in (3.1.1). But every subset λ[n]\lambda\subset[n] with #λ>1\#\lambda>1 appears as a branch in at least one reduced splitting chain of order nn, so the intersection in (3.1.1) is precisely Ω[n]\Omega_{[n]}. Therefore the claim holds for I=[n]I=[n], and hence for any non-singleton subset I[N]I\subset[N].

The v=1v=1 case of Proposition 3.3 has an important relationship with the main result of this section, which is the following theorem.

Theorem 3.4.

For each N2N\geq 2, the integral 𝒵N(1(K),𝒔)\mathcal{Z}_{N}(\mathbb{P}^{1}(K),\bm{s}) converges absolutely if and only if 𝒔ΩN\bm{s}\in\Omega_{N}, and in this case

𝒵N(1(K),𝒔)=(qq+1)N(I0,,Iq)[N]k=0q𝒵Ik(P,𝒔).\mathcal{Z}_{N}(\mathbb{P}^{1}(K),\bm{s})=\left(\frac{q}{q+1}\right)^{N}\sum_{(I_{0},\dots,I_{q})\vdash[N]}\prod_{k=0}^{q}\mathcal{Z}_{I_{k}}(P,\bm{s})~{}.
Proof.

The partition of 1(K)\mathbb{P}^{1}(K) in (2.1.4) can be rewritten in the form

1(K)=k=0qϕk(B1[0:1]),\mathbb{P}^{1}(K)=\bigsqcup_{k=0}^{q}\phi_{k}(B_{1}[0:1])~{},

where ϕkPGL2(R)\phi_{k}\in PGL_{2}(R) is the element represented by (1001)\left(\begin{smallmatrix}1&0\\ 0&1\end{smallmatrix}\right) if k=0k=0, (1ξk101)\left(\begin{smallmatrix}1&\xi^{k-1}\\ 0&1\end{smallmatrix}\right) if 0<k<q0<k<q, or (0110)\left(\begin{smallmatrix}0&1\\ 1&0\end{smallmatrix}\right) if k=qk=q. This leads to a partition of the NN-fold product,

(1(K))N=(I0,,Iq)[N]C(I0,,Iq),(\mathbb{P}^{1}(K))^{N}=\bigsqcup_{(I_{0},\dots,I_{q})\vdash[N]}C(I_{0},\dots,I_{q})~{},

where each part is a “cell” of the form

C(I0,,Iq)\displaystyle C(I_{0},\dots,I_{q}) :={([x1,0:x1,1],,[xN,0:xN,1])(1(K))N:[xi,0:xi,1]ϕk(B1[0:1])iIk}\displaystyle:=\{([x_{1,0}:x_{1,1}],\dots,[x_{N,0}:x_{N,1}])\in(\mathbb{P}^{1}(K))^{N}:[x_{i,0}:x_{i,1}]\in\phi_{k}(B_{1}[0:1])\iff i\in I_{k}\}
=k=0q(ϕk(B1[0:1]))Ik.\displaystyle=\prod_{k=0}^{q}(\phi_{k}(B_{1}[0:1]))^{I_{k}}~{}.

Accordingly, the integral 𝒵N(1(K),𝒔)\mathcal{Z}_{N}(\mathbb{P}^{1}(K),\bm{s}) breaks into a sum of integrals of the form

C(I0,,Iq)1i<jNδ([xi,0:xi,1],[xj,0:xj,1])sijdνN,\int_{C(I_{0},\dots,I_{q})}\prod_{1\leq i<j\leq N}\delta([x_{i,0}:x_{i,1}],[x_{j,0}:x_{j,1}])^{s_{ij}}\,d\nu^{N}~{}, (3.1.2)

summed over all (I0,,Iq)[N](I_{0},\dots,I_{q})\vdash[N]. Since every cell C(I0,,Iq)C(I_{0},\dots,I_{q}) has positive measure, the integral 𝒵N(1(K),𝒔)\mathcal{Z}_{N}(\mathbb{P}^{1}(K),\bm{s}) converges absolutely if and only if the integral in (3.1.2) converges absolutely for all (I0,,Iq)[N](I_{0},\dots,I_{q})\vdash[N]. Fix one (I0,,Iq)(I_{0},\dots,I_{q}) for the moment. By (2.1.4) and the definition of the ϕk\phi_{k}’s above, note that the entries of each tuple ([x1,0:x1,1],,[xN,0:xN,1])C(I0,,Iq)([x_{1,0}:x_{1,1}],\dots,[x_{N,0}:x_{N,1}])\in C(I_{0},\dots,I_{q}) satisfy δ([xi,0:xi,1],[xj,0:xj,1])sij=1\delta([x_{i,0}:x_{i,1}],[x_{j,0}:x_{j,1}])^{s_{ij}}=1 if and only if ii and jj belong to different parts of (I0,,Iq)(I_{0},\dots,I_{q}). Therefore the integrand in (3.1.2) factors as

