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

Present address: ]Los Alamos National Laboratory, Los Alamos, New Mexico 87545

Simulation of a generalized asset exchange model with economic growth and wealth distribution

Kang K. L. Liu Department of Physics, Boston University, Boston, Massachusetts 02215    N. Lubbers [ Department of Physics, Boston University, Boston, Massachusetts 02215    W. Klein Department of Physics, Boston University, Boston, Massachusetts 02215 Center for Computational Science, Boston University, Boston, Massachusetts 02215    J. Tobochnik Department of Physics, Kalamazoo College, Kalamazoo, Michigan 49006 Department of Physics, Boston University, Boston, Massachusetts 02215    B. M. Boghosian Department of Mathematics, Tufts University, Medford, Massachusetts 02155    Harvey Gould [email protected] Department of Physics, Boston University, Boston, Massachusetts 02215 Department of Physics, Clark University, Worcester, Massachusetts 01610
Abstract

The agent-based Yard-Sale model of wealth inequality is generalized to incorporate exponential economic growth and its distribution. The distribution of economic growth is nonuniform and is determined by the wealth of each agent and a parameter λ\lambda. Our numerical results indicate that the model has a critical point at λ=1\lambda=1 between a phase for λ<1\lambda<1 with economic mobility and exponentially growing wealth of all agents and a non-stationary phase for λ1\lambda\geq 1 with wealth condensation and no mobility. We define the energy of the system and show that the system can be considered to be in thermodynamic equilibrium for λ<1\lambda<1. Our estimates of various critical exponents are consistent with a mean-field theory (see following paper). The exponents do not obey the usual scaling laws unless a combination of parameters that we refer to as the Ginzburg parameter is held fixed as the phase transition is approached. The model illustrates that both poorer and richer agents benefit from economic growth if its distribution does not favor the richer agents too strongly. This work and the following theory paper contribute to our understanding of whether the methods of equilibrium statistical mechanics can be applied to economic systems.

asset exchange, growth, equality, driven system

I Introduction and the GED model

Although economies are complex systems that are difficult to understand [1], the consideration of simple models can provide insight if the questions are of a statistical nature and about the economy as a whole. One such question is whether economic growth can benefit all members of society [1, 3, 2]. Another question is to what degree is wealth accumulation a natural consequence of the way that wealth is exchanged and distributed. Whether these questions and others can be treated using methods appropriate to equilibrium systems is not settled [4].

In this paper we approach these questions by simulating an agent-based model that incorporates wealth exchange, economic growth, and its distribution. Agent-based wealth-exchange models exhibit a rich phenomenology [5, 6, 7, 8, 9, 10, 11, 12]. The agent-based asset-exchange model we generalize belongs to a class of wealth exchange models that has been of considerable interest and has provided insight into how economies function [13, 14, 15, 16, 17, 22, 23, 18, 19, 20, 21]. In these models the amount of wealth transfer is determined stochastically to represent the uncertainty of the value of an agent’s assets.

A particularly interesting agent-based model that incorporates wealth exchange is the Yard-Sale model [8, 10, 7, 14, 24, 25, 22, 21, 13, 19, 9, 17, 20, 18, 11, 12, 15, 16, 5, 6, 23]. In this model two agents are chosen at random and a fraction ff of the wealth of the poorer agent is transferred to the winning agent. The latter is determined at random with probability 1/2. If an equal amount of wealth is initially assigned to the NN agents, the result is that after many exchanges, the wealth is concentrated among increasingly fewer agents, culminating, in the limit of infinitely many wealth exchanges and NN\to\infty, to a single agent holding a finite fraction of the total wealth [14, 5, 23, 22, 9, 15, 17, 21, 25].

To investigate the effect of economic growth on the wealth distribution, we generalize the Yard-Sale model so that the wealth μW(t)\mu W(t) is added to the system after NN exchanges, where W(t)W(t) is the total wealth at time tt and the time tt is defined such that one unit of time corresponds to NN exchanges; W(t)W(t) grows exponentially with the rate μ>0\mu>0. The motivation for this type of growth is the expected annual increase in the gross domestic product [26, 27].

To distribute the increase of the total wealth due to growth to individual agents, we introduce the distribution parameter λ0\lambda\geq 0 and assign the added wealth to agent ii according to their wealth wi(t)w_{i}(t) at time tt as

Δwi(t)=μW(t)wiλ(t)i=1Nwiλ(t).\Delta w_{i}(t)=\mu W(t)\dfrac{w^{\lambda}_{i}(t)}{\sum_{i=1}^{N}w^{\lambda}_{i}(t)}. (1)

The form of Eq. (1) implies that as λ\lambda increases, the allocation of the added wealth is weighted more toward agents with greater wealth. We will refer to the model with the incorporation of exponential growth of the total wealth and the λ\lambda-dependent distribution mechanism in Eq. (1) as the Growth, Exchange, and Distribution (GED) model.

The motivation for the form of the wealth distribution in Eq. (1) is that in practice, not all agents benefit equally from economic growth, and that agents with more assets and resources are able to take more advantage of the growth of the economy. We argue in the Appendix that the allocation of growth according to Eq. (1) is consistent with economic data.

The distribution parameter λ\lambda, exchange parameter ff, growth parameter μ\mu, and the number of agents NN determine the wealth distribution in the model. Our primary results can be grouped into two categories – the implications for economic systems and the implications for our understanding of the statistical mechanics of systems that are near-mean-field. We find that there is a phase transition at λ=1\lambda=1 such that for λ<1\lambda<1, all agents benefit from economic growth, there is economic mobility, and the system is in thermodynamic equilibrium. In contrast, for λ1\lambda\geq 1, the system is non-stationary, there is no economic mobility, and there is wealth condensation as in the original Yard-Sale model. In the context of statistical mechanics we note that we can define an energy that satisfies the Boltzmann distribution for λ<1\lambda<1. The existence of the latter is consistent with the assumption that that the system is not just in a steady state, but is in thermodynamic equilibrium for λ<1\lambda<1.

The remainder of the paper is structured as follows: In Sec. II we show that the wealth distribution reaches a steady state and that wealth condensation is avoided for λ<1\lambda<1. In Sec. III we show that the GED model is effectively ergodic and that there is economic mobility for λ<1\lambda<1. In Sec. IV we find that we can define an energy that satisfies the Boltzmann distribution. We characterize the phase transition at λ=1\lambda=1 for a fixed number of agents in Sec. V.1 and estimate the critical exponents β\beta, γ\gamma, and α\alpha associated with the behavior of the order parameter, its variance, and the energy, respectively. The result is that the exponents determined for fixed NN do not satisfy the usual scaling laws. The consequences of the system being describable by a mean-field theory [28] are discussed in Sec. V.2, where we introduce the Ginzburg parameter and show that scaling is restored if the transition is approached at fixed Ginzburg parameter rather than for fixed NN. In Sec. VI we show that there is critical slowing down consistent with the mean-field theory predictions of Ref. [28]. We summarize and discuss our results in Sec. VII. In the appendix we argue from economic data that our method of distributing the growth is a reasonable zeroth order approximation.

II Steady state wealth distribution

Refer to caption
Figure 1: The time dependence of the rescaled wealth distribution as a function of the rank of N=2500N=2500 agents for λ=0\lambda=0 at t=100t=100 (\color[rgb]{1,0,0}\bullet), t=400t=400 (\color[rgb]{0,1,0}\blacktriangle), t=1600t=1600 (\square), and t=6400t=6400 (×\color[rgb]{0,0,1}\times). For t>1600t>1600, the rescaled wealth distributions collapse onto the same curve indicating that the rescaled wealth distribution has reached a steady state. Each agent is initially assigned wealth one, so that the total wealth is NN (N=2500N=2500, f=0.1f=0.1, μ=0.001\mu=0.001).

