-
PDF
- Split View
-
Views
-
Cite
Cite
Yves Achdou, Jiequn Han, Jean-Michel Lasry, Pierre-Louis Lions, Benjamin Moll, Income and Wealth Distribution in Macroeconomics: A Continuous-Time Approach, The Review of Economic Studies, Volume 89, Issue 1, January 2022, Pages 45–86, https://doi.org/10.1093/restud/rdab002
- Share Icon Share
Abstract
We recast the Aiyagari–Bewley–Huggett model of income and wealth distribution in continuous time. This workhorse model—as well as heterogeneous agent models more generally—then boils down to a system of partial differential equations, a fact we take advantage of to make two types of contributions. First, a number of new theoretical results: (1) an analytic characterization of the consumption and saving behaviour of the poor, particularly their marginal propensities to consume; (2) a closed-form solution for the wealth distribution in a special case with two income types; (3) a proof that there is a unique stationary equilibrium if the intertemporal elasticity of substitution is weakly greater than one. Second, we develop a simple, efficient and portable algorithm for numerically solving for equilibria in a wide class of heterogeneous agent models, including—but not limited to—the Aiyagari–Bewley–Huggett model.
1. Introduction
One of the key developments in macroeconomics research over the last three decades has been the incorporation of explicit heterogeneity into models of the macroeconomy. Fuelled by the increasing availability of high-quality micro data, the advent of more powerful computing methods as well as rising inequality in many advanced economies, such heterogeneous agent models have proliferated and are now ubiquitous. This is a welcome development for a number of reasons. First, it opens up the door to bringing micro data to the table in order to empirically discipline macro theories. Second, macroeconomists often want to analyse the welfare implications of particular shocks or policies. This is impossible without asking “who gains and who loses?”, that is, distributional considerations often cannot be ignored. Third, models with heterogeneity often deliver strikingly different aggregate implications than do representative agent models, for example, with respect to monetary and fiscal policies.1
Despite the continuously increasing popularity of macroeconomic models with rich heterogeneity, the literature has suffered from a dearth of theoretical and analytical results. Little is known about the properties of consumption and saving behaviour in the presence of borrowing constraints, those of the resulting wealth distribution, and equilibrium uniqueness (or lack thereof). Instead, most studies rely on purely numerical analyses to characterize the implications of such theories. But even such computational approaches are often difficult and costly, particularly if the question at hand requires solving for the economy’s transition dynamics or if the model features non-differentiabilities or non-convexities.
In this article, we make some progress on these issues by recasting the standard incomplete market model of Aiyagari (1994), Bewley (1986), and Huggett (1993) in continuous time.2 Our main contributions are twofold. First, we prove a number of new theoretical results about this workhorse model.3 Second, we develop a simple, efficient, and portable algorithm for numerically solving both stationary equilibria and transition dynamics of a wide class of heterogeneous agent models, including—but not limited to—the Aiyagari–Bewley–Huggett model.
Both types of contributions make use of an important property: when recast in continuous time, heterogeneous agent models boil down to systems of two coupled partial differential equations. The first of these is a Hamilton–Jacobi–Bellman (HJB) equation for the optimal choices of a single atomistic individual who takes the evolution of the distribution and hence prices as given. And the second is a Kolmogorov Forward (KF) equation characterizing the evolution of the distribution, given optimal choices of individuals.4 More generally, our approach is to cast heterogeneous agent models in terms of the mathematical theory of “Mean Field Games” (MFG) initiated by Lasry and Lions (2007).5 The system of coupled HJB and KF equations is known as the “backward–forward MFG system.”
In the context of the Aiyagari–Bewley–Huggett model, the HJB equation characterizes individuals’ optimal consumption and saving behaviour given a stochastic process for income; and the KF equation characterizes the evolution of the joint distribution of income and wealth. The two equations are coupled because optimal consumption and saving depend on the interest rate which is determined in equilibrium and hence depends on the wealth distribution. We start with a particularly parsimonious case: a Huggett (1993) economy in which idiosyncratic income risk takes the form of exogenous endowment shocks that follow a two-state Poisson process and in which individuals save in unproductive bonds that are in fixed supply. Later in the article, we extend many of our results to more general stochastic processes and to an Aiyagari (1994) economy in which individuals save in productive capital.
We prove three new theoretical results about the Aiyagari–Bewley–Huggett model. First, we provide an analytic characterization of the consumption and saving behaviour of the poor. We show that, under natural assumptions, an individual’s saving policy function behaves like
In addition to these results, which are new also relative to the existing discrete-time literature, we extend some useful existing discrete-time results and concepts to continuous time. First, we adapt a number of results from Aiyagari (1994), e.g., that the wealth distribution has a finite upper bound and that a stationary equilibrium exists. Second, we characterize the saving behaviour of the wealthy and show that, with constant relative risk aversion (CRRA) utility, consumption and saving policy functions become linear for high wealth (Benhabib et al., 2015; Benhabib and Bisin, 2018). Third, we show how to define in continuous time marginal propensities to consume and save over discrete time intervals. This is not obvious and, at the same time, important for bringing the model to the data. Finally, a methodological contribution of our article is to show how to handle borrowing constraints in continuous time: conveniently, the borrowing constraint never binds in the interior of the state space and only shows up in a boundary condition. The consumption first-order condition always holds with equality, thereby sidestepping any complications due to “occasionally binding constraints.” Many of our proofs exploit this fact.
As already mentioned, our second main contribution is the development of a simple, efficient, and portable numerical algorithm for computing a wide class of heterogeneous agent models. The algorithm is based on a finite difference method and applies to the computation of both stationary and time-varying equilibria.9 We explain this algorithm in the context of the Aiyagari–Bewley–Huggett model. But the algorithm is, in fact, considerably more general and applies to any heterogeneous agent model with a continuum of atomistic agents (and without aggregate shocks). In Sections 6 and 7, we demonstrate the algorithm’s generality by applying it to other theories that feature non-convexities, a fat-tailed wealth distribution and multiple assets. Codes for these applications (and many more) are available at https://benjaminmoll.com/codes/ in Matlab as well as Python, Julia, and C++.
The first step of the algorithm is to solve the HJB equation for a given time path of prices. The second step is to solve the KF equation for the evolution of the joint distribution of income and wealth. Conveniently, after having solved the HJB equation, one obtains the time path of the distribution essentially “for free,” i.e., with very few lines of code. This is because the KF equation is the “transpose problem” of the HJB equation.10 The third step is to iterate and repeat the first two steps until an equilibrium time path of prices is found. For the first step, we make use of the theory of “viscosity solutions” to HJB equations (Crandall and Lions, 1983), and the corresponding theory for their numerical solution using finite difference methods (Barles and Souganidis, 1991). While much of our paper can be read without knowledge of viscosity solutions, we provide a brief introduction in Section 6.
