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

Limiting distributions of generalized money exchange models

Hironobu Sakagawa Department of Mathematics, Faculty of Science and Technology, Keio University, 3-14-1 Hiyoshi, Kouhoku-ku, Yokohama 223-8522, JAPAN. E-mail address: [email protected]
Abstract

The “Money Exchange Model” is a type of agent-based simulation model used to study how wealth distribution and inequality evolve through monetary exchanges between individuals. The primary focus of this model is to identify the limiting wealth distributions that emerge at the macroscopic level, given the microscopic rules governing the exchanges among agents. In this paper, we formulate generalized versions of the immediate exchange model and the uniform saving model both of which are types of money exchange models, as discrete-time interacting particle systems and characterize their stationary distributions. Furthermore, we prove that under appropriate scaling, the asymptotic wealth distribution converges to a Gamma distribution or an exponential distribution for both models. The limiting distribution depends on the weight function that affects the probability distribution of the number of coins exchanged by each agent. In particular, our results provide a mathematically rigorous formulation and generalization of the assertions previously predicted in studies based on numerical simulations and heuristic arguments.

Key words. Econophysics, interacting particle system, stationary distribution, equivalence of ensembles, local limit theorem.

2020 Mathematics Subject Classification: 60K35, 60F99, 91B80.

1 Introduction

1.1 Money exchange models

The “Money Exchange Model” is a type of agent-based simulation model used to study how wealth distribution and inequality evolve through monetary exchanges between individuals. This model has been extensively studied in the field of econophysics, particularly by applying ideas from statistical physics, where the money exchange is viewed as analogous to the transfer of energy or particles.

Consider an economy consisting of a finite number of agents. The typical process proceeds as follows:

  1. (1)

    Initial conditions: Assign each agent a random or equal amount of money.

  2. (2)

    Selecting pairs of agents: Randomly select pairs of agents who will exchange money.

  3. (3)

    Money exchange: The selected pair exchanges money according to specific rules.

  4. (4)

    Iteration: Repeat (2) and (3) multiple times and observe how the distribution of money in the system changes.

The primary focus of this model is to identify the limiting wealth distributions that emerge at the macroscopic level, given the microscopic rules governing the exchanges among agents. Depending on the rules of exchange, the limiting distribution is often expected to take the form of a Gamma or an exponential distribution. In the field of econophysics, various studies have been conducted through numerical simulations and other methods (cf. [2], [24] and references theirin). However, mathematically rigorous studies of these models remain relatively few.

We begin by describing several specific models. The first one is the immediate exchange model proposed in [10]. In this model, the wealth of each agent is represented by a real-valued variable. At each time step, two agents are randomly chosen and give a random fraction of their wealth to each other. The fraction is determined by independent uniformly distributed random variables on the interval [0,1][0,1]. [10] studied the model through numerical simulations, and later, [11] analytically explored the infinite population version. As a more realistic microscopic model that includes spatial structure in the form of local interactions, [14] formulated the corresponding discrete version as an interacting particle system. We briefly explain their model and result. Consider a finite connected graph 𝒢=(𝒱,)\mathcal{G}=(\mathcal{V},\mathcal{E}). Each site x𝒱x\in\mathcal{V} corresponds to an agent, and the economy is represented by the set of agents 𝒱\mathcal{V}. The population size is given by |𝒱|=N|\mathcal{V}|=N. The amount of money each agent holds is represented by the number of coins, where Mn(x)M_{n}(x) represents the number of coins that agent x𝒱x\in\mathcal{V} possesses at time nn. The edge set \mathcal{E} represents a social network in which only agents connected by an edge can interact to exchange coins. At each time step, an edge e={x,y}e=\{x,y\}\in\mathcal{E} is chosen uniformly at random from \mathcal{E}. Agents xx and yy independently and uniformly select a random number of their coins to give to each other. Thus, {(Mn(x))x𝒱}n0\{(M_{n}(x))_{x\in\mathcal{V}}\}_{n\geq 0} constitutes a time-homogeneous Markov chain. In particular, the total number of coins is conserved in this process, and we denote this total by LL. [14] proved that the stationary distribution μN,L\mu_{N,L} for this Markov chain uniquely exists and gave the following explicit representation.

limnP(Mn(x)=c)=μN,L({ξ+𝒱;ξ(x)=c})=(c+1)(Lc+2N32N3)(L+2N12N1),\displaystyle\begin{split}\lim_{n\to\infty}P(M_{n}(x)=c)&=\mu_{N,L}\bigl{(}\{\xi\in\mathbb{Z}_{+}^{\mathcal{V}};\xi(x)=c\}\bigr{)}\\ &=\frac{(c+1)\Bigl{(}\begin{array}[]{c}L-c+2N-3\\ 2N-3\end{array}\Bigr{)}}{\Bigl{(}\begin{array}[]{c}L+2N-1\\ 2N-1\end{array}\Bigr{)}},\end{split} (1.1)

for every x𝒱x\in\mathcal{V} and c{0,1,2,,L}c\in\{0,1,2,\cdots,L\}. The first equality is a consequence of the Markov chain convergence theorem. Then, by applying formal calculations to the right-hand side, the authors obtained the following approximation.

limnP(Mn(x)=c)4cT2e2cT,\lim_{n\to\infty}P(M_{n}(x)=c)\approx\frac{4c}{T^{2}}e^{-\frac{2c}{T}}, (1.2)

for large enough NN and T=LNT=\frac{L}{N}. In this context TT represents the average number of coins per agent and is called money temperature by analogy with the temperature in physics and the limit NN\to\infty and T=LNT=\frac{L}{N}\to\infty are called large population and large money temperature limit. The right-hand side of (1.2) is a probability density function of the Gamma distribution with mean TT and shape parameter two, and this is consistent with the predicted result by [10] and [11]. As related results, [4] and [9] examined the duality between the real-valued model and the discrete state version in a continuous time setting. In addition, [9] considered the case where the exchange fraction is determined by a Beta distribution.

We also present two models with different exchange rules that we address in this paper.

Uniform saving model: Agents xx and yy independently save a random number of their coins according to a uniform distribution. The remaining coins are then pooled and uniformly redistributed between the two agents. As for the immediate exchange model, the limiting distribution for this model is predicted to be a Gamma distribution with shape parameter two (cf. [3], [19]), and [14] obtained the same approximation as (1.2).

Uniform reshuffling model: All the coins agents xx and yy possess are pooled and uniformly redistributed between the two agents. Different from the above two models, the limiting distribution for this model is predicted to be an exponential distribution with mean TT (cf. [6]). [14] obtained the corresponding approximation of the same form as (1.2).

So far, [14] has formulated several money exchange models as Markov chains with spatial structure, characterized their stationary distributions and further obtained their formal approximations. The conclusions explain the distribution of wealth in a sense, as predicted by numerical simulations and other methods. However, the approximation ``"``\approx" in (1.2) is not mathematically valid. The left-hand side is defined only for non-negative integers cc, while the right-hand represents a probability density function on +\mathbb{R}_{+}. The derivation of the approximation (1.2), although merely a formal calculation, seems to depend on assumptions that are not fully clarified. Specifically, in the proofs of Theorems 1, 2 and 3 in [14], L+1,L+2,,L+NL+1,L+2,\cdots,L+N are replaced with LL for sufficiently large LL and NN satisfying NLN\ll L. On the other hand, Lc+1,Lc+2,,Lc+NL-c+1,L-c+2,\cdots,L-c+N are replaced with LcL-c, rather than LL, even when cLc\ll L. Additionally, c+1c+1 has been conveniently replaced with cc. As a matter of fact, it seems unnatural to consider the approximation of μN,L({ξ+𝒱;ξ(x)=c})\mu_{N,L}(\{\xi\in\mathbb{Z}_{+}^{\mathcal{V}};\xi(x)=c\}) for each cc since the average number of coins per agent diverges in the limit T=LNT=\frac{L}{N}\to\infty. The money exchange model describes the microscopic movement of money, however, our goal is to derive the macroscopic distribution of wealth in the limit NN\to\infty and LN\frac{L}{N}\to\infty. In order to do that, we should analyze the convergence of μN,L\mu_{N,L} under appropriate scaling. Therefore, the primary objective of this paper is to provide a mathematically rigorous justification of the approximation (1.2) and to clearly demonstrate the convergence of the wealth distribution. Furthermore, we generalize the rules of money exchange as follows:

  • Randomly select the number of coins to pass or save based on a probability distribution that depends on the number of coins, rather than using the uniform distribution.

  • Allow the exchange or redistribution of coins among randomly selected groups of three or more agents.

These generalizations appear natural from both mathematical and economic perspectives. We note that as a generalization of the exchange rules in the immediate exchange model, [23] considered a broader class of models where mass is split, exchanged and merged. In [13], [15] and [16], the authors formulated other money exchange models as Markov chains and studied their stationary distributions and formal approximations in a manner similar to (1.1) and (1.2). Also, the mixing time has been studied recently for the binomial splitting model and the symmetric beta-binomial splitting model, which are variations of the uniform reshuffling model (cf. [21], [22]).

Before introducing our models and results we prepare several notations. In the following, +={0,1,2,}\mathbb{Z}_{+}=\{0,1,2,\cdots\} denotes the set of non-negative integers and ={1,2,3,}\mathbb{N}=\{1,2,3,\cdots\} denotes the set of positive integers. [a][a] denotes the integral part of a>0a>0 and we set ab:=max{a,b}a\vee b:=\max\{a,b\} and ab:=min{a,b}a\wedge b:=\min\{a,b\} for a,ba,b\in\mathbb{R}. For a finite set AA, |A||A| denotes its cardinality. For two sequences of positive numbers {an}\{a_{n}\} and {bn}\{b_{n}\}, anbna_{n}\sim b_{n} means that limnanbn=1\lim\limits_{n\to\infty}\frac{a_{n}}{b_{n}}=1. A function ff on \mathbb{R}^{\mathbb{Z}} is called local if it depends only on finitely many coordinates. For a probability measure μ\mu, Eμ[]E^{\mu}[\,\cdot\,] denotes the expectation with respect to μ\mu.

1.2 Model description and results

Let us state our model precisely. We adopt the standard notations commonly used in the study of interacting particle systems. For AA\subset\mathbb{Z} and LL\in\mathbb{N}, we define the configuration space

Ω(A,L)={η={η(x)}xA+A;xAη(x)=L}.\Omega(A,L)=\bigl{\{}\eta=\{\eta(x)\}_{x\in A}\in\mathbb{Z}_{+}^{A};\sum\limits_{x\in A}\eta(x)=L\bigr{\}}.

When A=ΛN:={1,2,,N}A=\Lambda_{N}:=\{1,2,\cdots,N\}, Ω(A,L)\Omega(A,L) is denoted by ΩN(L)\Omega_{N}(L). Consider now an economy populated by many agents. Each site xΛNx\in\Lambda_{N} corresponds to an agent and we assume that the economy can be represented by the set of agents ΛN\Lambda_{N}. NN corresponds to the population size. For each ηΩN(L)\eta\in\Omega_{N}(L), we interpret η(x)\eta(x), xΛNx\in\Lambda_{N} not as the number of particles, but as the number of coins held by agent xx. ρ\rho denotes a probability distribution on 𝒟N={AΛN;|A|2}\mathcal{D}_{N}=\{A\subset\Lambda_{N};|A|\geq 2\}, namely ρ(A)0\rho(A)\geq 0 for every AΛNA\subset\Lambda_{N} with |A|2|A|\geq 2 and A𝒟Nρ(A)=1\sum\limits_{A\in\mathcal{D}_{N}}\rho(A)=1. This represents the distribution that determines which agent handles the money exchange at each time step. We also take a non-negative function g(0)g(\not\equiv 0) defined on +\mathbb{Z}_{+}. Now, we introduce three money exchange models.

Immediate exchange model: Let {Xn}n0\{X_{n}\}_{n\geq 0} be a time-homogeneous Markov chain on the state space ΩN(L)\Omega_{N}(L). For given Xn=ξΩN(L)X_{n}=\xi\in\Omega_{N}(L), the configuration Xn+1X_{n+1} is determined from the following rule.

  1. (1)

    Choose a set A𝒟NA\in\mathcal{D}_{N} according to the distribution ρ\rho.

  2. (2)

    For given ξ\xi, let {c(x)}xA\{c(x)\}_{x\in A} be independent random variables whose distributions are given by

    P(c(x)=k)=1G(ξ(x))g(k),k{0,1,2,,ξ(x)},P(c(x)=k)=\frac{1}{G(\xi(x))}g(k),\ k\in\{0,1,2,\cdots,\xi(x)\}, (1.3)

    for xAx\in A where we set G(k)=j=0kg(j)G(k)=\sum\limits_{j=0}^{k}g(j), k+k\in\mathbb{Z}_{+}.

  3. (3)

    Choose a permutation σ𝒮A\sigma\in\mathcal{S}_{A} uniformly random, namely with probability 1|𝒮A|=1|A|!\frac{1}{|\mathcal{S}_{A}|}=\frac{1}{|A|!} where 𝒮A\mathcal{S}_{A} denotes the set of all permutations of AA.

  4. (4)

    For given ξ\xi and realizations AA, {c(x)}xA\{c(x)\}_{x\in A} and σ\sigma, define Xn+1X_{n+1} by

    Xn+1(z)={ξ(z)c(z)+c(σ1(z)) if zA,ξ(z) if zA.X_{n+1}(z)=\begin{cases}\xi(z)-c(z)+c(\sigma^{-1}(z))&\text{ if }z\in A,\\ \xi(z)&\text{ if }z\notin A.\\ \end{cases}

This dynamics can be interpreted as follows: At each time step, a money exchange occurs between agents in a randomly chosen set AA. c(x)c(x) represents the number of coins that agent xx transfers, which is determined by a probability distribution dependent on the weight function gg and ξ(x)\xi(x), the number of coins that agent xx possesses. According to a randomly chosen permutation σ𝒮A\sigma\in\mathcal{S}_{A}, each agent xAx\in A passes c(x)c(x) coins to agent σ(x)A\sigma(x)\in A. In this process, the total number of coins remains conserved. It is worth noting that the model studied in [14] corresponds to the case where gg is a constant function and ρ\rho is the uniform distribution on an edge set of ΛN\Lambda_{N}. In this case, the distribution (1.3) matches the uniform distribution on {0,1,2,,ξ(x)}\{0,1,2,\cdots,\xi(x)\} and the money exchange occurs between two agents connected by an edge in ΛN\Lambda_{N}. To be more precise, since the permutation σ𝒮A\sigma\in\mathcal{S}_{A} can include the identity permutation, our model can be viewed as the lazy version of their model in this case.