1i<jNδ([xi,0:xi,1],[xj,0:xj,1])sij\displaystyle\prod_{1\leq i<j\leq N}\delta([x_{i,0}:x_{i,1}],[x_{j,0}:x_{j,1}])^{s_{ij}} =k=0qi<ji,jIkδ([xi,0:xi,1],[xj,0:xj,1])sij,\displaystyle=\prod_{k=0}^{q}\prod_{\begin{subarray}{c}i<j\\ i,j\in I_{k}\end{subarray}}\delta([x_{i,0}:x_{i,1}],[x_{j,0}:x_{j,1}])^{s_{ij}}~{},

and the measure on C(I0,,Iq)C(I_{0},\dots,I_{q}) factors in a similar way, namely k=0qdνIk\prod_{k=0}^{q}d\nu^{I_{k}} where νIk\nu^{I_{k}} is the product measure on (1(K))Ik(\mathbb{P}^{1}(K))^{I_{k}}. Now Fubini’s Theorem for positive functions and PGL2(R)PGL_{2}(R)-invariance give

C(I0,,Iq)\displaystyle\int_{C(I_{0},\dots,I_{q})} |1i<jNδ([xi,0:xi,1],[xj,0:xj,1])sij|dνN\displaystyle\Bigg{|}\prod_{1\leq i<j\leq N}\delta([x_{i,0}:x_{i,1}],[x_{j,0}:x_{j,1}])^{s_{ij}}\Bigg{|}_{\mathbb{C}}\,d\nu^{N}
=k=0q(ϕk(B1[0:1]))Ik|i<ji,jIkδ([xi,0:xi,1],[xj,0:xj,1])sij|dνIk\displaystyle=\prod_{k=0}^{q}\int_{(\phi_{k}(B_{1}[0:1]))^{I_{k}}}\Bigg{|}\prod_{\begin{subarray}{c}i<j\\ i,j\in I_{k}\end{subarray}}\delta([x_{i,0}:x_{i,1}],[x_{j,0}:x_{j,1}])^{s_{ij}}\Bigg{|}_{\mathbb{C}}\,d\nu^{I_{k}}
=k=0q(B1[0:1])Ik|i<ji,jIkδ([xi,0:xi,1],[xj,0:xj,1])sij|dνIk,\displaystyle=\prod_{k=0}^{q}\int_{(B_{1}[0:1])^{I_{k}}}\Bigg{|}\prod_{\begin{subarray}{c}i<j\\ i,j\in I_{k}\end{subarray}}\delta([x_{i,0}:x_{i,1}],[x_{j,0}:x_{j,1}])^{s_{ij}}\Bigg{|}_{\mathbb{C}}\,d\nu^{I_{k}}~{},

so the integral in (3.1.2) converges absolutely if and only if all q+1q+1 of the integrals of the form

(B1[0:1])Iki<ji,jIkδ([xi,0:xi,1],[xj,0:xj,1])sijdνIk\int_{(B_{1}[0:1])^{I_{k}}}\prod_{\begin{subarray}{c}i<j\\ i,j\in I_{k}\end{subarray}}\delta([x_{i,0}:x_{i,1}],[x_{j,0}:x_{j,1}])^{s_{ij}}\,d\nu^{I_{k}} (3.1.3)

converge absolutely. The change of variables PIk(B1[0:1])IkP^{I_{k}}\to(B_{1}[0:1])^{I_{k}} given by ι:PB1[0:1]\iota:P\to B_{1}[0:1] in each coordinate, along with (2.1.1), (2.1.2), and (2.1.3), allows the integral in (3.1.3) to be rewritten as (qq+1)#Ik𝒵Ik(P,𝒔)(\frac{q}{q+1})^{\#I_{k}}\mathcal{Z}_{I_{k}}(P,\bm{s}), and thus Proposition 3.3 implies that it converges absolutely if and only if 𝒔ΩIk\bm{s}\in\Omega_{I_{k}}. Therefore the integral over C(I0,,Iq)C(I_{0},\dots,I_{q}) in (3.1.2) converges absolutely if and only if 𝒔ΩI0ΩIq\bm{s}\in\Omega_{I_{0}}\cap\dots\cap\Omega_{I_{q}}, and in this case Fubini’s Theorem for absolutely integrable functions, PGL2(R)PGL_{2}(R)-invariance, and the change of variables above allow it to be rewritten as

C(I0,,Iq)\displaystyle\int_{C(I_{0},\dots,I_{q})} 1i<jNδ([xi,0:xi,1],[xj,0:xj,1])sijdνN\displaystyle\prod_{1\leq i<j\leq N}\delta([x_{i,0}:x_{i,1}],[x_{j,0}:x_{j,1}])^{s_{ij}}\,d\nu^{N}
=k=0q(B1[0:1])Iki<ji,jIkδ([xi,0:xi,1],[xj,0:xj,1])sijdνIk\displaystyle=\prod_{k=0}^{q}\int_{(B_{1}[0:1])^{I_{k}}}\prod_{\begin{subarray}{c}i<j\\ i,j\in I_{k}\end{subarray}}\delta([x_{i,0}:x_{i,1}],[x_{j,0}:x_{j,1}])^{s_{ij}}\,d\nu^{I_{k}}
=k=0q(qq+1)#Ik𝒵Ik(P,𝒔)\displaystyle=\prod_{k=0}^{q}\left(\frac{q}{q+1}\right)^{\#I_{k}}\mathcal{Z}_{I_{k}}(P,\bm{s})
=(qq+1)Nk=0q𝒵Ik(P,𝒔).\displaystyle=\left(\frac{q}{q+1}\right)^{N}\prod_{k=0}^{q}\mathcal{Z}_{I_{k}}(P,\bm{s})~{}.