Continuous time imparts a number of computational advantages relative to discrete time. As explained in more detail in Section 5.1, these relate to the handling of borrowing constraints, the numerical solution of first-order conditions and the fact that continuous-time problems with discretized state space are, by construction, very “sparse.” These computational advantages are reflected in the algorithm’s efficiency which we showcase in Section 5.6. At the same time, the algorithm is simple. Implementing it requires only some basic knowledge of matrix algebra and access to a software package that can solve sparse linear systems (e.g. Matlab). Finally, the algorithm is portable. For example, it applies without change to problems that involve non-differentiabilities and non-convexities. These are difficult to handle with standard discrete-time methods.11 In contrast, viscosity solutions and finite difference methods are designed to handle non-differentiable and non-convex problems. To illustrate this, we use the same algorithm to compute equilibria of an economy in which the interplay of indivisible housing and mortgages with a down-payment constraint causes a non-convexity which can result in individual poverty traps and multiple stationary distributions, an idea going back to Galor and Zeira (1993) among others.
Besides hopefully being useful in their own right, our paper’s contributions are also the foundation for a number of generalizations that go beyond the setup that we consider in the present paper (as well as the extensions in Section 7). First, building on the finite difference method developed here, Ahn et al. (2017) and Fernández-Villaverde et al. (2019) develop computational methods for solving heterogeneous agent models with aggregate uncertainty in addition to idiosyncratic risk (as in Den Haan, 1997; Krusell and Smith, 1998). Second, Olivi (2018) leverages our continuous-time formulation to obtain powerful comparative statics and sufficient statistics in incomplete markets models—results we build on in our uniqueness proof. Third, Parra-Alvarez et al. (2017) discuss how to identify and estimate continuous-time Aiyagari-Bewley-Huggett models. Fourth, Nuño and Moll (2017) and Nuño and Thomas (2017) devise a method for computing social optima. Fifth, Shaker Akhtekhane (2017) and Moll (2018) extend our computational approach to economies with heterogeneous firms à la Hopenhayn (1992). Sixth, Ruttscheidt (2018), building on Ahn (2017), computes equilibria in economies with a large number of individual state variables (four or more) by marrying our finite difference method with a sparse grid approach (Bungartz and Griebel, 2004; Gerstner and Griebel, 2010). All of these generalizations build on the tools developed in the present article.
A large theoretical and quantitative literature studies environments in which heterogeneous households are subject to uninsurable idiosyncratic shocks. See Heathcote et al. (2009), Guvenen (2011), Quadrini and Ríos-Rull (2015), and Krueger et al. (2015) for recent surveys, and the textbook treatment in Ljungqvist and Sargent (2004). All of these are set in discrete time.
Much fewer papers have studied equilibrium models with heterogeneous households in continuous time.12 All of these papers make “just the right assumptions” about the environment being studied so that equilibria can be solved explicitly (or at least characterized tightly).13 In contrast, our aim is to develop tools for solving and analysing models that do not permit closed-form solutions. Our methods apply as long as the model under consideration can be boiled down to an HJB equation and a KF equation, a feature shared by a wide class of heterogeneous agent models. These two approaches are clearly complementary: on the one hand, having explicit solutions is often extremely valuable for gaining intuition; on the other hand, restricting attention to environments for which these can be found may represent a sort of “analytic straitjacket” for some applications and the availability of more general methods may prove useful in such contexts.
One other paper by Bayer et al. (2019) also studies a continuous-time version of the standard Aiyagari–Bewley–Huggett model. The main differences between their paper and ours are: (1) they analyse a partial equilibrium framework whereas we consider a general equilibrium framework and (2) we develop a numerical algorithm for solving both stationary and time-varying equilibria.14 Barczyk and Kredler (2014a,b, 2018) study quantitative continuous-time Aiyagari–Bewley–Huggett models with imperfectly altruistic overlapping generations. We instead focus on the simpler workhorse version with infinitely lived individuals, allowing us to prove a number of new theoretical results. Finally, Rocheteau et al. (2015) propose an elegant alternative general equilibrium model with incomplete markets in continuous time. Market incompleteness in their framework stems from lumpy consumption expenditure shocks rather than idiosyncratic income risk. As a result their model features only one individual state variable and many results can be derived in closed form. The tradeoff is that their theory is further from the standard Aiyagari–Bewley–Huggett model that forms the backbone of much of modern macroeconomics.15
Section 2 lays out our continuous-time version of the workhorse macroeconomic model of income and wealth distribution in the parsimonious form due to Huggett (1993) while Section 3 briefly sketches the general approach of casting heterogeneous agent models as Mean Field Games. Section 4 contains our new theoretical results. Section 5 describes our computational algorithm for both stationary and time-varying equilibria and discusses computational advantages relative to existing discrete-time methods. Section 6 applies the algorithm to a problem with a non-convexity and provides a brief introduction to viscosity solutions. Section 7 discusses a number of generalizations and extensions and Section 8 concludes.
2. The Workhorse Model of Income and Wealth Distribution in Macroeconomics
To explain the logic of our approach in the simplest possible fashion, we present it in a context that should be very familiar to many economists: a general equilibrium model with incomplete markets and uninsured idiosyncratic labour income risk as in Aiyagari (1994), Bewley (1986), and Huggett (1993). We first do this in the context of an economy in which individuals save in unproductive bonds that are in fixed supply as in Huggett (1993). We later consider different ways of closing the model.
2.1. Setup
Individuals maximize (1) subject to (2), (3) and the process for
2.2. Stationary equilibrium
The domain of the differential equations (7) and (8) is
To see why this is the appropriate boundary condition, note that the first-order condition
The two ordinary differential equations (7) and (8) together with (9), (10), and the equilibrium relationship (11) fully characterize the stationary equilibrium of our economy. In the Mean Field Games (MFG) literature in mathematics this system of coupled HJB and KF equations is called a “backward–forward MFG system,” here in its stationary form.
2.3. Transition dynamics
The two partial differential equations (12) and (13) together with (14), the equilibrium relationship (4) and the boundary conditions (15) to (17) fully characterize the evolution of our economy. This is the time-dependent version of a “backward–forward MFG system.” It has two properties that are worth emphasizing. First, the two equations (12) and (13) are coupled: on one hand, an individual’s consumption-saving decision depends on the evolution of the interest rate which is in turn determined by the evolution of the distribution; on the other hand, the evolution of the distribution depends on individuals’ saving decisions. Second, the two equations run in opposite directions in time: the Kolmogorov Forward equation (13) runs forward (as indicated by its name) and looks backwards—it answers the question “given the wealth distribution today, savings decisions and the random evolution of income, what is the wealth distribution tomorrow?” In contrast, the HJB equation (12) runs backwards and looks forward—it answers the question “given an individual’s valuation of income and wealth tomorrow, how much will she save today and what is the corresponding value function today?”
3. General Heterogeneous Agent Models as MEAN FIELD GAMES
That a heterogeneous agent model boils down to a system of coupled HJB and KF equations is not special to the Aiyagari–Bewley–Huggett model. We here briefly sketch the general approach of casting such models in terms of the mathematical theory of Mean Field Games (MFGs) and provide some references. We also comment on an issue concerning HJB and KF equations, namely what the correct notion of a solution to these partial differential equations (PDEs)is.