Random saving model: Let {Yn}n0\{Y_{n}\}_{n\geq 0} be a time-homogeneous Markov chain on the state space ΩN(L)\Omega_{N}(L). For given Yn=ξY_{n}=\xi, the configuration Yn+1Y_{n+1} is determined from the following rule.

  1. (1)

    Choose a set A𝒟NA\in\mathcal{D}_{N} according to the distribution ρ\rho.

  2. (2)

    For given ξ\xi, let {c(x)}xA\{c(x)\}_{x\in A} be independent random variables whose distributions are given by

    P(c(x)=k)=1G(ξ(x))g(k),k{0,1,2,,ξ(x)},P(c(x)=k)=\frac{1}{G(\xi(x))}g(k),\ k\in\{0,1,2,\cdots,\xi(x)\},

    for xAx\in A.

  3. (3)

    For given ξ\xi and c={c(x)}xAc=\{c(x)\}_{x\in A}, choose a configuration d={d(x)}xAΩ(A,SA(ξ)SA(c))d=\{d(x)\}_{x\in A}\in\Omega(A,S_{A}(\xi)-S_{A}(c)) uniformly random, namely with probability 1|Ω(A,SA(ξ)SA(c))|\frac{1}{|\Omega(A,S_{A}(\xi)-S_{A}(c))|} where we set SA(ξ)=xAξ(x)S_{A}(\xi)=\sum\limits_{x\in A}\xi(x) for {ξ(x)}xA\{\xi(x)\}_{x\in A}.

  4. (4)

    For given ξ\xi and realizations AA, {c(x)}xA\{c(x)\}_{x\in A} and {d(x)}xA\{d(x)\}_{x\in A}, define Yn+1Y_{n+1} by

    Yn+1(z)={c(z)+d(z) if zA,ξ(z) if zA.Y_{n+1}(z)=\begin{cases}c(z)+d(z)&\text{ if }z\in A,\\ \xi(z)&\text{ if }z\notin A.\\ \end{cases}

Note that in contrast to the immediate exchange model, the random variable c(x)c(x) represents how many coins the agent xx to save. When a money exchange occurs within set AA, each agent xAx\in A offers ξ(x)c(x)\xi(x)-c(x) coins. These coins are then pooled and redistributed among the agents in AA according to the uniform distribution. Similar to the immediate exchange model, [14] studied the case where gg is a constant function and ρ\rho is the uniform distribution on an edge set. In that model, c(x)c(x) is drawn from the uniform distribution on {0,1,2,,ξ(x)}\{0,1,2,\cdots,\xi(x)\}, and it is referred to as the uniform saving model. Since our model considers a more general distribution for c(x)c(x), we refer to it as the random saving model. As a special case, when the weight function gg is defined by g(k)=δ0(k)g(k)=\delta_{0}(k) for k+k\in\mathbb{Z}_{+}, each agent saves no money. When a money exchange occurs within set AA, all coins held by the agents in AA are pooled and redistributed. This model is referred to as the uniform reshuffling model, and we denote it by {Zn}n0\{Z_{n}\}_{n\geq 0}.

As the first result, we characterize the stationary distributions of these Markov chains.

Proposition 1.1.

Let N,LN,L\in\mathbb{N} be fixed. We assume that the hypergraph (ΛN,𝒟N,ρ)(\Lambda_{N},\mathcal{D}_{N,\rho}) is connected where the hyperedge set 𝒟N,ρ\mathcal{D}_{N,\rho} is defined by 𝒟N,ρ={AΛN;|A|2,ρ(A)>0}\mathcal{D}_{N,\rho}=\{A\subset\Lambda_{N};|A|\geq 2,\rho(A)>0\}.

  1. (i)(i)

    Assume that the weight function g:+[0,)g:\mathbb{Z}_{+}\to[0,\infty) satisfies g(0)>0g(0)>0 and g(1)>0g(1)>0. Then, there is a unique stationary distribution μN,L\mu_{N,L} for {Xn}n0\{X_{n}\}_{n\geq 0} and it is given by

    μN,L(ξ)=1ZN,LxΛNG(ξ(x)),ξΩN(L),\mu_{N,L}(\xi)=\frac{1}{Z_{N,L}}\prod_{x\in\Lambda_{N}}\!\!G(\xi(x)),\ \xi\in\Omega_{N}(L), (1.4)

    where G(k)=j=0kg(j),k+G(k)=\sum\limits_{j=0}^{k}g(j),\ k\in\mathbb{Z}_{+} and ZN,L=ξΩN(L)xΛNG(ξ(x))Z_{N,L}=\sum\limits_{\xi\in\Omega_{N}(L)}\prod\limits_{x\in\Lambda_{N}}G(\xi(x)) is the normalization factor. In particular, for every ηΩN(L)\eta\in\Omega_{N}(L), xΛNx\in\Lambda_{N} and k{0,1,2,,L}k\in\{0,1,2,\cdots,L\}, it holds that

    limnPη(Xn(x)=k)=μN,L(ξ(x)=k)=ZN1,LkG(k)ZN,L,\displaystyle\lim_{n\to\infty}P_{\eta}\bigl{(}X_{n}(x)=k\bigr{)}=\mu_{N,L}\bigl{(}\xi(x)=k\bigr{)}=\frac{Z_{N-1,L-k}\,G(k)}{Z_{N,L}},

    where PηP_{\eta} represents the law of {Xn}n0\{X_{n}\}_{n\geq 0} with the initial condition X0=ηX_{0}=\eta.

  2. (ii)(ii)

    Assume that the weight function g:+[0,)g:\mathbb{Z}_{+}\to[0,\infty) satisfies g(0)>0g(0)>0. Then, the exact same statement as (i)(i) applies to {Yn}n0\{Y_{n}\}_{n\geq 0} instead of {Xn}n0\{X_{n}\}_{n\geq 0}. In particular, for the uniform reshuffling model {Zn}n0\{Z_{n}\}_{n\geq 0}, there is a unique stationary distribution πN,L\pi_{N,L} and it is given by the uniform distribution on ΩN(L)\Omega_{N}(L), namely,

    πN,L(ξ)=1|ΩN(L)|,ξΩN(L).\pi_{N,L}(\xi)=\frac{1}{|\Omega_{N}(L)|},\ \xi\in\Omega_{N}(L).

By this proposition we can see that both the immediate exchange model and the random saving model have the same stationary distribution under the condition g(0)>0g(0)>0 and g(1)>0g(1)>0. Also, the choice of ρ\rho does not affect the stationary distribution.

Remark 1.1.

The condition g(0)>0g(0)>0 is always needed to make the measure (1.4) well-defined.

Remark 1.2.

Consider the case where the weight function gg is a constant function gγ>0g\equiv\gamma>0. Then, it holds that G(k)=γ(k+1)G(k)=\gamma(k+1), k+k\in\mathbb{Z}_{+} and the constant γ\gamma is canceled by the normalization factor in the definition of μN,L\mu_{N,L}. Therefore, we can take GG as G(k)=k+1G(k)=k+1, k+k\in\mathbb{Z}_{+} in (1.4) and in this case, the above result matches that of [14]. [14] also gave the explicit representation for ZN,LZ_{N,L} when g1g\equiv 1 and obtained (1.1).

Remark 1.3.

For the random saving model we can change the role of random variable c(x)c(x) to represent the number of coins to offer instead of the number of coins to save. Namely, we modify (3)(3) and (4)(4) in the definition of the random saving model as follows:

  1. (3)(3)^{\prime}

    For given ξ\xi and c={c(x)}xAc=\{c(x)\}_{x\in A}, choose a configuration d={d(x)}xAΩ(A,SA(c))d=\{d(x)\}_{x\in A}\in\Omega(A,S_{A}(c)) uniformly random, namely with probability 1|Ω(A,SA(c))|\frac{1}{|\Omega(A,S_{A}(c))|}.

  2. (4)(4)^{\prime}

    For given ξ\xi and realizations AA, {c(x)}xA\{c(x)\}_{x\in A} and {d(x)}xA\{d(x)\}_{x\in A}, define Yn+1Y_{n+1} by

    Yn+1(z)={ξ(z)c(z)+d(z) if zA,ξ(z) if zA.Y_{n+1}(z)=\begin{cases}\xi(z)-c(z)+d(z)&\text{ if }z\in A,\\ \xi(z)&\text{ if }z\notin A.\\ \end{cases}

By symmetry, the completely same proof for Proposition 1.1 below works well in this setting and the same result holds for this modified model.

Now, we are in the position to state the main result of this paper. To justify the limit NN\to\infty and LN\frac{L}{N}\to\infty, we assume that the total number of coins L=LNL=L_{N} satisfies limNLNNaN=T\lim\limits_{N\to\infty}\frac{L_{N}}{Na_{N}}=T for some T>0T>0 and divergent sequence {aN}N1\{a_{N}\}_{N\geq 1}. Then, we can prove that the law of the scaled field {1aNη(x)}xΛN\bigl{\{}\frac{1}{a_{N}}\eta(x)\bigr{\}}_{x\in\Lambda_{N}} under μN,L\mu_{N,L} or πN,L\pi_{N,L} converges to the i.i.d. product of probability distributions on +\mathbb{R}_{+}. Its marginal distribution depends on the asymptotic behavior of the weight function gg.

Theorem 1.1.

Let {LN}N1\{L_{N}\}_{N\geq 1} be a sequence of positive integers that satisfies limNLNNaN=T\lim\limits_{N\to\infty}\frac{L_{N}}{Na_{N}}=T for some positive constant T>0T>0 and a sequence {aN}N1\{a_{N}\}_{N\geq 1} which satisfies limNaN=\lim\limits_{N\to\infty}a_{N}=\infty. Assume that g(0)>0g(0)>0 and the following condition holds: There exist α\alpha\in\mathbb{R} and cα(0,)c_{\alpha}\in(0,\infty) such that limkg(k)kα=cα\lim\limits_{k\to\infty}\frac{g(k)}{k^{\alpha}}=c_{\alpha}. Then, for every bounded continuous local function f:Af:\mathbb{R}^{A}\to\mathbb{R}, it holds that

limNEμN,LN[f(aN)]=Eμ¯α,TA[f()],\displaystyle\lim_{N\to\infty}E^{\mu_{N,L_{N}}}\Bigl{[}f\bigl{(}\frac{\cdot}{a_{N}}\bigr{)}\Bigr{]}=E^{\overline{\mu}_{\alpha,T}^{A}}\bigl{[}f(\,\cdot\,)\bigr{]},

where AA is a finite subset of +\mathbb{Z}_{+} and μ¯α,TA\overline{\mu}_{\alpha,T}^{A} denotes the product probability measure on +A\mathbb{R}_{+}^{A} whose one site marginal distribution on +\mathbb{R}_{+} is given by

μα,T(dr)={1Γ(α+2)(α+2T)α+2rα+1eα+2Trdr if α>1,1Te1Trdr if α1.\displaystyle\mu_{\alpha,T}(dr)=\begin{cases}\frac{1}{\Gamma(\alpha+2)}\bigl{(}\frac{\alpha+2}{T}\bigr{)}^{\alpha+2}r^{\alpha+1}e^{-\frac{\alpha+2}{T}r}dr&\text{ if }\alpha>-1,\\ \frac{1}{T}e^{-\frac{1}{T}r}dr&\text{ if }\alpha\leq-1.\end{cases} (1.5)

Γ(β)=0rβ1er𝑑r\Gamma(\beta)=\int_{0}^{\infty}r^{\beta-1}e^{-r}dr is the Gamma function with parameter β>0\beta>0.

Also, if g:+[0,)g:\mathbb{Z}_{+}\to[0,\infty) satisfies g(0)>0g(0)>0 and j0g(j)<\sum\limits_{j\geq 0}g(j)<\infty, then the same conclusion as in the case α1\alpha\leq-1 above holds. In particular, for every bounded continuous local function f:Af:\mathbb{R}^{A}\to\mathbb{R}, it holds that

limNEπN,LN[f(aN)]=Eπ¯TA[f()],\displaystyle\lim_{N\to\infty}E^{\pi_{N,L_{N}}}\Bigl{[}f\bigl{(}\frac{\cdot}{a_{N}}\bigr{)}\Bigr{]}=E^{{\overline{\pi}}_{T}^{A}}\bigl{[}f(\,\cdot\,)\bigr{]},

where AA is a finite subset of +\mathbb{Z}_{+} and π¯TA\overline{\pi}_{T}^{A} denotes the product probability measure on +A\mathbb{R}_{+}^{A} whose one site marginal distribution on +\mathbb{R}_{+} is given by πT(dr)=1Te1Trdr\pi_{T}(dr)=\frac{1}{T}e^{-\frac{1}{T}r}dr.

As an easy consequence of Proposition 1.1 and Theorem 1.1, we obtain the following.

Corollary 1.1.

Let {Xn(N)}n0\{X_{n}^{(N)}\}_{n\geq 0}, {Yn(N)}n0\{Y_{n}^{(N)}\}_{n\geq 0} and {Zn(N)}n0\{Z_{n}^{(N)}\}_{n\geq 0} denote the immediate exchange model, the random saving model and the uniform reshuffling model on the state space ΩN(LN)\Omega_{N}(L_{N}), respectively. Under the same conditions as in Proposition 1.1 and Theorem 1.1, we have

limNlimnEη[1N|{xΛN;1aNXn(N)(x)(b,c)}|]=μα,T((b,c)),\displaystyle\lim_{N\to\infty}\lim_{n\to\infty}E_{\eta}\Bigl{[}\frac{1}{N}\bigm{|}\!\!\bigl{\{}x\in\Lambda_{N};\frac{1}{a_{N}}{X_{n}^{(N)}(x)}\in(b,c)\bigr{\}}\!\!\bigm{|}\Bigr{]}=\mu_{\alpha,T}((b,c)),
limNlimnEη[1N|{xΛN;1aNYn(N)(x)(b,c)}|]=μα,T((b,c)),\displaystyle\lim_{N\to\infty}\lim_{n\to\infty}E_{\eta}\Bigl{[}\frac{1}{N}\bigm{|}\!\!\bigl{\{}x\in\Lambda_{N};\frac{1}{a_{N}}{Y_{n}^{(N)}(x)}\in(b,c)\bigr{\}}\!\!\bigm{|}\Bigr{]}=\mu_{\alpha,T}((b,c)),

and

limNlimnEη[1N|{xΛN;1aNZn(N)(x)(b,c)}|]=πT((b,c)),\displaystyle\lim_{N\to\infty}\lim_{n\to\infty}E_{\eta}\Bigl{[}\frac{1}{N}\bigm{|}\!\!\bigl{\{}x\in\Lambda_{N};\frac{1}{a_{N}}{Z_{n}^{(N)}(x)}\in(b,c)\bigr{\}}\!\!\bigm{|}\Bigr{]}=\pi_{T}((b,c)),

for every ηΩN(LN)\eta\in\Omega_{N}(L_{N}) and every 0b<c0\leq b<c\leq\infty where Eη[]E_{\eta}[\ \cdot\ ] denotes the expectation with respect to the law of the Markov chain with the initial condition η\eta.