Finally, since 𝒵N(1(K),𝒔)\mathcal{Z}_{N}(\mathbb{P}^{1}(K),\bm{s}) is the sum of these integrals over all (I0,,Iq)[N](I_{0},\dots,I_{q})\vdash[N], it converges absolutely if and only if

𝒔(I1,,Iq)[N](ΩI1ΩIq)=I[N]#I>1ΩI.\bm{s}\in\bigcap_{(I_{1},\dots,I_{q})\vdash[N]}\left(\Omega_{I_{1}}\cap\dots\cap\Omega_{I_{q}}\right)=\bigcap_{\begin{subarray}{c}I\subset[N]\\ \#I>1\end{subarray}}\Omega_{I}~{}.

The last equality of intersections holds because each subset I[N]I\subset[N] with #I>1\#I>1 appears as a part in at least one of the ordered partitions (I1,,Iq)[N](I_{1},\dots,I_{q})\vdash[N], and because none of the parts with #Ik1\#I_{k}\leq 1 affect the intersection (because ΩIk=N(N1)/2\Omega_{I_{k}}=\mathbb{C}^{N(N-1)/2} for such IkI_{k}). The intersection of ΩI\Omega_{I} over all I[N]I\subset[N] with #I>1\#I>1 is clearly equal to Ω[N]=ΩN\Omega_{[N]}=\Omega_{N} by Definition 3.2, so the proof is complete.

3.2 Finishing the proof of Theorem 2.1

Theorem 3.4 established that the integral 𝒵N(1(K),𝒔)\mathcal{Z}_{N}(\mathbb{P}^{1}(K),\bm{s}) converges absolutely if and only if 𝒔ΩN\bm{s}\in\Omega_{N}, and for such 𝒔\bm{s} it gave

𝒵N(1(K),𝒔)=(qq+1)N(I0,,Iq)[N]k=0q𝒵Ik(P,𝒔).\mathcal{Z}_{N}(\mathbb{P}^{1}(K),\bm{s})=\left(\frac{q}{q+1}\right)^{N}\sum_{(I_{0},\dots,I_{q})\vdash[N]}\prod_{k=0}^{q}\mathcal{Z}_{I_{k}}(P,\bm{s})~{}. (3.2.1)

It remains to show that the righthand sum can be converted into the sum over N{\bm{\pitchfork}}\in\mathcal{R}_{N} proposed in Theorem 2.1.

Proof of Theorem 2.1.

We begin by breaking the terms of the sum in (3.2.1) into two main groups. The simpler group is indexed by those (I0,,Iq)(I_{0},\dots,I_{q}) with Ij=[N]I_{j}=[N] for some jj and Ik=I_{k}=\varnothing for all kjk\neq j, in which case 𝒵Ij(P,𝒔)=𝒵N(P,𝒔)\mathcal{Z}_{I_{j}}(P,\bm{s})=\mathcal{Z}_{N}(P,\bm{s}) and 𝒵Ik(P,𝒔)=1\mathcal{Z}_{I_{k}}(P,\bm{s})=1 for all kjk\neq j. Therefore each of the group’s q+1q+1 terms (one for each j{0,,q}j\in\{0,\dots,q\}) contributes the quantity k=0q𝒵Ik(P,𝒔)=𝒵N(P,𝒔)\prod_{k=0}^{q}\mathcal{Z}_{I_{k}}(P,\bm{s})=\mathcal{Z}_{N}(P,\bm{s}) to the sum in (3.2.1) for a total contribution of

(q+1)𝒵N(P,𝒔)=q+1qNNλ()(q1)deg(λ)1qeλ(𝒔)1(q+1)\mathcal{Z}_{N}(P,\bm{s})=\frac{q+1}{q^{N}}\sum_{{\bm{\pitchfork}}\in\mathcal{R}_{N}}\prod_{\lambda\in\mathcal{B}({\bm{\pitchfork}})}\frac{(q-1)_{\deg_{\bm{\pitchfork}}(\lambda)-1}}{q^{e_{\lambda}(\bm{s})}-1} (3.2.2)