Because the total wealth increases exponentially, it is convenient to introduce the rescaled wealth of an agent,

w~i(t)=NW(t)wi(t),\widetilde{w}_{i}(t)=\frac{N}{W(t)}w_{i}(t), (2)

and consider the rescaled wealth distribution of the NN agents. That is, after the increased wealth due to growth is distributed to the agents, their wealth is scaled so that the total rescaled wealth equals NN, the initial total wealth. In the following all references to the wealth of the agents will be to their rescaled wealth, and we will omit the tilde for simplicity.

The simulation of the GED model proceeds as follows:

  1. 0.

    Usually, we assign the wealth of each agent at random and then rescale wiw_{i} so that the total wealth iwi(t=0)\sum_{i}w_{i}(t=0) is equal to NN. The results for this section only are for wi(t=0)=1w_{i}(t=0)=1.

  2. 1.

    Choose agents ii and jj at random regardless of their wealth and determine the amount fmin[wi(t),wj(t)]f\min[w_{i}(t),w_{j}(t)] to be exchanged. Choose at random which agent gains and which agent loses.

  3. 2.

    Implement step one NN times. Because the NN exchanges conserve the total wealth, the latter remains equal to NN.

  4. 3.

    Assign the additional wealth due to growth to the agents according to Eq. (1).

  5. 4.

    Rescale wi(t)w_{i}(t) so that W(t)=wi(t)=NW(t)=\sum w_{i}(t)=N.

  6. 5.

    Set t=t+1t=t+1.

  7. 6.

    Repeat steps 2–5 until a steady state wealth distribution is attained (for λ<1\lambda<1) and then determine the average values of the desired quantities of interest.

The simulations in this section are for N=2500N=2500, f=0.1f=0.1, μ=0.001\mu=0.001, and various values of the distribution parameter λ\lambda. The qualitative results discussed here do not depend on the values of the number of agents NN, the fraction of the poorer agents’s wealth that is exchanged ff, and the growth parameter μ\mu.

We show in Fig. 1 the time dependence of the rescaled wealth distribution for λ=0\lambda=0, starting from the initial condition wi(t=0)=1w_{i}(t=0)=1 for all ii. The wealth disparity between richer and poorer agents initially increases until a steady state is established. Once a steady state is reached, the rescaled wealth distribution remains fixed, and the wealth in every rank increases as eμte^{\mu t}.

Refer to caption
Figure 2: The rescaled wealth distribution versus the rank of N=2500N=2500 agents for λ=0.2\lambda=0.2 (\color[rgb]{1,0,0}\bullet), λ=0.4\lambda=0.4 (\color[rgb]{0,1,0}\blacktriangle), λ=0.6\lambda=0.6 (\blacksquare), and λ=0.8\lambda=0.8 (\color[rgb]{0,0,1}\blacklozenge) at t=106t=10^{6} after a steady state has been reached. The wealth distribution becomes less equal as λ1\lambda\to 1^{-} (f=0.1f=0.1, μ=0.001\mu=0.001).

According to Eq. (1), the growth allocation is weighted more toward the richer agents as λ1\lambda\to 1^{-}, thus leading to a less equal steady state rescaled wealth distribution (see Fig. 2). The time to reach a steady state increases as λ\lambda approaches 11, and as for λ=0\lambda=0, the wealth of all agents increases exponentially after a steady state has been reached. (The limit λ1\lambda\to 1 will always be from below unless otherwise specified.) We find that for 0λ<10\leq\lambda<1, “a rising tide lifts all boats” and all agents benefit from economic growth.

Refer to caption
Figure 3: The rescaled wealth distribution at t=500t=500 (\color[rgb]{1,0,0}\bullet), t=10000t=10000 (\blacksquare), and t=50000t=50000 (\color[rgb]{0,0,1}\blacklozenge) for λ=1\lambda=1. In contrast to the behavior for λ<1\lambda<1, the slope of the rescaled wealth distribution decreases with time and a steady state is not reached. The time for a single agent to gain almost all of the wealth is a decreasing function of λ\lambda for λ>1\lambda>1 (N=2500N=2500, f=0.1f=0.1, μ=0.001\mu=0.001).

The time-dependence of the rescaled wealth distribution is shown for λ=1\lambda=1 in Fig. 3. A steady state is not reached, and the slope of the rescaled wealth distribution increases with time, corresponding to the accumulation of wealth by fewer and fewer agents until eventually a single agent gains almost all the wealth. Similar results are found for λ>1\lambda>1. The time for a single agent to dominate decreases as λ\lambda increases for λ>1\lambda>1.

Note that λ=1\lambda=1 is a special case for which the increase in an agent’s wealth due to growth is proportional to the agent’s wealth. Hence, for λ=1\lambda=1, the ratio of the rescaled wealth of any two agents does not change after the distribution of the growth in wealth according to Eq. (1). Consequently, aside from the exponential growth of the total wealth, the entire dynamics of the system is driven only by the wealth exchange mechanism, and the evolution of the wealth distribution for λ=1\lambda=1 is identical to its evolution in the original Yard-Sale model; that is, the model with no economic growth.

III Effective ergodicity and economic mobility

The numerical results in Sec. II indicate that the GED model exhibits distinct behavior for λ<1\lambda<1 and λ1\lambda\geq 1. In particular, for λ<1\lambda<1, all agents benefit from economic growth, whereas for λ1\lambda\geq 1 only the richest agent becomes richer. In Sec. III.1 we show that the GED model is effectively ergodic for λ<1\lambda<1, but is not ergodic for λ1\lambda\geq 1. In Sec. III.2 we find that the agents have nonzero economic mobility for λ<1\lambda<1, but have zero mobility for λ1\lambda\geq 1. Unlike in Sec. II, we randomly assign the wealth of each agent at t=0t=0 from a uniform distribution and then rescale the agents’ wealth so that the initial total wealth equals NN.

III.1 Effective ergodicity

To determine whether the system is effectively ergodic, we define the (rescaled) wealth metric as [29]

Ω(t)=1Ni=1N[w¯i(t)w¯(t)]2,\Omega(t)=\frac{1}{N}\sum_{i=1}^{N}\big{[}{\overline{w}}_{i}(t)-\overline{w}(t)\big{]}^{2}, (3)

where w¯i(t)\overline{w}_{i}(t) is the time averaged wealth of agent ii at time tt,

w¯i(t)\displaystyle\overline{w}_{i}(t) =1t0twi(t)𝑑t,\displaystyle=\frac{1}{t}\!\int_{0}^{t}w_{i}(t^{\prime})\,dt^{\prime}, (4)
w¯(t)\displaystyle\overline{w}(t) =1Ni=1Nw¯i(t).\displaystyle=\frac{1}{N}\sum_{i=1}^{N}{\overline{w}}_{i}(t). (5)

The metric Ω(t)\Omega(t) in Eq. (3) is a measure of how the time averaged wealth of each agent approaches the wealth averaged over all agents. If the system is effectively ergodic, Ω(t)1/t\Omega(t)\propto 1/t [29]. Effective ergodicity is a necessary, but not a sufficient condition for ergodicity.

The linear time-dependence of Ω(0)/Ω(t)\Omega(0)/\Omega(t) shown in Fig. 4(a) for λ<1\lambda<1 implies that the system is effectively ergodic for λ<1\lambda<1. In contrast, Ω(0)/Ω(t)\Omega(0)/\Omega(t) for λ=1\lambda=1 does not increase linearly with tt [see Fig. 4(b)], and hence the system is not ergodic. Similar results are found for λ>1\lambda>1.

