Theoretical Foundations of
Buffer Stock Saving

April 24, 2023
 
Christopher D. Carroll1


_____________________________________________________________________________________

Abstract
This paper builds foundations for rigorous and intuitive understanding of ‘buffer stock’ saving models (Bewley (1977)-like models with a wealth target), pairing each theoretical result with quantitative illustrations. After describing conditions under which a consumption function exists, the paper articulates stricter ‘Growth Impatience’ conditions that guarantee alternative forms of stability — either at the population level, or for individual consumers. Together, the numerical tools and analytical results constitute a comprehensive toolkit for understanding buffer stock models.

            Keywords 

Precautionary saving, buffer stock saving, marginal propensity to consume, permanent income hypothesis, income fluctuation problem

            JEL codes 

D81, D91, E21

PIC

   REMARK:  https://econ-ark.org/materials/bufferstocktheory

Dashboard:  Click ‘Launch Dashboard’ Button

     html:  https://econ-ark.github.io/BufferStockTheory/

      PDF:  BufferStockTheory.pdf

   Slides:  BufferStockTheory-Slides.pdf

   GitHub:  https://github.com/econ-ark/BufferStockTheory

1Contact: ccarroll@jhu.edu, Department of Economics, 590 Wyman Hall, Johns Hopkins University, Baltimore, MD 21218, https://www.econ2.jhu.edu/people/ccarroll, and National Bureau of Economic Research.    

1 Introduction

For consumers whose income is affected by realistic transitory and permanent shocks (a la Friedman (1957) and Muth (1960)), only one further ingredient is required to construct a microeconomically testable model of optimal consumption: A description of preferences. Zeldes (1989) was the first to calibrate a quantitatively plausible example, spawning a literature showing that such models’ predictions can match household life cycle data reasonably well, whether or not explicit liquidity constraints are imposed.1

A connected literature, starting with Bewley (1977), has derived limiting properties of related infinite-horizon problems – but only in models more complex than the case with just Friedman-Muth shocks and preferences, because standard contraction mapping theorems (from Bellman (1957) through Stokey et al. (1989) and beyond) cannot be applied when utility and/or marginal utility are unbounded, and many proof methods rule out Friedman-Muth permanent shocks.2

This paper’s first contribution is to articulate conditions under which the infinite-horizon Friedman-Muth(-Zeldes) problem (with permanent shocks, and without complications like a consumption floor or liquidity constraints) defines a contraction mapping whose limit is useful (neither c = 0  nor c = ∞ everywhere) as the horizon recedes. A ‘Finite Value of Autarky Condition’ is mostly sufficient (the ‘Weak Return Impatience Condition’3 is unlikely to bind). Because the infinite-horizon solution is the limit of finite-horizon recursions, many results are also applicable to finite-horizon problems.

But the main theoretical contribution is to identify, for the infinite-horizon case, conditions under which ‘stable’ points exist (the ratio of wealth to permanent income will move toward a ‘target’; alternatively, there is a ‘balanced growth’ equilibrium) either for individual consumers or for the aggregate. The requirement for stability is always that the model’s parameters satisfy a ‘Growth Impatience Condition’ whose details depend on the quantity whose stability is of interest. A model with any such stable point(s) qualifies as a ‘buffer stock’ model. (Buffer stock models are neither a subset nor a superset of Bewley (1977) models – see below).4

Even without a formal proof of its existence, buffer stock saving has been intuitively understood to underlie key results in heterogeneous agent macroeconomics; for example, the logic of target saving is central to the claim by Krueger, Mitman, and Perri (2016) in the Handbook of Macroeconomics that such models explain why, during the Great Recession, middle-class consumers cut their spending more than the poor or the rich. Theory below provides the rigorous basis for this claim: Learning that the future has become more uncertain does not change the urgent imperatives of the poor (their high u′(c)  means they — optimally — have little room to maneuver). And, increased labor income uncertainty does not much change the behavior of the rich because it poses little risk to their consumption. Only people in the middle have both the motivation and the wiggle-room to respond to uncertainty by substantially reducing their spending. Analytical derivations required for the proofs also provide intuition for many other results familiar from the numerical literature.

The paper begins by describing sufficient conditions for the problem to define a sensible (nondegenerate) limiting consumption function (while explaining how the model relates to those previously considered). The conditions are interestingly parallel to those required for the liquidity constrained perfect foresight model; that parallel is explored and explained. The limiting properties of the consumption function as resources approach infinity, and as they approach their lower bound, are then used to prove the contraction mapping theorem.

The next theoretical contribution is to show that a model with an ‘artificial’ liquidity constraint (it prohibits borrowing by consumers who could certainly repay) is a limiting case of the unconstrained model. The analytical appeal of the unconstrained model is that it is both mathematically convenient (the consumption function is everywhere twice continuously differentiable), and arbitrarily close (cf. Section 2.11) to less tractable models. This congenial environment makes proofs easier (if we define a proposition as holding in the infinite-horizon limit when it holds as a finite horizon extends to infinity).

In proving the remaining theorems, the next section examines key properties of the model. First, as cash approaches infinity the expected growth rate of consumption and the marginal propensity to consume (MPC) converge to their values in the perfect foresight case. Second, as cash approaches zero the expected growth rate of consumption approaches infinity, and the MPC approaches a simple analytical limit. Next, the central theorems articulate conditions under which different measures of ‘growth impatience’ imply conclusions about points of stability (‘target’ or ‘balanced growth’ points).

The final section elaborates the conditions under which, even with a fixed aggregate interest rate that differs from the time preference rate, a small open economy populated by buffer stock consumers has a balanced growth equilibrium in which growth rates of consumption, income, and wealth match the exogenous growth rate of permanent income (equivalent, here, to productivity growth). In the terms of Schmitt-Grohé and Uribe (2003), buffer stock saving is an appealing method of ‘closing’ a small open economy, because it requires no ad-hoc assumptions. Not even liquidity constraints.5

2 The Problem

2.1 Setup

The infinite horizon solution is the (limiting) first-period solution to a sequence of finite-horizon problems as the horizon (the last period of life) becomes arbitrarily distant.

That is, for the value function, fixing a terminal date T  , we are interested in vT− n  in the sequence of value functions {v ,v    ,...,v    }
  T   T−1      T −n . We will say that the problem has a ‘nondegenerate’ infinite horizon solution if, corresponding to that v  , as n ↑ ∞ there is a limiting consumption function c(m, p) = limn ↑∞ cT− n(m, p)  which is neither c = 0  everywhere (for all m, p  ) nor c =  ∞ everywhere (a ‘sensible’ solution).

Concretely, a consumer born n  periods before date T  solves the problem6

pict

where the Constant Relative Risk Aversion (CRRA) utility function

pict

exhibits relative risk aversion ρ > 1  .7 The consumer’s initial condition is defined by market resources mt  and permanent noncapital income pt  , which both are positive,

{pt,mt } ∈ (0,∞ ),
(1)

and the consumer cannot die in debt,

cT ≤ mT  .
(2)

In the usual exposition, a dynamic budget constraint (DBC) combines several distinct events that jointly determine next period’s mt+1   (given this period’s choices); for the detailed analysis here, we disarticulate and describe every separate step:8 ,  9

   at = mt − ct
 kt+1 = at
 b   =  k   R
  t+1    t+1
 pt+1 = pt𝒢-Ψt+1
          ◟≡◝𝒢◜ ◞
             t+1
mt+1 =  bt+1 + pt+1ξξξt+1,

where a
 t  measures the consumer’s assets at the end of t  , which translate one-for-one into capital kt+1   at the beginning of the next period. Before the consumption choice, kt+1   is augmented by a fixed interest factor R = (1 + r)  to become bt+1   – the consumer’s financial (‘bank’) balances before next period’s consumption choice; mt+1   (‘market resources’) is the sum of financial wealth bt+1   and noncapital income pt+1ξξξt+1   (permanent noncapital income p
 t+1   multiplied by the transitory-income-shock factor ξξξt+1 described below). pt+1   is derived from pt  via application of a growth factor 𝒢 ,10 modified by a mean-one iid shock Ψt+1   , 𝔼t[Ψt+n ] = 1 ∀ n ≥ 1  satisfying Ψ ∈ [Ψ,-¯Ψ ]  for 0 < Ψ  ≤ 1 ≤  ¯Ψ <  ∞
    --- (and Ψ  = Ψ¯ = 1
---  is the degenerate case with no permanent shocks).