by the v=1v=1 and I=[N]I=[N] case of Proposition 3.3. The other group of terms is indexed by the ordered partitions (I0,,Iq)[N](I_{0},\dots,I_{q})\vdash[N] satisfying I0,,Iq[N]I_{0},\dots,I_{q}\subsetneq[N]. To deal with them carefully, we fix one such (I0,,Iq)(I_{0},\dots,I_{q}) for the moment, and note that the number dd of nonempty parts IkI_{k} must be at least 2. Thus we have indices k1,,kd{0,,q}k_{1},\dots,k_{d}\in\{0,\dots,q\} with IkjI_{k_{j}}\neq\varnothing, and for every k{0,,q}{k1,,kd}k\in\{0,\dots,q\}\setminus\{k_{1},\dots,k_{d}\} we have Ik=I_{k}=\varnothing and hence 𝒵Ik(P,𝒔)=1\mathcal{Z}_{I_{k}}(P,\bm{s})=1. For the nonempty sets IkjI_{k_{j}}, Proposition 3.3 expands 𝒵Ikj(P,𝒔)\mathcal{Z}_{I_{k_{j}}}(P,\bm{s}) as a sum over Ikj\mathcal{R}_{I_{k_{j}}} (whose elements shall be denoted j{\bm{\pitchfork}}_{j} instead of {\bm{\pitchfork}}) and hence

k=0q𝒵Ik(P,𝒔)\displaystyle\prod_{k=0}^{q}\mathcal{Z}_{I_{k}}(P,\bm{s}) =j=1d1q#IkjjIkjλ(j)(q1)degj(λ)1qeλ(𝒔)1\displaystyle=\prod_{j=1}^{d}\frac{1}{q^{\#I_{k_{j}}}}\sum_{{\bm{\pitchfork}}_{j}\in\mathcal{R}_{I_{k_{j}}}}\prod_{\lambda\in\mathcal{B}({\bm{\pitchfork}}_{j})}\frac{(q-1)_{\deg_{{\bm{\pitchfork}}_{j}}(\lambda)-1}}{q^{e_{\lambda}(\bm{s})}-1}
=1qN(1,,d)Ik1××Ikdj=1dλ(j)(q1)degj(λ)1qeλ(𝒔)1\displaystyle=\frac{1}{q^{N}}\sum_{({\bm{\pitchfork}}_{1},\dots,{\bm{\pitchfork}}_{d})\in\mathcal{R}_{I_{k_{1}}}\times\cdots\times\mathcal{R}_{I_{k_{d}}}}\prod_{j=1}^{d}\prod_{\lambda\in\mathcal{B}({\bm{\pitchfork}}_{j})}\frac{(q-1)_{\deg_{{\bm{\pitchfork}}_{j}}(\lambda)-1}}{q^{e_{\lambda}(\bm{s})}-1}
=1qN(1,,d)Ik1××Ikdλ(1)(d)(q1)degj(λ)1qeλ(𝒔)1.\displaystyle=\frac{1}{q^{N}}\sum_{({\bm{\pitchfork}}_{1},\dots,{\bm{\pitchfork}}_{d})\in\mathcal{R}_{I_{k_{1}}}\times\cdots\times\mathcal{R}_{I_{k_{d}}}}\prod_{\lambda\in\mathcal{B}({\bm{\pitchfork}}_{1})\sqcup\cdots\sqcup\mathcal{B}({\bm{\pitchfork}}_{d})}\frac{(q-1)_{\deg_{{\bm{\pitchfork}}_{j}}(\lambda)-1}}{q^{e_{\lambda}(\bm{s})}-1}~{}.

We now make use of a simple correspondence between the tuples (1,,d)Ik1××Ikd({\bm{\pitchfork}}_{1},\dots,{\bm{\pitchfork}}_{d})\in\mathcal{R}_{I_{k_{1}}}\times\cdots\times\mathcal{R}_{I_{k_{d}}} and the reduced splitting chains =(0,1,,L)N{\bm{\pitchfork}}=({\pitchfork}_{0},{\pitchfork}_{1},\dots,{\pitchfork}_{L})\in\mathcal{R}_{N} satisfying 1={Ik1,,Ikd}{\pitchfork}_{1}=\{I_{k_{1}},\dots,I_{k_{d}}\}. To establish it, note that each N{\bm{\pitchfork}}\in\mathcal{R}_{N} corresponds uniquely to its branch set ()\mathcal{B}({\bm{\pitchfork}}) (Lemma 2.5(b) of [Web21b]), which generalizes in an obvious way to reduced II-splitting chains (for any nonempty I[N]I\subset[N]). Now if =(0,1,,L)N{\bm{\pitchfork}}=({\pitchfork}_{0},{\pitchfork}_{1},\dots,{\pitchfork}_{L})\in\mathcal{R}_{N} satisfies 1={Ik1,,Ikd}{\pitchfork}_{1}=\{I_{k_{1}},\dots,I_{k_{d}}\}, the corresponding branch set ()\mathcal{B}({\bm{\pitchfork}}) decomposes as

()={[N]}j=1d{λ():λIkj}.\mathcal{B}({\bm{\pitchfork}})=\{[N]\}\sqcup\bigsqcup_{j=1}^{d}\{\lambda\in\mathcal{B}({\bm{\pitchfork}}):\lambda\subset I_{k_{j}}\}~{}.