We note that it is natural to consider the scaling of the process by a factor of 1aN\frac{1}{a_{N}}. This is because, under the condition on LNL_{N}, we have 1NxΛN(1aNXn(N)(x))=LNNaN=T(1+o(1))\frac{1}{N}\sum\limits_{x\in\Lambda_{N}}\bigl{(}\frac{1}{a_{N}}X_{n}^{(N)}(x)\bigr{)}=\frac{L_{N}}{Na_{N}}=T(1+o(1)) as NN\to\infty, which means that the asymptotic average number of coins per agent for the scaled process is given by TT. This value corresponds to the money temperature in our model. Corollary 1.1 provides a precise formulation and generalization of earlier studies in the physics literature, which were based on numerical simulations and heuristic arguments. As time approaches infinity, and in the large population and large money temperature limit, the asymptotic wealth distribution, i.e. the proportion of agents holding a specific number of coins converges to a Gamma distribution or an exponential distribution for both the immediate exchange model and the random saving model, while it converges to an exponential distribution for the uniform reshuffling model. The parameters of the Gamma distribution depend on the asymptotic behavior of the weight function gg. When α>1\alpha>-1 the limiting wealth distribution is given by a Gamma distribution with mean TT and shape parameter α+2\alpha+2. While, when α1\alpha\leq-1, the limiting wealth distribution is given by an exponential distribution with mean TT and this does not depend on the parameter α\alpha. In particular, if gg is a constant function then α=0\alpha=0 and the limiting distribution is a Gamma distribution with mean TT and shape parameter two, which corresponds to (1.2). Additionally, when g(k)=(k+1)αg(k)=(k+1)^{\alpha}, k+k\in\mathbb{Z}_{+}, the above results are consistent with numerical simulations; see Figures 1 and 2.

Refer to caption Refer to caption
Refer to caption Refer to caption
Figure 1: Simulation results for a single realization of the immediate exchange model and the random saving model with g(k)=(k+1)αg(k)=(k+1)^{\alpha} where α=1,3\alpha=1,3. The number of agents is N=104N=10^{4} and the total number of coins is L=106L=10^{6}, namely the average number of coins per agent equals to 100100. The initial condition is set to a constant configuration X0100X_{0}\equiv 100 or Y0100Y_{0}\equiv 100. ρ\rho is distributed uniformly over the edge set {{x,y};x,yΛN,xy}\{\{x,y\};x,y\in\Lambda_{N},x\neq y\}, and in the immediate exchange model, swapping shall always be performed between the selected edges. The gray histograms represent the wealth distribution, i.e. the proportion of agents holding a specific number of coins after n=105n=10^{5} updates. The dotted line is the graph of the probability density function of the Gamma distribution: fa,b(r)=1Γ(a)bara1e1brf_{a,b}(r)=\frac{1}{\Gamma(a)b^{a}}r^{a-1}e^{-\frac{1}{b}r} with the shape parameter a=α+2a=\alpha+2 and the scale parameter b=100α+2b=\frac{100}{\alpha+2}.
Refer to caption Refer to caption
Refer to caption Refer to caption
Refer to caption Refer to caption
Figure 2: Simulation results for a single realization of the immediate exchange model and the random saving model with g(k)=(k+1)αg(k)=(k+1)^{\alpha} where α=1,2\alpha=-1,-2. The settings of the simulations are the same as Figure 1. The dotted line is the graph of the probability density function of the exponential distribution: fλ(r)=λeλrf_{\lambda}(r)={\lambda}e^{-\lambda r} with the parameter λ=1100\lambda=\frac{1}{100}. When the weight function gg decays rapidly in the immediate exchange model, the probability of each agent exchanging only a small number of coins at a time increases, leading to a longer convergence time to reach a steady state. The simulation result for the case g(k)=(k+1)2g(k)=(k+1)^{-2} reflects this and n=105n=10^{5} updates are not sufficient for convergence (the middle left). However, after n=106n=10^{6} updates, the histograms approach the limiting probability density function (the lower left). Also, the result for the case N=103N=10^{3} and n=106n=10^{6} indicates that we need to take a large population NN (and hence a large LL) for convergence (the lower right).
Remark 1.4.

The limiting distribution (1.5) for the case α1\alpha\leq-1 can be interpreted as follows: If the weight function gg decays rapidly in the random saving model, the probability that each agent saves a large number of coins becomes very small. Moreover, since we are considering the scaled process {1aNYn(N)}n0\{\frac{1}{a_{N}}Y_{n}^{(N)}\}_{n\geq 0}, we can assume that each agent offers nearly all the coins they have in each money exchange. Consequently, similarly to the uniform reshuffling model, the limiting distribution becomes an exponential distribution with mean TT. In the immediate exchange model, when the weight function gg decays rapidly, the probability that each agent exchanges a large number of coins also becomes very small. For the scaled process, this situation can be regarded as similar to the one-coin model, where one agent gives only one coin to another agent at a time. In the one-coin model, the limiting distribution is expected to be exponential (cf. [6], [13]) and the result above aligns with this. Furthermore, a mathematically rigorous justification of the result for the one-coin model can be achieved by formulating and proving it in the same manner as demonstrated in this paper.

In the rest of the paper we provide the proofs of Proposition 1.1 and Theorem 1.1 in Sections 2 and 3, respectively. We give some comments about the strategy of the proof. The proof of Proposition 1.1 is standard. It is not difficult to see that our models are irreducible aperiodic Markov chains on the finite state space ΩN(L)\Omega_{N}(L). All we have to do is to characterize the unique stationary distribution, which is achieved by carefully verifying the detailed balance condition. Namely, we demonstrate that our models are reversible Markov chains, and the reversible distributions for {Xn}n0\{X_{n}\}_{n\geq 0} and {Yn}n0\{Y_{n}\}_{n\geq 0} are given by (1.4). With respect to the proof of Theorem 1.1, the convergence of the marginal distribution of μN,L\mu_{N,L} is closely related to the equivalence of ensembles (cf. [8], [12]). The sequence {G(k)}k+\{G(k)\}_{k\in\mathbb{Z}_{+}} is neither a probability distribution nor generally summable over kk. However, multiplying G(k)G(k) by the exponential factor sks^{k}, s[0,1)s\in[0,1) ensures the convergence of k+skG(k)\sum\limits_{k\in\mathbb{Z}_{+}}s^{k}G(k) under the assumption on gg. The stationary distribution (1.4) can then be interpreted as the microcanonical distribution of an i.i.d. product, where the one-site marginal distribution is of exponential type and proportional to GG. Instead of the usual condition 1NxΛNη(x)m(0,)\frac{1}{N}\sum\limits_{x\in\Lambda_{N}}\eta(x)\to m\in(0,\infty) as NN\to\infty, we consider the condition 1NaNxΛNη(x)T(0,)\frac{1}{Na_{N}}\sum\limits_{x\in\Lambda_{N}}\eta(x)\to T\in(0,\infty) and investigate the convergence of the law of the scaled field 1aNη\frac{1}{a_{N}}\eta under the corresponding microcanonical distribution. We adapt the proof of the equivalence of ensembles from [12, Appendix 2] to this unusual setting. In particular, the local limit theorem for a triangular array of random variables plays an important role in the argument.

Throughout the paper CC, CC^{\prime}, C′′C^{\prime\prime} represent positive constants that do not depend on the size of the system NN, but may depend on other parameters. These constants in various estimates may change from place to place in the paper.

2 Proof of Proposition 1.1

In this section we prove Proposition 1.1. Under the condition that the hypergraph (ΛN,𝒟N,ρ)(\Lambda_{N},\mathcal{D}_{N,\rho}) is connected, we have that for every x,yΛNx,y\in\Lambda_{N} there exists a sequence {Ak}0kl𝒟N,ρ\{A_{k}\}_{0\leq k\leq l}\subset\mathcal{D}_{N,\rho} such that xA0x\in A_{0}, yAly\in A_{l} and AkAk+1A_{k}\cap A_{k+1}\neq\emptyset for every 0kl10\leq k\leq l-1. Combining this fact with the assumption g(0)>0g(0)>0 and g(1)>0g(1)>0, it is easy to see that the following holds.

  • For every ξ,ηΩN(L)\xi,\eta\in\Omega_{N}(L) there exists m=m(ξ,η)0m=m(\xi,\eta)\geq 0 such that P(Xm=η|X0=ξ)>0P(X_{m}=\eta|X_{0}=\xi)>0.

  • P(Xn+1=η|Xn=η)>0P(X_{n+1}=\eta|X_{n}=\eta)>0 for every ηΩN(L)\eta\in\Omega_{N}(L) and n0n\geq 0.

The same statement also holds for {Yn}n0\{Y_{n}\}_{n\geq 0} under the condition g(0)>0g(0)>0. Namely, {Xn}n0\{X_{n}\}_{n\geq 0} and {Yn}n0\{Y_{n}\}_{n\geq 0} are irreducible aperiodic Markov chains on the finite state space ΩN(L)\Omega_{N}(L). Then, the stationary distribution of {Xn}n0({Yn}n0)\{X_{n}\}_{n\geq 0}\ (\{Y_{n}\}_{n\geq 0}) uniquely exists and the law of Xn(Yn)X_{n}\ (Y_{n}) converges to it in the limit nn\to\infty by the Markov chain convergence theorem. Therefore, all we have to show is that the measure μN,L\mu_{N,L} given by (1.4) is the stationary distribution for both of {Xn}n0\{X_{n}\}_{n\geq 0} and {Yn}n0\{Y_{n}\}_{n\geq 0}. Actually, by verifying the detailed balance condition:

μN,L(ξ)P(Xn+1=η|Xn=ξ)=μN,L(η)P(Xn+1=ξ|Xn=η) for every ξ,ηΩN(L),\mu_{N,L}(\xi)P(X_{n+1}=\eta|X_{n}=\xi)=\mu_{N,L}(\eta)P(X_{n+1}=\xi|X_{n}=\eta)\ \text{ for every }\xi,\eta\in\Omega_{N}(L),

and

μN,L(ξ)P(Yn+1=η|Yn=ξ)=μN,L(η)P(Yn+1=ξ|Yn=η) for every ξ,ηΩN(L),\mu_{N,L}(\xi)P(Y_{n+1}=\eta|Y_{n}=\xi)=\mu_{N,L}(\eta)P(Y_{n+1}=\xi|Y_{n}=\eta)\ \text{ for every }\xi,\eta\in\Omega_{N}(L),

we prove that (1.4) is the reversible distribution for both of {Xn}n0\{X_{n}\}_{n\geq 0} and {Yn}n0\{Y_{n}\}_{n\geq 0}.

Proof for the immediate exchange model: Take arbitrary ξ,ηΩN(L)\xi,\eta\in\Omega_{N}(L). There exists A0=A0(ξ,η)ΛNA_{0}=A_{0}(\xi,\eta)\subset\Lambda_{N} such that ξη\xi\neq\eta on A0A_{0} and ξ=η\xi=\eta on ΛNA0\Lambda_{N}\setminus A_{0}. Such set A0A_{0} is uniquely determined from ξ\xi and η\eta. We have that

P(Xn+1=η|Xn=ξ)\displaystyle P(X_{n+1}=\eta|X_{n}=\xi) =A𝒟NAA0ρ(A)σ𝒮A1|𝒮A|P(η(z)=ξ(z)c(z)+c(σ1(z)) for every zA|ξ),\displaystyle=\sum_{{\begin{subarray}{c}A\in\mathcal{D}_{N}\\ A\supset A_{0}\end{subarray}}}\rho(A)\sum_{\sigma\in\mathcal{S}_{A}}\frac{1}{|\mathcal{S}_{A}|}P\bigl{(}\eta(z)=\xi(z)-c(z)+c(\sigma^{-1}(z))\text{ for every }z\in A\bigm{|}\!\xi\bigr{)},

where P(|ξ)P(\ \cdot\ |\xi) denotes the law of {c(x)}\{c(x)\} in the dynamics (2)(2) for given the configuration ξ\xi. If ρ(A)=0\rho(A)=0 for every A𝒟NA\in\mathcal{D}_{N} so that AA0A\supset A_{0}, then the right-hand side is equal to 0. For a finite set AA, we label its elements as A={x1,x2,,x|A|}A=\{x_{1},x_{2},\cdots,x_{|A|}\}. Then,

P(η(z)=ξ(z)c(z)+c(σ1(z)) for every zA|ξ)\displaystyle P\bigl{(}\eta(z)=\xi(z)-c(z)+c(\sigma^{-1}(z))\text{ for every }z\in A\bigm{|}\!\xi\bigr{)}
=t1=0ξ(x1)t2=0ξ(x2)t|A|=0ξ(x|A|)i=1|A|g(ti)G(ξ(xi))I(η(xi)=ξ(xi)ti+tσ1(i) for every 1i|A|)\displaystyle=\sum_{t_{1}=0}^{\xi(x_{1})}\sum_{t_{2}=0}^{\xi(x_{2})}\cdots\sum_{t_{|A|}=0}^{\xi(x_{|A|})}\prod_{i=1}^{|A|}\frac{g(t_{i})}{G(\xi(x_{i}))}\cdot I\bigl{(}\eta(x_{i})=\xi(x_{i})-t_{i}+t_{\sigma^{-1}(i)}\text{ for every }1\leq i\leq|A|\bigr{)}
=t1=0ξ(x1)t2=0ξ(x2)t|A|=0ξ(x|A|)i=1|A|{g(ti)G(ξ(xi))I(η(xi)=ξ(xi)ti+tσ1(i))},\displaystyle=\sum_{t_{1}=0}^{\xi(x_{1})}\sum_{t_{2}=0}^{\xi(x_{2})}\cdots\sum_{t_{|A|}=0}^{\xi(x_{|A|})}\prod_{i=1}^{|A|}\Bigl{\{}\frac{g(t_{i})}{G(\xi(x_{i}))}I\bigl{(}\eta(x_{i})=\xi(x_{i})-t_{i}+t_{\sigma^{-1}(i)}\bigr{)}\Bigr{\}},

where for each σ𝒮A\sigma\in\mathcal{S}_{A} and 1i|A|1\leq i\leq|A|, we identify σ1(i)\sigma^{-1}(i) with the label 1k|A|1\leq k\leq|A| which satisfies xk=σ1(xi)x_{k}=\sigma^{-1}(x_{i}). Therefore,