Following Zeldes (1989), in future periods t + n ∀ n ≥ 1  there is a small probability ℘  that income will be zero (a ‘zero-income event’),

       {
ξξξt+n =   0             with probability ℘ > 0
         𝜃𝜃𝜃t+n∕(1 − ℘)  with probability (1 − ℘)
(3)

where 𝜃𝜃𝜃t+n  is an iid mean-one random variable (𝔼t[𝜃𝜃𝜃t+n] = 1 ∀ n > 0  ) whose distribution satisfies 𝜃𝜃𝜃 ∈ [𝜃𝜃𝜃,¯𝜃𝜃𝜃]  where 0 < 𝜃𝜃𝜃 ≤  1 ≤ ¯𝜃𝜃𝜃 < ∞
    -- .11 Call the cumulative distribution functions ℱ Ψ  and ℱ𝜃𝜃𝜃  (where ℱξξξ  is derived trivially from (3) and ℱ 𝜃𝜃𝜃  ). For quick identification in tables and graphs, we will call this the ‘Friedman/Muth’ model because it is a specific implementation of Friedman (1957)’s ideas as interpreted by Muth (1960).

The model looks more special than it is. In particular, a positive probability of zero-income events may seem objectionable (despite empirical support).12 But a nonzero minimum value of ξξξ  (motivated, say, by the existence of unemployment insurance) could be handled by capitalizing the PDV of minimum income into current market assets,13 transforming that model back into this one. And no key results would change if the transitory shocks were persistent but mean-reverting (instead of IID). Also, the assumption of a positive point mass for the worst realization of the transitory shock is inessential, but simplifies the proofs and is a powerful aid to intuition.

2.2 Comparison to Existing Literature

This model differs from Bewley’s (1977) classic formulation in several ways. The CRRA utility function does not satisfy Bewley’s assumption that u (0 )  is well-defined, or that u′(0)  is well-defined and finite; indeed, neither the value function nor the marginal value function will be bounded. It differs from Schectman and Escudero (1977) in that they impose liquidity constraints and positive minimum income. It differs from both of these in that it permits permanent growth in income, and also permanent shocks to income, which a large empirical literature finds to be of dominant importance in microdata.14 It differs from Deaton (1991) because liquidity constraints are absent; there are separate transitory and permanent shocks (a la Muth (1960)); and the transitory shocks here can occasionally cause income to reach zero.

It differs from models found in Stokey et. al. (1989) because neither liquidity constraints nor bounds on utility or marginal utility are imposed.15 ,  16  Li and Stachurski (2014) show how to allow unbounded returns by using policy function iteration, but also impose constraints.

The paper with perhaps the most commonalities is Ma, Stachurski, and Toda (2020), henceforth MST, who establish existence and uniqueness of a solution to a general income fluctuation problem in a Markovian setting. The most important differences are that MST impose liquidity constraints and that expected marginal utility of income is finite (𝔼[u′(Y )] < ∞ )  . These assumptions are not consistent with the combination of CRRA utility and income dynamics used here, whose joint properties are key to the results.17

2.3 The Problem Can Be Normalized By Permanent Income

We establish a bit more notation by reviewing the familiar result that in such problems (CRRA utility, permanent shocks) the number of states can be reduced from two (m  and p  ) to one (m  = m ∕p )  . Value in the last period is u(m  )
    T  ; using (in the last line in (4)) the fact that for our CRRA utility function,           1−ρ
u(xy ) = x   u(y)  , and generically defining nonbold variables as the boldface counterpart normalized by pt  (as with m  = m  ∕p  ), consider the problem in the second-to-last period,

vT −1(mT − 1,pT −1) = mcaTx−1 u (pT −1cT− 1) + β 𝔼T −1[u (pTmT )]
                          {                                 }
                   = p1− ρ  max  u(cT− 1) + β 𝔼T −1[u(𝒢TmT )] .
                       T−1  cT−1
(4)

Now, in a one-time deviation from the notational convention established in the last sentence, define nonbold ‘normalized value’ not as vt∕pt  but as vt = vt∕p1t− ρ  , because this allows us to exploit features of the related problem,

vt(mt ) = max  u(ct) + β 𝔼t[𝒢1− ρvt+1 (mt+1 )]
         {c}Tt               t+1

      s.t.
    at = mt − ct

  kt+1 = at∕𝒢t+1
  bt+1 = kt+1R = (R∕ 𝒢t+1)at =  ℛt+1at
 m    =  b   + ξξξ   ,
   t+1    t+1    t+1
(5)

where ℛt+1 ≡  (R ∕𝒢t+1)  is a ‘permanent-income-growth-normalized’ return factor, and the reformulated problem’s first order condition is18

c−tρ = R β 𝔼t [𝒢 −ρc−ρ ].
              t+1 t+1
(6)

Since v  (m  ) = u(m  )
  T   T        T  , defining v    (m    )
 T −1   T−1  from (5), we obtain

pict

This logic induces to earlier periods; if we solve the normalized one-state-variable problem (5), we will have solutions to the original problem for any t < T  from:

pict

2.4 Definition of a Nondegenerate Solution

The problem has a nondegenerate solution if, as the horizon n  gets arbitrarily large, the solution in the first period of life cT− n(m)  gets arbitrarily close to a limiting c(m )  :

pict

that satisfies

0 < c(m) < ∞
(7)

for every 0 < m <  ∞ .

2.5 Perfect Foresight Benchmarks

The familiar analytical solution to the perfect foresight model, obtained by setting ℘ =  0  and 𝜃𝜃𝜃-= ¯𝜃𝜃𝜃 = Ψ--= Ψ¯ =  1  , allows us to define remaining notation and terminology.

2.5.1 The Finite Human Wealth Condition (FHWC)

The dynamic budget constraint, strictly positive marginal utility, and the can’t-die-in-debt condition (2) imply an exactly-holding intertemporal budget constraint (IBC):

             --bt-    ◜--◞h◟t--◝
PDVT  (c) = ◜m ◞◟− p◝ + PDVT  (p),
    t         t    t       t
(8)

where b  is beginning-of-period ‘market’ balances; with ℛ ≡  R∕𝒢 ‘human wealth’ is

ht = pt + ℛ −1pt + ℛ −2pt + ⋅⋅⋅ + ℛt−T pt
     (       −(T− t+1))
       1-−-ℛ---------
   =      1 − ℛ −1     pt.
     ◟-------◝◜-------◞
             ≡ht

For h ≡ limn→ ∞ hT −n  to be finite, we need the Finite Human Wealth Condition:

FHWC:    𝒢 ∕R <  1         .
         ◟◝◜◞
         ≡ℛ −1
(9)

where we call 𝒢 ∕R  the ‘Finite Human Wealth Factor’ (FHWF) because if this factor is less than one human wealth will be finite because (noncapital) income growth is smaller than the interest rate at which that income is being discounted.