3.1. General backward–forward MFG systems
Any heterogeneous agent model with a continuum of atomistic agents (and without aggregate shocks) can be written as a “backward–forward MFG system” of coupled HJB and KF equations. The system from the Aiyagari–Bewley–Huggett model generalizes in two obvious ways. First, the notion of an agent is abstract and also covers firms. Second, general heterogeneous agent models may feature
The mathematical theory of MFGs was initiated by Lasry and Lions (2007). Cardaliaguet (2013) and Ryzhik (2018) provide excellent and relatively accessible accounts of its current state. A natural question is whether this literature contains any “off-the-shelf” results on backward–forward MFG systems that apply to the economic models we want to study, say with regard to existence and uniqueness of solutions. As we explain in Supplementary Appendix C.5, the answer is “no” unfortunately: several of the typical features of heterogeneous agent models in economics mean that they are not special cases of the MFGs treated in mathematics.
3.2. Classical versus weak solutions of HJB and KF equations
A classical solution to a PDE or ordinary differential equation (ODE) is a solution that is differentiable as many times as needed to satisfy the corresponding equation. For example, classical solutions to the first-order HJB and KF equations (12) and (13) would need to be once differentiable. Similarly, classical solutions to second-order equations that arise for example if
Most of our article employs classical methods and the preceding paragraph equips the reader sufficiently well for those parts that do not. An exception is the model with a non-convexity due to indivisible housing in Section 6. We there discuss in more detail the usefulness of viscosity solutions.
4. Theoretical Results for the Aiyagari–Bewley–Huggett Model: Consumption, Saving, and Inequality
This Section presents theoretical results about our continuous-time version of the Aiyagari–Bewley–Huggett model, including the three new results emphasized in the introduction. Sections 4.1 to 4.5 analyse the HJB and KF equations (7) and (8) in partial equilibrium, i.e., taking as given a fixed interest rate
4.1. An Euler equation
Our first few theoretical results concern the consumption and saving behaviour of individuals. Our characterization of individual behaviour uses the following Lemma.
Differentiate the HJB equation (7) with respect to
4.2. Consumption and saving behaviour of the poor
Our first main result is obtained by analysing the Euler equation (18) close to the borrowing constraint. The interesting case is when the behaviour at the constraint differs qualitatively from that of rich individuals. Whether this is the case depends crucially on two factors: the tightness of the borrowing constraint
The next proposition shows that the borrowing constraint “matters” if this assumption holds. Standard utility functions satisfy
In what follows as well as elsewhere in the article, we use the following asymptotic notation: for any two functions
Let
- $$s_1(\underline{a})=0$$but$$s_1(a)<0$$all$$a>\underline{a}$$. That is, only individuals exactly at$$\underline{a}$$are constrained, whereas those with wealth$$a>\underline{a}$$are unconstrained and decumulate assets.
- as$$a\rightarrow \underline{a}$$, the saving and consumption policy functions of the low-income type and the corresponding instantaneous marginal propensity to consume satisfy(19)$$$$ \begin{align}\label{eq:approx} s_1(a) &\sim - \sqrt{2\nu_1} \sqrt{a - \underline{a}},\\ \notag c_1(a) &\sim y_1 + r a + \sqrt{2\nu_1}\sqrt{a - \underline{a}},\\ \end{align} $$$$(20)$$$$ \begin{align} \label{eq:MPC_inst} c_1'(a) &\sim r + \sqrt{\frac{\nu_1}{2(a - \underline{a})}},\\ \end{align} $$$$where(21)$$$$ \begin{align} \label{eq:nu} \begin{split} \nu_1 :&= \frac{(\rho - r)u'(\underline{c}_{1}) + \lambda_1 (u'(\underline{c}_{1}) - u'(\underline{c}_{2}))}{-u''(\underline{c}_{1})}\\ &\approx (\rho-r)\text{IES}(\underline{c}_1)\underline{c}_1 + \lambda_1(\underline{c}_2 - \underline{c}_1), \end{split} \end{align} $$$$$$\underline{c}_j = c_j(\underline{a}),j=1,2$$and$$\textrm{IES}(c):=-u'(c)/(u''(c)c)$$.25 The derivatives of$$c_1$$and$$s_1$$are unbounded at the borrowing constraint,$$c_1'(a) \rightarrow \infty$$and$$s_1'(a) \rightarrow -\infty$$as$$a\rightarrow \underline{a}$$.
The proof of the proposition, like those of all others, is in the Supplementary Appendix. The proof of the first part follows straight from the state constraint boundary condition (10). The second part of the proof follows from characterizing the limiting behaviour of the squared saving policy function
The consumption and saving behaviour in the Proposition is illustrated in Figure 1.

Importantly, the derivatives of type
The result that the borrowing constraint is reached in finite time bears some similarity to optimal stopping time problems (see e.g. Stokey, 2009). Just like in stopping time problems, continuous time avoids a type of integer problem arising in discrete time: the borrowing constraint would be reached after a non-integer time period, but discrete time forces this to occur after an integer number of periods.26
Proposition 1 features an intuitive formula (21) for the speed at which individuals hit the borrowing constraint,
Intuition for Proposition 1 and Corollary 1: two useful special cases. To understand the intuition for the square root in Proposition 1, the saving behaviour in Corollary 1 and the role of Assumption 1, we now consider two special cases for which analytic solutions are available. Both abstract from income uncertainty which is inessential to this point.27

First special case in which borrowing constraint binds in finite time
To understand the square root in the saving and consumption policy functions in Proposition 1, consider an individual at
It is easy to show that

4.3. Consumption and saving behaviour of the wealthy
Proposition 1 characterizes consumption and saving behaviour close to the borrowing constraint. The following Proposition 2 characterizes this behaviour for large wealth levels. This will be useful below, when we characterize the upper tail of the wealth distribution.
Assume that
Then there exists
$$a_{\max}<\infty$$such that$$s_j(a)<0$$for all$$a > a_{\max},j=1,2$$, and$$s_2(a)\sim \zeta_2(a_{\max}-a)$$as$$a \rightarrow a_{\max}$$for some constant$$\zeta_2$$.- In the special case of CRRA utility (5) individual policy functions are asymptotically linear in$$a$$:(25)$$$$ \begin{equation} \label{eq:CRRAlim}s_j(a) \sim \frac{r- \rho}{\gamma} a, \quad c_j(a) \sim \frac{\rho - (1-\gamma) r}{\gamma} a \quad \mbox{as} \ a \rightarrow \infty \end{equation} $$$$
The first part of the Proposition is the analogue of Proposition 4 in Aiyagari (1993). The condition that
As
The asymptotic linearity of consumption and saving policy functions with CRRA utility has played a key role in the literature. For instance, Krusell and Smith (1998) argue that this linearity explains their finding that the business cycle properties of a baseline heterogeneous agent model are virtually indistinguishable from its representative agent counterpart. Future studies may want to gauge the robustness of this result to relaxing the CRRA assumption.
4.4. Marginal propensities to consume and save
We now characterize marginal propensities to consume and save, defined as the changes in consumption and saving in response to a windfall increase in available funds
Differentiating this expression and using the budget constraint, we get the following result.
We make three observations. First, in contrast to the instantaneous MPC
Summarizing, when people hit the borrowing constraint in finite time, MPCs depend on wealth and, in particular, are higher for poorer people.
When individuals experience new income draws within the time interval
This follows from a direct application of the Feynman–Kac formula.
Figure 4(a) plots the MPC computed in this way for the two income types and assuming that individuals have CRRA utility (5). For comparison, Figure 4(b) plots the “instantaneous MPC,” i.e. the slope of the consumption function.