{xΛNG(ξ(x))}P(Xn+1=η|Xn=ξ)=A𝒟NAA0[xΛNAG(ξ(x))ρ(A)|𝒮A|σ𝒮At1=0ξ(x1)t2=0ξ(x2)t|A|=0ξ(x|A|)i=1|A|{g(ti)I(η(xi)=ξ(xi)ti+tσ1(i))}]=A𝒟NAA0[xΛNA{G(ξ(x))+G(η(x))2}ρ(A)|𝒮A|σ𝒮At1=0ξ(x1)t2=0ξ(x2)t|A|=0ξ(x|A|)i=1|A|{g(ti)I(η(xi)=ξ(xi)ti+tσ1(i))}],\displaystyle\begin{split}&\Bigl{\{}\prod_{x\in\Lambda_{N}}G(\xi(x))\Bigr{\}}\cdot P(X_{n+1}=\eta|X_{n}=\xi)\\ &=\sum_{{\begin{subarray}{c}A\in\mathcal{D}_{N}\\ A\supset A_{0}\end{subarray}}}\Bigl{[}\prod_{x\in\Lambda_{N}\setminus A}\!\!G(\xi(x))\cdot\frac{\rho(A)}{|\mathcal{S}_{A}|}\sum_{\sigma\in\mathcal{S}_{A}}\sum_{t_{1}=0}^{\xi(x_{1})}\sum_{t_{2}=0}^{\xi(x_{2})}\cdots\sum_{t_{|A|}=0}^{\xi(x_{|A|})}\prod_{i=1}^{|A|}\Bigl{\{}{g(t_{i})}I\bigl{(}\eta(x_{i})=\xi(x_{i})-t_{i}+t_{\sigma^{-1}(i)}\bigr{)}\Bigr{\}}\Bigr{]}\\ &=\sum_{{\begin{subarray}{c}A\in\mathcal{D}_{N}\\ A\supset A_{0}\end{subarray}}}\Bigl{[}\prod_{x\in\Lambda_{N}\setminus A}\!\!\Bigl{\{}\frac{G(\xi(x))+G(\eta(x))}{2}\Bigr{\}}\cdot\frac{\rho(A)}{|\mathcal{S}_{A}|}\sum_{\sigma\in\mathcal{S}_{A}}\sum_{t_{1}=0}^{\xi(x_{1})}\sum_{t_{2}=0}^{\xi(x_{2})}\cdots\sum_{t_{|A|}=0}^{\xi(x_{|A|})}\prod_{i=1}^{|A|}\Bigl{\{}{g(t_{i})}I\bigl{(}\eta(x_{i})=\xi(x_{i})-t_{i}+t_{\sigma^{-1}(i)}\bigr{)}\Bigr{\}}\Bigr{]},\end{split} (2.1)

where the last equality follows from ξ=η\xi=\eta on ΛNA0\Lambda_{N}\setminus A_{0}. For AΛNA\subset\Lambda_{N} and σ𝒮A\sigma\in\mathcal{S}_{A}, we set

hA(σ):=t1=0ξ(x1)t2=0ξ(x2)t|A|=0ξ(x|A|)i=1|A|{g(ti)I(η(xi)=ξ(xi)ti+tσ1(i))}.h_{A}(\sigma):=\sum_{t_{1}=0}^{\xi(x_{1})}\sum_{t_{2}=0}^{\xi(x_{2})}\cdots\sum_{t_{|A|}=0}^{\xi(x_{|A|})}\prod_{i=1}^{|A|}\Bigl{\{}{g(t_{i})}I\bigl{(}\eta(x_{i})=\xi(x_{i})-t_{i}+t_{\sigma^{-1}(i)}\bigr{)}\Bigr{\}}.

To show that (2) is symmetric with respect to ξ\xi and η\eta, it is sufficient to show that σ𝒮AhA(σ)\sum\limits_{\sigma\in\mathcal{S}_{A}}h_{A}(\sigma) is symmetric with respect to ξ\xi and η\eta for every A𝒟NA\in\mathcal{D}_{N} so that AA0A\supset A_{0}. Now, each permutation σ𝒮A\sigma\in\mathcal{S}_{A} can be decomposed as the product of cyclic permutations and the summand in hA(σ)h_{A}(\sigma) is given by a product form. Accordingly, the following two statements are sufficient for the symmetry of σ𝒮AhA(σ)\sum\limits_{\sigma\in\mathcal{S}_{A}}h_{A}(\sigma).

  • When |A|=2|A|=2, hA(σ)h_{A}(\sigma) is symmetric with respect to {ξ(x)}xA\{\xi(x)\}_{x\in A} and {η(x)}xA\{\eta(x)\}_{x\in A} for every σ𝒮A\sigma\in\mathcal{S}_{A}.

  • When |A|3|A|\geq 3, hA(σ)+hA(σ1)h_{A}(\sigma)+h_{A}(\sigma^{-1}) is symmetric with respect to {ξ(x)}xA\{\xi(x)\}_{x\in A} and {η(x)}xA\{\eta(x)\}_{x\in A} for every cyclic permutation σ𝒮A\sigma\in\mathcal{S}_{A}.

These can be reformulated as follows:

Lemma 2.1.

Let g:+[0,)g:\mathbb{Z}_{+}\to[0,\infty) and a={ai}i=1n,b={bi}i=1n+na=\{a_{i}\}_{i=1}^{n},b=\{b_{i}\}_{i=1}^{n}\in\mathbb{Z}_{+}^{n} be two sequences of non-negative integers which satisfy i=1nai=i=1nbi\sum\limits_{i=1}^{n}a_{i}=\sum\limits_{i=1}^{n}b_{i}.

  1. (i)(i)

    Let n=2n=2 and define

    S=S(a,b):=t1=0a1t2=0a2i=12{g(ti)I(titi+1=aibi)},\displaystyle S=S(a,b):=\sum_{t_{1}=0}^{a_{1}}\sum_{t_{2}=0}^{a_{2}}\prod_{i=1}^{2}\Bigl{\{}{g(t_{i})}I\bigl{(}t_{i}-t_{i+1}=a_{i}-b_{i}\bigr{)}\Bigr{\}},

    where we identify t3t_{3} as t1t_{1}. Then, SS is symmetric with respect to aa and bb.

  2. (ii)(ii)

    Let n3n\geq 3 and define

    S=S+(a,b)+S(a,b)\displaystyle S=S_{+}(a,b)+S_{-}(a,b) :=t1=0a1t2=0a2tn=0ani=1n{g(ti)I(titi+1=aibi)}\displaystyle:=\sum_{t_{1}=0}^{a_{1}}\sum_{t_{2}=0}^{a_{2}}\cdots\sum_{t_{n}=0}^{a_{n}}\prod_{i=1}^{n}\Bigl{\{}{g(t_{i})}I\bigl{(}t_{i}-t_{i+1}=a_{i}-b_{i}\bigr{)}\Bigr{\}}
    +t1=0a1t2=0a2tn=0ani=1n{g(ti)I(titi1=aibi)},\displaystyle\qquad\qquad+\sum_{t_{1}=0}^{a_{1}}\sum_{t_{2}=0}^{a_{2}}\cdots\sum_{t_{n}=0}^{a_{n}}\prod_{i=1}^{n}\Bigl{\{}{g(t_{i})}I\bigl{(}t_{i}-t_{i-1}=a_{i}-b_{i}\bigr{)}\Bigr{\}},

    where we identify tn+1t_{n+1} as t1t_{1} and t0t_{0} as tnt_{n}. Then, SS is symmetric with respect to aa and bb.

Proof.
  1. (i)(i)

    Under the condition a1+a2=b1+b2a_{1}+a_{2}=b_{1}+b_{2}, we have

    S(a,b)\displaystyle S(a,b) =t1=0a1t2=0a2g(t1)g(t1a1+b1)I(t2=t1a1+b1)\displaystyle=\sum_{t_{1}=0}^{a_{1}}\sum_{t_{2}=0}^{a_{2}}g(t_{1})g(t_{1}-a_{1}+b_{1})I\bigl{(}t_{2}=t_{1}-a_{1}+b_{1}\bigr{)}
    =t1=0a1g(t1)g(t1a1+b1)I(0t1a1+b1a2)\displaystyle=\sum_{t_{1}=0}^{a_{1}}g(t_{1})g(t_{1}-a_{1}+b_{1})I\bigl{(}0\leq t_{1}-a_{1}+b_{1}\leq a_{2}\bigr{)}
    =t1=0(a1b1)a1b2g(t1)g(t1a1+b1)\displaystyle=\sum_{t_{1}=0\vee(a_{1}-b_{1})}^{a_{1}\wedge b_{2}}g(t_{1})g(t_{1}-a_{1}+b_{1})
    =t1=a1b1(a1+b1)(b1+b2)g(t1b1)g(t1a1).\displaystyle=\sum_{t_{1}=a_{1}\vee b_{1}}^{(a_{1}+b_{1})\wedge(b_{1}+b_{2})}g(t_{1}-b_{1})g(t_{1}-a_{1}).

    This is symmetric with respect to aa and bb because b1+b2=12(a1+a2+b1+b2)b_{1}+b_{2}=\frac{1}{2}(a_{1}+a_{2}+b_{1}+b_{2}).

  2. (ii)(ii)

    We compute that

    S+(a,b)\displaystyle S_{+}(a,b) =t10t20tn0[i=1n{g(ti)I(titi+1=aibi)}i=1n{I(tiai)I(ti+1bi)}]\displaystyle=\sum_{t_{1}\geq 0}\sum_{t_{2}\geq 0}\cdots\sum_{t_{n}\geq 0}\Bigl{[}\prod_{i=1}^{n}\Bigl{\{}{g(t_{i})}I\bigl{(}t_{i}-t_{i+1}=a_{i}-b_{i}\bigr{)}\Bigr{\}}\prod_{i=1}^{n}\Bigl{\{}I\bigl{(}t_{i}\leq a_{i})I(t_{i+1}\leq b_{i}\bigr{)}\Bigr{\}}\Bigr{]}
    =t10t20tn0i=1ng(ti)i=1n{I(titi+1=aibi)I(tiaibi1)}.\displaystyle=\sum_{t_{1}\geq 0}\sum_{t_{2}\geq 0}\cdots\sum_{t_{n}\geq 0}\prod_{i=1}^{n}{g(t_{i})}\prod_{i=1}^{n}\Bigl{\{}I\bigl{(}t_{i}-t_{i+1}=a_{i}-b_{i}\bigr{)}I\bigl{(}t_{i}\leq a_{i}\wedge b_{i-1}\bigr{)}\Bigr{\}}.

    Similarly,

    S(a,b)\displaystyle S_{-}(a,b) =t10t20tn0i=1ng(ti)i=1n{I(titi1=aibi)I(tiaibi+1)}\displaystyle=\sum_{t_{1}\geq 0}\sum_{t_{2}\geq 0}\cdots\sum_{t_{n}\geq 0}\prod_{i=1}^{n}{g(t_{i})}\prod_{i=1}^{n}\Bigl{\{}I\bigl{(}t_{i}-t_{i-1}=a_{i}-b_{i}\bigr{)}I\bigl{(}t_{i}\leq a_{i}\wedge b_{i+1}\bigr{)}\Bigr{\}}
    =t10t20tn0i=1ng(ti)i=1n{I(ti+1ti=aibi)I(ti+1aibi+1)}\displaystyle=\sum_{t_{1}\geq 0}\sum_{t_{2}\geq 0}\cdots\sum_{t_{n}\geq 0}\prod_{i=1}^{n}{g(t_{i})}\prod_{i=1}^{n}\Bigl{\{}I\bigl{(}t_{i+1}-t_{i}=a_{i}-b_{i}\bigr{)}I\bigl{(}t_{i+1}\leq a_{i}\wedge b_{i+1}\bigr{)}\Bigr{\}}
    =t10t20tn0i=1ng(ti)i=1n{I(titi+1=biai)I(tiai1bi)}\displaystyle=\sum_{t_{1}\geq 0}\sum_{t_{2}\geq 0}\cdots\sum_{t_{n}\geq 0}\prod_{i=1}^{n}{g(t_{i})}\prod_{i=1}^{n}\Bigl{\{}I\bigl{(}t_{i}-t_{i+1}=b_{i}-a_{i}\bigr{)}I\bigl{(}t_{i}\leq a_{i-1}\wedge b_{i}\bigr{)}\Bigr{\}}
    =S+(b,a),\displaystyle=S_{+}(b,a),

    where the second equality follows from rewriting the variable tit_{i} by ti+1t_{i+1}, 1in1\leq i\leq n. Therefore, S+(a,b)+S(a,b)S_{+}(a,b)+S_{-}(a,b) is symmetric with respect to aa and bb.

Proof for the random saving model: Take arbitrary ξ,ηΩN(L)\xi,\eta\in\Omega_{N}(L). We use the similar notation as the proof for the immediate exchange model.