2.5.2 The Return Impatience Condition (RIC)

Without constraints, the consumption Euler equation always holds; with   ′      −ρ
u (c) = c  ,

               1∕ρ
ct+1∕ct = (◟Rβ◝ )◜-◞
            ≡ÞÞÞ
(10)

where the archaic letter ‘thorn’ represents what we will call the ‘Absolute Patience Factor’ because, if the ‘absolute impatience condition’ holds,19

AIC:  ÞÞÞ <  1,
(11)

the consumer’s level of spending will be too large to sustain indefinitely. We call such a consumer ‘absolutely impatient.’

A ‘Return Patience Factor’ relates absolute patience to the return factor:

RPF:   ÞÞÞ  ≡ ÞÞÞ ∕R
        R
(12)

and since consumption is growing by ÞÞÞ  but discounted by R  :

pict

from which the IBC (8) implies

          ≡ κt
    ◜(------◞◟----)◝
ct =  --1-−-ÞÞÞR---  (bt + ht)
      1 − ÞÞÞTR−t+1
(13)

defining a normalized finite-horizon perfect foresight consumption function:

pict

where κ-
 t  is the marginal propensity to consume (MPC) — answering the question ‘if the consumer had an extra unit of resources, how much more spending would occur?’ (The overbar signifies that ¯c  will be an upper bound as we modify the problem to incorporate constraints and uncertainty; analogously, κ-  is the MPC’s lower bound).

The horizon-exponentiated term in the denominator of (13) is why, for κ-  to be strictly positive as n = T −  t  goes to infinity, we must impose the Return Impatience Condition:

      Þ
RIC:  ÞÞR < 1,
(14)

so that

            Þ
0 < κ-≡ 1 − ÞÞR =  lnim→∞ κT −n.
(15)

The RIC thus implies that the consumer cannot be so pathologically patient as to wish, in the limit as the horizon approaches infinity, to spend nothing today out of an increase in current wealth (the RIC rules out the degenerate limiting solution ¯c(m ) = 0  ). We call a consumer who satisfies the RIC ‘return impatient.’

2.5.3 Nondegenerate PF Unconstrained Solution Requires FHWC

Given that the RIC holds, and (as before) defining limiting objects by the absence of a time subscript, the limiting upper bound consumption function will be

¯c(m ) = (m + h − 1)κ,
(16)

and so in order to rule out the degenerate limiting solution ¯c(m) = ∞ we need h  to be finite; that is, we must impose the Finite Human Wealth Condition (FHWC), eq. (9).

2.5.4 The Growth Impatience Condition (GIC)

By analogy to the RPF, we define a ‘growth patience factor’ as

GPF:     ÞÞÞ𝒢 = ÞÞÞ ∕𝒢,
(17)

which defines a ‘growth impatience condition’:

GIC:    ÞÞÞ𝒢 < 1.
(18)

2.5.5 Perfect Foresight Finite Value of Autarky Condition (PFFVAC)

Under ‘autarky,’ capital markets do not exist; the consumer has no choice but to spend permanent noncapital income p  in every period. Because u(xy ) = x1 −ρu(y)  , the value the consumer would achieve is

vautarky=  u(p ) + βu(p 𝒢 ) + β2u (p 𝒢2) + ...
 t           t (      t           t)
                 1 − (β𝒢1 −ρ)T−t+1
       =  u(pt)  ----------1−-ρ---
                     1 − β𝒢

which (for 𝒢 > 0  ) asymptotes to a finite number as n =  T − t  approaches + ∞ if any of these equivalent conditions holds:

            ◜-≡◞ℶ◟-◝
PF -FVAC:    β𝒢1 −ρ <1
                −ρ
           βR 𝒢    <R ∕𝒢 ≡ ℛ
               Þ      1−1∕ρ
               ÞÞR  <ℛ      ,
(19)

where we call ℶ  20 the ‘Perfect Foresight Value Of Autarky Factor’ (PF-VAF), and the variants of (19) constitute alternative formulations of the Perfect Foresight Finite Value of Autarky Condition (‘PF-FVAC’); they guarantee that a perfect-foresight consumer who always spends all permanent income has ‘finite autarky value.’21

If the FHWC is satisfied, the PF-FVAC implies that the RIC is satisfied.22 Likewise, if the FHWC and the GIC are both satisfied, PF-FVAC follows:

 ÞÞÞ <  𝒢 < R
                  1− 1∕ρ
ÞÞÞR <  𝒢∕R <  (𝒢∕R )     < 1
(21)

(the last line holds because FHWC ⇒  0 ≤ (𝒢∕R ) < 1  and ρ >  1 ⇒ 0 < 1 − 1∕ ρ < 1  ).

The first panel of Table 4 summarizes: The PF-Unconstrained model has a nondegenerate limiting solution if we impose the RIC and FHWC (these conditions are necessary as well as sufficient). Together the PF-FVAC and the FHWC imply the RIC. If we impose the GIC and the FHWC, both the PF-FVAC and the RIC follow, so GIC+FHWC are also sufficient. But there are circumstances under which the RIC and FHWC can hold while the PF-FVAC fails (‘¬ PF-FVAC’). For example, if 𝒢 = 0  , the problem is a standard ‘cake-eating’ problem with a nondegenerate solution under the RIC (when the consumer has access to capital markets).

The easiest way to grasp the relations among these conditions is by studying Figure 1. Each node represents a quantity defined above. The arrow associated with each inequality imposes that condition. For example, one way we wrote the PF-FVAC in equation (19) is ÞÞÞ <  R1∕ρ𝒢1−1∕ρ  , so imposition of the PF-FVAC is captured by the diagonal arrow connecting ÞÞÞ  and R1∕ρ𝒢1−1∕ρ  . Traversing the boundary of the diagram clockwise starting at ÞÞÞ  involves imposing first the GIC then the FHWC, and the consequent arrival at the bottom right node tells us that these two conditions jointly imply the PF-FVAC. Reversal of a condition reverses the arrow’s direction; so, for example, the bottom-most arrow going to R1∕ρ𝒢1−1∕ρ  imposes ¬ FHWC; but we can cancel the cancellation and reverse the arrow. This would allow us to traverse the diagram clockwise from ÞÞÞ  through 𝒢 to  1∕ρ  1− 1∕ρ
R   𝒢  to R  , revealing that imposition of GIC and FHWC (and, redundantly, FHWC again) let us conclude that the RIC holds because the starting point is ÞÞÞ  and the endpoint is R  . (Consult Appendix I for an exposition of diagrams of this type, which are a simple application of Category Theory (Riehl (2017))).

PIC
Figure 1:Perfect Foresight Relation of GIC, FHWC, RIC, and PFFVAC

An arrowhead points to the larger of the two quantities being compared. For example, the diagonal arrow indicates that ÞÞÞ < R1∕ρ𝒢1−1∕ρ  , which is one way of writing the PF-FVAC, equation (19)

2.5.6 PF Constrained Solution Exists Under RIC or {¬ RIC, GIC}

We next sketch the perfect foresight constrained solution because it defines a benchmark (and limit) for the unconstrained problem with uncertainty (our ultimate interest).

If a liquidity constraint requiring b ≥ 0  is ever to be relevant, it must be relevant at the lowest possible level of market resources, mt = 1  , defined by the lower bound, bt = 0  (if it were relevant at any higher level of m  , it would certainly be relevant here, because u′ > 0  ). The constraint is ‘relevant’ if it prevents the choice that would otherwise be optimal; at mt =  1  it is relevant if the marginal utility from spending all of today’s resources ct = mt =  1  , exceeds the marginal utility from doing the same thing next period, ct+1 = 1  ; that is, if such choices would violate the Euler equation (6):

 − ρ       −ρ −ρ
1   >  Rβ𝒢   1  ,
(22)

which is just a restatement of the GIC.