As expected, the former is a smoother version of the latter and, in contrast to the latter, does not exceed
As an aside, in some applications a slightly altered version of the MPCs in Definition 1 is easier to map to the data. Empirical studies do not typically estimate MPCs out of an infinitesimal increase in resources. Instead, they estimate the increase in consumption in response to a discrete increase, say by
Using the analytic expression for

Figure 5(b) instead graphs the dependence of the low-income type’s MPC on the realization of the high income
4.5. The stationary wealth distribution
We now present the paper’s second main theoretical result: an analytic solution to the Kolmogorov Forward equation characterizing the stationary distribution with two income types (8) for given individual saving policy functions. This analytic solution yields a number of insights about properties of the stationary wealth distribution, particularly at the borrowing constraint and in the right tail.
- (Close to the borrowing constraint) The stationary distribution of low-income types has a Dirac point mass at the borrowing constraint$$\underline{a}$$, i.e., its CDF satisfies$$G_1(\underline{a})=m_1>0$$. The Dirac point mass$$m_1$$can be found from the constants of integration$$\kappa_1,\kappa_2$$and is expressed in terms$$\lambda_1,\lambda_2,s_1,s_2$$in Supplementary Appendix equation (A.71). The CDF further satisfies(34)$$$$ \begin{equation}\label{eq:CDF1} G_1(a) \sim m_1 \exp\left(\lambda_1 \sqrt{2(a-\underline{a})/\nu_1} \right) \quad \mbox{as} \ a \downarrow \underline{a} \end{equation} $$$$
The stationary distribution of high-income types does not have a Dirac point mass at
$$\underline{a}$$, i.e., its CDF satisfies$$G_2(\underline{a})= m_2 =0$$, and its density is in fact finite,$$g_2(\underline{a})<\infty$$. (In the right tail) The support of the stationary wealth distribution is bounded above at some
$$a_{\max}<\infty$$defined in Proposition 2. It does not have a Dirac point mass at$$a_{\max}$$.(Smoothness) In contrast to the analogous discrete-time economy, the density of wealth is continuous and differentiable for all
$$a>\underline{a}$$, i.e., everywhere except at the borrowing constraint.
Corollary 3 in the Supplementary Appendix lists some additional but less central properties of the wealth distribution that follow from (33).

Saving behaviour and stationary wealth distribution with
Part 2 of the Proposition states that the stationary wealth distribution in our economy is bounded above. Like discrete-time versions of Aiyagari–Bewley–Huggett economies with idiosyncratic labour income risk only, our model therefore has difficulties explaining the high observed wealth concentration in developed economies like the U.S. In particular, empirical wealth distributions seem to feature fat Pareto tails. Section 7 extends the model to feature such a fat-tailed stationary distribution by introducing a second, risky asset.
Part 3, which can also be seen in Figure 1, highlights an important difference between our continuous-time formulation and the traditional discrete-time one: except for the Dirac mass exactly at the borrowing constraint
4.6. Stationary equilibrium: existence and uniqueness
We construct stationary equilibria along the same lines as in Aiyagari (1994). That is, we fix an interest rate
Figure 7 illustrates the typical effect of an increase in

Effect of an increase in
An increase in

A stationary equilibrium is then an interest rate
If relative risk aversion
We next turn to our third main theoretical result: uniqueness of a stationary equilibrium.
Individual consumption
$$c_j(a;r)$$is strictly decreasing in$$r$$for all$$a > 0$$and$$j=1,2$$.Individual saving
$$s_j(a;r)$$is strictly increasing in$$r$$for all$$a > 0$$and$$j=1,2$$.An increase in the interest rate leads to a rightward shift in the stationary distribution in the sense of first-order stochastic dominance:
$$G(a;r) = G_1(a;r) + G_2(a;r)$$is decreasing in$$r$$for all$$a$$in its support.Aggregate saving
$$S(r)$$is strictly increasing and hence our continuous-time version of Huggett’s economy has at most one stationary equilibrium.
Part 2 uses the budget constraint
Part 3, first-order stochastic dominance, and Part 4, that aggregate saving is increasing in
5. Computation
We now describe our algorithm for numerically computing equilibria of continuous-time heterogeneous agent models. We use a finite difference (FD) method based on work by Achdou and Capuzzo-Dolcetta (2010) and Achdou (2013) which is simple, efficient and easily extended to other environments. We explain our method in the context of the baseline heterogeneous agent model of Section 2. But the algorithm is, in fact, considerably more general and applies to any heterogeneous agent model with a continuum of atomistic agents (and without aggregate shocks). It is particularly well-suited for computing transition dynamics and solving problems with non-convexities, a fact we illustrate in Section 7 by computing equilibria of such economies. Codes for these applications (and many more) are available from https://benjaminmoll.com/codes/ in Matlab as well as Python, Julia and C++.
5.1. Computational advantages relative to discrete time
Before explaining our algorithm, we provide a brief overview of some of its computational advantages relative to traditional discrete-time methods. We here list four computational advantages that we consider crucial and that contribute notably to the efficiency gains over traditional methods. The first of these advantages is special to the solution of problems with borrowing constraints. The second to fourth advantages concern the solution of heterogeneous agent models more broadly (e.g. models with heterogeneous firms).
Second and related, the first-order condition in (36) is “static” in the sense that it only involves contemporaneous variables. Given (a guess for) the value function
The third advantage of continuous time is a form of “sparsity.” To solve the HJB and KF equations (7) and (8), we discretize these so that their solution boils down to solving systems of linear equations. The resulting matrices are typically extremely sparse, namely “tridiagonal” or at least “block-tridiagonal.” This sparsity generates considerable efficiency gains because there are well-developed routines for solving sparse linear systems, either implemented as part of commercial software packages like Matlab or open-source libraries like SuiteSparse. The reason that tridiagonal matrices arise is that a discretized continuous-time process either stays at the current grid point, takes one step to the left or one step to the right. But, it never jumps.38
Fourth, in all heterogeneous agent models, there is a tight link between solving the HJB and KF equations. One can typically “kill two birds with one stone” in the sense that, having computed the solution to the HJB equation one gets the solution to the KF equation “for free”: the matrix in the discretized version of the latter is the transpose of the matrix in that of the former. The underlying mathematical reason is that the KF equation is the “transpose problem” of the HJB equation or, more precisely, that the differential operator in the KF equation is the adjoint of the operator in the HJB equation.39
5.2. Bird’s eye view of algorithm for stationary equilibria
Our aim is to calculate stationary equilibria—functions
Iterating on the equilibrium system. From a bird’s eye perspective our algorithm for solving the stationary equilibrium shares many similarities with algorithms typically used to solve discrete-time heterogeneous agent models. In the context of our Huggett economy, we use a bisection algorithm on the stationary interest rate. We begin an iteration with an initial guess
Given
$$r^\ell$$, solve the HJB equation (7) using a FD method and calculate$$s_{j}^{\ell}(a)$$.Given
$$s_{j}^{\ell}(a)$$, solve the KF equation (8) for$$g_{j}^{\ell}(a)$$using a FD method.Given
$$g_{j}^{\ell}(a)$$, compute the net supply of bonds$$S(r_\ell) = \int_{\underline{a}}^\infty a (g_{1}^{\ell}(a) + g_{2}^{\ell}(a))da$$and update the interest rate: if$$S(r^{\ell})>B$$, decrease it to$$r^{\ell+1}<r^{\ell}$$and vice versa.