Refer to caption
Refer to caption
Figure 4: (a) The linear time-dependence of the inverse wealth metric indicates that the system is effectively ergodic for λ=0.5\lambda=0.5 (upper line) and λ=0.9\lambda=0.9 (bottom line). The inverse slopes are 37773777 and 4430044300 for λ=0.5\lambda=0.5 and λ=0.9\lambda=0.9, respectively. (b) The inverse wealth metric for λ=1.0\lambda=1.0 does not increase linearly and the system is not ergodic (N=5000N=5000, f=0.01f=0.01, μ=0.1\mu=0.1).

III.2 Economic mobility

In a system with economic mobility, poorer agents can become wealthier and richer agents can become poorer. In contrast, agents in a system with very low economic mobility rarely change their rank and the very rich become richer [30].

To determine the mobility, we rank the agents according to their wealth at various times and compute the correlation function C(t)C(t) of the ranks of the agents once a steady state has been reached for λ<1\lambda<1. The rank correlation function Ci(t)C_{i}(t) of agent ii is defined as

Ci(t)=Ri(t)Ri(0)Ri2Ri2Ri2,C_{i}(t)=\frac{\langle R_{i}(t)R_{i}(0)\rangle-\langle R_{i}\rangle^{2}}{\langle R_{i}^{2}\rangle-\langle R_{i}\rangle^{2}}, (6)

where Ri(t)R_{i}(t) is the rank of agent ii at time tt and Ri=N/2\langle R_{i}\rangle=N/2. The corresponding quantity for λ1\lambda\geq 1, where a steady state is not reached, is the Pearson correlation function given by [31]

Ci(t)=[Ri(t)R(t)][Ri(0)R(0)][(Ri(t)Ri(t))2][(Ri(0)Ri(0))2],C_{i}(t)=\frac{\big{[}R_{i}(t)-\langle R(t)\rangle\big{]}\big{[}R_{i}(0)-\langle R(0)\rangle\big{]}}{\sqrt{\big{[}\big{(}R_{i}(t)-\langle R_{i}(t)\rangle\big{)}^{2}\big{]}\big{[}\big{(}R_{i}(0)-\langle R_{i}(0)\rangle\big{)}^{2}\big{]}}}, (7)

The correlation function averaged over all agents is C(t)=(1/N)iCi(t)C(t)=(1/N)\sum_{i}C_{i}(t). As can be seen from Fig. 5(a), C(t)0C(t)\rightarrow 0 as tt\rightarrow\infty for λ<1\lambda<1, which indicates that the rank of an agent as tt\rightarrow\infty is not correlated with its rank at t=0t=0 and the agents have a nonzero economic mobility for λ<1\lambda<1. The λ\lambda dependence of the average time that the richest agent remains the richest is discussed in Sec. VI. In contrast, in Fig. 5(b) we see that C(t)C(t) approaches a constant for λ1\lambda\geq 1, indicating that the ranks are strongly correlated at different times, and there is no economic mobility.

Refer to caption
Refer to caption
Figure 5: (a) The rank correlation decays exponentially (red line) for λ=0.9\lambda=0.9, indicating that the mobility is nonzero. (b) The Pearson correlation function for λ=1.0\lambda=1.0 (\bullet) and λ=1.1\lambda=1.1 (\square). For λ1\lambda\geq 1, C(t)C(t) remains nonzero even as tt\rightarrow\infty, which indicates that the rank of an agent remains correlated and there is no mobility (N=2500N=2500, f=0.1f=0.1, μ=0.001\mu=0.001).

IV Equilibrium, not just steady state

We have seen that the GED model approaches a steady state and is effectively ergodic for λ<1\lambda<1. In the following, we will show that a reasonable definition of the total energy yields an energy distribution that is consistent with the Boltzmann distribution. The existence of the latter is consistent with the idea that that the system is not just in a steady state, but is in thermodynamic equilibrium for λ<1\lambda<1.

As discussed in Ref. [28] (following paper), a mean-field theory treatment of the GED model yields a quantity that can be interpreted as the total energy of the system at time tt

E(t)=i=1N[1wi(t)]2.E(t)=\sum_{i=1}^{N}[1-w_{i}(t)]^{2}.\\ (8)

Note that the energy is zero if all agents have the same wealth (wi=1w_{i}=1). Equation (8) yields a mean energy that is extensive, that is, EN\langle E\rangle\propto N for fixed values of λ\lambda, ff, and μ\mu. For example, for λ=0.99\lambda=0.99, f=0.01f=0.01, and μ=0.1\mu=0.1, we find that EN=4000/EN=1000=336.4/84.2=3.9954\langle E_{N=4000}\rangle/\langle E_{N=1000}\rangle=336.4/84.2=3.995\approx 4.

The probability density P(E)P(E) is shown in Fig. 6(a) for N=5000N=5000, λ=0.8\lambda=0.8, f=0.01f=0.01, and μ=0.1\mu=0.1. As expected, P(E)P(E) is fit well by a Gaussian. Fits of P(E)P(E) to a Gaussian become less robust as λ1\lambda\to 1 for fixed NN, ff, and μ\mu. We will discuss the behavior of P(E)P(E) for larger values of λ\lambda in Sec. V.2.

If the system is in thermodynamic equilibrium, we expect that the probability density P(E,β)P(E,\beta) to be proportional to g(E)eβEg(E)e^{-\beta E}, where β\beta is an effective inverse temperature that depends on λ\lambda, ff, and μ\mu, and g(E)g(E) is the density of states, which is independent of λ\lambda, ff, and μ\mu and hence independent of β\beta. Because g(E)g(E) is independent of β\beta, the ratio P(E,β1)/P(E,β2)P(E,\beta_{1})/P(E,\beta_{2}) is an exponential proportional to exp[(β1β2)]E\exp[-(\beta_{1}-\beta_{2})]E if the system is characterized by the Boltzmann distribution with the energy given by Eq. (8). The range of values of EE over which this ratio is nonzero and finite is limited by the overlap of the two probabilities, which becomes smaller as NN is increased. In Fig. 6(b) we see that the ratio P(E,β2)/P(E,β1)P(E,\beta_{2})/P(E,\beta_{1}) is consistent with the Boltzmann distribution e(β2β1)Ee^{-(\beta_{2}-\beta_{1})E} for N=5000N=5000, λ=0.8\lambda=0.8, and μ=0.1\mu=0.1, with βf1\beta\propto f^{-1}, f2=0.0095f_{2}=0.0095 and f1=0.01f_{1}=0.01; a larger value of ff (for fixed values of λ\lambda and μ\mu) corresponds to a smaller value of β\beta and hence a higher value of the effective temperature. The dependence of the effective temperature β1\beta^{-1} on ff is consistent with the association of ff with the presence of both additive and multiplicative noise in the system [28]. Similar results hold for two similar values of λ\lambda for fixed ff and μ\mu. The existence of an energy, and the observation that its probability is proportional to the Boltzmann distribution implies that the system can be considered to be in thermodynamic equilibrium, at least for λ=0.8\lambda=0.8.

Refer to caption
Refer to caption
Figure 6: (a) The energy probability density P(E)P(E) for N=5000N=5000, λ=0.8\lambda=0.8, f=0.01f=0.01, and μ=0.1\mu=0.1 is consistent with the Gaussian distribution P(E)exp(EE¯)2/σE2)P(E)\propto\exp(E-\overline{E})^{2}/\sigma_{E}^{2}), with E¯=24.8\overline{E}=24.8 and σE=0.44\sigma_{E}=0.44 [red curve]. (b) The ratio P(E,f1=0.01)/P(E,f2=0.0095)P(E,f_{1}=0.01)/P(E,f_{2}=0.0095) is consistent with an exponential of the form eΔβEe^{-\Delta\beta E}, with Δβ1.08\Delta\beta\approx 1.08 (red curve).