We now examine implications of possible configurations of the conditions. (Tables 3 and 4 (and the table in appendix E) codify.)

¬ GIC and RIC. If the GIC fails but the RIC (14) holds, Appendix E shows that, for some 0 < m   < 1
      #  , an unconstrained consumer behaving according to the perfect foresight solution (16) would choose c < m  for all m >  m#   . In this case the solution to the constrained consumer’s problem is simple: For any m  ≥ m#   the constraint does not bind (and will never bind in the future); for such m  the constrained consumption function is identical to the unconstrained one. If the consumer were somehow23 to arrive at an m  < m#  < 1  the constraint would bind and the consumer would consume c = m  . Using `∙ for the version of a function ∙ in the presence of constraints (and recalling that ¯c(m )  is the unconstrained perfect foresight solution):

       {
`c(m) =   m      if m < m#
         ¯c(m )  if m ≥ m#.

GIC and RIC. When the RIC and GIC both hold, Appendix E shows that the limiting constrained consumption function is piecewise linear, with `c(m ) = m  up to a first ‘kink point’ at   0
m # >  1  , and with discrete declines in the MPC at a set of kink points    1    2
{m #, m #,...} . As m  ↑ ∞ the constrained consumption function `c(m )  becomes arbitrarily close to the unconstrained ¯c(m )  , and the marginal propensity to consume function `κκκ(m ) ≡ `c′(m )  limits to κ- .24 Similarly, the value function `v (m)  is nondegenerate and limits into the value function of the unconstrained consumer.

This logic holds even when the finite human wealth condition fails (¬ FHWC), because the constraint prevents the (limiting) consumer25 from borrowing against unbounded human wealth to finance unbounded current consumption. Under these circumstances, the consumer who starts with any b > 0
 t  will, over time, run those resources down so that after some finite number of periods τ  the consumer will reach bt+τ = 0  , and thereafter will set c = p  for eternity (which the PF-FVAC says yields finite value). Using the same steps as for equation (19), value of the interim program is also finite:

pict

So, even under ¬ FHWC, the limiting consumer’s value for any finite m  will be the sum of two finite numbers: One due to the unconstrained choice made over the finite horizon leading up to bt+τ = 0  , and one reflecting the value of consuming pt+τ  thereafter.

GIC and ¬ RIC. The most peculiar possibility occurs when the RIC fails. Under these circumstances the FHWC must also fail (Appendix E), and the constrained consumption function is nondegenerate. (See appendix Figure 8 for a numerical example). While limm ↑∞ `κκκ(m ) = 0  , nevertheless the limiting constrained consumption function `c(m )  is finite, strictly positive, and strictly increasing in m  . This result interestingly reconciles the conflicting intuitions from the unconstrained case, where ¬ RIC would suggest a degenerate limit of `c(m ) = 0  while ¬ FHWC would suggest a degenerate limit of `c(m ) = ∞ .

We now examine the case with uncertainty but without constraints, which turns out to be a close parallel to the model with constraints but without uncertainty.

2.6 Uncertainty-Modified Conditions

2.6.1 The Growth Impatience Condition (GIC-Mod))

When noncapital income uncertainty is introduced, the expectation of beginning-of-period bank balances bt+1   can be rewritten as:

pict

where Jensen’s inequality guarantees that the expectation of the inverse of the permanent shock is greater than one. Now define

          −1 − 1
Ψ-≡  (𝔼[Ψ   ])
(23)

which satisfies Ψ--< 1  (thanks again to Mr. Jensen), so it is convenient to define

𝒢-≡  𝒢Ψ--< 𝒢

because this allows us to write uncertainty-modified versions of earlier equations and conditions in a manner exactly parallel to those for the perfect foresight case; for example, we define a modified Growth Patience Factor:

GPF -Mod:   ÞÞÞ𝒢-= ÞÞÞ∕𝒢-=  𝔼[ÞÞÞ∕ (𝒢 Ψ )]
(24)

with a corresponding modified version of the Growth Impatience Condition:

GIC -Mod:   ÞÞÞ  < 1,
             𝒢-
(25)

that is stronger than the perfect foresight version (18) because 𝒢 < 𝒢
-- .

Table 1:Microeconomic Model Calibration
|--------------------------------------------------------------------------------|
|Calibrated-Parameters----------------|------------------------------------------|
|            Description              |Parameter   Value          Source         |
|-------------------------------------|------------------------------------------|
| Permanent  Income  Growth  Factor   |    𝒢        1.03   PSID:  Carroll (1992 )|
|          Interest Factor            |    R        1.04       Conventional      |
|                                     |                                          |
|      Time  Preference Factor        |    β        0.96       Conventional      |
|Coe fficient of Relative Risk Aversion |    ρ          2        Conventional      |
|                                     |                                          |
|    Probability of Zero Income       |    ℘        0.005   PSID:  Carroll (1992 )|
| Std Dev  of Log Permanent  Shock    |    σΨ        0.1    PSID:  Carroll (1992 )|
|                                     |                                          |
--Std--Dev-of-Log-Transitory-Shock---------σ𝜃--------0.1----PSID:--Carroll (1992-)-

2.6.2 The Finite Value of Autarky Condition (FVAC)

Analogously to (19), value for a consumer who spent exactly their permanent income every period would reflect the product of the expectation of the (independent) future shocks to permanent income:

pict

suggesting the definition of a utility-compensated equivalent of the permanent shock,

Ψ  = (𝔼 [Ψ1 −ρ])1∕(1−ρ),
---
(26)

which will satisfy Ψ-<  1
---  for ρ > 1  and nondegenerate Ψ  . Defining

𝒢-=  𝒢Ψ,-
(27)

v
  t  will be positive and finite as T  approaches ∞ if

            ≡ ℶ = VAF
              ◜-◞1◟− ◝ρ
FVAC:   0 <   β𝒢-    <  1
               0 < β <  𝒢ρ−1,
                        --
(28)

and (28) is the ‘finite value of autarky condition’ because it guarantees value is finite for a consumer who always consumes (now stochastic) permanent income (ℶ-
--  is the ‘Value of Autarky Factor’ (‘VAF’)).26 For nondegenerate Ψ  , this condition is stronger (harder to satisfy – requiring lower β  ) – than the perfect foresight version (19) because 𝒢-<  𝒢
-- .27

2.7 The Baseline Numerical Solution

Figure 2, familiar from the literature, depicts the successive consumption rules that apply in the last period of life (c  (m ))
  T  , the second-to-last period, and earlier periods under baseline parameter values listed in Table 2. (The 45 degree line is cT(m ) = m  because in the last period of life it is optimal to spend all remaining resources.)

In the figure, the consumption rules appear to converge to a nondegenerate c(m )  . Our next purpose is to show that this appearance is not deceptive.