When
The first equation is the discretized HJB equation (7), the second equation is the discretized KF equation (8) and the third equation is the discretized market clearing condition (11). The
5.3. Step 1: Solving the HJB equation
For Step 1, we solve the HJB equation (7) using a FD method. We now explain this approach. The Supplementary Appendix contains a more detailed explanation.
Theory for numerical solution of HJB equations (Barles–Souganidis). Before we explain our approach, we note that there is a well-developed theory concerning the numerical solution of HJB equations using FD schemes in the same way as there is a well-developed theory concerning the numerical solution of discrete-time Bellman equations. The key result is due to Barles and Souganidis (1991) who have proven that, under certain conditions, the solution to a FD scheme converges to the (unique viscosity) solution of the HJB equation. The interested reader should consult Barles and Souganidis’ original (and relatively accessible) paper or the introduction by Tourin (2013). In short, for their result to hold, the FD scheme needs to satisfy three conditions: (1) “monotonicity”, (2) “stability” and (3) “consistency.” Here, it suffices to note that (2) and (3) are typically easy to satisfy and, in practice, the main difficulty is to design a FD scheme that is “monotone.”
The upwind FD scheme for the HJB equation (43) can be conveniently written in matrix notation. Denoting by
Boundary conditions and handling the borrowing constraint. Besides guaranteeing that the Barles–Souganidis monotonicity condition holds, an upwind scheme like (43) has an additional advantage: the handling of boundary conditions. First, consider the upper end of the state space
5.4. Step 2: Solving the Kolmogorov forward equation
For Step 2, consider the stationary KF equation (8). We again discretize the equation using a FD scheme. In contrast to the HJB equation which is non-linear in
The deep underlying reason for this choice of discretization is that the KF equation actually is the “transpose” problem of the HJB equation. More precisely, the differential operator in the KF equation (8) is the adjoint of the operator in the HJB equation (7), the “infinitesimal generator.”41 Our transpose discretization of the KF equation (39) is not only well-founded mathematically; it is also extremely convenient: having solved the HJB equation, the solution of the Kolmogorov Forward equation is essentially “for free.”
The same numerical method—building the matrix
Does the presence of a Dirac point mass in the stationary wealth distribution
5.5. Computing transition dynamics
The algorithm to calculate time-varying equilibria—functions
Given
$$r^\ell(t)$$and (17), solve the HJB equation (12), marching backward in time. Calculate the saving policy function$$s_j^\ell(a,t)$$.Given
$$s_j^\ell(a,t)$$and (16), solve the KF equation, marching forward in time, for$$g_{j}^{\ell}(a,t)$$.Given
$$g_{j}^\ell(a,t)$$, compute the net supply of bonds$$S^\ell(t) = \int_{\underline{a}}^\infty a (g_{1}^{\ell}(a,t) + g_{1}^{\ell}(a,t))da$$and update the interest rate as$$r^{\ell+1}(t) = r^{\ell}(t) - \xi(t) \frac{d S^\ell(t)}{dt}$$where$$\xi(t)>0$$.42
When
An alternative to Step 3 in the algorithm above is to view Steps 1 and 2 as defining excess supply as a function of the entire time path
Our computational method for transition dynamics can, of course, also be used to compute (non-linear) impulse responses to unanticipated aggregate shocks (“MIT shocks”). It should be straightforward to use a linearized counterpart to compute linear impulse responses to small MIT shocks along the lines of Boppart et al. (2018) and Auclert et al. (2019) to obtain further speed gains.
5.6. Performance and comparison to a discrete-time method
Section 5.1 listed a number of computational advantages of continuous time relative to discrete time. We here substantiate these claims and compare the computational performance of our method to that of a state-of-the-art discrete-time method. We also explain how one may assess the accuracy of our solution method and discuss the relation to traditional discrete-time accuracy metrics like Euler equation errors.
A test problem. As a basis for these comparisons, we use a standard partial-equilibrium income fluctuation problem with a fixed interest rate. We focus on a partial-equilibrium problem because our strategy for iterating on equilibrium prices and computing transition dynamics is identical to that of standard discrete-time methods, i.e., any speed gain due to continuous time will necessarily occur in partial equilibrium. However, we now consider a version with a richer income process. We consider both a continuous-time and a discrete-time version and specify these to be as comparable as possible. For example, the discrete-time version features an AR(1) process for the logarithm of income and the continuous-time version features an Ornstein–Uhlenbeck process, a diffusion process that is the natural continuous-time analogue of an AR(1) process, both with comparable persistence and standard deviation and with a stationary mean of one. Other details about the specification and parameterization are in Supplementary Appendix F.1. We solve the continuous-time version using our FD method and the discrete-time version using the endogenous grid method (Carroll, 2006).
Assessing accuracy. When comparing different computational methods, we want to understand which method is faster while keeping the accuracy of the numerical solution constant. We therefore require a metric for assessing this accuracy. A challenge in this regard is that, standard discrete-time accuracy metrics such as Euler equations errors are not applicable in continuous time. To see this, consider the analogous discrete-time problem in Section 5.1 with Euler equation (37). As explained by Santos (2000), the rationale for examining the residuals in this Euler equation is that it is the first-order condition of the maximization problem in the Bellman equation. And by bounding the error in this first-order condition, one can bound the error in the policy function and more importantly the value function, i.e. the welfare loss from suboptimal behaviour due to numerical error. But for HJB equations like (41) and the associated finite-difference approximation (42), there is no error in the first-order condition (36) because it can be solved by hand. Instead, any error in the numerical solution of this PDE stems only from the finite-difference approximation of its derivatives. This is explained in more detail in Supplementary Appendix F.1 where we also briefly discuss other candidate accuracy metrics from the mathematics literature on HJB equations.
Given this, we use the following pragmatic approach. We first solve the two income fluctuation problems using an extremely fine grid with
Speed-accuracy tradeoff. Figure 9(a) plots speed-accuracy tradeoffs for the continuous- and discrete-time solution methods of our test problem, using the policy function error as our accuracy metric. Each blue circle corresponds to a continuous-time computation but with a different number of grid points ranging from a very coarse discretization with 25 grid points to an extremely fine one with 10,000 grid points (the “true” solution). Similarly, each red cross corresponds to a discrete-time computation. For each computation, the figure plots the time until the algorithm converged measured in seconds (vertical axis) against the policy function error (horizontal axis). Starting with a coarse grid in the lower right and increasing the number of grid points, the computations become slower but more accurate and we move toward the upper left. By varying the number of grid points, the blue line with circles therefore traces out a continuous-time speed-accuracy tradeoff. Similarly, the red line with crosses traces out the analogous discrete-time tradeoff.

Computational speed and accuracy: continuous versus discrete time
Notes: The figure reports speed and accuracy measures for the numerical solution of an income fluctuation problem in both continuous and discrete time. See Supplementary Appendix F.1 for a detailed description of the exercise. Each blue circle correspond to a continuous-time computation with a different number of grid points ranging from 25 to 10,000 grid points (the “true” solution). Similarly, each red cross corresponds to a discrete-time computation. (a) uses the policy function error as the accuracy metric and (b) uses the percentage error in aggregate consumption. The code is available at https://benjaminmoll.com/comparison/.