Each of the sets {λ():λIkj}\{\lambda\in\mathcal{B}({\bm{\pitchfork}}):\lambda\subset I_{k_{j}}\} is the branch set (j)\mathcal{B}({\bm{\pitchfork}}_{j}) for a unique jIkj{\bm{\pitchfork}}_{j}\in\mathcal{R}_{I_{k_{j}}}, so in this sense {\bm{\pitchfork}} “breaks” into a unique tuple (1,,d)Ik1××Ikd({\bm{\pitchfork}}_{1},\dots,{\bm{\pitchfork}}_{d})\in\mathcal{R}_{I_{k_{1}}}\times\cdots\times\mathcal{R}_{I_{k_{d}}}. On the other hand, any tuple (1,,d)Ik1××Ikd({\bm{\pitchfork}}_{1},\dots,{\bm{\pitchfork}}_{d})\in\mathcal{R}_{I_{k_{1}}}\times\cdots\times\mathcal{R}_{I_{k_{d}}} can be “assembled” as follows. Since {Ik1,,Ikd}\{I_{k_{1}},\dots,I_{k_{d}}\} is a partition of [N][N], taking the union of the dd branch sets (1),,(d)\mathcal{B}({\bm{\pitchfork}}_{1}),\dots,\mathcal{B}({\bm{\pitchfork}}_{d}) and the singleton {[N]}\{[N]\} forms the branch set ()\mathcal{B}({\bm{\pitchfork}}) for a unique N{\bm{\pitchfork}}\in\mathcal{R}_{N}. It is clear that “breaking” and “assembling” are inverses, giving a correspondence NIk1××Ikd\mathcal{R}_{N}\longleftrightarrow\mathcal{R}_{I_{k_{1}}}\times\cdots\times\mathcal{R}_{I_{k_{d}}} under which each identification (1,,d){\bm{\pitchfork}}\longleftrightarrow({\bm{\pitchfork}}_{1},\dots,{\bm{\pitchfork}}_{d}) amounts to a branch set equation, i.e.,

(){[N]}=(1)(d).\mathcal{B}({\bm{\pitchfork}})\setminus\{[N]\}=\mathcal{B}({\bm{\pitchfork}}_{1})\sqcup\cdots\sqcup\mathcal{B}({\bm{\pitchfork}}_{d})~{}.

In particular, each λ(){[N]}\lambda\in\mathcal{B}({\bm{\pitchfork}})\setminus\{[N]\} is contained in exactly one (j)\mathcal{B}({\bm{\pitchfork}}_{j}), and deg(λ)=degj(λ)\deg_{\bm{\pitchfork}}(\lambda)=\deg_{{\bm{\pitchfork}}_{j}}(\lambda) by Definition 1.6 in this case. These facts allow the sum over I1××Id\mathcal{R}_{I_{1}}\times\cdots\times\mathcal{R}_{I_{d}} above to be rewritten as a sum over all N{\bm{\pitchfork}}\in\mathcal{R}_{N} with 1={Ik1,,Ikq}{\pitchfork}_{1}=\{I_{k_{1}},\dots,I_{k_{q}}\}, and each product over λ(k1)(kd)\lambda\in\mathcal{B}({\bm{\pitchfork}}_{k_{1}})\sqcup\cdots\sqcup\mathcal{B}({\bm{\pitchfork}}_{k_{d}}) inside it is simply a product over λ(){[N]}\lambda\in\mathcal{B}({\bm{\pitchfork}})\setminus\{[N]\}. We conclude that an ordered partition (I0,,Iq)[N](I_{0},\dots,I_{q})\vdash[N] with I0,,Iq[N]I_{0},\dots,I_{q}\subsetneq[N] contributes the quantity

k=0q𝒵Ik(P,𝒔)=1qNN1={Ik1,,Ikd}λ(){[N]}(q1)deg(λ)1qeλ(𝒔)1\prod_{k=0}^{q}\mathcal{Z}_{I_{k}}(P,\bm{s})=\frac{1}{q^{N}}\sum_{\begin{subarray}{c}{\bm{\pitchfork}}\in\mathcal{R}_{N}\\ {\pitchfork}_{1}=\{I_{k_{1}},\dots,I_{k_{d}}\}\end{subarray}}\prod_{\lambda\in\mathcal{B}({\bm{\pitchfork}})\setminus\{[N]\}}\frac{(q-1)_{\deg_{\bm{\pitchfork}}(\lambda)-1}}{q^{e_{\lambda}(\bm{s})}-1} (3.2.3)

to the sum in (3.2.1), where {Ik1,,Ikd}\{I_{k_{1}},\dots,I_{k_{d}}\} is the (unordered) subset of nonempty parts in that particular ordered partition. We must now total the contribution in (3.2.3) over all possible (I0,,Iq)[N](I_{0},\dots,I_{q})\vdash[N] with I0,,Iq[N]I_{0},\dots,I_{q}\subsetneq[N]. Given a partition {λ1,,λd}[N]\{\lambda_{1},\dots,\lambda_{d}\}\vdash[N] with d2d\geq 2, note that there are precisely (q+1)d=(q+1)(q)d1(q+1)_{d}=(q+1)\cdot(q)_{d-1} ordered partitions (I0,,Iq)[N](I_{0},\dots,I_{q})\vdash[N] such that {Ik1,,Ikd}={λ1,,λd}\{I_{k_{1}},\dots,I_{k_{d}}\}=\{\lambda_{1},\dots,\lambda_{d}\}. Therefore summing (3.2.3) over all (I0,,Iq)[N](I_{0},\dots,I_{q})\vdash[N] with I0,,Iq[N]I_{0},\dots,I_{q}\subsetneq[N] gives