V Characterization of the phase transition

V.1 Fixed number of agents

The numerical results discussed in this section are for N=5000N=5000, f=0.01f=0.01, and μ=0.1\mu=0.1. Averages are taken over a time of 10610^{6} (N×106N\times 10^{6} exchanges) after a transient time of 10610^{6}. The major source of uncertainty in our estimations of the values of the various critical exponents is the choice of the range of values to be retained in the least squares fits.

Because we will characterize the approach to the phase transition at λ=1\lambda=1 in terms of power laws, it is convenient to define the quantity

ϵ1λ,\epsilon\equiv 1-\lambda, (9)

and will assume that ϵ>0\epsilon>0.

To characterize the phase transition at ϵ=0\epsilon=0, we need to identify an order parameter. A common measure of income or wealth inequality is the Gini coefficient GnG_{n} [32]. If all agents have the same wealth, Gn=0G_{n}=0. In contrast, if one agent has all the wealth, Gn=1G_{n}=1, corresponding to the maximum degree of inequality. These characteristics appear to make 1Gn1-G_{n} a reasonable choice of the order parameter. However, because the wealth distribution reaches a steady state for λ<1\lambda<1, the fluctuations of the Gini coefficient are zero in the limit NN\to\infty, and hence the susceptibility, which would be associated with the variance of GnG_{n}, would be zero, making GnG_{n} an inappropriate choice of the order parameter.

The order parameter and the value of β\beta. Another choice of the order parameter is the fraction of the wealth held by all the agents except the richest agent, that is,

ϕ=NwmaxN,\phi=\frac{N-w_{\max}}{N}, (10)

where wmaxw_{\max} is the wealth of the richest agent. For λ<1\lambda<1 we find that wmaxNw_{\max}\ll N and depends weakly on NN for fixed values of λ\lambda, ff, and μ\mu. For example, wmax=2.0w_{\max}=2.0 for N=1000N=1000 and wmax=2.1w_{\max}=2.1 for N=4000N=4000, with λ=0.99\lambda=0.99, f=0.01f=0.01, and μ=0.1\mu=0.1. The weak dependence of wmaxw_{\max} on NN implies that ϕ\phi defined in Eq. (10) approaches one as NN\to\infty for λ<1\lambda<1 independently of the value of λ\lambda (see Fig. 7). Hence, the value of the critical exponent β\beta associated with the order parameter is β=0\beta=0. We also point out that for λ1\lambda\geq 1, the order parameter is zero in the limit NN\to\infty. If this branch is continued to λ<1\lambda<1, ϕ\phi will remain zero (because all agents except the richest have zero wealth), indicating that there is hysteresis. This behavior of ϕ\phi raises the question of the order of the transition at λ=1\lambda=1. One possibility is that the transition is a spinodal [33].

Refer to caption
Figure 7: The λ\lambda-dependence of the order parameter ϕ\phi, the fraction of wealth held by all agents except the richest agent, for N=5000N=5000 (top curve) and N=1000N=1000 (bottom curve) with f=0.01f=0.01 and μ=0.1\mu=0.1. In the limit NN\to\infty the order parameter becomes a step function with ϕ=1\phi=1 for λ<1\lambda<1 and ϕ=0\phi=0 for λ1\lambda\geq 1.

The susceptibility and the value of γ\gamma. The ϵ\epsilon-dependence of the variance of wmaxw_{\max} is shown in Fig. 8(a) and can be fit to a power law to give an effective exponent for the susceptibility close to one for ϵ2×103\epsilon\geq 2\times 10^{-3}; fits for smaller values of ϵ\epsilon give an effective exponent approximately equal to 1.7. Both estimates indicate that the variance of the wealth of the richest agent diverges strongly, and hence we choose the order parameter to be as defined in Eq. (10).

More consistent results for the susceptibility can be found from C2C_{2}, the variance of the wealth of a single agent averaged over all agents, and not just the variance of the wealth of the richest agent. In Fig. 8(b) we see that there is less curvature in the plot of logC2\log C_{2} versus logϵ\log\epsilon, and we find an effective exponent close to one if fits are made for ϵ<0.0015\epsilon<0.0015. If values of C2C_{2} are included for larger values of ϵ\epsilon, the effective exponent from the least squares fits is in the range [0.93,1.0][0.93,1.0]. Given these much better fits, we associate the susceptibility with NC2NC_{2}:

χ=NC2,\chi=NC_{2}, (11)

and conclude that the critical exponent associated with the susceptibility is γ1\gamma\approx 1.

Refer to caption
Refer to caption
Figure 8: (a) The ϵ\epsilon-dependence of the variance of the wealth of the richest agent, shows crossover behavior. Fits of the variance to a power law for ϵ=(1λ)2×103\epsilon=(1-\lambda)\geq 2\times 10^{-3} give an effective exponent close to one (red curve); fits for smaller values of ϵ\epsilon give an effective exponent equal to 1.7\approx 1.7 (blue curve). (b) The ϵ\epsilon-dependence of C2C_{2}, the variance of the wealth of the individual agents, shows less curvature, and gives an effective exponent close to 1.0 (red curve) if fits are made for ϵ<0.0015\epsilon<0.0015; the effective exponent is in the range [0.93,1.0][0.93,1.0] depending on the values of ϵ\epsilon that are included in the least squares fits (N=5000N=5000, f=0.01f=0.01, μ=0.1\mu=0.1).

The mean energy, heat capacity, and the value of α\alpha. Although the total rescaled wealth is a constant, the energy depends on the way the wealth is distributed. We use Eq. (8) to determine the mean energy by averaging the quantity i=1N(1wi)2\sum_{i=1}^{N}(1-w_{i})^{2} over many realizations. Our results for the mean energy E\langle E\rangle are shown in Fig. 9(a). We see that the ϵ\epsilon-dependence of E\langle E\rangle is consistent with ϵ1α\epsilon^{1-\alpha} as ϵ0\epsilon\to 0 with α2\alpha\approx 2. This divergence of E\langle E\rangle is inconsistent with equilibrium statistical mechanics, which requires that the total energy be finite for finite values of NN.

Refer to caption
Refer to caption
Figure 9: (a) The divergent ϵ\epsilon-dependence of the mean total energy E\langle E\rangle for fixed NN, ff, and μ\mu is consistent with the power law ϵ1\epsilon^{-1} (red line). Least squares fit of E\langle E\rangle yield values of the divergence in the range [0.93,1.03][0.93,1.03]. (b) The ϵ\epsilon-dependence of the heat capacity CC is consistent with the power law ϵ2\epsilon^{-2} (red line). Least squares fits of CC yield values of the effective exponent in the range [1.8,2.1][1.8,2.1] (N=5000N=5000, f=0.01f=0.01, μ=0.1\mu=0.1).

We define the heat capacity CC to be the variance of the total energy defined in Eq. (8). The ϵ\epsilon-dependence of of CC is consistent with the dependence ϵα\epsilon^{-\alpha} with the exponent α=2\alpha=2 as shown in Fig. 9(b). Our determination of the value of α\alpha depends on the range of values of ϵ\epsilon that are included in the least squares fits and are in the range [1.8,2.1][1.8,2.1].