Table 2:Model Characteristics Calculated from Parameters
|-----------------------------------------|----------------------------|-Approximate--|
|                                         |                            |              |
|                                         |                            |  Calculated  |
|              Description                |   Symbol  and  Formula     |    Value     |
|-----------------------------------------|----------------------------|--------------|
|      Finite Human  Wealth  Factor       | ℛ −1   ≡   𝒢∕R             |    0.990     |
|                                         |               1−ρ          |              |
|      PF  Value of Autarky  Factor       |  ℶ     ≡   β𝒢              |    0.932     |
|Growth  Compensated   Permanent   Shock  |  Ψ--   ≡   (𝔼[Ψ −1])−1     |    0.990     |
|                                         |                            |              |
|     Uncertainty -Adjusted Growth        |  𝒢-    ≡   𝒢Ψ--            |    1.020     |
| Utility Compensated   Permanent  Shock   |  Ψ--   ≡   (𝔼[Ψ1 −ρ])1∕(1− ρ) |    0.990     |
|                                         |  ---                       |              |
|      Utility Compensated   Growth        |  𝒢-    ≡   𝒢Ψ--            |    1.020     |
|        Absolute Patience Factor         | ÞÞÞ      ≡   (Rβ )1∕ρ         |    0.999     |
|                                         |                            |              |
|         Return Patience Factor          | ÞÞÞR     ≡   ÞÞÞ∕R             |    0.961     |
|                                         |                            |              |
|             Growth  Patience Factor     | ÞÞÞ 𝒢    ≡   ÞÞÞ∕ 𝒢            |    0.970     |
|    Modi fied Growth  Patience Factor     | ÞÞÞ 𝒢    ≡   ÞÞÞ∕ 𝒢-           |    0.980     |
|                                         |   --          1−ρ  1−ρ     |              |
|        Value of Autarky  Factor         |  ℶ-    ≡   β𝒢    Ψ--       |    0.941     |
|    Weak  Return  Impatience Factor      |℘1∕ρÞÞÞ   ≡   (℘βR )1∕ρ        |    0.071     |
---------------------------------------------------------------------------------------

PIC
Figure 2:Convergence of the Consumption Rules

2.8 Concave Consumption Function Characteristics

A precondition for the main theorem is that the maximization problem defines a sequence of continuously differentiable strictly increasing strictly concave28 functions {cT,cT− 1,...} . The straightforward but tedious proof is relegated to Appendix A. (Carroll and Kimball (1996) proved concavity but not continuous differentiability, which our theorem will use). For present purposes, the important point is that the income process induces what Aiyagari (1994) dubbed a ‘natural borrowing constraint’: ct(m ) < m  for all periods t < T  because a consumer who spent all available resources would arrive in period t + 1  with balances bt+1   of zero, and then might earn zero income over the remaining horizon, risking the possibility of a requirement to spend zero, yielding negative infinite utility. To avoid this calamity, the consumer never spends everything.  Zeldes (1989) seems to have been the first to argue, based on his numerical results, that the natural borrowing constraint was a quantitatively plausible alternative to ‘artificial’ or ‘ad hoc’ borrowing constraints.29

Strict concavity and continuous differentiability of the consumption function are key elements in many of the arguments below, but are not characteristics of models with ‘artificial’ borrowing constraints (though the analytical convenience of these features is considerable – see below).

2.9 Bounds for the Consumption Functions

The consumption functions depicted in Figure 2 appear to have limiting slopes as m ↓ 0  and as m  ↑ ∞ . This section confirms that impression and derives those slopes, which will be needed in the contraction mapping proof.30

Assume (as justified above) that a continuously differentiable concave consumption function exists in period t + 1,  with an origin at ct+1 (0 ) = 0  , a minimal MPC ˆκt+1 > 0  , and maximal MPC ¯κt+1 ≤ 1  . (If t + 1 = T  these will be ˆκT = ¯κT  = 1  ; for earlier periods they will exist by recursion.)

Under our imposed assumption that human wealth is finite, the MPC bound as wealth approaches infinity is easy to understand: As the proportion of consumption that will be financed out of human wealth approaches zero, the proportional difference between the solution to the model with uncertainty and the perfect foresight model shrinks to zero. In the course of proving this, Appendix G provides a recursive expression (used below) for the (inverse of the) limiting MPC as wealth approaches ∞ :

κ−t1 = 1 + ÞÞÞR κ−t1+1.
(29)

2.9.1 The Weak RIC Condition (WRIC)

Appendix equation (86) reflects the derivation of a parallel expression for the limiting maximal MPC as m  ↓ 0
  t  :

¯κ−t 1=  1 + ℘1 ∕ρÞÞÞR ¯κ−t+11
(30)

where {     } ∞
 ¯κ −T1−n  n=0   is a decreasing convergent sequence if the ‘weak return patience factor’ ℘1∕ρÞÞÞ
      R   satisfies:

WRIC:    1 ≥ ℘1∕ρÞÞÞR,
(31)

a condition we dub the ‘Weak Return Impatience Condition’ because with ℘  < 1  it will hold more easily (for a larger set of parameter values) than the RIC (ÞÞÞR < 1  ). The essence of the argument is that as wealth approaches zero, the overriding consideration that limits consumption is the (recursive) fear of the zero-income events. (That is why the probability of the zero income event ℘  appears in the expression.)

We are now in position to observe that the optimal consumption function must satisfy

κtmt  ≤  ct(mt ) ≤ κ¯tmt
(32)

because consumption starts at zero and is continuously differentiable, is strictly concave,31 and always exhibits a slope between κt  and ¯κt  (formal proof: Appendix C).

2.10 The Problem Is a Contraction Mapping Under WRIC and FVAC

As mentioned above, standard theorems in the contraction mapping literature before and after Stokey et. al. (1989) require utility or marginal utility to be bounded over the space of possible values of m  ; that requirement does not apply here because the possibility (however unlikely) of an unbroken string of zero-income events through the end of the horizon means that utility (and marginal utility) are unbounded as m ↓ 0  . Although a recent literature examines the existence and uniqueness of solutions to Bellman equations in the presence of ‘unbounded returns’ (see, e.g., Matkowski and Nowak (2011)), it appears that the techniques in that literature cannot be used to solve the problem here because the required conditions are violated by a problem that incorporates permanent shocks.32

Fortunately, Boyd (1990) provided a weighted contraction mapping theorem that Alvarez and Stokey (1998) showed could be used to address the homogeneous case (of which CRRA is an example) in a deterministic framework; later, Durán (2003) showed how to extend the Boyd (1990) approach to the stochastic case.

Definition 1. Consider any function ∙ ∈ 𝒞(𝒜, ℬ)  where 𝒞(𝒜, ℬ )  is the space of continuous functions from 𝒜  to ℬ  . Suppose г  ∈ 𝒞(𝒜, ℬ )  with ℬ ⊆  ℝ  and г >  0  . Then ∙ is г  -bounded if the г  -norm of ∙ ,

            [        ]
              | ∙-(m-)|
∥ ∙ ∥г = supm    г(m )   ,
(33)

is finite.

For 𝒞  (𝒜, ℬ )
  г  defined as the set of functions in 𝒞 (𝒜,ℬ )  that are г  -bounded; w  , x  , y  , and z  as examples of г  -bounded functions; and using 0(m ) = 0  to indicate the function that returns zero for any argument, Boyd (1990) proves the following.

Boyd’s Weighted Contraction Mapping Theorem. Let T  : 𝒞г (𝒜,ℬ ) → 𝒞 (𝒜, ℬ )  such that33 ,34

pict

Then T  defines a contraction with a unique fixed point.

For our problem, take 𝒜  as ℝ >0   and ℬ  as ℝ  , and define

pict

Using this, we introduce the mapping 𝒯 : 𝒞г(𝒜, ℬ ) → 𝒞 (𝒜, ℬ)  .35 Our operator 𝒯  satisfies the conditions that Boyd requires of his operator T  if we impose two restrictions on parameter values.36 The first is the WRIC necessary for convergence of the maximal MPC, equation (31) above. More serious is the Finite Value of Autarky Condition, equation (28). (The interpretation of these restrictions is elaborated in Section 2.12 below.) Imposing these restrictions, we are now in position to state the central theorem of the paper.

Theorem 1. 𝒯  is a contraction mapping if the weak return impatience condition (31) and the finite value of autarky condition (28) hold.