(I0,,Iq)[N]I0,,Iq[N]k=0q𝒵Ik(P,𝒔)\displaystyle\sum_{\begin{subarray}{c}(I_{0},\dots,I_{q})\vdash[N]\\ I_{0},\dots,I_{q}\subsetneq[N]\end{subarray}}\prod_{k=0}^{q}\mathcal{Z}_{I_{k}}(P,\bm{s}) =1qN(I0,,Iq)[N]I0,,Iq[N]N1={Ik1,,Ikd}λ(){[N]}(q1)deg(λ)1qeλ(𝒔)1\displaystyle=\frac{1}{q^{N}}\sum_{\begin{subarray}{c}(I_{0},\dots,I_{q})\vdash[N]\\ I_{0},\dots,I_{q}\subsetneq[N]\end{subarray}}\sum_{\begin{subarray}{c}{\bm{\pitchfork}}\in\mathcal{R}_{N}\\ {\pitchfork}_{1}=\{I_{k_{1}},\dots,I_{k_{d}}\}\end{subarray}}\prod_{\lambda\in\mathcal{B}({\bm{\pitchfork}})\setminus\{[N]\}}\frac{(q-1)_{\deg_{\bm{\pitchfork}}(\lambda)-1}}{q^{e_{\lambda}(\bm{s})}-1}
=q+1qN{λ1,,λd}[N]d2(q)d1N1={λ1,,λd}λ(){[N]}(q1)deg(λ)1qeλ(𝒔)1.\displaystyle=\frac{q+1}{q^{N}}\sum_{\begin{subarray}{c}\{\lambda_{1},\dots,\lambda_{d}\}\vdash[N]\\ d\geq 2\end{subarray}}(q)_{d-1}\sum_{\begin{subarray}{c}{\bm{\pitchfork}}\in\mathcal{R}_{N}\\ {\pitchfork}_{1}=\{\lambda_{1},\dots,\lambda_{d}\}\end{subarray}}\prod_{\lambda\in\mathcal{B}({\bm{\pitchfork}})\setminus\{[N]\}}\frac{(q-1)_{\deg_{\bm{\pitchfork}}(\lambda)-1}}{q^{e_{\lambda}(\bm{s})}-1}~{}.

Given a partition {λ1,,λd}[N]\{\lambda_{1},\dots,\lambda_{d}\}\vdash[N], those splitting chains N{\bm{\pitchfork}}\in\mathcal{R}_{N} with 1={λ1,,λd}{\pitchfork}_{1}=\{\lambda_{1},\dots,\lambda_{d}\} all have deg([N])=#1=d\deg_{\bm{\pitchfork}}([N])=\#{\pitchfork}_{1}=d by Definition 1.6. Moreover, no N{\bm{\pitchfork}}\in\mathcal{R}_{N} is missed or repeated in the sum of sums above, so it can be rewritten as

(I0,,Iq)[N]I0,,Iq[N]k=0q𝒵Ik(P,𝒔)\displaystyle\sum_{\begin{subarray}{c}(I_{0},\dots,I_{q})\vdash[N]\\ I_{0},\dots,I_{q}\subsetneq[N]\end{subarray}}\prod_{k=0}^{q}\mathcal{Z}_{I_{k}}(P,\bm{s}) =q+1qNN(q)deg([N])1λ(){[N]}(q1)deg(λ)1qeλ(𝒔)1\displaystyle=\frac{q+1}{q^{N}}\sum_{{\bm{\pitchfork}}\in\mathcal{R}_{N}}(q)_{\deg_{\bm{\pitchfork}}([N])-1}\prod_{\lambda\in\mathcal{B}({\bm{\pitchfork}})\setminus\{[N]\}}\frac{(q-1)_{\deg_{\bm{\pitchfork}}(\lambda)-1}}{q^{e_{\lambda}(\bm{s})}-1}
=q+1qNN(q)deg([N])1(q1)deg([N])1(qe[N](𝒔)1)λ()(q1)deg(λ)1qeλ(𝒔)1\displaystyle=\frac{q+1}{q^{N}}\sum_{{\bm{\pitchfork}}\in\mathcal{R}_{N}}\frac{(q)_{\deg_{\bm{\pitchfork}}([N])-1}}{(q-1)_{\deg_{\bm{\pitchfork}}([N])-1}}\cdot(q^{e_{[N]}(\bm{s})}-1)\prod_{\lambda\in\mathcal{B}({\bm{\pitchfork}})}\frac{(q-1)_{\deg_{\bm{\pitchfork}}(\lambda)-1}}{q^{e_{\lambda}(\bm{s})}-1}
=q+1qNNqN+i<jsijqq+1deg([N])λ()(q1)deg(λ)1qeλ(𝒔)1.\displaystyle=\frac{q+1}{q^{N}}\sum_{{\bm{\pitchfork}}\in\mathcal{R}_{N}}\frac{q^{N+\sum_{i<j}s_{ij}}-q}{q+1-\deg_{\bm{\pitchfork}}([N])}\prod_{\lambda\in\mathcal{B}({\bm{\pitchfork}})}\frac{(q-1)_{\deg_{\bm{\pitchfork}}(\lambda)-1}}{q^{e_{\lambda}(\bm{s})}-1}~{}.