In summary, our numerical results for the critical exponents α\alpha, β\beta, and γ\gamma, determined as ϵ\epsilon is varied for a given value of NN are consistent with

β=0, γ=1, and α=2(fixed Nf, and μ).\beta=0,\mbox{ }\gamma=1,\mbox{ and }\alpha=2\qquad\mbox{(fixed $N$, $f$, and $\mu$).} (12)

These numerical values are inconsistent with the usual scaling law

α+2β+γ=2.\alpha+2\beta+\gamma=2. (13)

However, the value of α\alpha determined from the power law behavior of the heat capacity is consistent with the λ\lambda-dependence of the mean total energy, that is, C=E/λC=\partial\langle E\rangle/\partial\lambda.

V.2 Mean-field theory and fixed Ginzburg parameter

Although it is natural to determine the critical behavior of the GED model as the critical point is approached for a fixed number of agents, our numerical results for the critical exponents are not consistent with the scaling law, Eq. (13), nor consistent with equilibrium statistical mechanics because the mean energy per agent diverges as the critical point is approached even for finite values of NN and fixed values of ff and μ\mu. Simulations also show that E\langle E\rangle and the heat capacity CC are proportional to NN for fixed values of λ\lambda, ff, and μ\mu so that the divergent behavior of E\langle E\rangle is not removed by first taking the limit NN\to\infty before taking the limit ϵ0\epsilon\to 0.

In Ref. [28] (following paper) a mean-field treatment of the GED model is developed based on the random exchange of wealth between an agent chosen at random and an agent whose wealth is assigned to be equal to the mean wealth of the remaining agents. The mean-field treatment predicts that the critical exponents are given by

β=0, γ=1, and α=1.\beta=0,\mbox{ }\gamma=1,\mbox{ and }\alpha=1. (14)

The predicted mean-field values of the exponents in Eq. (14) are consistent with Eq. (13). The mean-field theory [28] also predicts that the mean energy per agent approaches a constant as ϵ0\epsilon\to 0.

To compare the mean-field predictions with the simulations for different values of NN we need to account for the fact ff and μ\mu are rates and depend on the definition of the unit of time. It is shown in the mean-field theory of Ref. [28] that to achieve a consistent thermodynamic description of the GED model, we need to rescale ff and μ\mu as we change NN so that

f=f0/N and μ=μ0/N.f=f_{0}/N\mbox{ and }\mu=\mu_{0}/N. (15)

Another condition for the applicability of the mean-field treatment of the GED model is that the Ginzburg parameter GG, defined as

Gμ0N(1λ)f02,G\equiv\frac{\mu_{0}N(1-\lambda)}{f_{0}^{2}}, (16)

be much greater than one and be held fixed as ϵ0\epsilon\to 0. As has been found for the long-range and fully connected Ising models [33, 34, 35], we will find that the mean-field theory predictions for the critical behavior of the energy and heat capacity of the GED model are consistent with equilibrium statistical mechanics only if the Ginzburg parameter is held fixed as the critical point is approached. We will also see that the results for β\beta and γ\gamma do not depend on keeping GG fixed as has been found for other fully connected models [33, 34, 35].

We emphasize that if μ\mu and ff are not rescaled in the simulations, the energy per agent would diverge as NN\rightarrow\infty even for fixed Ginzburg parameter.

To compare the mean-field theory predictions to the simulations, we choose N0=5000N_{0}=5000, f0=0.01f_{0}=0.01, and μ0=0.1\mu_{0}=0.1, with f=(N0/N)f0f=(N_{0}/N)f_{0} and μ=(N0/N)μ0\mu=(N_{0}/N)\mu_{0}. For a particular choice of the value of λ\lambda, we determine the value of NN needed to keep the value of GG in Eq. (16) fixed at G=106G=10^{6}. Our simulations are for 0.80λ0.9980.80\leq\lambda\leq 0.998 and 5×103N5×1055\times 10^{3}\leq N\leq 5\times 10^{5}.

Because our results for E\langle E\rangle and CC depend on keeping GG fixed, we first discuss their ϵ\epsilon dependence. In Fig. 10 we see that E/N\langle E\rangle/N approaches a constant as ϵ0\epsilon\to 0, in contrast to its divergent ϵ\epsilon behavior for fixed NN, ff, and μ\mu. Because Eϵ1α\langle E\rangle\sim\epsilon^{1-\alpha}, we find that simulations for fixed Ginzburg parameter are consistent with α=1\alpha=1.

Refer to caption
Figure 10: The ϵ\epsilon dependence of E/N\langle E\rangle/N, the mean energy per agent, for G=106G=10^{6} is consistent with the linear ϵ\epsilon-dependence a0+a1ϵa_{0}+a_{1}\epsilon as ϵ0\epsilon\to 0, with a00.005a_{0}\approx 0.005 and a10.002a_{1}\approx 0.002 (red line). The ϵ\epsilon dependence of E/N\langle E\rangle/N is given by Eq. (17).

Our numerical results for E\langle E\rangle are consistent with the relation

ENG.\langle E\rangle\sim\frac{N}{G}. (17)

Equation (17) is predicted by mean-field theory [28] for fixed GG. For fixed GG, Eq. (17) implies that E/N\langle E\rangle/N approaches a constant as ϵ0\epsilon\to 0, consistent with our simulations. In contrast, for fixed NN, ff, and μ\mu we find

ENG=Nf02Nμ0ϵ=N2f2Nμϵ=Nf2μϵ(fixed Nf, and μ),\langle E\rangle\sim\frac{N}{G}=\dfrac{Nf_{0}^{2}}{N\mu_{0}\epsilon}=\dfrac{N^{2}f^{2}}{N\mu\epsilon}=\dfrac{Nf^{2}}{\mu\epsilon}\qquad\mbox{(fixed $N$, $f$, and $\mu$)}, (18)

where f0=Nff_{0}=Nf and μ0=Nμ\mu_{0}=N\mu. Equation (18) implies that for fixed values of λ\lambda, ff, and μ\mu, E\langle E\rangle is proportional to NN and diverges as ϵ1\epsilon^{-1} for fixed values of NN, ff, and μ\mu; both behaviors are consistent with our simulations.

The ϵ\epsilon-dependence of CC, the variance of the total energy, for fixed GG is shown in Fig. 11. We see that CϵαC\sim\epsilon^{-\alpha}, with α1\alpha\approx 1, consistent with the prediction of mean-field theory [28].

Refer to caption
Figure 11: The ϵ\epsilon-dependence of CC, the variance of the total energy, for fixed Ginzburg parameter is consistent with the power law ϵα\epsilon^{-\alpha}, with α=1\alpha=1 (red line). A least square fit gives an exponent of 0.91\approx 0.91.

Our numerical results for the variance of the total energy are consistent with the relation

CNG2.C\sim\frac{N}{G^{2}}. (19)

For fixed GG, we have Nϵ1N\propto\epsilon^{-1} and hence Cϵ1C\sim\epsilon^{-1} and α=1\alpha=1. For fixed values of ff and μ\mu we have

CNf04N2μ02ϵ2=N4f4N3μ2ϵ2Nϵ2(fixed f and μ).C\sim\dfrac{Nf_{0}^{4}}{N^{2}\mu_{0}^{2}\epsilon^{2}}=\dfrac{N^{4}f^{4}}{N^{3}\mu^{2}\epsilon^{2}}\sim\frac{N}{\epsilon^{2}}\qquad(\mbox{fixed $f$ and $\mu$}). (20)

In this case α=2\alpha=2 and CC is proportional to NN for fixed values of λ\lambda, ff, and μ\mu, consistent with the simulations.