Intuitively, Boyd’s theorem shows that if you can find a г  that is everywhere finite but goes to infinity ‘as fast or faster’ than the function you are normalizing with г  , the normalized problem defines a contraction mapping. The intuition for the FVAC condition is that, with an infinite horizon, with any strictly positive initial amount of bank balances b0   , in the limit your value can always be made greater than you would get by consuming exactly the sustainable amount (say, by consuming (r∕R )b0 − 𝜖  for some arbitrarily small 𝜖 > 0  ).

We use the bounding function

              1−ρ
г(m ) = η + m    ,
(34)

for some real scalar η >  0  whose value is determined in the course of the proof.37 Under this definition of г  , {𝒯0} (mt ) = u(¯κmt )  is clearly г  -bounded.

The cumbersome details of the proof are relegated to Appendix C. Given that the value function converges, Appendix D.2 shows that the consumption functions converge.38

2.11 The Liquidity Constrained Solution as a Limit

A related problem commonly considered in the literature (e.g., by Deaton (1991)), with a liquidity constraint and a positive minimum value of income, is the limit of the problem considered here as the probability ℘  of the zero-income event approaches zero.

The ‘related’ problem makes two changes to the problem defined above:

  1. An ‘artificial’ liquidity constraint is imposed: at ≥ 0
  2. The probability of zero-income events is zero: ℘ =  0

The essence of the argument is simple. Imposing the artificial constraint without changing ℘  would not change behavior at all: The possibility of earning zero income over the remaining horizon already prevents the consumer from ending the period with zero assets. So, for precautionary reasons, the consumer will save something.

But the extent to which the consumer feels the need to make this precautionary provision depends on the probability that it will turn out to matter. As ℘ ↓ 0  , that probability becomes arbitrarily small, so the amount of precautionary saving induced by the zero-income events approaches zero as ℘ ↓ 0  . But “zero” is the amount of precautionary saving that would be induced by a zero-probability event for the impatient liquidity constrained consumer.

Another way to understand this is just to think of the liquidity constraint reflecting a component of the utility function that is zero whenever the consumer ends the period with (strictly) positive assets, but negative infinity if the consumer ends the period with (weakly) negative assets.

Table 3:Definitions and Comparisons of Conditions
|--------Perfect-ForesigFhitniVteersHiuomnasn---Wealth|-ConditionUn(cFeHrWtCaint)y--Versions-----------|
|                 𝒢∕R <  1                 |                 𝒢∕R <  1                 |
|    The growth factor for permanent income   |     The model’s risks are mean -preserving  |
|    𝒢 must be smaller than the discounting  |    spreads, so the PDV of future income is   |
|     factor R for human wealth to be finite.   |       unchanged by their introduction.      |
|                                          |                                          |
|-------------------------------------------------------------------------------------|
|------------------------Absolute-Impatience-Condition--(AIC-)------------------------|
|                  ÞÞÞ < 1                   |                  ÞÞÞ < 1                   |
|                                          |                                          |
|        The unconstrained consumer is        |    If wealth is large enough, the expectation|
|     sufficiently impatient that the level of|      of consumption next period will be    |
|    consumption will be declining over time: |    smaller than this period’s consumption:   |
|                                          |                                          |
|                ct+1 < ct                 |             lim  𝔼t [ct+1] < ct            |
|                                          |            mt→∞                          |
|-------------------------------------------------------------------------------------|
|---------------------------Return--Impatience-Conditions-----------------------------|
|   Return  Impatience  Condition  (RIC )    |           Weak  RIC  (WRIC  )            |
|----------------Þ-------------------------|----------------1∕ρÞ-----------------------|
|                ÞÞ∕R <  1                 |               ℘  ÞÞ∕R <  1               |
|     The growth factor for consumption ÞÞÞ     |      If the probability of the zero-income     |
|     must be smaller than the discounting   |   event is ℘ = 1 then income is always zero |
|    factor R, so that the PDV of current and |     and the condition becomes identical to   |
|       future consumption will be finite:     |         the RIC. Otherwise, weaker.       |
|                                          |                                          |
|           ′                              |         ′            1∕ρ                 |
|          c (m ) = 1 − ÞÞÞ∕R < 1           |        c (m ) < 1 − ℘  ÞÞÞ∕R <  1         |
|-------------------------------------------------------------------------------------|
|                           Growth  Impatience  Conditions                            |
|-------------------GIC--------------------|----------------GIC--Mod------------------|
|------------------------------------------|-------------------−1---------------------|
|                ÞÞÞ∕𝒢 <  1                 |             ÞÞÞ 𝔼[Ψ   ]∕𝒢 < 1              |
|    For an unconstrained PF consumer, the    |                                          |
|     ratio of c to p will fall over time. For|   By Jensen’s inequality stronger than GIC.  |
|     constrained, guarantees the constraint  |     Ensures consumers will not expect to  |
|        eventually binds. Guarantees        |         accumulate m unboundedly.         |
|           lim  𝔼t[Ψt+1mt+1 ∕mt] = ÞÞÞ 𝒢      |                                          |
|          mt↑∞                            |         mltim→ ∞ 𝔼t[mt+1∕mt ] = ÞÞÞ𝒢-         |
|                                          |                                          |
|-------------------------------------------------------------------------------------|
|-------------------------Finite Value-of Autarky-Conditions--------------------------|
|----------------PF--FVAC-------------------|------------------FVAC--------------------|
|                β𝒢1− ρ < 1                |            β𝒢1− ρ𝔼[Ψ1 −ρ] < 1            |
|       equivalently ÞÞÞ < R1∕ρ𝒢1− 1∕ρ       |                                          |
|                                          |                                          |
|     The discounted utility of constrained     |   By Jensen’s inequality, stronger than the   |
|     consumers who spend their permanent    |       PF-FVAC  because for ρ1 >−ρ1 and       |
|      income each period should be finite.   |       nondegenerate Ψ, 𝔼[Ψ    ] > 1.       |
|                                          |                                          |
---------------------------------------------------------------------------------------

See Appendix F for the formal proof justifying the foregoing intuitive discussion.39

The conditions required for convergence and nondegeneracy are thus strikingly similar between the liquidity constrained perfect foresight model and the model with uncertainty but no explicit constraints: The liquidity constrained perfect foresight model is just the limiting case of the model with uncertainty as the degree of all three kinds of uncertainty (zero-income events, other transitory shocks, and permanent shocks) approaches zero.

2.12 Relations Between Parametric Restrictions

The full relationship among conditions is represented in Figure 3. Though the diagram looks complex, it is merely a modified version of the earlier simple diagram (Figure 1) with further (mostly intermediate) inequalities inserted. (Arrows with a “because” now label relations that always hold under the model’s assumptions.)40

PIC
Figure 3:Relation of All Inequality Conditions
See Table 2 for Numerical Values of Nodes Under Baseline Parameters

The ‘weakness’ of the additional condition sufficient for contraction beyond the FVAC, the WRIC, can be seen by asking ‘under what circumstances would the FVAC hold but the WRIC fail?’ Algebraically, the requirement is

β 𝒢1−ρΨ-1−ρ <  1  < (℘β )1∕ρ∕R1− 1∕ρ.
      ---
(35)

If we require R  ≥ 1  , the WRIC is redundant because now β <  1 < Rρ−1   , so that (with ρ > 1  and β <  1  ) the RIC (and WRIC) must hold. But neither theory nor evidence demand that R ≥ 1  . We can therefore approach the question of the WRIC’s relevance by asking just how low R  must be for the condition to be relevant. Suppose for illustration that ρ = 2  , Ψ1-− ρ = 1.01  , 𝒢1 −ρ = 1.01−1   and ℘ = 0.10  . In that case (35) reduces to

pict

but since β < 1  by assumption, the binding requirement is that

pict

so that for example if β  = 0.96  we would need R <  0.096  (that is, a perpetual riskfree rate of return of worse than -90 percent a year) in order for the WRIC to bind.