The key takeaway from the figure is that the continuous-time speed-accuracy tradeoff strictly dominates its discrete-time counterpart: for any given policy function error, the continuous-time method is always faster; conversely, for any given computational speed, the continuous-time method is always more accurate. Supplementary Appendix Figure A.1(a) reports the ratio of the computational times for different accuracy levels. It shows that the continuous-time method is at least twice as fast as the discrete-time method but can be more than 30 times faster if high accuracy (low error) is required.
Figure 9(b) repeats this exercise but instead using the percentage error in aggregate consumption as the accuracy metric. The computations are now somewhat more time-intensive because they require computing stationary distributions in addition to policy functions. The difference in computational performance is even more striking: for a given level of accuracy the continuous-time method is between 10 and 500 times faster—see Supplementary Appendix Figure A.1(a) which plots the relative speed.
General equilibrium and transition dynamics. As already noted, our strategy for iterating on equilibrium prices and computing transition dynamics is the same as in standard discrete time problems and we therefore do not conduct a comparison between the two. We nevertheless briefly comment on our method’s performance for computing these. All of Figures 1, 6, and 8 for the Huggett economy earlier in the paper were computed using a Matlab implementation of the algorithm in Section 5.2. Even though we work with a fine wealth grid with
5.7. Finite difference methods in economics and alternatives
Candler (1999) has previously used a FD method to solve HJB equations arising in economics and also discusses upwinding. Our numerical method adds to his in three dimensions. First and most obviously, we consider coupled HJB and KF equations rather than just the HJB equation in isolation: the system (7), (8), and (11) rather than just (7). Second, even when considered in isolation, our HJB equation differs from Candler’s because it features a borrowing constraint—a ubiquitous feature of heterogeneous agent models—and we design an upwind method that respects this constraint. Third, we show that our solution method has well-developed theoretical underpinnings by making the connection to the Barles and Souganidis (1991) theory.
Another method that is closely related to FD methods and has been previously used in economics is the Markov-chain approximation method (MCAM) of Kushner and Dupuis (2013). See, e.g., Golosov and Lucas (2007) and Barczyk and Kredler (2014a,b, 2018). One way of viewing our FD method is as a simple special case of MCAM. As we explained in Section 5.3, our FD method effectively approximates the law of motion for continuous state variables with a discrete-state Poisson process; that is, we use the FD method to build an approximating Markov chain. MCAM is a more general approach to building such approximating Markov chains, e.g., via trinomial trees.
Besides the FD and Markov-chain approximation (MCA) methods, there are many alternative methods for solving partial differential equations in general and HJB and KF equations in particular. Examples include finite-element, finite-volume, and semi-Lagrangean methods as well as approximation via orthogonal (e.g. Chebyshev) polynomials. In principle, these other methods can also be used to solve heterogeneous agent models of the type discussed here; in particular by following the same Steps 1 to 3 laid out in Section 5.2 but simply exchanging the solution method used within Steps 1 and 2. There is no sense in which our FD method dominates these other methods, some of which may even be more accurate for coarse discretizations. We nevertheless prefer the FD method for two reasons. First, it is transparent and easy to implement: in case the algorithm spits out junk, it is usually easy to track down the problem. Second, it delivers a useful symmetry for the HJB and KF equations: the transpose property discussed in Section 5.4 which is typically not shared by other methods. Finally, because the FD method is fast, choosing fine grids is usually sufficiently cheap and the potentially higher accuracy of alternative methods for coarse discretizations is therefore not relevant.
6. Non-convexities and the Power of Viscosity Solutions
Many important economic problems involve non-convexities. These are difficult to handle with standard discrete-time methods. In contrast, viscosity solutions and finite difference methods are designed to handle such problems. To illustrate this, we now present an economy in which the interplay of indivisible housing and mortgages with a down-payment constraint causes a non-convexity. We show that the same algorithm that we used in the standard Aiyagari–Bewley–Huggett model can be used without change. We also use this problem to provide a brief introduction to viscosity solutions and to comment on their usefulness.
6.1. Non-convexities: indivisible housing, mortgages, poverty traps
We here provide a parsimonious example of a theory that features a non-convexity: individuals can take out a mortgage to buy houses subject to a down-payment constraint and housing is indivisible. The purpose of this subsection is not to propose a quantitatively realistic model of housing; rather it is to showcase what kind of models can be solved with our computational algorithm.
Equivalently, the down-payment needs to be at least a fraction
The function

Model with indivisible housing: policy and value functions and multiple stationary distributions.
The vertical line in the two graphs is at
Figure 10(c) plots the resulting saving policy function. The black, dashed horizontal line is at zero, i.e., saving is positive above that line and negative below. Optimal saving has the typical feature of problems with non-convexities: for each income type, there is a threshold wealth level (the “Skiba point”) below which individuals decumulate assets and above which they accumulate assets. In Figure 10(c), these points are where the saving policy functions intersect zero while sloping upward. As usual, each “Skiba point” is strictly below the point of the non-convexity
Since the theory features classic poverty trap dynamics, there can be multiple stationary wealth distributions. Figures 10(e) and (f) confirm this possibility: they plot two possible stationary wealth distributions. In fact, there is a continuum of stationary wealth distributions. In results not shown here due to space constraints, we have also computed the model’s transition dynamics. Not surprisingly given the discussion thus far, the economy features history dependence in the sense that initial conditions determine where the economy ends up in the long run. As already noted, the point of this subsection is not to argue quantitatively that the presence of indivisible housing and down-payment constraints creates history dependence. Rather it is to showcase the possibilities of our computational algorithm.
6.2. Viscosity Solutions
Because the value function features kinks the HJB equation does not have a classical solution (Section 3.2). Instead, the solution is a particular type of weak solution: a viscosity solution (Crandall and Lions, 1983). Supplementary Appendix D defines this solution concept and provides an overview of the corresponding theory. We here provide a brief summary.
As discussed in Section 5.1 a computational advantage of continuous time is that first-order conditions are static and can often be solved by hand. This is still true in the presence of non-convexities. For instance, the first-order condition in (48) is simply
Finally, the Barles–Souganidis theory still applies in the presence of kinks and therefore our computational algorithm can be applied without change.
7. Generalizations and Extensions
We here outline a number of generalizations and extensions of the baseline Huggett model in Sections 2 to 5. Details are in Supplementary Appendix G. The purpose of this brief section is to showcase the generality of our methods and in particular the portability of our computational algorithm.
More general income processes. Our baseline model assumed that income
Aiyagari model. Supplementary Appendix G.2 extends our results to the case where individuals save in productive capital and income takes the form of labour income as in Aiyagari (1994).
Soft borrowing constraints. Supplementary Appendix G.3 considers a wedge between borrowing and saving rates and characterize the implication for saving behaviour and the wealth distribution. This form of “soft constraint” can explain the empirical observation that wealth distributions typically have a spike at zero net worth and mass both to the left and the right of zero.