Note that the summand for each N{\bm{\pitchfork}}\in\mathcal{R}_{N} is still defined for any prime power qq since the denominators (q1)deg([N])1(q-1)_{\deg_{\bm{\pitchfork}}([N])-1} and q+1deg([N])q+1-\deg_{\bm{\pitchfork}}([N]) (which vanish when q=deg([N])1q=\deg_{\bm{\pitchfork}}([N])-1) are cancelled by the numerator (q1)deg([N])1(q-1)_{\deg_{\bm{\pitchfork}}([N])-1} appearing in the product over λ()\lambda\in\mathcal{B}({\bm{\pitchfork}}). Finally, we evaluate the righthand side of (3.2.1) by combining the sum directly above with that in (3.2.2) and multiplying through by (qq+1)N(\frac{q}{q+1})^{N}. This yields the desired formula for 𝒵N(1(K),𝒔)\mathcal{Z}_{N}(\mathbb{P}^{1}(K),\bm{s}):

𝒵N(1(K),𝒔)\displaystyle\mathcal{Z}_{N}(\mathbb{P}^{1}(K),\bm{s}) =1(q+1)N1N(1+qN+i<jsijqq+1deg([N]))λ()(q1)deg(λ)1qeλ(𝒔)1\displaystyle=\frac{1}{(q+1)^{N-1}}\sum_{{\bm{\pitchfork}}\in\mathcal{R}_{N}}\left(1+\frac{q^{N+\sum_{i<j}s_{ij}}-q}{q+1-\deg_{\bm{\pitchfork}}([N])}\right)\prod_{\lambda\in\mathcal{B}({\bm{\pitchfork}})}\frac{(q-1)_{\deg_{\bm{\pitchfork}}(\lambda)-1}}{q^{e_{\lambda}(\bm{s})}-1}
=1(q+1)N1NqN+i<jsij+1deg([N])q+1deg([N])λ()(q1)deg(λ)1qeλ(𝒔)1\displaystyle=\frac{1}{(q+1)^{N-1}}\sum_{{\bm{\pitchfork}}\in\mathcal{R}_{N}}\frac{q^{N+\sum_{i<j}s_{ij}}+1-\deg_{\bm{\pitchfork}}([N])}{q+1-\deg_{\bm{\pitchfork}}([N])}\prod_{\lambda\in\mathcal{B}({\bm{\pitchfork}})}\frac{(q-1)_{\deg_{\bm{\pitchfork}}(\lambda)-1}}{q^{e_{\lambda}(\bm{s})}-1}

3.3 Finishing the proof of Theorem 2.3

Our final task is to prove the (q+1)(q+1)th Power Law, which we noted in Section 2.2 is equivalent to the equations in (2.2.3). That is, it remains to prove

ZN(1(K),β)N!=N0++Nq=NN0,,Nq0k=0q(qq+1)NkZNk(P,β)Nk!for all β>0 and N0.\frac{Z_{N}(\mathbb{P}^{1}(K),\beta)}{N!}=\sum_{\begin{subarray}{c}N_{0}+\dots+N_{q}=N\\ N_{0},\dots,N_{q}\geq 0\end{subarray}}\prod_{k=0}^{q}\left(\frac{q}{q+1}\right)^{N_{k}}\frac{Z_{N_{k}}(P,\beta)}{N_{k}!}\qquad\text{for all $\beta>0$ and $N\geq 0$}.
Proof.

Fix N0N\geq 0 and β>0\beta>0, and fix 𝒔\bm{s} via sij=βs_{ij}=\beta for all i<ji<j, so that 𝒵N(1(K),𝒔)=ZN(1(K),β)\mathcal{Z}_{N}(\mathbb{P}^{1}(K),\bm{s})=Z_{N}(\mathbb{P}^{1}(K),\beta) and 𝒵I(P,𝒔)=Z#I(P,β)\mathcal{Z}_{I}(P,\bm{s})=Z_{\#I}(P,\beta) for any subset I[N]I\subset[N]. The formula in Theorem 3.4 relates these functions of β\beta via