The ϵ1\epsilon^{-1} dependence of CC for fixed GG is consistent with the mean-field prediction in Ref. [28]. Note that the variance of the total energy is proportional to NN for fixed λ\lambda, ff and μ\mu, but is independent of NN for fixed Ginzburg parameter. This seemingly inconsistent dependence on NN is due to the dependence of NN on ϵ\epsilon for fixed Ginzburg parameter. This confusion is a consequence of the fully connected nature of the GED model. (Recall that an agent can exchange wealth with equal probability with any other agent in the system.) If the mean-field limit is taken according to the prescription of Kac et al. [36], then the agents would be placed, for example, on a two-dimensional lattice and the range RR over which the agents could exchange wealth would be finite [37]. In this case the Ginzburg parameter would be a function of RR rather than NN, and the specific heat would be the heat capacity divided by NN.

The ϵ1\epsilon^{-1} dependence of the variance of the total energy near ϵ=0\epsilon=0 implies that the mean energy must include a logarithmic dependence on λ\lambda. For example, the form, Ea0+a1ϵ+aL/lnϵ\langle E\rangle\sim a_{0}+a_{1}\epsilon+a_{L}/\ln\epsilon, implies that C=E/λC=\partial\langle E\rangle/\partial\lambda scales as ϵ1\epsilon^{-1} with a logarithmic correction. Mean-field theory is incapable of finding logarithmic corrections, and our data for E\langle E\rangle is not sufficiently accurate to detect the presence of logarithmic factors.

Although our numerical results for α\alpha depend on whether NN or GG is held fixed as λ1\lambda\to 1, our results for the order parameter and susceptibility exponents β\beta and γ\gamma are independent of the nature of the approach to the phase transition. We find that wmaxNw_{\max}\ll N and is independent of λ\lambda for fixed GG. Hence ϕ=1\phi=1 for λ<1\lambda<1, implying that β=0\beta=0 as was found for fixed NN. The divergence of the susceptibility χ=NC2\chi=NC_{2} shown in Fig. 12(b) is consistent with χϵγ\chi\sim\epsilon^{-\gamma} with γ=1\gamma=1.

Refer to caption
Figure 12: The divergence of the susceptibility NC2NC_{2} is consistent with χϵγ\chi\sim\epsilon^{-\gamma} with γ=1\gamma=1 (red line).

VI Critical slowing down

We find that various time scales increase rapidly as ϵ0\epsilon\to 0, which limits how close the simulations can approach the transition. One time scale of interest is τra\tau_{\rm ra}, the mean lifetime of the richest agent. We expect that if the mobility of the agents is nonzero, then the richest agent at t=0t=0 will no longer be the richest after some time has elapsed. We define τra\tau_{\rm ra} as the mean time that a particular agent remains the richest and assume that τra\tau_{\rm ra} is a simple measure of the decorrelation time of the wealth of individual agents.

Another time scale of interest is the mixing time τm\tau_{m} associated with the time-dependence of the wealth metric; τm\tau_{m} is related to the inverse slope of the wealth metric as

Ω(0)/Ω(t)=t/τm.\Omega(0)/\Omega(t)=t/\tau_{m}. (21)

We also computed the time-displaced energy autocorrelation function given by

CE(t)=E(t)E(0)E2E2E2,C_{E}(t)=\frac{\langle E(t)E(0)\rangle-\langle E\rangle^{2}}{\langle E^{2}\rangle-\langle E\rangle^{2}}, (22)

where E(t)E(t) is the value of the energy of the system at time tt. We find that CE(t)C_{E}(t) relaxes exponentially, and hence we can extract the energy decorrelation time τE\tau_{E}. Our simulation results for these various times are for fixed Ginzburg parameter (G=106G=10^{6}).

For fixed GG the simulation results for the ϵ\epsilon-dependence of τra\tau_{\rm ra} are shown in Fig. 13(a) and are consistent with the power law ϵ1\epsilon^{-1}. To obtain accurate results for the wealth metric Ω(t)\Omega(t), we averaged Ω(t)\Omega(t) over ten origins and found that the linear dependence of Ω(0)/Ω(t)\Omega(0)/\Omega(t) holds over a wide range of tt and yields robust values of τm\tau_{m}. The exponential dependence of CE(t)C_{E}(t) holds for t5τEt\lesssim 5\tau_{E}, yielding some uncertainty in the fitted values of τE\tau_{E}. The ϵ\epsilon-dependence of τm\tau_{m} and τE\tau_{E} are shown in Fig. 13(b). We see that both τm\tau_{m} and τE\tau_{E} increase rapidly as ϵ0\epsilon\to 0 and that their ϵ\epsilon-dependence is consistent with ϵ2\epsilon^{-2}.

Although the mean-field theory [28] makes no direct predictions for the mixing time, the ϵ\epsilon-dependence of τm\tau_{m} can be understood by noting that the metric measures the time it takes for the average wealth of each agent to equal the global average. Because the time it takes for the richest agent to cease being the richest and for another agent to assume that role diverges as approximately ϵ1\epsilon^{-1}, the time for a system of NN agents to mix is Nϵ1N\epsilon^{-1}. Because Nϵ1N\propto\epsilon^{-1} for fixed GG [see Eq. (16)], we find that τmϵ2\tau_{m}\sim\epsilon^{-2} in agreement with the simulations.

Refer to caption
Refer to caption
Figure 13: (a) The ϵ\epsilon-dependence for constant Ginzburg parameter (G=106G=10^{6}) of τra\tau_{\rm ra}, the lifetime of the richest agent, diverges as τraϵ0.9\tau_{\rm ra}\sim\epsilon^{-0.9} (red line). (b) The ϵ\epsilon-dependence of the mixing time τm\tau_{m} (top red line) and τE\tau_{E} (bottom blue line) diverge as ϵ2\sim\epsilon^{-2}. Least squares fits give an exponent in the range [1.95,2.1][1.95,2.1].

The mean-field theory of Ref. [28] predicts that the decorrelation time for fixed GG scales as

τmf1μ0ϵ.\tau_{\rm mf}\sim\frac{1}{\mu_{0}\epsilon}. (23)

The reason for the apparent discrepancy between the ϵ\epsilon-dependence predicted by Eq. (23) and our simulation results for τm\tau_{m} and τE\tau_{E} is that the time unit in the simulations corresponds to NN exchanges, during which one agent exchanges wealth with only one agent on the average. In contrast, the applicability of mean-field theory requires that in one time unit, one agent exchanges wealth with NN agents on the average. Hence, the simulation and mean-field time units differ by a factor of NN or 1/ϵ1/\epsilon for fixed Ginzburg parameter.

The divergent behavior of τm\tau_{m} and τE\tau_{E} are examples of critical slowing down, which is associated with a cooperative effect and is not a property of a single agent. In contrast, τra\tau_{\rm ra} is a property of a single agent rather than of the system as a whole and becomes independent of ϵ\epsilon if we define the time as required by the applicability of mean-field theory.

Although the results of our simulations are consistent with the ϵ\epsilon-dependence in Eq. (23) for constant Ginzburg parameter, Eq. (23) also predicts that τE\tau_{E} is independent of the values of ff and NN. As discussed in Ref. [28], the derivation of Eq. (23) neglects the effects of both additive and multiplicative noise. The weak dependence of τm\tau_{m} and τE\tau_{E} on NN and ff, as well as their dependence on μ\mu, is discussed in Ref. [28].

VII Discussion

We have generalized the Yard-Sale model to incorporate economic growth and its distribution according to the wealth of the agents as determined by Eq. (1) and the parameter λ\lambda. Our numerical results suggest that there are two phases. For λ<1\lambda<1 the system reaches a steady state with economic mobility, is effectively ergodic, and can be considered to be in thermodynamic equilibrium. In contrast, for λ1\lambda\geq 1 there is no economic mobility, the system does not reach a steady state, and in the limit tt\to\infty, there is condensation of a finite fraction of the system’s wealth in a vanishingly small number of agents. In addition, the system is not ergodic and shares some of the characteristics of the geometric random walk which is also not ergodic and cannot be treated by equilibrium methods [4, 38].

