Limiting distributions of generalized money exchange models
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)
Initial conditions: Assign each agent a random or equal amount of money.
-
(2)
Selecting pairs of agents: Randomly select pairs of agents who will exchange money.
-
(3)
Money exchange: The selected pair exchanges money according to specific rules.
-
(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 . [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 . Each site corresponds to an agent, and the economy is represented by the set of agents . The population size is given by . The amount of money each agent holds is represented by the number of coins, where represents the number of coins that agent possesses at time . The edge set represents a social network in which only agents connected by an edge can interact to exchange coins. At each time step, an edge is chosen uniformly at random from . Agents and independently and uniformly select a random number of their coins to give to each other. Thus, constitutes a time-homogeneous Markov chain. In particular, the total number of coins is conserved in this process, and we denote this total by . [14] proved that the stationary distribution for this Markov chain uniquely exists and gave the following explicit representation.
(1.1) |
for every and . 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.
(1.2) |
for large enough and . In this context represents the average number of coins per agent and is called money temperature by analogy with the temperature in physics and the limit and 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 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 and 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 and 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 (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 in (1.2) is not mathematically valid. The left-hand side is defined only for non-negative integers , while the right-hand represents a probability density function on . 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], are replaced with for sufficiently large and satisfying . On the other hand, are replaced with , rather than , even when . Additionally, has been conveniently replaced with . As a matter of fact, it seems unnatural to consider the approximation of for each since the average number of coins per agent diverges in the limit . The money exchange model describes the microscopic movement of money, however, our goal is to derive the macroscopic distribution of wealth in the limit and . In order to do that, we should analyze the convergence of 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, denotes the set of non-negative integers and denotes the set of positive integers. denotes the integral part of and we set and for . For a finite set , denotes its cardinality. For two sequences of positive numbers and , means that . A function on is called local if it depends only on finitely many coordinates. For a probability measure , denotes the expectation with respect to .
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 and , we define the configuration space
When , is denoted by . Consider now an economy populated by many agents. Each site corresponds to an agent and we assume that the economy can be represented by the set of agents . corresponds to the population size. For each , we interpret , not as the number of particles, but as the number of coins held by agent . denotes a probability distribution on , namely for every with and . This represents the distribution that determines which agent handles the money exchange at each time step. We also take a non-negative function defined on . Now, we introduce three money exchange models.
Immediate exchange model: Let be a time-homogeneous Markov chain on the state space . For given , the configuration is determined from the following rule.
-
(1)
Choose a set according to the distribution .
-
(2)
For given , let be independent random variables whose distributions are given by
(1.3) for where we set , .
-
(3)
Choose a permutation uniformly random, namely with probability where denotes the set of all permutations of .
-
(4)
For given and realizations , and , define by
This dynamics can be interpreted as follows: At each time step, a money exchange occurs between agents in a randomly chosen set . represents the number of coins that agent transfers, which is determined by a probability distribution dependent on the weight function and , the number of coins that agent possesses. According to a randomly chosen permutation , each agent passes coins to agent . 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 is a constant function and is the uniform distribution on an edge set of . In this case, the distribution (1.3) matches the uniform distribution on and the money exchange occurs between two agents connected by an edge in . To be more precise, since the permutation can include the identity permutation, our model can be viewed as the lazy version of their model in this case.
Random saving model: Let be a time-homogeneous Markov chain on the state space . For given , the configuration is determined from the following rule.
-
(1)
Choose a set according to the distribution .
-
(2)
For given , let be independent random variables whose distributions are given by
for .
-
(3)
For given and , choose a configuration uniformly random, namely with probability where we set for .
-
(4)
For given and realizations , and , define by
Note that in contrast to the immediate exchange model, the random variable represents how many coins the agent to save. When a money exchange occurs within set , each agent offers coins. These coins are then pooled and redistributed among the agents in according to the uniform distribution. Similar to the immediate exchange model, [14] studied the case where is a constant function and is the uniform distribution on an edge set. In that model, is drawn from the uniform distribution on , and it is referred to as the uniform saving model. Since our model considers a more general distribution for , we refer to it as the random saving model. As a special case, when the weight function is defined by for , each agent saves no money. When a money exchange occurs within set , all coins held by the agents in are pooled and redistributed. This model is referred to as the uniform reshuffling model, and we denote it by .
As the first result, we characterize the stationary distributions of these Markov chains.
Proposition 1.1.
Let be fixed. We assume that the hypergraph is connected where the hyperedge set is defined by .
-
Assume that the weight function satisfies and . Then, there is a unique stationary distribution for and it is given by
(1.4) where and is the normalization factor. In particular, for every , and , it holds that
where represents the law of with the initial condition .
-
Assume that the weight function satisfies . Then, the exact same statement as applies to instead of . In particular, for the uniform reshuffling model , there is a unique stationary distribution and it is given by the uniform distribution on , namely,
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 and . Also, the choice of does not affect the stationary distribution.
Remark 1.1.
The condition is always needed to make the measure (1.4) well-defined.
Remark 1.2.