ZN(1(K),β)\displaystyle Z_{N}(\mathbb{P}^{1}(K),\beta) =𝒵N(1(K),𝒔)\displaystyle=\mathcal{Z}_{N}(\mathbb{P}^{1}(K),\bm{s})
=(qq+1)N(I0,,Iq)[N]k=0q𝒵#Ik(P,𝒔)\displaystyle=\left(\frac{q}{q+1}\right)^{N}\sum_{(I_{0},\dots,I_{q})\vdash[N]}\prod_{k=0}^{q}\mathcal{Z}_{\#I_{k}}(P,\bm{s})
=(I0,,Iq)[N]k=0q(qq+1)#IkZ#Ik(P,β).\displaystyle=\sum_{(I_{0},\dots,I_{q})\vdash[N]}\prod_{k=0}^{q}\left(\frac{q}{q+1}\right)^{\#I_{k}}Z_{\#I_{k}}(P,\beta)~{}.

For each choice of q+1q+1 ordered integers N0,,Nq0N_{0},\dots,N_{q}\geq 0 satisfying N0++Nq=NN_{0}+\dots+N_{q}=N, there are precisely

(NN0,,Nq)=N!N0!Nq!\binom{N}{N_{0},\dots,N_{q}}=\frac{N!}{N_{0}!\cdots N_{q}!}

ordered partitions (I0,,Iq)[N](I_{0},\dots,I_{q})\vdash[N] satisfying #I0=N0,,#Iq=Nq\#I_{0}=N_{0},\dots,\#I_{q}=N_{q}. Finally, grouping ordered partitions according to all possible ordered integer choices establishes the desired equation:

ZN(1(K),𝒔)N!\displaystyle\frac{Z_{N}(\mathbb{P}^{1}(K),\bm{s})}{N!} =1N!(I0,,Iq)[N]k=0q(qq+1)#IkZ#Ik(P,β)\displaystyle=\frac{1}{N!}\cdot\sum_{(I_{0},\dots,I_{q})\vdash[N]}\prod_{k=0}^{q}\left(\frac{q}{q+1}\right)^{\#I_{k}}Z_{\#I_{k}}(P,\beta)
=1N!N0++Nq=NN0,,Nq0(NN0,,Nq)k=0q(qq+1)NkZNk(P,β)\displaystyle=\frac{1}{N!}\cdot\sum_{\begin{subarray}{c}N_{0}+\dots+N_{q}=N\\ N_{0},\dots,N_{q}\geq 0\end{subarray}}\binom{N}{N_{0},\dots,N_{q}}\prod_{k=0}^{q}\left(\frac{q}{q+1}\right)^{N_{k}}Z_{N_{k}}(P,\beta)
=N0++Nq=NN0,,Nq0k=0q(qq+1)NkZNk(P,β)Nk!.\displaystyle=\sum_{\begin{subarray}{c}N_{0}+\dots+N_{q}=N\\ N_{0},\dots,N_{q}\geq 0\end{subarray}}\prod_{k=0}^{q}\left(\frac{q}{q+1}\right)^{N_{k}}\frac{Z_{N_{k}}(P,\beta)}{N_{k}!}~{}.

Acknowledgements: I would like to thank Clay Petsche for confirming several details about the measure and metric on 1(K)\mathbb{P}^{1}(K), and I would like to thank Chris Sinclair for many useful suggestions regarding the Power Laws and the Quadratic Recurrence.

References

  • [BGMR06] Joe Buhler, Daniel Goldstein, David Moews, and Joel Rosenberg, The probability that a random monic pp-adic polynomial splits, Experiment. Math. 15 (2006), no. 1, 21–32. MR 2229382
  • [DM91] Jan Denef and Diane Meuser, A functional equation of Igusa’s local zeta function, Amer. J. Math. 113 (1991), no. 6, 1135–1152. MR 1137535
  • [FP15] Paul Fili and Clayton Petsche, Energy integrals over local fields and global height bounds, Int. Math. Res. Not. IMRN (2015), no. 5, 1278–1294. MR 3340356
  • [FW08] Peter J. Forrester and S. Ole Warnaar, The importance of the Selberg integral, Bull. Amer. Math. Soc. (N.S.) 45 (2008), no. 4, 489–534. MR 2434345
  • [Sin20] Christopher D. Sinclair, Non-Archimedean Electrostatics, arXiv e-prints (2020), arXiv:2002.07121.
  • [Vol10] Christopher Voll, Functional equations for zeta functions of groups and rings, Ann. of Math. (2) 172 (2010), no. 2, 1181–1218. MR 2680489
  • [Web21a] J. Webster, The combinatorics of log-Coulomb gases in pp-fields, Ph.D. thesis, Eugene, OR, 2021.
  • [Web21b] Joe Webster, log-Coulomb gas with norm-density in pp-fields, P-Adic Num Ultrametr Anal Appl 13 (2021), no. 1, 1–43.
  • [Wei95] André Weil, Basic number theory, Classics in Mathematics, Springer-Verlag, Berlin, 1995, Reprint of the second (1973) edition. MR 1344916

 

Joe Webster
Department of Mathematics, University of Virginia, Charlottesville, VA 22903

email: [email protected]