Fat tails. As shown in Proposition 3, the stationary wealth distribution in the Huggett economy with a bounded income process is bounded above. More generally, any income process with a thin-tailed stationary distribution generates a thin-tailed wealth distribution. This property of the model is problematic vis-à-vis empirical wealth distributions which typically feature fat upper tails. In Supplementary Appendix G.4, we extend the Huggett model of Section 2 to feature a fat-tailed stationary wealth distribution by introducing idiosyncratic investment risk (Nardi and Fella, 2017; Benhabib and Bisin, 2018).
Multiple assets with adjustment costs. The model in Sections 6 featured two assets: bonds and housing. But portfolio adjustment between the two assets was costless so that they collapsed into one state variable, net worth. Such costless portfolio adjustment is often a bad assumption, in particular when modeling illiquid assets such as housing. On our computational website, we therefore extend our algorithm to multiple assets with kinked (but convex) adjustment costs. Kaplan et al. (2018) argue that such a two-asset structure is important for understanding the monetary transmission mechanism.
Stopping time problems. Problems with multiple assets may also feature non-convex adjustment costs like fixed costs. Individuals then solve stopping time problems (Stokey, 2009). Instead of solving a standard HJB equation, the value function solves a “HJB Variational Inequality” (HJBVI, Øksendal, 1995; Tourin, 2013).45 On our computational website, we also generalize our algorithm to such stopping time problems.46 McKay and Wieland (2019), Guerrieri et al. (2020), and Laibson et al. (2020) use this algorithm to study durables and housing investment as well as mortgage refinancing decisions. The method also promises to be useful in other applications, e.g. problems involving default by individuals (see e.g. Livshits et al., 2007) or by sovereign states (see e.g. Aguiar et al., 2013; Bornstein, 2020).
8. Conclusion
This article makes two types of contributions. First, we prove a number of new theoretical results about the Aiyagari–Bewley–Huggett model, the workhorse theory of income and wealth distribution in macroeconomics: (1) an analytic characterization of the consumption and saving behaviour of the poor, particularly their marginal propensities to consume; (2) a closed-form solution for the wealth distribution in a special case with two income types; (3) a proof that there is a unique stationary equilibrium if the intertemporal elasticity of substitution is weakly greater than one. Second, we develop a simple, efficient and portable algorithm for numerically solving both stationary equilibria and transition dynamics of a wide class of heterogeneous agent models, including—but not limited to—this model. Both types of contributions were made possible by recasting the Aiyagari–Bewley–Huggett model in continuous time, thereby transforming it into a system of partial differential equations.
It is our hope that the methods developed in this article, particularly the numerical algorithm, will also prove useful in other applications. One potential application is to spatial theories of trade and development as in Rossi-Hansberg (2005) and Allen and Arkolakis (2014). These theories typically feature a continuum of producers and households distributed over a continuum of locations. In dynamic versions, space would simply be an additional variable in the HJB and KF equations. A challenge would be how to solve for equilibrium prices which are typically functions of space rather than a small number of (potentially time-varying) scalars. Related, a second avenue for future research is to explore richer interactions between individuals. In the class of theories, we have considered here, individuals interact only through prices. But for many questions of interest, richer interactions may be important: for instance there may be more “local” interactions in the form of knowledge spillovers or diffusion (see e.g. Perla and Tonetti, 2014; Lucas and Moll, 2014; Burger et al., 2016; Benhabib et al., 2017; Papanicolaou et al., 2020). In principle, the apparatus put forward in this article—the backward–forward MFG system of coupled HJB and KF equations—is general enough to encompass such richer models.
The editor in charge of this paper was Christian Hellwig.
Acknowledgments
This version supersedes an earlier version of the paper entitled “Heterogeneous Agent Models in Continuous Time.” It is supplemented by two online appendices https://benjaminmoll.com/HACT_appendix/ and https://benjaminmoll.com/HACT_Numerical_Appendix/ as well as a website with codes https://benjaminmoll.com/codes/. We are grateful to Fernando Alvarez, Adrien Auclert, Dave Backus, Daniel Barczyk, Roland Bénabou, Jess Benhabib, Jocelyn Boussard, Paco Buera, Lorenzo Caliendo, Dan Cao, Gabe Chodorow-Reich, Wouter Den Haan, Xavier Gabaix, Fatih Guvenen, Mark Huggett, Mariacristina De Nardi, Greg Kaplan, Tatiana Kirsanova, Nobu Kiyotaki, Matthias Kredler, Ellen McGrattan, Giuseppe Moscarini, Galo Nuño, Ezra Oberfield, Alan Olivi, Jesse Perla, Matt Rognlie, Tony Smith, Ivan Werning, Wei Xiong, Stan Zin, and seminar participants at various institutions for useful comments. We also thank Déborah Sanchez for stimulating discussions in early stages of this project and SeHyoun Ahn, Riccardo Cioffi, Adrien Couturier, Xiaochen Feng, Soroush Sabet, Rui Sousa, and Max Vogler for outstanding research assistance.
Supplementary Data
Supplementary data are available at Review of Economic Studies online. And the replication packages are available at https://doi.org/10.5281/zenodo.4357052.
Data Availability Statement
The data underlying this article are available in Zenodo, at https://doi.org/10.5281/zenodo.4357052. A dataset used for the appendix was also derived from sources in the public domain: 2007 Survey of Consumer Finances, accessible at https://www.federalreserve.gov/econres/scf_2007.htm.
Footnotes
Deaton (2016) succinctly summarizes the second and third reasons: “Aggregation needs to be seen, not as a nuisance, but as a hallmark of seriousness [...] While we often must focus on aggregates for macroeconomic policy, it is impossible to think coherently about national well-being while ignoring inequality and poverty, neither of which is visible in aggregate data. Indeed, and except in exceptional cases, macroeconomic aggregates themselves depend on distribution.”
Another important early reference is Imrohoroğlu (1989) who studies a model with both idiosyncratic and aggregate risk. The presence of a storage technology means that the model is set in partial equilibrium (the interest rate is exogenous and equals zero). Because of this difference, we refer to the Aiyagari–Bewley–Huggett model throughout the paper, but without wanting to diminish Imrohoroğlu’s important contribution.
Of course and as is well-known, the unadorned Aiyagari–Bewley–Huggett model is not sufficiently rich to be an empirically realistic theory of income and wealth distribution. Understanding its theoretical properties is nevertheless important, simply because it forms the backbone of much of modern macroeconomics.
The “Kolmogorov Forward equation” is also often called “Fokker–Planck equation.” Because the term “Kolmogorov Forward equation” seems to be somewhat more widely used in economics, we will use this convention throughout the article. But these are really two different names for the same equation.
The theory of “Mean Field Games” is a general and rigorous framework for the analysis of dynamic, stochastic games with a continuum of players. The name comes from an analogy to the continuum limit taken in “Mean Field theory” which approximates large systems of interacting particles by assuming that these interact only with the statistical mean of other particles. In general, MFGs can be written in terms of a so-called “Master equation” which reduces to the “backward–forward MFG system” in the case without aggregate uncertainty. For more on MFGs, see e.g. Guéant et al. (2011) and Cardaliaguet (2013).