Consider the case where the weight function is a constant function . Then, it holds that , and the constant is canceled by the normalization factor in the definition of . Therefore, we can take as , in (1.4) and in this case, the above result matches that of [14]. [14] also gave the explicit representation for when and obtained (1.1).
Remark 1.3.
For the random saving model we can change the role of random variable to represent the number of coins to offer instead of the number of coins to save. Namely, we modify and in the definition of the random saving model as follows:
-
For given and , choose a configuration uniformly random, namely with probability .
-
For given and realizations , and , define by
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 and , we assume that the total number of coins satisfies for some and divergent sequence . Then, we can prove that the law of the scaled field under or converges to the i.i.d. product of probability distributions on . Its marginal distribution depends on the asymptotic behavior of the weight function .
Theorem 1.1.
Let be a sequence of positive integers that satisfies for some positive constant and a sequence which satisfies . Assume that and the following condition holds: There exist and such that . Then, for every bounded continuous local function , it holds that
where is a finite subset of and denotes the product probability measure on whose one site marginal distribution on is given by
(1.5) |
is the Gamma function with parameter .
Also, if satisfies and , then the same conclusion as in the case above holds. In particular, for every bounded continuous local function , it holds that
where is a finite subset of and denotes the product probability measure on whose one site marginal distribution on is given by .
Corollary 1.1.
Let , and denote the immediate exchange model, the random saving model and the uniform reshuffling model on the state space , respectively. Under the same conditions as in Proposition 1.1 and Theorem 1.1, we have
and
for every and every where denotes the expectation with respect to the law of the Markov chain with the initial condition .
We note that it is natural to consider the scaling of the process by a factor of . This is because, under the condition on , we have as , which means that the asymptotic average number of coins per agent for the scaled process is given by . 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 . When the limiting wealth distribution is given by a Gamma distribution with mean and shape parameter . While, when , the limiting wealth distribution is given by an exponential distribution with mean and this does not depend on the parameter . In particular, if is a constant function then and the limiting distribution is a Gamma distribution with mean and shape parameter two, which corresponds to (1.2). Additionally, when , , the above results are consistent with numerical simulations; see Figures 1 and 2.
![]() |
![]() |
![]() |
![]() |
![]() |
![]() |
![]() |
![]() |
![]() |
![]() |
Remark 1.4.
The limiting distribution (1.5) for the case can be interpreted as follows: If the weight function 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 , 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 . In the immediate exchange model, when the weight function 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 . 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 and are given by (1.4). With respect to the proof of Theorem 1.1, the convergence of the marginal distribution of is closely related to the equivalence of ensembles (cf. [8], [12]). The sequence is neither a probability distribution nor generally summable over . However, multiplying by the exponential factor , ensures the convergence of under the assumption on . 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 . Instead of the usual condition as , we consider the condition and investigate the convergence of the law of the scaled field 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 , , represent positive constants that do not depend on the size of the system , 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 is connected, we have that for every there exists a sequence such that , and for every . Combining this fact with the assumption and , it is easy to see that the following holds.
-
•
For every there exists such that .
-
•
for every and .
The same statement also holds for under the condition . Namely, and are irreducible aperiodic Markov chains on the finite state space . Then, the stationary distribution of uniquely exists and the law of converges to it in the limit by the Markov chain convergence theorem. Therefore, all we have to show is that the measure given by (1.4) is the stationary distribution for both of and . Actually, by verifying the detailed balance condition:
and
we prove that (1.4) is the reversible distribution for both of and .
Proof for the immediate exchange model: Take arbitrary . There exists such that on and on . Such set is uniquely determined from and . We have that
where denotes the law of in the dynamics for given the configuration . If for every so that , then the right-hand side is equal to . For a finite set , we label its elements as . Then,
where for each and , we identify with the label which satisfies . Therefore,
(2.1) |
where the last equality follows from on . For and , we set
To show that (2) is symmetric with respect to and , it is sufficient to show that is symmetric with respect to and for every so that . Now, each permutation can be decomposed as the product of cyclic permutations and the summand in is given by a product form. Accordingly, the following two statements are sufficient for the symmetry of .
-
•
When , is symmetric with respect to and for every .
-
•
When , is symmetric with respect to and for every cyclic permutation .
These can be reformulated as follows:
Lemma 2.1.
Let and be two sequences of non-negative integers which satisfy .
-
Let and define
where we identify as . Then, is symmetric with respect to and .
-
Let and define
where we identify as and as . Then, is symmetric with respect to and .
Proof.
-
Under the condition , we have
This is symmetric with respect to and because .
-
We compute that
Similarly,
where the second equality follows from rewriting the variable by , . Therefore, is symmetric with respect to and .
∎
Proof for the random saving model: Take arbitrary . We use the similar notation as the proof for the immediate exchange model.
where for . We have for . Also, for given so that for every , there exists unique which satisfies for every . Therefore,
This is symmetric with respect to and .
If we define as , , then and the above argument yields that is symmetric with respect to and . Therefore, the uniform reshuffling model is doubly stochastic and its unique stationary distribution is give by the uniform distribution on . Actually, this matches when 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 and there exist and such that . We prepare several notations. Define , and , . By the assumption on , we have the following asymptotics of as .