It is remarkable that it is possible to define a thermodynamic energy for a system that involves wealth and has no obvious energy analogue. The interpretation of the energy and its variance is subtle and thermodynamic consistency is achieved only if the mean-field limit is taken appropriately.

We showed in Sec. IV that P(E)P(E), the probability density of the energy of the system, is well fit by a Gaussian function for N=5000N=5000 and λ=0.8\lambda=0.8. Simulations for N=5000N=5000 and values of λ\lambda much closer to one show departures from a Gaussian, even though the wealth fluctuation metric still indicates that the system is effectively ergodic. The deviation of P(E)P(E) from a Gaussian for fixed NN (and fixed μ\mu and ff) is due to the fact that GG decreases as λ1\lambda\to 1 and eventually becomes too small for mean-field theory to be applicable. Simulations for fixed Ginzburg parameter begin to show deviations from a Gaussian distribution for λ\lambda much closer to one. Although the Ginzburg parameter is large and fixed, the importance of multiplicative noise increases as λ1\lambda\to 1, and eventually the effect of the multiplicative noise can no longer be ignored [28].

Our numerical values of the various exponents are consistent with the mean-field theory of Ref. [28], but their estimated numerical values must be viewed with caution because they are obtained by extrapolation over a limited range of λ\lambda and for finite values of GG and NN. Much larger values of GG would be needed to obtain more accurate numerical results for λ\lambda closer to one.

The transition at λ=1\lambda=1 is from a system in thermodynamic equilibrium for λ<1\lambda<1 to a system that undergoes wealth condensation for λ1\lambda\geq 1. In Ref. [28] it is shown that the evolution of the model for λ1\lambda\geq 1 is the same as unstable state evolution in the fully connected Ising model for model A dynamics [39]. In Fig. 14 we show the evolution of the wealth of the richest agent after the value of λ\lambda has been changed from λ=0.8\lambda=0.8 to λ=1.01\lambda=1.01 and to λ=1.05\lambda=1.05. We see that the wealth of the richest agent initially increases exponentially. We also find that the duration of exponential growth decreases for larger values of λ\lambda after the change (not shown).

Refer to caption
Figure 14: The time dependence of the wealth of the richest agent after an instantaneous change of λ\lambda from λ=0.8\lambda=0.8 to λ=1.01\lambda=1.01 (red) and from λ=0.8\lambda=0.8 to λ=1.05\lambda=1.05 (blue). Each change is averaged over five runs. The time tt is measured from the instantaneous change of λ\lambda after the system has reached equilibrium. The wealth increases exponentially as et/τqe^{t/\tau_{q}}, with τq390\tau_{q}\approx 390 for λ=1.01\lambda=1.01 and τq230\tau_{q}\approx 230 for λ=1.05\lambda=1.05, consistent with unstable state evolution [28] (N=20000N=20000, f=0.01f=0.01, and μ=0.1\mu=0.1).

Our results have possible important economic implications. As λ\lambda is increased, the benefits of growth are weighted more toward the wealthy, and wealth inequality increases. Nevertheless, as long as λ<1\lambda<1, the wealth of agents of all ranks grows at the same rate once a steady state is reached, and all agents benefit from economic growth. However, if the benefits of growth are skewed too much toward the wealthy (λ1\lambda\geq 1), poor and middle rank agents no longer benefit from economic growth, and wealth condensation occurs. For λ=1\lambda=1 the model reduces to the geometric random walk with resulting wealth condensation [4], as shown in the following paper [28].

There is some question whether economic systems can be treated as being in equilibrium or even exhibit effective ergodicity [4, 38, 10]. Our results suggest that ergodicity and equilibrium may depend on various system parameters. Because parameters such as λ\lambda and μ\mu are not temporal constants in real economies, our results also suggest that the applicability of equilibrium methods may be situational and vary with time.

Acknowledgements.
We thank Timothy Khouw, Ole Peters, John Ogren, Alan Gabel, Bill Gibson, Karen Smith, and Louis Colonna-Romano for useful conversations.

*

Appendix A Comparison to some economic data

Modeling the economy of a country as large and diverse as the United States by compressing economic growth and transactions into three parameters, λ\lambda, ff, and μ\mu is a gross simplification. The assumption that these parameters are independent of time also is unrealistic. In the following, we analyze the growth data [40] and wealth distribution data [41] compiled by Karen Smith and published by the Urban Institute. Our analysis suggests that the assumptions that the distribution of growth can be modeled as in Eq. (1) and that the parameters are independent of time is a reasonable zeroth order approximation to the distribution of wealth in the real economy. We will discuss the exchange term and its relevance to the real economy in the following paper [28].

From the growth rate of the gross domestic product shown in constant dollars in Ref. [40], we note that the temporal fluctuations of the (inflation adjusted) growth rate of the gross domestic product exhibits large swings that appear to be damped as a function of time. The decline in the growth caused by the great recession starting in 2008 is an example of a large fluctuation, but the growth rate has remained close to the mean rate of roughly 3% over the last 30 years, thus implying that μ0.03\mu\approx 0.03.

By using the wealth distribution chart in Ref. [41], we can calculate the change of the wealth of people in various percentiles. From the relation [see Eq. (1)], we can estimate λ\lambda as

λ=log(Wr(t2)Wr(t1)Wr(t1)),\lambda=\log\Bigg{(}\frac{W_{r}(t_{2})-W_{r}(t_{1})}{W_{r}(t_{1})}\Bigg{)}, (24)

where Wr(t)W_{r}(t) is the wealth of people of economic rank (percentile) rr at time tt.

We estimated λ\lambda for the 50th, 90th and 95th wealth percentiles in the intervals 1983–1989, 1995 –1998 and 2013–2016 (see Table 1). Although the values of λ\lambda are not constant for different time intervals and percentiles, they vary by only a few percentage points as a function of percentile. They vary more as a function of time, with the 50% percentile having the greatest variation. The change of λ\lambda appears to decrease for later times, consistent with the damping of the variation of the growth. The variation of λ\lambda is more pronounced for even lower rankings. However, because the wealth of the lower rankings is considerably smaller, the variation of the value of λ\lambda has less effect on the wealth of the poor.

percentile 1983–1989 1995–1998 2013–2016
95% 0.85 0.90 0.90
90% 0.84 0.89 0.90
50% 0.75 0.90 0.84
Table 1: The calculated values of λ\lambda for the percentiles and time intervals indicated using economic data from Refs. [40] and [41].

We conclude from the growth and wealth distribution data [40, 41] that the distribution of economic growth assumed in the GED model is a reasonable zeroth order approximation, particularly for the upper half of the wealth ranks of the United States over the past 30 years. Of course, there is much that it is not included in the model, such as the effects of wars, famines, storms, and recessions, which are not obtainable from the simple GED model. However, it appears from the data that the model is a reasonable approximation over time scales of the order of decades and yields insights into the importance of how economic growth is distributed.