P(Yn+1=η|Yn=ξ)\displaystyle P(Y_{n+1}=\eta|Y_{n}=\xi)
=A𝒟NAA0ρ(A)P(η(z)=c(z)+d(z) for every zA|ξ)\displaystyle=\sum_{{\begin{subarray}{c}A\in\mathcal{D}_{N}\\ A\supset A_{0}\end{subarray}}}\rho(A)P\bigl{(}\eta(z)=c(z)+d(z)\text{ for every }z\in A\bigm{|}\xi\bigr{)}
=A𝒟NAA0ρ(A)t1=0ξ(x1)t2=0ξ(x2)t|A|=0ξ(x|A|)[i=1|A|g(ti)G(ξ(xi))ζΩ(A,SA(ξ)SA(t))1|Ω(A,SA(ξ)SA(t))|\displaystyle=\sum_{{\begin{subarray}{c}A\in\mathcal{D}_{N}\\ A\supset A_{0}\end{subarray}}}\rho(A)\sum_{t_{1}=0}^{\xi(x_{1})}\sum_{t_{2}=0}^{\xi(x_{2})}\cdots\sum_{t_{|A|}=0}^{\xi(x_{|A|})}\Bigl{[}\prod_{i=1}^{|A|}\frac{g(t_{i})}{G(\xi(x_{i}))}\sum\limits_{\zeta\in\Omega(A,S_{A}(\xi)-S_{A}(t))}\frac{1}{|\Omega(A,S_{A}(\xi)-S_{A}(t))|}
×I(η(xi)=ti+ζ(xi) for every 1i|A|)]\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\times I\bigl{(}\eta(x_{i})=t_{i}+\zeta(x_{i})\text{ for every }1\leq i\leq|A|\bigr{)}\Bigr{]}
=A𝒟NAA0ρ(A)t1=0ξ(x1)η(x1)t2=0ξ(x2)η(x2)t|A|=0ξ(x|A|)η(x|A|)[i=1|A|g(ti)G(ξ(xi))ζΩ(A,SA(ξ)SA(t))1|Ω(A,SA(ξ)SA(t))|\displaystyle=\sum_{{\begin{subarray}{c}A\in\mathcal{D}_{N}\\ A\supset A_{0}\end{subarray}}}\rho(A)\sum_{t_{1}=0}^{\xi(x_{1})\wedge\eta(x_{1})}\sum_{t_{2}=0}^{\xi(x_{2})\wedge\eta(x_{2})}\cdots\sum_{t_{|A|}=0}^{\xi(x_{|A|})\wedge\eta(x_{|A|})}\Bigl{[}\prod_{i=1}^{|A|}\frac{g(t_{i})}{G(\xi(x_{i}))}\sum\limits_{\zeta\in\Omega(A,S_{A}(\xi)-S_{A}(t))}\frac{1}{|\Omega(A,S_{A}(\xi)-S_{A}(t))|}
×I(η(xi)=ti+ζ(xi) for every 1i|A|)],\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\times I\bigl{(}\eta(x_{i})=t_{i}+\zeta(x_{i})\text{ for every }1\leq i\leq|A|\bigr{)}\Bigr{]},

where SA(t)=i=1|A|tiS_{A}(t)=\sum\limits_{i=1}^{|A|}t_{i} for t={ti}i=1|A|t=\{t_{i}\}_{i=1}^{|A|}. We have SA(ξ)=SA(η)S_{A}(\xi)=S_{A}(\eta) for AA0A\supset A_{0}. Also, for given t={ti}i=1|A|t=\{t_{i}\}_{i=1}^{|A|} so that 0tiξ(xi)η(xi)0\leq t_{i}\leq\xi(x_{i})\wedge\eta(x_{i}) for every 1i|A|1\leq i\leq|A|, there exists unique ζΩ(A,SA(ξ)SA(t))\zeta\in\Omega(A,S_{A}(\xi)-S_{A}(t)) which satisfies η(xi)=ti+ζ(xi)\eta(x_{i})=t_{i}+\zeta(x_{i}) for every 1i|A|1\leq i\leq|A|. Therefore,

{xΛNG(ξ(x))}P(Yn+1=η|Yn=ξ)\displaystyle\Bigl{\{}\prod_{x\in\Lambda_{N}}G(\xi(x))\Bigr{\}}\cdot P(Y_{n+1}=\eta|Y_{n}=\xi)
=A𝒟NAA0[xΛNA{G(ξ(x))+G(η(x))2}ρ(A)t1=0ξ(x1)η(x1)t2=0ξ(x2)η(x2)t|A|=0ξ(x|A|)η(x|A|)\displaystyle\quad=\sum_{{\begin{subarray}{c}A\in\mathcal{D}_{N}\\ A\supset A_{0}\end{subarray}}}\Bigl{[}\prod_{x\in\Lambda_{N}\setminus A}\!\!\Bigl{\{}\frac{G(\xi(x))+G(\eta(x))}{2}\Bigr{\}}\cdot{\rho(A)}\sum_{t_{1}=0}^{\xi(x_{1})\wedge\eta(x_{1})}\sum_{t_{2}=0}^{\xi(x_{2})\wedge\eta(x_{2})}\cdots\sum_{t_{|A|}=0}^{\xi(x_{|A|})\wedge\eta(x_{|A|})}
×i=1|A|g(ti)1|Ω(A,SA(ξ)+SA(η)2SA(t))|].\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\times\prod_{i=1}^{|A|}{g(t_{i})}\cdot\frac{1}{|\Omega(A,\frac{S_{A}(\xi)+S_{A}(\eta)}{2}-S_{A}(t))|}\Bigr{]}.

This is symmetric with respect to ξ\xi and η\eta.

If we define gg as g(k)=δ0(k)g(k)=\delta_{0}(k), k+k\in\mathbb{Z}_{+}, then G1G\equiv 1 and the above argument yields that P(Yn+1=η|Yn=ξ)P(Y_{n+1}=\eta|Y_{n}=\xi) is symmetric with respect to ξ\xi and η\eta. Therefore, the uniform reshuffling model is doubly stochastic and its unique stationary distribution is give by the uniform distribution on ΩN(L)\Omega_{N}(L). Actually, this matches when G1G\equiv 1 is set in (1.4). ∎

3 Proof of Theorem 1.1

For the proof of Theorem 1.1, we adapt the proof of the equivalence of ensembles for the i.i.d. product measure (cf. [12, Appendix 2]). In the following we assume that g(0)>0g(0)>0 and there exist α\alpha\in\mathbb{R} and cα(0,)c_{\alpha}\in(0,\infty) such that limkg(k)kα=cα\lim\limits_{k\to\infty}\frac{g(k)}{k^{\alpha}}=c_{\alpha}. We prepare several notations. Define G(k)=j=0kg(j)G(k)=\sum\limits_{j=0}^{k}g(j), k+k\in\mathbb{Z}_{+} and Qn(s)=k0knskG(k)Q_{n}(s)=\sum\limits_{k\geq 0}k^{n}s^{k}G(k), n+n\in\mathbb{Z}_{+}. By the assumption on gg, we have the following asymptotics of G(k)G(k) as kk\to\infty.

G(k){cαα+1kα+1 if α>1,cαlogk if α=1,C0 if α<1,G(k)\sim\begin{cases}\frac{c_{\alpha}}{\alpha+1}k^{\alpha+1}&\text{ if }\alpha>-1,\\ {c_{\alpha}}{\log k}&\text{ if }\alpha=-1,\\ C_{0}&\text{ if }\alpha<-1,\\ \end{cases} (3.1)

where C0=C0(g)>0C_{0}=C_{0}(g)>0 is a constant which depends on gg. In particular, the radius of convergence of Qn(s)Q_{n}(s) is 11 and it holds that lims1Qn(s)=\lim\limits_{s\uparrow 1}Q_{n}(s)=\infty for every n+n\in\mathbb{Z}_{+} and α\alpha\in\mathbb{R}. We define the exponential family of distributions {νs();s[0,1)}\{\nu_{s}(\,\cdot\,);s\in[0,1)\} on +\mathbb{Z}_{+} by νs(k)=skG(k)Q0(s)\nu_{s}(k)=\frac{s^{k}G(k)}{Q_{0}(s)}, k+k\in\mathbb{Z}_{+}. It is easy to see that Eνs[η(0)]=Q1(s)Q0(s)E^{\nu_{s}}[\eta(0)]=\frac{Q_{1}(s)}{Q_{0}(s)} is continuous, increasing in ss and diverges to infinity as s1s\uparrow 1. Hence, for every K>0K>0 there exists unique s=s(K)(0,1){s}^{*}={s}^{*}(K)\in(0,1) such that Eνs[η(0)]=KE^{\nu_{s^{*}}}[\eta(0)]=K. To examine the asymptotic behavior of s(K)s^{*}(K) as KK\to\infty, we use a Tauberian theorem of the following form (cf. [1, Corollary 1.7.3]).

Theorem 3.1.

Let {ak}k0\{a_{k}\}_{k\geq 0} be a sequence of non-negative numbers and assume that A(s)=k=0akskA(s)=\sum\limits_{k=0}^{\infty}a_{k}s^{k} converges for s[0,1)s\in[0,1) and {ak}k0\{a_{k}\}_{k\geq 0} is monotone. Then, the following are equivalent.

  • A(s)Γ(β+1)(1s)βh(11s)A(s)\sim\Gamma(\beta+1)(1-s)^{-\beta}h(\frac{1}{1-s}) as s1s\uparrow 1 for β>0\beta>0 and slowly varying function hh.

  • akβkβ1h(k)a_{k}\sim\beta k^{\beta-1}h(k) as kk\to\infty for β>0\beta>0 and slowly varying function hh.

By this theorem and (3.1), the following asymptotics holds in the limit s1s\uparrow 1.

Qn(s){cαΓ(n+α+2)α+1(1s)(α+n+2) if α>1,cαΓ(n+1)(1s)(n+1)log11s if α=1,C0Γ(n+1)(1s)(n+1) if α<1,Q_{n}(s)\sim\begin{cases}\frac{c_{\alpha}\Gamma(n+\alpha+2)}{\alpha+1}(1-s)^{-(\alpha+n+2)}&\text{ if }\alpha>-1,\\ {c_{\alpha}\Gamma(n+1)}(1-s)^{-(n+1)}\log\frac{1}{1-s}&\text{ if }\alpha=-1,\\ {C_{0}\Gamma(n+1)}(1-s)^{-(n+1)}&\text{ if }\alpha<-1,\end{cases} (3.2)

where we used the relation Γ(β+1)=βΓ(β)\Gamma(\beta+1)=\beta\Gamma(\beta) for every β>0\beta>0. Therefore,

Eνs[η(0)]=Q1(s)Q0(s){α+21s if α>1,11s if α1,E^{\nu_{s}}[\eta(0)]=\frac{Q_{1}(s)}{Q_{0}(s)}\sim\begin{cases}\frac{\alpha+2}{1-s}&\text{ if }\alpha>-1,\\ \frac{1}{1-s}&\text{ if }\alpha\leq-1,\\ \end{cases}

and this yields that

s(K)={1α+2K(1+o(1)) if α>1,11K(1+o(1)) if α1,s^{*}(K)=\begin{cases}1-\frac{\alpha+2}{K}(1+o(1))&\text{ if }\alpha>-1,\\ 1-\frac{1}{K}(1+o(1))&\text{ if }\alpha\leq-1,\\ \end{cases} (3.3)

as KK\to\infty. By these asymptotics we also have

Varνs(K)(η(0))=Q2(s(K))Q0(s(K))K2={1α+2K2(1+o(1)) if α>1,K2(1+o(1)) if α1,\displaystyle\mathrm{Var}_{\nu_{s^{*}(K)}}(\eta(0))=\frac{Q_{2}(s^{*}(K))}{Q_{0}(s^{*}(K))}-K^{2}=\begin{cases}\frac{1}{\alpha+2}{K^{2}}(1+o(1))&\text{ if }\alpha>-1,\\ {K^{2}}(1+o(1))&\text{ if }\alpha\leq-1,\\ \end{cases} (3.4)

as KK\to\infty.

Next, for each s[0,1)s\in[0,1) and BB\subset\mathbb{Z}, let ν¯sB\overline{\nu}_{s}^{B} be the product measure on +B\mathbb{Z}_{+}^{B} whose one site marginal distribution equals to νs\nu_{s}. ν¯sB(|Ω(B,L))\overline{\nu}_{s}^{B}\bigl{(}\,\cdot\,\bigm{|}\Omega(B,L)\bigr{)} denotes the conditioned probability of ν¯sB\overline{\nu}_{s}^{B} on the event that the total number of coins on BB equals to LL. Then, we have the following key identity for (1.4).

μN,L()=ν¯sΛN(|ΩN(L)).\mu_{N,L}(\,\cdot\,)=\overline{\nu}_{s}^{\Lambda_{N}}\bigl{(}\,\cdot\,\bigm{|}\Omega_{N}(L)\bigr{)}.

Notice that the right-hand side does not depend on the choice of the parameter ss. Let f:Af:\mathbb{R}^{A}\to\mathbb{R} be a bounded continuous local function where AA is a finite subset of +\mathbb{Z}_{+}. For every NN\in\mathbb{N} so that ΛNA\Lambda_{N}\supset A, we have

EμN,L[f(aN)]=η+Af(ηaN)ν¯sΛN(ξA=η|xΛNξ(x)=L)=η+Af(ηaN)ν¯sA(η)+η+Af(ηaN){ν¯sΛNA(xΛNAξ(x)=LSA(η))ν¯sΛN(xΛNξ(x)=L)1}ν¯sA(η)=:I1+I2,\displaystyle\begin{split}E^{\mu_{N,L}}\bigl{[}f\bigl{(}\frac{\cdot}{a_{N}}\bigr{)}\bigr{]}&=\sum\limits_{\eta\in\mathbb{Z}_{+}^{A}}f\bigl{(}\frac{\eta}{a_{N}}\bigr{)}\overline{\nu}_{s}^{\Lambda_{N}}\Bigl{(}\xi_{A}=\eta\bigm{|}\sum\limits_{x\in\Lambda_{N}}\xi(x)=L\Bigr{)}\\ &=\sum\limits_{\eta\in\mathbb{Z}_{+}^{A}}f\bigl{(}\frac{\eta}{a_{N}}\bigr{)}\overline{\nu}_{s}^{A}(\eta)+\sum\limits_{\eta\in\mathbb{Z}_{+}^{A}}f\bigl{(}\frac{\eta}{a_{N}}\bigr{)}\Bigl{\{}\frac{\overline{\nu}_{s}^{\Lambda_{N}\setminus A}\Bigl{(}\sum\limits_{x\in\Lambda_{N}\setminus A}\xi(x)=L-S_{A}(\eta)\Bigr{)}}{\overline{\nu}_{s}^{\Lambda_{N}}\Bigl{(}\sum\limits_{x\in\Lambda_{N}}\xi(x)=L\Bigr{)}}-1\Bigr{\}}\overline{\nu}_{s}^{A}(\eta)\\ &=:I_{1}+I_{2},\end{split} (3.5)

where ξA\xi_{A} denotes the configurations ξ\xi restricted on the set AA. Now, we set L=LNL=L_{N} where {LN}N1\{L_{N}\}_{N\geq 1} be a sequence of positive integers that satisfies limNLNNaN=T\lim\limits_{N\to\infty}\frac{L_{N}}{Na_{N}}=T for some constant T>0T>0 and divergent sequence {aN}N1\{a_{N}\}_{N\geq 1}. We also take ss in the right-hand side of (3.5) as sN:=s(LNN)s_{N}^{*}:=s^{*}(\frac{L_{N}}{N}). For this choice of ss we show that I1+Af(r)μ¯α,TA(dr)I_{1}\to\int_{\mathbb{R}_{+}^{A}}f(r)\overline{\mu}_{\alpha,T}^{A}(dr) and I20I_{2}\to 0 as NN\to\infty.