Perhaps the best way of thinking about this is to note that the space of parameter values for which the WRIC is relevant shrinks out of existence as ℘ →  0  , which Section 2.11 showed was the precise limiting condition under which behavior becomes arbitrarily close to the liquidity constrained solution (in the absence of other risks). On the other hand, when ℘ =  1  , the consumer has no noncapital income (so that the FHWC holds) and with ℘ = 1  the WRIC is identical to the RIC; but the RIC is the only condition required for a solution to exist for a perfect foresight consumer with no noncapital income. Thus the WRIC forms a sort of ‘bridge’ between the liquidity constrained and the unconstrained problems as ℘  moves from 0 to 1.

2.12.1 When the RIC Fails

In the perfect foresight problem (Section 2.5.3), the RIC was necessary for existence of a nondegenerate solution. It is surprising, therefore, that in the presence of uncertainty, the much weaker WRIC is sufficient for nondegeneracy (assuming that the FVAC holds). We can directly derive the features the problem must exhibit (given the FVAC) under ¬ RIC (that is,          1∕ρ
R <  (Rβ)   )  :

              implied by FVAC
       ◜-----------◞◟---------1∕◝ρ
   R < (R β)1∕ρ  <  (R(𝒢 Ψ-)ρ− 1)
             1∕ρ   1−1∕ρ
   R < (R ∕𝒢)   𝒢Ψ--
R ∕𝒢 < Ψ--
       ---
(36)

but since Ψ-<  1  (cf. the argument below (26)), this requires R∕𝒢  < 1  ; so, given the FVAC, the RIC can fail only if human wealth is unbounded. As an illustration of the convenience of our diagrams, note that this algebraically complicated conclusion could be easily reached diagrammatically in figure 3 by starting at the R  node and imposing ¬RIC  , which reverses the RIC arrow and lets us traverse the diagram along any clockwise path to the PF-VAF node at which point we realize that we cannot impose the FHWC because that would let us conclude R  > R  .

As in the perfect foresight constrained problem, unbounded limiting human wealth (¬ FHWC) here does not lead to a degenerate limiting consumption function (finite human wealth is not required for the convergence theorem). But, from equation (29) and the discussion surrounding it, an implication of ¬ RIC is that limm ↑∞ c′(m ) = 0  . Thus, interestingly, in the special {¬RIC,   ¬FHWC   } case (unavailable in the perfect foresight model) the presence of uncertainty both permits unlimited human wealth (in the n ↑ ∞ limit) and at the same time prevents unlimited human wealth from resulting in (limiting) infinite consumption (at any finite m  ). Intuitively, in the presence of uncertainty, pathological patience (which in the perfect foresight model results in a limiting consumption function of c(m ) = 0  ) plus unbounded human wealth (which the perfect foresight model prohibits because it leads to a limiting consumption function c(m ) = ∞ for any finite m  ) combine to yield a unique finite limiting (as n ↑ ∞ ) level of consumption and MPC for any finite value of m  . Note the close parallel to the conclusion in the perfect foresight liquidity constrained model in the {GIC¬ RIC} case. There, too, the tension between infinite human wealth and pathological patience was resolved with a nondegenerate consumption function whose limiting MPC was zero.41

2.12.2 When the RIC Holds

FHWC. If the RIC and FHWC both hold, a perfect foresight solution exists (see 2.5.3 above). As m  ↑ ∞ the limiting c  and v  functions become arbitrarily close to those in the perfect foresight model, because human wealth pays for a vanishingly small portion of spending. This will be the main case analyzed in detail below.

¬ FHWC. The more exotic case is where FHWC fails; in the perfect foresight model, {RIC,¬ FHWC} is the degenerate case with limiting ¯c(m ) = ∞ . Here, the FVAC implies that the PF-FVAC holds (traverse Figure 3 clockwise from Þ
ÞÞ  by imposing FVAC and continue to the PF-VAF node): Reversing the arrow connecting the R  and PF-VAF nodes implies that under ¬FHWC  :

pict

where the transition from the first to the second lines is justified because                   1∕ρ
¬FHWC    ⇒  (R ∕𝒢)   <  1  . So, {RIC, ¬ FHWC} implies the GIC holds. However, we are not entitled to conclude that the GIC-Mod holds: ÞÞÞ < 𝒢 does not imply ÞÞÞ <  Ψ-𝒢 where Ψ  < 1
---  .

We have now established the principal points of comparison between the perfect foresight solutions and the solutions under uncertainty; these are codified in the remaining parts of Tables 3 and 4.

Table 4:Sufficient Conditions for Nondegenerate‡ Solution
|--------------------------|---------------|-----------------------------------------|
|Consumption   Model (s)   | Conditions    |               Comments                  |
|--------------------------|------------∘--|-----------------------------------------|
|¯c(m ): PF Unconstrained   |RIC, FHWC      |RIC ⇒  |v(m )| < ∞; FHWC   ⇒  0 < |v(m )| |
|c(m ) = κm                |               |PF  model with  no human  wealth (h = 0) |
|                          |               |                                         |
|           Section 2.5.3: |               |RIC  prevents ¯c(m ) = c(m ) = 0          |
|           Section 2.5.3: |               |FHWC    prevents ¯c(m ) = ∞               |
|                          |               |                                         |
|                Eq  (22): |               |PF -FVAC+FHWC      ⇒  RIC                |
|----------------Eq--(21):-|---------------|GIC+FHWC------⇒--PF--FVAC-----------------|
|`c(m ): PF Constrained     |¬GIC,  RIC     |FHWC    holds (𝒢 < ÞÞÞ <  R ⇒  𝒢 < R )     |
|           Section 2.5.6: |               |`c(m ) = ¯c(m ) for m > m   < 1            |
|                          |               |                        #                |
|                          |---------------|(¬RIC--would--yield-m#--=--0 so-`c(m-)-=-0)|
|            Appendix  E:  |GIC,RIC        |limm  →∞ `c(m ) = ¯c(m ),limm → ∞ `κκκ(m ) = κ- |
|                          |---------------|kinks-where-horizon-to-b-=-0-changes∗----|
|            Appendix  E:  |GIC, ¬RIC      |limm  →∞ `κκκ (m ) = 0                        |
|                          |               |                                    ∗    |
|--------------------------|---------------|kinks-where-horizon-to-b-=-0-changes-----|
|c(m ): Friedman/Muth       |Section 3.1,    |c(m ) < c(m ) < ¯c(m )                    |
|                          |Section 3.2     |v(m ) < v(m ) < ¯v(m )                    |
|            Section 2.10: |FVAC,--WRIC----|Su-fficient-for Contraction----------------|
|                          |               |                                         |
|            Section 2.12: |               |WRIC   is weaker than RIC                |
|                Figure 3: |               |FVAC   is stronger than PF -FVAC         |
|          Section 2.12.2: |               |¬FHWC+RIC      ⇒GIC,   limm → ∞ κκκ(m ) = κ-|
|          Section 2.12.1: |               |¬RIC  ⇒  ¬FHWC,   limm  →∞ κκκ (m ) = 0     |
|             Section 3.3: |               |“Bu-ffer-Stock-Saving”-Conditions---------|
|                          |               |                                         |
|           Section 3.3.2: |               |     GIC  ⇒  ∃ mˇ s.t. 0 < ˇm <  ∞         |
------------Section-3.3.1:------------------GIC--Mod--⇒--∃-mˆ-s.t.-0 <-ˆm-<--∞----------