(3.1) |
where is a constant which depends on . In particular, the radius of convergence of is and it holds that for every and . We define the exponential family of distributions on by , . It is easy to see that is continuous, increasing in and diverges to infinity as . Hence, for every there exists unique such that . To examine the asymptotic behavior of as , we use a Tauberian theorem of the following form (cf. [1, Corollary 1.7.3]).
Theorem 3.1.
Let be a sequence of non-negative numbers and assume that converges for and is monotone. Then, the following are equivalent.
-
•
as for and slowly varying function .
-
•
as for and slowly varying function .
By this theorem and (3.1), the following asymptotics holds in the limit .
(3.2) |
where we used the relation for every . Therefore,
and this yields that
(3.3) |
as . By these asymptotics we also have
(3.4) |
as .
Next, for each and , let be the product measure on whose one site marginal distribution equals to . denotes the conditioned probability of on the event that the total number of coins on equals to . Then, we have the following key identity for (1.4).
Notice that the right-hand side does not depend on the choice of the parameter . Let be a bounded continuous local function where is a finite subset of . For every so that , we have
(3.5) |
where denotes the configurations restricted on the set . Now, we set where be a sequence of positive integers that satisfies for some constant and divergent sequence . We also take in the right-hand side of (3.5) as . For this choice of we show that and as .
For the proof of the convergence of , we assume that is a function of one variable for notational simplicity. The general case can be proven by the similar manner since is a product measure with the same marginal distribution. Firstly, we consider the case . Let . By (3.1) and (3.2),
as . We note that since in the limit , a finite sum in is negligible and we can replace with the right-hand side of (3.1) with an error of . Then, by (3.3) and the condition on ,
as where the convergence follows from Riemann integral. By taking the limit , the right-hand side converges to . For , we have
for every large enough where are positive constants independent of . By taking the limits and , we obtain .
Secondly, we consider the case . By (3.1) and (3.2) again,
as . and this yields that
In the limits and , the first term of the right-hand side converges to and the second term vanishes due to the extra factor . goes to in the same way as . The case also follows from the similar argument.
Next, for the convergence of , we use the following local limit theorem.
Theorem 3.2.
Let be a sequence of positive numbers which satisfies and set . For each , denotes a family of independent and identically distributed -valued random variables with common distribution . Then, for every finite set , it holds that
(3.6) |
where .
The proof of this theorem is given later. By applying this theorem for and , we have
as for every . Note that terms do not depend on . Set , . Then, the above asymptotics and (3.4) yield that
as for every and
for some constant . Therefore,
The first term of the right-hand side goes to as for every and the second term goes to as and by the similar computation as the estimate of above. Hence, we obtain and this completes the proof.
If we assume the condition: and for instead, then the proof for the case above can be applied as it is.
∎
Remark 3.1.
If in (3.6) then the exponential term converges to , 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 of into and , the estimate for the latter part was reduced to the estimate of which can be managed because we know the explicit form of .
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 depends on the number of random variables . Combining this with the fact that as , 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 , . 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 as . Therefore, we give the proof of the theorem according to the classical argument [20, Chapter VII]. For notational simplicity we only consider the case . The modification for the general finite set is straightforward.
Set , and . We define and , where . By the inversion formula, we have
Therefore,
for every and . We show that the right-hand side converges to as and .
For , assume that the law of converges to the standard normal distribution. Then, we have for every and we obtain as by the bounded convergence theorem. For the convergence of the law of , we have only to show that as for every by Lindberg’s central limit theorem (cf. [7, Theorem 3.4.10]). By (3.4), implies that for every large enough. Therefore,
where the second inequality follows from the fact that for every . We first consider the case for the estimate of .
for some constants and every large enough where we used (3.2), (3.3) and (3.4) for the first inequality. The right-hand side goes to as because
as for every . By the similar computation we obtain when . For the estimate of , we have
as where we used Markov’s inequality and (3.4).
Next, we consider . By Taylor’s theorem, there exists such that for every which satisfies , we have
Therefore, for every which satisfies . Taking in the definition of as , we obtain
as . Similarly, we have as .
The final task is the estimate of . We take in the definition of as above. Let be fixed and define . By the proof of [17, Lemma 5.4], we know that
We have as by (3.2) and (3.3). Also,
By the assumption on , (3.1) and (3.3), for every there exists and such that for every and . Moreover, for every and where is a constant independent of and . Therefore,
The first term of the right-hand side converges to if and if for some by the similar computation as before. The second term is less than and this goes to as . As a result, for every there exists such that as and
for every large enough and every . By taking the limit and using the estimate for every , we obtain
for every so that . Hence, by taking small enough, there exists such that for every large enough and so that . This yields that
as and we can complete the proof of Theorem 3.2.
∎
Proof of Corollary 1.1. By Proposition 1.1, we have
Therefore, it is sufficient to show that . This follows from Theorem 1.1 and the basic facts about the weak convergence of probability measures. The same is true for and .
∎
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 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).