For the proof of the convergence of I1I_{1}, we assume that ff is a function of one variable for notational simplicity. The general case can be proven by the similar manner since ν¯sA\overline{\nu}_{s}^{A} is a product measure with the same marginal distribution. Firstly, we consider the case α>1\alpha>-1. Let R>0R>0. By (3.1) and (3.2),

I1\displaystyle I_{1} =k0f(kaN)1Q0(sN)(sN)kG(k)\displaystyle=\sum\limits_{k\geq 0}f\bigl{(}\frac{k}{a_{N}}\bigr{)}\frac{1}{Q_{0}(s_{N}^{*})}(s_{N}^{*})^{k}G(k)
=1Γ(α+2)(1sN)k=0[RaN]f(kaN)(sN)k((1sN)k)α+1\displaystyle=\frac{1}{\Gamma(\alpha+2)}(1-s_{N}^{*})\sum\limits_{k=0}^{[Ra_{N}]}f\bigl{(}\frac{k}{a_{N}}\bigr{)}(s_{N}^{*})^{k}\bigl{(}(1-s_{N}^{*})k\bigr{)}^{\alpha+1}
+1Γ(α+2)(1sN)k=[RaN]+1f(kaN)(sN)k((1sN)k)α+1+o(1)\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad+\frac{1}{\Gamma(\alpha+2)}(1-s_{N}^{*})\sum\limits_{k=[Ra_{N}]+1}^{\infty}f\bigl{(}\frac{k}{a_{N}}\bigr{)}(s_{N}^{*})^{k}\bigl{(}(1-s_{N}^{*})k\bigr{)}^{\alpha+1}+o(1)
=:I3+I4+o(1),\displaystyle=:I_{3}+I_{4}+o(1),

as NN\to\infty. We note that since Q0(sN)Q_{0}(s_{N}^{*})\to\infty in the limit NN\to\infty, a finite sum in I1I_{1} is negligible and we can replace GG with the right-hand side of (3.1) with an error of o(1)o(1). Then, by (3.3) and the condition on LNL_{N},

I3\displaystyle I_{3} =1Γ(α+2)α+2T1aNk=0[RaN]f(kaN){(1α+2T1aN)aN}kaN(α+2TkaN)α+1+o(1)\displaystyle=\frac{1}{\Gamma(\alpha+2)}\frac{\alpha+2}{T}\frac{1}{a_{N}}\sum\limits_{k=0}^{[Ra_{N}]}f\bigl{(}\frac{k}{a_{N}}\bigr{)}\Bigl{\{}\bigl{(}1-\frac{\alpha+2}{T}\frac{1}{a_{N}}\bigr{)}^{a_{N}}\Bigr{\}}^{\frac{k}{a_{N}}}\bigl{(}\frac{\alpha+2}{T}\frac{k}{a_{N}}\bigr{)}^{\alpha+1}+o(1)
1Γ(α+2)(α+2T)α+20Rf(r)eα+2Trrα+1𝑑r,\displaystyle\to\frac{1}{\Gamma(\alpha+2)}\bigl{(}\frac{\alpha+2}{T}\bigr{)}^{\alpha+2}\int_{0}^{R}f(r)e^{-\frac{\alpha+2}{T}r}r^{\alpha+1}dr,

as NN\to\infty where the convergence follows from Riemann integral. By taking the limit RR\to\infty, the right-hand side converges to 0f(r)μα,T(dr)\int_{0}^{\infty}f(r)\mu_{\alpha,T}(dr). For I4I_{4}, we have

|I4|\displaystyle|I_{4}| 1Γ(α+2)α+2T1aNk=[RaN]+1|f(kaN)|(1α+2T1aN)k(α+2TkaN)α+1+o(1)\displaystyle\leq\frac{1}{\Gamma(\alpha+2)}\frac{\alpha+2}{T}\frac{1}{a_{N}}\sum\limits_{k=[Ra_{N}]+1}^{\infty}\bigm{|}\!\!f\bigl{(}\frac{k}{a_{N}}\bigr{)}\!\!\bigm{|}\bigl{(}1-\frac{\alpha+2}{T}\frac{1}{a_{N}}\bigr{)}^{k}\bigl{(}\frac{\alpha+2}{T}\frac{k}{a_{N}}\bigr{)}^{\alpha+1}+o(1)
C1aNk=[RaN]+1eα+2TkaN(kaN)α+1+o(1)\displaystyle\leq C\frac{1}{a_{N}}\sum\limits_{k=[Ra_{N}]+1}^{\infty}e^{-\frac{\alpha+2}{T}\frac{k}{a_{N}}}\bigl{(}\frac{k}{a_{N}}\bigr{)}^{\alpha+1}+o(1)
Cj=[R]eα+2Tjjα+1+o(1),\displaystyle\leq C^{\prime}\sum\limits_{j=[R]}^{\infty}e^{-\frac{\alpha+2}{T}j}j^{\alpha+1}+o(1),

for every NN large enough where C,CC,C^{\prime} are positive constants independent of NN. By taking the limits NN\to\infty and RR\to\infty, we obtain I40I_{4}\to 0.

Secondly, we consider the case α=1\alpha=-1. By (3.1) and (3.2) again,

I1\displaystyle I_{1} =1sNlog(1sN)k=0[RaN]f(kaN)(sN)klogk+1sNlog(1sN)k=[RaN]+1f(kaN)(sN)klogk+o(1)\displaystyle=\frac{1-s_{N}^{*}}{-\log(1-s_{N}^{*})}\sum\limits_{k=0}^{[Ra_{N}]}f\bigl{(}\frac{k}{a_{N}}\bigr{)}(s_{N}^{*})^{k}\log k+\frac{1-s_{N}^{*}}{-\log(1-s_{N}^{*})}\sum\limits_{k=[Ra_{N}]+1}^{\infty}f\bigl{(}\frac{k}{a_{N}}\bigr{)}(s_{N}^{*})^{k}\log k+o(1)
=:I3+I4+o(1),\displaystyle=:I^{\prime}_{3}+I^{\prime}_{4}+o(1),

as NN\to\infty. log(1sN)=(1+o(1))logaN{-\log(1-s_{N}^{*})}=(1+o(1))\log a_{N} and this yields that

I3\displaystyle I^{\prime}_{3} =1TaNlogaNk=0[RaN]f(kaN){(11TaN)aN}kaNlogk+o(1)\displaystyle=\frac{1}{Ta_{N}\log a_{N}}\sum\limits_{k=0}^{[Ra_{N}]}f\bigl{(}\frac{k}{a_{N}}\bigr{)}\Bigl{\{}\bigl{(}1-\frac{1}{Ta_{N}}\bigr{)}^{a_{N}}\Bigr{\}}^{\frac{k}{a_{N}}}{\log k}+o(1)
=1TaNk=0[RaN]f(kaN){(11TaN)aN}kaN+1TlogaN1aNk=0[RaN]f(kaN){(11TaN)aN}kaNlog(kaN)+o(1).\displaystyle=\frac{1}{Ta_{N}}\sum\limits_{k=0}^{[Ra_{N}]}f\bigl{(}\frac{k}{a_{N}}\bigr{)}\Bigl{\{}\bigl{(}1-\frac{1}{Ta_{N}}\bigr{)}^{a_{N}}\Bigr{\}}^{\frac{k}{a_{N}}}+\frac{1}{T\log a_{N}}\frac{1}{a_{N}}\sum\limits_{k=0}^{[Ra_{N}]}f\bigl{(}\frac{k}{a_{N}}\bigr{)}\Bigl{\{}\bigl{(}1-\frac{1}{Ta_{N}}\bigr{)}^{a_{N}}\Bigr{\}}^{\frac{k}{a_{N}}}\log\bigl{(}\frac{k}{a_{N}}\bigr{)}+o(1).

In the limits NN\to\infty and RR\to\infty, the first term of the right-hand side converges to 1T0f(r)erT𝑑r\frac{1}{T}\int_{0}^{\infty}f(r)e^{-\frac{r}{T}}dr and the second term vanishes due to the extra factor 1logaN\frac{1}{\log a_{N}}. I4I^{\prime}_{4} goes to 0 in the same way as I4I_{4}. The case α<1\alpha<-1 also follows from the similar argument.

Next, for the convergence of I2I_{2}, we use the following local limit theorem.

Theorem 3.2.

Let {bN}N1\{b_{N}\}_{N\geq 1} be a sequence of positive numbers which satisfies limNbN=\lim\limits_{N\to\infty}b_{N}=\infty and set sN:=s(bN)s_{N}^{*}:=s^{*}(b_{N}). For each NN\in\mathbb{N}, {Xj(N)}jΛN\{X_{j}^{(N)}\}_{j\in\Lambda_{N}} denotes a family of independent and identically distributed +\mathbb{Z}_{+}-valued random variables with common distribution νsN\nu_{s_{N}^{*}}. Then, for every finite set B+B\subset\mathbb{Z}_{+}, it holds that

limNsupL0|σN2(N|B|)P(jΛNBXj(N)=L)12πexp{(L(N|B|)bN)22σN2(N|B|)}|=0,\displaystyle\lim\limits_{N\to\infty}\sup\limits_{L\geq 0}\Bigm{|}\!\sqrt{\sigma_{N}^{2}(N-|B|)}P\Bigl{(}\sum_{j\in\Lambda_{N}\setminus B}X_{j}^{(N)}=L\Bigr{)}-\frac{1}{\sqrt{2\pi}}\exp\Bigl{\{}-\frac{(L-(N-|B|)b_{N})^{2}}{2\sigma_{N}^{2}(N-|B|)}\Bigr{\}}\!\Bigm{|}=0, (3.6)

where σN2=Var(X1(N))\sigma_{N}^{2}=\mathrm{Var}(X_{1}^{(N)}).

The proof of this theorem is given later. By applying this theorem for L=LNL=L_{N} and bN=LNNb_{N}=\frac{L_{N}}{N}, we have

ν¯sNΛNA(xΛNAξ(x)=LNSA(η))ν¯sNΛN(xΛNξ(x)=LN)\displaystyle\frac{\overline{\nu}_{s_{N}^{*}}^{\Lambda_{N}\setminus A}\Bigl{(}\sum\limits_{x\in\Lambda_{N}\setminus A}\xi(x)=L_{N}-S_{A}(\eta)\Bigr{)}}{\overline{\nu}_{s_{N}^{*}}^{\Lambda_{N}}\Bigl{(}\sum\limits_{x\in\Lambda_{N}}\xi(x)=L_{N}\Bigr{)}} =NN|A|exp{(LNSA(η)(N|A|)LNN)22σN2(N|A|)}+o(1)1+o(1),\displaystyle=\frac{\sqrt{N}}{\sqrt{N-|A|}}\frac{\exp\Bigl{\{}-\frac{(L_{N}-S_{A}(\eta)-(N-|A|)\frac{L_{N}}{N})^{2}}{{2\sigma^{2}_{N}}{(N-|A|)}}\Bigr{\}}+o(1)}{1+o(1)},

as NN\to\infty for every η+A\eta\in\mathbb{Z}_{+}^{A}. Note that o(1)o(1) terms do not depend on η\eta. Set DA(k)={0,1,,k}AD_{A}(k)=\{0,1,\cdots,k\}^{A}, k+k\in\mathbb{Z}_{+}. Then, the above asymptotics and (3.4) yield that

λN,R:=supηDA([RaN])|ν¯sNΛNA(xΛNAξ(x)=LNSA(η))ν¯sNΛN(xΛNξ(x)=LN)1|0,\displaystyle\lambda_{N,R}:=\sup\limits_{\eta\in D_{A}([Ra_{N}])}\Bigm{|}\!\frac{\overline{\nu}_{s_{N}^{*}}^{\Lambda_{N}\setminus A}\Bigl{(}\sum\limits_{x\in\Lambda_{N}\setminus A}\xi(x)=L_{N}-S_{A}(\eta)\Bigr{)}}{\overline{\nu}_{s_{N}^{*}}^{\Lambda_{N}}\Bigl{(}\sum\limits_{x\in\Lambda_{N}}\xi(x)=L_{N}\Bigr{)}}-1\!\Bigm{|}\to 0,

as NN\to\infty for every R>0R>0 and

supN1supη+A|ν¯sNΛNA(xΛNAξ(x)=LNSA(η))ν¯sNΛN(xΛNξ(x)=LN)1|C′′,\displaystyle\sup_{N\geq 1}\sup\limits_{\eta\in\mathbb{Z}_{+}^{A}}\Bigm{|}\!\frac{\overline{\nu}_{s_{N}^{*}}^{\Lambda_{N}\setminus A}\Bigl{(}\sum\limits_{x\in\Lambda_{N}\setminus A}\xi(x)=L_{N}-S_{A}(\eta)\Bigr{)}}{\overline{\nu}_{s_{N}^{*}}^{\Lambda_{N}}\Bigl{(}\sum\limits_{x\in\Lambda_{N}}\xi(x)=L_{N}\Bigr{)}}-1\!\Bigm{|}\,\leq C^{\prime\prime},

for some constant C′′>0C^{\prime\prime}>0. Therefore,

|I2|\displaystyle|I_{2}| ηDA([RaN])|f(ηaN)||ν¯sNΛNA(xΛNAξ(x)=LNSA(η))ν¯sNΛN(xΛNξ(x)=LN)1|ν¯sNA(η)\displaystyle\leq\sum\limits_{\eta\in D_{A}([Ra_{N}])}\bigm{|}\!f\bigl{(}\frac{\eta}{a_{N}}\bigr{)}\!\bigm{|}\Bigm{|}\!\frac{\overline{\nu}_{s_{N}^{*}}^{\Lambda_{N}\setminus A}\Bigl{(}\sum\limits_{x\in\Lambda_{N}\setminus A}\xi(x)=L_{N}-S_{A}(\eta)\Bigr{)}}{\overline{\nu}_{s_{N}^{*}}^{\Lambda_{N}}\Bigl{(}\sum\limits_{x\in\Lambda_{N}}\xi(x)=L_{N}\Bigr{)}}-1\!\Bigm{|}\overline{\nu}_{s_{N}^{*}}^{A}(\eta)
+η+ADA([RaN])|f(ηaN)||ν¯sNΛNA(xΛNAξ(x)=LNSA(η))ν¯sNΛN(xΛNξ(x)=LN)1|ν¯sNA(η)\displaystyle\qquad\quad+\sum\limits_{\eta\in\mathbb{Z}_{+}^{A}\setminus D_{A}([Ra_{N}])}\bigm{|}\!f\bigl{(}\frac{\eta}{a_{N}}\bigr{)}\!\bigm{|}\Bigm{|}\!\frac{\overline{\nu}_{s_{N}^{*}}^{\Lambda_{N}\setminus A}\Bigl{(}\sum\limits_{x\in\Lambda_{N}\setminus A}\xi(x)=L_{N}-S_{A}(\eta)\Bigr{)}}{\overline{\nu}_{s_{N}^{*}}^{\Lambda_{N}}\Bigl{(}\sum\limits_{x\in\Lambda_{N}}\xi(x)=L_{N}\Bigr{)}}-1\!\Bigm{|}\overline{\nu}_{s_{N}^{*}}^{A}(\eta)
fλN,R+C′′η+ADA([RaN])|f(ηaN)|ν¯sNA(η).\displaystyle\leq\|f\|_{\infty}\lambda_{N,R}+C^{\prime\prime}\sum\limits_{\eta\in\mathbb{Z}_{+}^{A}\setminus D_{A}([Ra_{N}])}\bigm{|}\!f\bigl{(}\frac{\eta}{a_{N}}\bigr{)}\!\bigm{|}\overline{\nu}_{s_{N}^{*}}^{A}(\eta).