The distribution of MPCs determines, for example, the efficacy of fiscal stimulus (e.g. Kaplan and Violante, 2014; Hagedorn et al., 2017), the transition mechanism of monetary policy (e.g. Auclert, 2019; Kaplan et al., 2018), the effect of a credit crunch or house price movements on consumer spending (e.g. Guerrieri and Lorenzoni, 2017; Berger et al., 2018), and the extent to which inequality affects aggregate demand (e.g. Auclert and Rognlie, 2018; Auclert and Rognlie, 2017).
The uniqueness result additionally assumes that individuals cannot borrow. A key step in our proof is an important result by Olivi (2018). Contemporaneous work by Light (2018) derives a uniqueness result under a similar condition on the IES in a more restrictive discrete-time setting: an Aiyagari economy with CRRA utility and Cobb–Douglas production (as well as no borrowing).
For the same reasons, one of the first results that every graduate student learns is that the neoclassical growth model—the representative-agent counterpart to the Aiyagari model—features a unique steady state.
Our numerical method is based on Achdou and Capuzzo-Dolcetta (2010) and Achdou (2013) but modified to handle the particular features of heterogeneous agent models, in particular borrowing constraints.
More precisely, the differential operator in the KF equation is the adjoint of the differential operator in the HJB equation. The adjoint of an operator is the infinite-dimensional analogue of a matrix transpose.
Because first-order conditions are no longer sufficient and standard envelope theorems do not apply.
See, for example, Jovanovic (1979), Moscarini (2005), Alvarez and Shimer (2011), Moll (2014), Stokey (2014), Vindigni et al. (2015), Jones and Kim (2018), Jones (2015), Toda and Walsh (2015), Benhabib et al. (2016), Cao and Luo (2017), and Kasa and Lei (2018). Miao (2005), Luttmer (2007, 2011, 2015, and Benhabib et al. (2017) analyse theories with heterogeneous producers. These papers, like ours, all study economies with a continuum of heterogeneous agents yielding a system of coupled HJB and KF equations. In contrast, other papers study environments with a finite number of heterogeneous agents (typically equal to two). For example, see Scheinkman and Weiss (1986) and applications of their framework by Conze et al. (1993) and Lippi et al. (2013).
Similarly, there are also several discrete-time approaches for retaining tractability in environments with heterogeneous households (e.g. Bénabou, 2002; Krebs, 2003; Heathcote et al., 2014).
Additionally, Bayer, Rendall, and Wälde assume a “natural borrowing constraint” implying that individuals never actually hit that constraint. Another difference is that they characterize individuals’ saving behaviour in terms of a differential equation for its consumption policy function whereas we work with the HJB equation. Lise (2013) is another paper studying a continuous-time partial-equilibrium setting.
Wang (2007) proposes an elegant continuous-time Aiyagari–Bewley–Huggett model that can be solved analytically but at the cost of making two non-standard assumptions on preferences: CARA utility and discount rates that are increasing in past consumption (in the absence of the second assumption, constant absolute risk aversion (CARA) utility implies exploding wealth inequality, and non-existence of a stationary distribution).
As discussed in detail in Aiyagari (1994), if the borrowing limit
Suppressing dependence on time
The system can be written more compactly as two non-linear partial differential equations in
This is in contrast to discrete-time formulations where there is a set
Note that this inequality has very little to do with the inequality in discrete-time first-order conditions due to occasionally binding borrowing constraints—see e.g. equation (37) later in the article. In fact, the two inequalities go in opposite directions. Even though both inequalities result from the presence of borrowing constraints, the logic behind them is completely different.
See Soner (1986a, b) and Capuzzo-Dolcetta and Lions (1990).
The standard notion of a measure-valued solution is only defined on the interior of the state space and therefore cannot be used to deal with a Dirac point mass at the boundary, a feature that arises in our application. We show in Supplementary Appendix E how to extend the standard notion to take into account this possibility.
This uses the extension of Ito’s formula to Poisson processes:
Assumption 1 is also what determines consumption and saving behaviour at the constraint for a wide class of less standard utility functions, say, with subsistence concerns. For example, in the Stone–Geary case
Type
A discrete-time analogue of Corollary 1 is derived in Huggett (1997) who proves that for a sufficiently long sequence of “bad” shocks, individuals hit the borrowing constraint (the argument is part of the proof of his Lemma 1). Unlike our formula, Huggett’s result does not include a characterization of the speed at which this happens, i.e., there is no analogue of our
We are indebted to Xavier Gabaix for suggesting the first special case. Also see Holm (2018) who characterizes consumption behaviour with deterministic income, a borrowing constraint and hyperbolic absolute risk aversion (HARA) utility.
In the case
The proof makes use of a simple homogeneity property: for all
Alternatively, we can take
Under Assumption 1, we expect
See for example Figure 1 in Imrohoroğlu (1989) and Figure 17.7.1 in Ljungqvist and Sargent (2004). To see why this must happen, consider a discrete-time Huggett economy with two income states. All individuals with wealth
Differentiating
Açıkgöz (2018) provides a numerical example of multiplicity in an Aiyagari model with an IES of
In this regard, it shares some similarities with the “endogenous grid method” of Carroll (2006). The difference is that in continuous time this also works with “exogenous grids.”
The first-order condition (36) also does not involve an expectation operator as in (37), i.e., a summation or costly numerical integral over future income states. Instead, the HJB equation (7) captures the stochastic evolution of income with an additive terms
Except of course if the process is a Poisson process, i.e., if jumps are “built in.” That being said, the sparsity property survives as long as there is at least one continuously moving state variable (like wealth), i.e. not all individual state variables follow discrete-state Poisson processes.
In principle, one can use an analogous approach in discrete time: form a Markov transition matrix over all states and use it to both iterate backward over value functions and forward over distributions. This method is less popular than it should be and researchers often solve for distributions by Monte-Carlo simulation.
In fact, in the special case without uncertainty analysed in the present section and under the assumption
The “infinitesimal generator” is the continuous-time analogue of a discrete-time transition matrix, and the adjoint of an operator is the infinite-dimensional analogue of a matrix transpose. In our context, the infinitesimal generator captures the evolution of the process in
Next, one can show that the operator in the KF equation (8) is the adjoint of this operator: denoting by
A good initial guess satisfies
In contrast, computing transitions for the Aiyagari model in Supplementary Appendix G.2, where prices are explicit functions of the aggregate capital stock, takes only 1 minute and 40 seconds. The Matlab code for the stationary equilibrium and transition dynamics of the Huggett model are available at https://benjaminmoll.com/huggett_equilibrium_iterate/ and http://benjaminmoll.com/huggett_transition/. The code for the Aiyagari model is at http://benjaminmoll.com/aiyagari_poisson_MITshock/.
Among the theoretical results, we extend Propositions 1, 2, 4, and 5. That is, all propositions from Section 4 except Propositions 3 (the analytic solution for the stationary distribution with two income types).
Economists typically tackle such problems by imposing a “smooth pasting” condition on the boundary between an inaction region and an adjustment region. While this approach is convenient in one dimension when this boundary is a simple threshold, it becomes impractical in multiple dimensions. The HJBVI approach instead avoids imposing a smooth pasting condition (which becomes a result rather than an imposed axiom) and multi-dimensional problems pose no conceptual problem over one-dimensional one.
See the applications labelled “Stopping Time Problem” at https://benjaminmoll.com/codes/.