References

  • [1] A. Chakraborti, I. M. Toke, M. Patriarca, and F. Abergel, “Econophysics review: II. Agent-based models,” Quant. Finance 11, 1013 (2011).
  • [2] J. F. Kennedy, “Remarks in Heber Springs Arkansas at the dedication of the Greers Ferry Dam,” The American Presidency Project, 3 October 1963, <https://www.jfklibrary.org/asset-viewer/archives/JFKPOF/047/JFKPOF-047-015>.
  • [3] E. Gudreis, “Unequal America,” Harvard Magazine, July–August (2008).
  • [4] O. Peters, “Optimal leverage from non-ergodicity,” Quant. Finance 11, 1593 (2011).
  • [5] J. Angle, “The surplus theory of social stratification and the size distribution of personal wealth,” Social Forces 65, 293 (1986).
  • [6] J. Angle, “Deriving the size distribution of personal wealth from the rich get richer, the poor get poorer,” J. Math. Sociology 18, 27 (1993).
  • [7] A. Chakraborti and B. K. Chakrabarti, “Statistical mechanics of money: how saving propensity affects its distribution,” Eur. Phys. J. B 17, 167 (2000).
  • [8] A. Drägulescu and V. M. Yakovenko, “Statistical mechanics of money,” Eur. Phys. J. B 17, 723 (2000).
  • [9] C. F. Moukarzel, S. Goncalves, J. R. Iglesias, M. Rodriguez-Achach, and R. Huerta-Quintanilla, “Wealth condensation in a multiplicative random asset exchange model,” Eur. Phys. J.-Spec. Top. 143, 75 (2007).
  • [10] V. M. Yakovenko and J. B. Rosser Jr., “Colloquium: Statistical mechanics of money, wealth, and income,” Rev. Mod. Phys. 81, 1703 (2009).
  • [11] M. Patriarca and A. Chakraborti, “Kinetic exchange models: From molecular physics to social science,” Am. J. Phys. 81, 618 (2013).
  • [12] J.-P. Bouchaud and M. Mézard, “Wealth condensation in a simple model of economy,” Physica A 282, 536 (2000).
  • [13] Z. Burda, D. Johnston, J. Jurkiewicz, M. Kamiński, M. A. Nowak, G. Papp, and I. Zahed, “Wealth condensation in Pareto macroeconomies,” Phys. Rev. E 65, 026102 (2002).
  • [14] B. M. Boghosian, “Kinetics of wealth and the Pareto law,” Phys. Rev. E 89, 042804 (2014).
  • [15] B. M. Boghosian, “Fokker-Planck description of wealth dynamics and the origin of Pareto’s Law,” Int. J. Mod. Phys. C 25, 1441008 (2014).
  • [16] B. M. Boghosian, M. Johnson, and J. A. Marcq, “An H theorem for Boltzmann’s equation for the yard-sale model of asset exchange,” J. Stat. Phys. 161, 1339 (2015).
  • [17] C. Chorro, “A simple probabilistic approach of the Yard-Sale model,” Statistics and Probability Letters 112, 35 (2016).
  • [18] B. M. Boghosian, A. Devitt-Lee, M. Johnson, J. Li, J. A. Marcq, and H. Wang, “Oligarchy as a phase transition: The effect of wealth-attained advantage in a Fokker-Planck description of asset exchange,” Physica A 476, 15 (2017). “Wealth-attained advantage” is implemented by biasing the coin flip for the exchange of wealth in favor of the wealthier agent. In contrast, we implement a form of wealth-attained advantage by an uneven redistribution of wealth due to growth.
  • [19] A Devitt-Lee, H. Wang, J. Ii, and B. Boghosian, “A nonstandard description of wealth concentration in large-scale economies,” Siam J. Appl. Math 78, 996 (2018).
  • [20] J. Li, B. M. Boghosian, and C. Li, “The affine wealth model: An agent-based model of asset exchange that allows for negative-wealth agents and its empirical validation,” Physica A 516, 423 (2019).
  • [21] B. M. Boghosian, “The inescapable casino,” Sci. Amer. 321 (11), 72 (2019).
  • [22] P. L. Krapivsky and S. Redner, “Wealth distributions in asset exchange models,” Science and Culture 76, 424 (2010).
  • [23] S. Ispolatov, P. L. Krapivsky, and S. Redner, “Wealth distributions in asset exchange models,” J. Eur. Phys. B 2, 267 (1998).
  • [24] A. Chakraborti, “Distributions of money in model markets of economy,” Int J. Mod. Phys. C 13, 1315 (2002).
  • [25] B. Hayes, “Follow the money,” Am. Sci. 90, 400 (2002).
  • [26] F. Slanina, “Inelastically scattering particles and wealth distribution in an open economy,” Phys. Rev. E 69, 046102 (2004). This paper was one of the first to consider increasing wealth, although the exchange mechanism was additive rather than multiplicative.
  • [27] The Yard-Sale model and its generalizations incorporate the exchange of wealth not income. The distinction between wealth and income is important in more detailed models of the economy. The distribution of income and wealth show similar trends, but the distribution of wealth is more unequal than the distribution of income. See, for example, M. Cragg and R. Ghayad, “Growing apart: The evolution of income vs. wealth inequality,” The Economists’ Voice 12(1), 1–12 (2005), A. Drägulescu, V. M. Yakovenko, “Exponential and power-law probability distributions of wealth and income in the United Kingdom and the United States,” Physica A 299, 213 (2001), and Y. Berman, E. Ben-Jacob, and Y. Shapira, “The dynamics of wealth inequality and the effect of income distribution,” PLoS ONE 11(4), e0154196 (2016).
  • [28] W. Klein, N. Lubbers, K. K. L. Liu, and H. Gould, “Mean-field theory of an asset exchange model with economic growth and wealth distribution,” arXiv:2102.01274.
  • [29] D. Thirumalai and R. D. Mountain, “Activated dynamics, loss of ergodicity, and transport in supercooled liquids,” Phys. Rev. A 42, 4574 (1990) and Phys. Rev. E 47, 479 (1993).
  • [30] B.-H. F. Cardoso, S. Gonçalves, J. R. Iglesias, “Wealth distribution models with regulations: Dynamics and equilibria,” Physica A 551, 124201 (2020). The authors investigate a modification of the Yard-Sale model that favors the poorer agent (social protection). One consequence is that the agents in their modified Yard-Sale model have economic mobility under certain conditions. The authors also investigate the hysteresis of the wealth distribution due to the removal of social protection.
  • [31] J. L. Rodgers and W. A. Nicewander, “Thirteen ways to look at the correlation coefficient,” Amer. Stat. 42, 59 (1988).
  • [32] For a discussion of the Gini coefficient, see, for example, <https://en.wikipedia.org/wiki/Gini_coefficient>.
  • [33] W. Klein, H. Gould, N. Gulbahce, J. B. Rundle, and K. F. Tiampo, “The structure of fluctuations near mean-field critical points and spinodals and its implication for physical processes,” Phys. Rev. E 75, 031114 (2007).
  • [34] L. Colonna-Romano, H. Gould and W. Klein, “Anomalous mean-field behavior of the fully connected Ising model,” Phys. Rev. E 90, 042111 (2014).
  • [35] Kang K. L. Liu, J. B. Silva, W. Klein, and H. Gould, “Anomalous behavior of the specific heat in systems with long-range interactions,” manuscript in preparation.
  • [36] M. Kac, G. E. Uhlenbeck and P. Hemmer, “On the van der Waals theory of the vapor-liquid equilibrium. I. Discussion of a one-dimensional model,” J. Math. Phys. 4, 216 (1963).
  • [37] T. Khouw, W. Klein, and H. Gould, “Globalization and inequality in an agent-based asset exchange model,” manuscript in preparation.
  • [38] O. Peters and W. Klein, “Ergodicity breaking in geometric Brownian motion,” Phys. Rev. Lett. 110, 100603 (2013).
  • [39] P. C. Hohenberg and B. I. Halperin, “Theory of dynamic critical phenomena,” Rev. Mod. Phys. 49, 435 (1977).
  • [40] The U.S. real GDP growth rate adjusted for inflation, https://www.multpl.com/us-real-gdp-growth-rate>.
  • [41] Summary of wealth inequality in the U.S., <http://apps.urban.org/features/wealth-inequality-charts/>.