The first term of the right-hand side goes to 0 as NN\to\infty for every R>0R>0 and the second term goes to 0 as NN\to\infty and RR\to\infty by the similar computation as the estimate of I4I_{4} above. Hence, we obtain I20I_{2}\to 0 and this completes the proof.

If we assume the condition: g(0)>0g(0)>0 and j0g(j)<\sum\limits_{j\geq 0}g(j)<\infty for g:+[0,)g:\mathbb{Z}_{+}\to[0,\infty) instead, then the proof for the case α<1\alpha<-1 above can be applied as it is.

Remark 3.1.

If (L(N|B|)bN)22σN2(N|B|)(L-(N-|B|)b_{N})^{2}\gg{2\sigma_{N}^{2}(N-|B|)} in (3.6) then the exponential term converges to 0, rendering the local limit theorem ineffective. For this reason, the local limit theorem in the form of Theorem 3.2 was insufficient to prove the equivalence of ensembles in general settings, and a more refined version such as the Edgeworth expansion of at least the second order was necessary (cf. the proofs of Corollary 1.4 and Corollary 1.7 in [12, Appendix 2]). On the other hand, such an expansion is not necessary and Theorem 3.2 is sufficient in our case. Because we divided the summation η+A\sum\limits_{\eta\in\mathbb{Z}_{+}^{A}} of I2I_{2} into ηDA([RaN])\sum\limits_{\eta\in D_{A}([Ra_{N}])} and η+ADA([RaN])\sum\limits_{\eta\in\mathbb{Z}_{+}^{A}\setminus D_{A}([Ra_{N}])}, the estimate for the latter part was reduced to the estimate of ν¯sNA(+ADA([RaN]))\overline{\nu}_{s_{N}^{*}}^{A}(\mathbb{Z}_{+}^{A}\setminus D_{A}([Ra_{N}])) which can be managed because we know the explicit form of νsN{\nu}_{s_{N}^{*}}.

Proof of Theorem 3.2. First of all, we note that Theorem 3.2 corresponds to the local limit theorem for a triangular array of random variables since νsN\nu_{s_{N}^{*}} depends on the number of random variables NN. Combining this with the fact that sN:=s(bN)1s_{N}^{*}:=s^{*}(b_{N})\to 1 as NN\to\infty, we cannot directly apply Theorem 1.3 or Theorem 1.5 in [12, Appendix 2] which studied the refined version of the local limit theorem for the i.i.d random variables with common distribution νs\nu_{s}, s(0,1)s\in(0,1). Also, the known criteria of the local limit theorem for a triangular array of integer-valued random variables (e.g. [5], [18]) do not hold in our setting since σN2\sigma_{N}^{2}\to\infty as NN\to\infty. Therefore, we give the proof of the theorem according to the classical argument [20, Chapter VII]. For notational simplicity we only consider the case B=B=\emptyset. The modification for the general finite set B+B\subset\mathbb{Z}_{+} is straightforward.

Set X~j(N):=1NσN2(Xj(N)bN)\widetilde{X}_{j}^{(N)}:=\frac{1}{\sqrt{N\sigma^{2}_{N}}}(X_{j}^{(N)}-b_{N}), jΛNj\in\Lambda_{N} and tN(L):=1NσN2(LNbN)t_{N}(L):=\frac{1}{\sqrt{N\sigma^{2}_{N}}}(L-Nb_{N}). We define ϕN(θ)=E[exp{iθX1(N)}]\phi_{N}(\theta)=E\bigl{[}\exp\{i\theta X_{1}^{(N)}\}\bigr{]} and ψN(θ)=E[exp{iθ(jΛNX~j(N))}]\psi_{N}(\theta)=E\bigl{[}\exp\bigl{\{}i\theta(\sum\limits_{j\in\Lambda_{N}}\widetilde{X}_{j}^{(N)})\bigr{\}}\bigr{]}, θ\theta\in\mathbb{R} where i=1i=\sqrt{-1}. By the inversion formula, we have

P(jΛNXj(N)=L)=12πNσN2πNσN2πNσN2eiθtN(L)ψN(θ)𝑑θ.\displaystyle P\bigl{(}\sum\limits_{j\in\Lambda_{N}}X_{j}^{(N)}=L\bigr{)}=\frac{1}{2\pi\sqrt{N\sigma_{N}^{2}}}\int_{-\pi\sqrt{N\sigma_{N}^{2}}}^{\pi\sqrt{N\sigma_{N}^{2}}}e^{-i\theta t_{N}(L)}\psi_{N}(\theta)d\theta.

Therefore,

2π|NσN2P(jΛNXj(N)=L)12πe12tN(L)2|\displaystyle 2\pi\Bigm{|}\!\sqrt{N\sigma_{N}^{2}}P\bigl{(}\sum\limits_{j\in\Lambda_{N}}X_{j}^{(N)}=L\bigr{)}-\frac{1}{\sqrt{2\pi}}e^{-\frac{1}{2}t_{N}(L)^{2}}\!\Bigm{|}
=|πNσN2πNσN2eiθtN(L)ψN(θ)dθeiθtN(L)e12θ2dθ|\displaystyle=\Bigm{|}\!\int_{-\pi\sqrt{N\sigma_{N}^{2}}}^{\pi\sqrt{N\sigma_{N}^{2}}}e^{-i\theta t_{N}(L)}\psi_{N}(\theta)d\theta-\int_{-\infty}^{\infty}e^{-i\theta t_{N}(L)}e^{-\frac{1}{2}\theta^{2}}d\theta\!\Bigm{|}
|θ|R|ψN(θ)e12θ2|𝑑θ+R|θ|γNσN2|ψN(θ)|𝑑θ+γNσN2|θ|πNσN2|ψN(θ)|𝑑θ+|θ|Re12θ2𝑑θ\displaystyle\leq\int_{|\theta|\leq R}|\psi_{N}(\theta)-e^{-\frac{1}{2}\theta^{2}}|d\theta+\int_{R\leq|\theta|\leq\gamma\sqrt{N\sigma_{N}^{2}}}|\psi_{N}(\theta)|d\theta+\int_{\gamma\sqrt{N\sigma_{N}^{2}}\leq|\theta|\leq\pi\sqrt{N\sigma_{N}^{2}}}|\psi_{N}(\theta)|d\theta+\int_{|\theta|\geq R}e^{-\frac{1}{2}\theta^{2}}d\theta
=:I1+I2+I3+I4,\displaystyle=:I_{1}+I_{2}+I_{3}+I_{4},

for every R>0R>0 and 0<γ<π0<\gamma<\pi. We show that the right-hand side converges to 0 as NN\to\infty and RR\to\infty.

For I1I_{1}, assume that the law of jΛNX~j(N)\sum\limits_{j\in\Lambda_{N}}\widetilde{X}_{j}^{(N)} converges to the standard normal distribution. Then, we have limNψN(θ)=e12θ2\lim\limits_{N\to\infty}\psi_{N}(\theta)=e^{-\frac{1}{2}\theta^{2}} for every θ\theta\in\mathbb{R} and we obtain I10I_{1}\to 0 as NN\to\infty by the bounded convergence theorem. For the convergence of the law of jΛNX~j(N)\sum\limits_{j\in\Lambda_{N}}\widetilde{X}_{j}^{(N)}, we have only to show that jΛNE[(X~j(N))2;|X~j(N)|ε]0\sum\limits_{j\in\Lambda_{N}}E\bigl{[}(\widetilde{X}_{j}^{(N)})^{2};|\widetilde{X}_{j}^{(N)}|\geq\varepsilon\bigr{]}\to 0 as NN\to\infty for every ε>0\varepsilon>0 by Lindberg’s central limit theorem (cf. [7, Theorem 3.4.10]). By (3.4), |X~j(N)|ε|\widetilde{X}_{j}^{(N)}|\geq\varepsilon implies that Xj(N)12εNσN2X_{j}^{(N)}\geq\frac{1}{2}\varepsilon\sqrt{N\sigma_{N}^{2}} for every NN large enough. Therefore,

jΛNE[(X~j(N))2;|X~j(N)|ε]\displaystyle\sum\limits_{j\in\Lambda_{N}}E\bigl{[}(\widetilde{X}_{j}^{(N)})^{2};|\widetilde{X}_{j}^{(N)}|\geq\varepsilon\bigr{]} 1σN2E[(X1(N)bN)2;X1(N)12εNσN2]\displaystyle\leq\frac{1}{\sigma_{N}^{2}}E\Bigl{[}(X_{1}^{(N)}-b_{N})^{2};X_{1}^{(N)}\geq\frac{1}{2}\varepsilon\sqrt{N\sigma_{N}^{2}}\Bigr{]}
2σN2E[(X1(N))2;X1(N)12εNσN2]+2bN2σN2P(X1(N)12εNσN2)\displaystyle\leq\frac{2}{\sigma_{N}^{2}}E\Bigl{[}({X}_{1}^{(N)})^{2};{X}_{1}^{(N)}\geq\frac{1}{2}\varepsilon\sqrt{N\sigma_{N}^{2}}\Bigr{]}+\frac{2b_{N}^{2}}{\sigma_{N}^{2}}P\Bigl{(}{X}_{1}^{(N)}\geq\frac{1}{2}\varepsilon\sqrt{N\sigma_{N}^{2}}\Bigr{)}
=:J1+J2,\displaystyle=:J_{1}+J_{2},

where the second inequality follows from the fact that (a+b)22a2+2b2(a+b)^{2}\leq 2a^{2}+2b^{2} for every a,ba,b\in\mathbb{R}. We first consider the case α>1\alpha>-1 for the estimate of J1J_{1}.

J1\displaystyle J_{1} =2σN2k12εNσN2k21Q0(sN)(sN)kG(k)\displaystyle=\frac{2}{\sigma_{N}^{2}}\sum\limits_{k\geq\frac{1}{2}\varepsilon\sqrt{N\sigma_{N}^{2}}}k^{2}\frac{1}{Q_{0}(s_{N}^{*})}(s_{N}^{*})^{k}G(k)
CbN2kε4α+2NbN2k2(α+2bN)α+2{(1α+2bN)bN}kbNkα+1\displaystyle\leq\frac{C}{b_{N}^{2}}\sum\limits_{k\geq\frac{\varepsilon}{4\sqrt{\alpha+2}}\sqrt{Nb_{N}^{2}}}k^{2}\Bigl{(}\frac{\alpha+2}{b_{N}}\Bigr{)}^{\alpha+2}\Bigl{\{}\bigl{(}1-\frac{\alpha+2}{b_{N}}\bigr{)}^{b_{N}}\Bigr{\}}^{\frac{k}{b_{N}}}k^{\alpha+1}
C1bNkbNε4α+2N(kbN)α+3{(1α+2bN)bN}kbN,\displaystyle\leq{C^{\prime}}\frac{1}{b_{N}}\sum\limits_{\frac{k}{b_{N}}\geq\frac{\varepsilon}{4\sqrt{\alpha+2}}\sqrt{N}}\Bigl{(}\frac{k}{b_{N}}\Bigr{)}^{\alpha+3}\Bigl{\{}\bigl{(}1-\frac{\alpha+2}{b_{N}}\bigr{)}^{b_{N}}\Bigr{\}}^{\frac{k}{b_{N}}},

for some constants C,C>0C,C^{\prime}>0 and every NN large enough where we used (3.2), (3.3) and (3.4) for the first inequality. The right-hand side goes to 0 as NN\to\infty because

1bNkbNa(kbN)α+3{(1α+2bN)bN}kbNarα+3e(α+2)r𝑑r<,\displaystyle\frac{1}{b_{N}}\sum\limits_{\frac{k}{b_{N}}\geq a}\Bigl{(}\frac{k}{b_{N}}\Bigr{)}^{\alpha+3}\Bigl{\{}\bigl{(}1-\frac{\alpha+2}{b_{N}}\bigr{)}^{b_{N}}\Bigr{\}}^{\frac{k}{b_{N}}}\to\int_{a}^{\infty}r^{\alpha+3}e^{-(\alpha+2)r}dr<\infty,

as NN\to\infty for every a0a\geq 0. By the similar computation we obtain J10J_{1}\to 0 when α1\alpha\leq-1. For the estimate of J2J_{2}, we have

J2CP(X1(N)12εNσN2)\displaystyle J_{2}\leq CP\Bigl{(}X_{1}^{(N)}\geq\frac{1}{2}\varepsilon\sqrt{N\sigma_{N}^{2}}\Bigr{)} C2εNσN2E[X1(N)]CN0,\displaystyle\leq C\frac{2}{\varepsilon\sqrt{N\sigma^{2}_{N}}}E[X_{1}^{(N)}]\leq\frac{C^{\prime}}{\sqrt{N}}\to 0,

as NN\to\infty where we used Markov’s inequality and (3.4).

Next, we consider I2I_{2}. By Taylor’s theorem, there exists γ0>0\gamma_{0}>0 such that for every θ\theta\in\mathbb{R} which satisfies |θNσN2|γ0|\frac{\theta}{\sqrt{N\sigma^{2}_{N}}}|\leq\gamma_{0}, we have

|E[eiθX~1(N)]|114(θNσN2)2E[(X1(N)bN)2]=1θ24Neθ24N.\displaystyle|E\bigl{[}e^{i\theta\widetilde{X}_{1}^{(N)}}\bigr{]}|\leq 1-\frac{1}{4}\bigl{(}\frac{\theta}{\sqrt{N\sigma^{2}_{N}}}\bigr{)}^{2}E\bigl{[}(X_{1}^{(N)}-b_{N})^{2}\bigr{]}=1-\frac{\theta^{2}}{4N}\leq e^{-\frac{\theta^{2}}{4N}}.

Therefore, |ψN(θ)|eθ24|\psi_{N}(\theta)|\leq e^{-\frac{\theta^{2}}{4}} for every θ\theta\in\mathbb{R} which satisfies |θ|γ0NσN2|{\theta}|\leq\gamma_{0}{\sqrt{N\sigma^{2}_{N}}}. Taking γ\gamma in the definition of I2I_{2} as γ0\gamma_{0}, we obtain

I2R|θ|γ0NσN2eθ24𝑑θ2Reθ24𝑑θ0,\displaystyle I_{2}\leq\int_{R\leq|\theta|\leq\gamma_{0}\sqrt{N\sigma_{N}^{2}}}e^{-\frac{\theta^{2}}{4}}d\theta\leq 2\int_{R}^{\infty}e^{-\frac{\theta^{2}}{4}}d\theta\to 0,

as RR\to\infty. Similarly, we have I40I_{4}\to 0 as RR\to\infty.

The final task is the estimate of I3I_{3}. We take γ\gamma in the definition of I3I_{3} as γ0\gamma_{0} above. Let 0<δ<10<\delta<1 be fixed and define ϕNδ(θ):=E[(δeiθ)X1(N)]=k0δkeiθkνsN(k)\phi_{N}^{\delta}(\theta):=E\bigl{[}(\delta e^{i\theta})^{X_{1}^{(N)}}\bigr{]}=\sum\limits_{k\geq 0}\delta^{k}e^{i\theta k}{{\nu}_{s_{N}^{*}}(k)}. By the proof of [17, Lemma 5.4], we know that

|ϕNδ(θ)|1|1δeiθ|{νsN(0)+k0|νsN(k+1)νsN(k)|}.\displaystyle|\phi_{N}^{\delta}(\theta)|\leq\frac{1}{|1-\delta e^{i\theta}|}\Bigl{\{}{\nu}_{s_{N}^{*}}(0)+\sum\limits_{k\geq 0}|{{\nu}_{s_{N}^{*}}(k+1)}-{{\nu}_{s_{N}^{*}}(k)}|\Bigr{\}}.

We have νsN(0)=G(0)Q0(sN)0{\nu}_{s_{N}^{*}}(0)=\frac{G(0)}{Q_{0}(s_{N}^{*})}\to 0 as NN\to\infty by (3.2) and (3.3). Also,

|νsN(k+1)νsN(k)|\displaystyle|{{\nu}_{s_{N}^{*}}(k+1)}-{{\nu}_{s_{N}^{*}}(k)}| =νsN(k)|νsN(k+1)νsN(k)1|\displaystyle={\nu}_{s_{N}^{*}}(k)\bigm{|}\!\frac{{\nu}_{s_{N}^{*}}(k+1)}{{\nu}_{s_{N}^{*}}(k)}-1\!\bigm{|}
=νsN(k)|sN{g(k+1)G(k)+1}1|νsN(k){(1sN)+g(k+1)G(k)}.\displaystyle={\nu}_{s_{N}^{*}}(k)\bigm{|}\!s_{N}^{*}\bigl{\{}\frac{g(k+1)}{G(k)}+1\bigr{\}}-1\!\bigm{|}\,\leq{\nu}_{s_{N}^{*}}(k)\Bigl{\{}(1-s_{N}^{*})+\frac{g(k+1)}{G(k)}\Bigr{\}}.

By the assumption on gg, (3.1) and (3.3), for every ε>0\varepsilon>0 there exists N01N_{0}\geq 1 and C1>0C_{1}>0 such that (1sN)+g(k+1)G(k)C1bN(1-s_{N}^{*})+\frac{g(k+1)}{G(k)}\leq\frac{C_{1}}{b_{N}} for every NN0N\geq N_{0} and kεbNk\geq\varepsilon b_{N}. Moreover, (1sN)+g(k+1)G(k)C2(1-s_{N}^{*})+\frac{g(k+1)}{G(k)}\leq{C_{2}} for every k+k\in\mathbb{Z}_{+} and NN\in\mathbb{N} where C2>0C_{2}>0 is a constant independent of NN and kk. Therefore,

k0|νsN(k+1)νsN(k)|k=0εbNC2νsN(k)+kεbNC1bNνsN(k).\displaystyle\sum\limits_{k\geq 0}|{{\nu}_{s_{N}^{*}}(k+1)}-{{\nu}_{s_{N}^{*}}(k)}|\leq\sum\limits_{k=0}^{\varepsilon b_{N}}C_{2}{\nu}_{s_{N}^{*}}(k)+\sum\limits_{k\geq\varepsilon b_{N}}\frac{C_{1}}{b_{N}}{\nu}_{s_{N}^{*}}(k).

The first term of the right-hand side converges to C30εrα+1e(α+2)r𝑑rC_{3}\int_{0}^{\varepsilon}r^{\alpha+1}e^{-(\alpha+2)r}dr if α>1\alpha>-1 and C30εer𝑑rC_{3}\int_{0}^{\varepsilon}e^{-r}dr if α1\alpha\leq-1 for some C3>0C_{3}>0 by the similar computation as before. The second term is less than C1bN\frac{C_{1}}{b_{N}} and this goes to 0 as NN\to\infty. As a result, for every ε>0\varepsilon>0 there exists C4(ε)>0C_{4}(\varepsilon)>0 such that C4(ε)0C_{4}(\varepsilon)\to 0 as ε0\varepsilon\to 0 and

|ϕNδ(θ)|1|1δeiθ|{C4(ε)+1πγ0},\displaystyle|\phi_{N}^{\delta}(\theta)|\leq\frac{1}{|1-\delta e^{i\theta}|}\bigl{\{}C_{4}(\varepsilon)+\frac{1}{\pi}\gamma_{0}\bigr{\}},

for every NN large enough and every θ\theta\in\mathbb{R}. By taking the limit δ1\delta\uparrow 1 and using the estimate |1eiθ|2π|θ||1-e^{i\theta}|\geq\frac{2}{\pi}|\theta| for every |θ|π|\theta|\leq\pi, we obtain

|ϕN(θNσN2)|NσN2|θ|π2{C4(ε)+1πγ0}1γ0π2{C4(ε)+1πγ0}=π2γ0C4(ε)+12,\displaystyle|\phi_{N}\bigl{(}\frac{\theta}{\sqrt{N\sigma_{N}^{2}}}\bigr{)}|\leq\frac{\sqrt{N\sigma_{N}^{2}}}{|\theta|}\frac{\pi}{2}\bigl{\{}C_{4}(\varepsilon)+\frac{1}{\pi}\gamma_{0}\bigr{\}}\leq\frac{1}{\gamma_{0}}\frac{\pi}{2}\bigl{\{}C_{4}(\varepsilon)+\frac{1}{\pi}\gamma_{0}\bigr{\}}=\frac{\pi}{2\gamma_{0}}C_{4}(\varepsilon)+\frac{1}{2},

for every θ\theta\in\mathbb{R} so that γ0NσN2|θ|πNσN2\gamma_{0}\sqrt{N\sigma_{N}^{2}}\leq|\theta|\leq\pi\sqrt{N\sigma_{N}^{2}}. Hence, by taking ε>0\varepsilon>0 small enough, there exists r<1r<1 such that |ϕN(θNσN2)|r|\phi_{N}\bigl{(}\frac{\theta}{\sqrt{N\sigma_{N}^{2}}}\bigr{)}|\leq r for every NN large enough and θ\theta\in\mathbb{R} so that γ0NσN2|θ|πNσN2\gamma_{0}\sqrt{N\sigma_{N}^{2}}\leq|\theta|\leq\pi\sqrt{N\sigma_{N}^{2}}. This yields that

I3=γ0NσN2|θ|πNσN2|ψN(θ)|𝑑θ\displaystyle I_{3}=\int_{\gamma_{0}\sqrt{N\sigma_{N}^{2}}\leq|\theta|\leq\pi\sqrt{N\sigma_{N}^{2}}}|\psi_{N}(\theta)|d\theta =γ0NσN2|θ|πNσN2|ϕN(θNσN2)|N𝑑θ\displaystyle=\int_{\gamma_{0}\sqrt{N\sigma_{N}^{2}}\leq|\theta|\leq\pi\sqrt{N\sigma_{N}^{2}}}|\phi_{N}(\frac{\theta}{\sqrt{N\sigma_{N}^{2}}})|^{N}d\theta
2πNσN2rN0,\displaystyle\leq 2\pi\sqrt{N\sigma_{N}^{2}}r^{N}\to 0,

as NN\to\infty and we can complete the proof of Theorem 3.2.

Proof of Corollary 1.1. By Proposition 1.1, we have

limnEη[1N|{xΛN;1aNXn(N)(x)(b,c)}|]\displaystyle\lim\limits_{n\to\infty}E_{\eta}\Bigl{[}\frac{1}{N}\bigm{|}\!\!\bigl{\{}x\in\Lambda_{N};\frac{1}{a_{N}}{X_{n}^{(N)}(x)}\in(b,c)\bigr{\}}\!\!\bigm{|}\Bigr{]} =limn1NxΛNEη[I(1aNXn(N)(x)(b,c))]\displaystyle=\lim\limits_{n\to\infty}\frac{1}{N}\sum\limits_{x\in\Lambda_{N}}E_{\eta}\Bigl{[}I\bigl{(}\frac{1}{a_{N}}X_{n}^{(N)}(x)\in(b,c)\bigr{)}\Bigr{]}
=1NxΛNμN,LN(1aNξ(x)(b,c))\displaystyle=\frac{1}{N}\sum\limits_{x\in\Lambda_{N}}\mu_{N,L_{N}}\bigl{(}\frac{1}{a_{N}}\xi(x)\in(b,c)\bigr{)}
=μN,LN(1aNξ(1)(b,c)).\displaystyle=\mu_{N,L_{N}}\bigl{(}\frac{1}{a_{N}}\xi(1)\in(b,c)\bigr{)}.

Therefore, it is sufficient to show that limNμN,LN(1aNξ(1)(b,c))=μα,T((b,c))\lim\limits_{N\to\infty}\mu_{N,L_{N}}\bigl{(}\frac{1}{a_{N}}\xi(1)\in(b,c)\bigr{)}=\mu_{\alpha,T}((b,c)). This follows from Theorem 1.1 and the basic facts about the weak convergence of probability measures. The same is true for {Yn(N)}n0\{Y_{n}^{(N)}\}_{n\geq 0} and {Zn(N)}n0\{Z_{n}^{(N)}\}_{n\geq 0}.

Acknowledgement

The author thanks Mai Aihara for valuable discussions and her help on numerical simulations. This work was partially supported by JSPS KAKENHI Grant Number 22K03359.

References

  • [1] N. Bingham, C. M. Goldie, J. L. Teugels, Regular variation, Cambridge University Press, Cambridge, (1987).
  • [2] B.K. Chakrabarti, A. Chakraborti, S. R. Chakravarty, A. Chatterjee, Econophysics of Income and Wealth Distributions, Cambridge University Press, Cambridge, (2013).
  • [3] A. Chakraborti, B. K. Chakrabarti, Statistical mechanics of money: how saving propensity affects its distribution, Eur. Phys. J. B 17, 167–170 (2000).
  • [4] P. Cirillo, F. Redig, W. Ruszel, Duality and stationary distributions of wealth distribution models, J. Phys. A 47, 085203 (2014).
  • [5] B. Davis, D. McDonald, An elementary proof of the local central limit theorem, J. Theoret. Probab. 8, 693–701 (1995).
  • [6] A.A. Dragulescu, V.M. Yakovenko, Statistical mechanics of money, Eur. Phys. J. B 17, 723–729 (2000).
  • [7] R. Durrett, Probability: Theory and Examples, 5th ed., Cambridge University Press, Cambridge, (2019).
  • [8] S. Friedli, Y. Velenik, Statistical mechanics of lattice systems. A concrete mathematical introduction, Cambridge University Press, Cambridge, (2018).
  • [9] B. van Ginkel, F. Redig, F. Sau, Duality and stationary distributions of the “Immediate exchange model” and its generalizations, J. Stat. Phys. 163, 92–112 (2016).
  • [10] E. Heinsalu, M. Patriarca, Kinetic models of immediate exchange, Eur. Phys. J. B 87, 170–179 (2014).
  • [11] G. Katriel, The immediate exchange model: an analytical investigation, Eur. Phys. J. B 88, 19–24 (2015).
  • [12] C. Kipnis, C. Landim, Scaling limits of interacting particle systems, Grundlehren Math. Wiss. 320, Springer-Verlag, Berlin, 1999.
  • [13] N. Lanchier, Rigorous proof of the Boltzmann-Gibbs distribution of money on connected graphs, J. Stat. Phys. 167, 160–172 (2017).
  • [14] N. Lanchier, S. Reed, Rigorous results for the distribution of money on connected graphs, J. Stat. Phys. 171, 727–743 (2018).
  • [15] N. Lanchier, S. Reed, Rigorous results for the distribution of money on connected graphs (models with debts), J. Stat. Phys. 176, 1115–1137 (2019).
  • [16] N. Lanchier, S. Reed, Distribution of money on connected graphs with multiple banks, Math. Model. Nat. Phenom. 19, Paper No.10 (2024).
  • [17] C. Landim, S. Sethuraman, S. Varadhan, Spectral gap for zero-range dynamics Ann. Probab. 24, 1871–1902 (1996).
  • [18] A. B. Mukhin, Local limit theorems for lattice random variables, Theory Probab. Appl. 36, 698–713 (1991).
  • [19] M. Patriarca, A. Chakraborti, K. Kaski, Statistical model with standard Γ\Gamma distribution, Phys. Rev. E 70, 016104 (2004).
  • [20] V. V. Petrov, Sums of independent random variables, Springer, Springer-Verlag, Berlin, (1975).
  • [21] R. Pymar, N. Rivera, Mixing of the symmetric beta-binomial splitting process on arbitrary graphs, arXiv preprint arXiv:2307.02406, (2023).
  • [22] M. Quattropani, F. Sau, Mixing of the averaging process and its discrete dual on finite-dimensional geometries, Ann. Appl. Probab. 33, 1136–1171 (2023).
  • [23] F. Redig, F. Sau, Generalized immediate exchange models and their symmetries, Stochastic Process. Appl. 127, 3251–3267 (2017).
  • [24] V. M. Yakovenko, J.J. Barkley Rosser, Colloquium: statistical mechanics of money, wealth, and income, Rev. Mod. Phys. 81, 1703–1725 (2009).