‡ For feasible m  satisfying 0 < m < ∞ , a nondegenerate limiting consumption function defines a unique optimal value of c  satisfying 0 < c(m ) < ∞ ; a nondegenerate limiting value function defines a corresponding unique value of − ∞ < v(m) < 0  .
  ∘ RIC, FHWC are necessary as well as sufficient for the perfect foresight case.  ∗ That is, the first kink point in c(m)  is m#  s.t. for m < m#  the constraint will bind now, while for m > m#  the constraint will bind one period in the future. The second kink point corresponds to the m  where the constraint will bind two periods in the future, etc.
  ∗∗ In the Friedman/Muth model, the RIC+FHWC are sufficient, but not necessary for nondegeneracy

3 Analysis of the Converged Consumption Function

Figures 4-6 capture the main properties of the converged consumption rule when the RIC, GIC-Mod, and FHWC all hold.42

Figure 4 shows the expected growth factors for consumption, the level of market resources, and the market resources ratio, 𝔼t[ct+1∕ct]  and 𝔼t[mt+1 ∕mt ]  , and 𝔼t[mt+1 ∕mt ]  , for a consumer behaving according to the converged consumption rule, while Figures 56 illustrate theoretical bounds for the consumption function and the MPC.

Three points are worth highlighting.

First, as mt  ↑ ∞ the expected consumption growth factor goes to ÞÞÞ  , indicated by the lower bound in Figure 4, and the marginal propensity to consume approaches κ-=  (1 − ÞÞÞR )  (see Figure 5) — the same as the perfect foresight MPC. Second, as mt  approaches zero the consumption growth factor approaches ∞ (Figure 4) and the MPC approaches           1∕ρ
¯κ = (1 − ℘   ÞÞÞR)  (Figure 5). Third, there is a value of the market resources ratio mt = mˇ  at which the expected growth rate of the level of market resources m  matches the expected growth rate of permanent income 𝒢 , and a different (larger) target ratio mˆ  where 𝔼 [mt+1∕mt ] = 1  (and the expected growth rate of consumption is lower than 𝒢 ). Thus, at the individual level, this model does not have a single m  at which p, m,  and c  all are expected to grow at the same rate.

pict
Figure 4:‘Stable’ m  Values and Expected Growth Rates

3.1 Limits as m  Approaches Infinity

Define

pict

which is the solution to an infinite-horizon problem with no noncapital income (ξξξt+n = 0 ∀ n ≥ 1  ); clearly c(m ) < c(m )  , since allowing the possibility of future noncapital income cannot reduce current consumption. Our imposition of the RIC guarantees that κ->  0  , so this solution satisfies our definition of nondegeneracy, and because this solution is always available it defines a lower bound on both the consumption and value functions.

Assuming the FHWC holds, the infinite horizon perfect foresight solution (16) constitutes an upper bound on consumption in the presence of uncertainty, since the introduction of uncertainty strictly decreases the level of consumption at any m  (Carroll and Kimball (1996)). Thus,

c(m ) <c (m ) < ¯c(m )
   1 <c (m )∕c-(m ) < ¯c(m )∕c(m ).
(37)

But

pict

so as m  ↑ ∞, c(m )∕c(m ) →  1  , and the continuous differentiability and strict concavity of c(m )  therefore implies

      ′       ′       ′
mlim↑∞ c(m ) = c(m ) = ¯c (m ) = κ

because any other fixed limit would eventually lead to a level of consumption either exceeding ¯c(m )  or lower than c(m )  . Figure 5 illustrates these limits by plotting the numerical solution.

PIC
Figure 5:Limiting MPC’s

PIC

Figure 6:Upper and Lower Bounds on the Consumption Function

Next we establish the limit of the expected consumption growth factor as mt  ↑ ∞ :

pict

But

pict

and

 lim  𝒢t+1c(mt+1 )∕¯c(mt) =  lim  𝒢t+1¯c(mt+1 )∕c(mt) =  lim  𝒢t+1mt+1 ∕mt,
mt↑∞                      mt↑∞                      mt↑∞

while (for convenience defining a(mt ) = mt − c(mt )  ),

                         (               ξ   )
 lim  𝒢t+1mt+1 ∕mt =  lim   Ra-(mt)-+-𝒢t+1ξξt+1
mt↑∞                 mt↑∞           mt
                  =  (R β)1∕ρ = ÞÞÞ

because lim      a′(m ) = ÞÞÞ
   mt↑∞           R   43 and                  ¯ ¯
𝒢t+1ξξξt+1 ∕mt ≤ (𝒢 Ψ 𝜃𝜃𝜃∕(1 − ℘))∕mt  which goes to zero as mt  goes to infinity.

Hence we have

ÞÞÞ  ≤ mlitm↑∞ 𝔼t[ct+1 ∕ct] ≤ ÞÞÞ

so as cash goes to infinity, consumption growth approaches its value ÞÞÞ  in the perfect foresight model.

3.2 Limits as m Approaches Zero

Equation (30) shows that the limiting value of ¯κ  is

pict

Defining e(m ) = c(m )∕m  we have

pict

Now using the continuous differentiability of the consumption function along with L’Hôpital’s rule, we have

pict

Figure 5 visually confirms that the numerical solution obtains this limit for the MPC as m  approaches zero.

For consumption growth, as m ↓ 0  we have

pict

where the second-to-last line follows because           [( c(ℛ   a(m )))     ]
limmt ↓0 𝔼t   ---t+κ1¯mt--t-  𝒢t+1 is positive, and the last line follows because the minimum possible realization of 𝜃𝜃𝜃
  t+1   is 𝜃𝜃𝜃 > 0  so the minimum possible value of expected next-period consumption is positive.44

3.3 Unique ‘Stable’ Points

Theorems whose substance is described here (and whose details are in an appendix) articulate alternative (but closely related) stability criteria.

3.3.1 ‘Individual Target Wealth’ ˆm

One kind of ‘stable’ point is a ‘target’ value ˆm  such that if m  = mˆ
  t  , then 𝔼  [m    ] = m
  t  t+1      t  . Existence of such a target turns out to require the GIC-Mod condition.

Theorem 2. For the nondegenerate solution to the problem defined in Section 2.1 when FVAC, WRIC, and GIC-Mod all hold, there exists a unique cash-on-hand-to-permanent-income ratio mˆ > 0  such that

𝔼t[mt+1 ∕mt] = 1 if mt = ˆm.
(38)

Moreover, ˆm  is a point of ‘stability’ in the sense that

 ∀mt ∈  (0, ˆm), 𝔼t[mt+1 ] > mt
∀mt ∈ (mˆ, ∞ ), 𝔼t[mt+1 ] < mt.
(39)

Since mt+1  = (mt − c(mt ))ℛt+1 + ξξξt+1   , the implicit equation for mˆ  is

𝔼t
(40)

This is the most restictive among several competing definitions of stability.

3.3.2 Individual Balanced-Growth ‘pseudo steady state’ ˇm

Our least restrictive definition of ‘stability’ derives from a traditional question in macro models: whether there is a ‘balanced growth’ equilibrium in which aggregate variables (income, consumption, market resources) all grow forever by the same factor 𝒢 . For our model, Figure 4 showed that there is no single m  for which 𝔼 [m    ]∕m   = 𝔼 [c   ]∕c  = 𝒢
 t   t+1    t    t  t+1    t  for an individual consumer. Nevertheless, the next section will show that economies populated by heterogeneous collections of such consumers can exhibit balanced growth in the aggregate, and in the cross-section of households.

As an input to that analysis, we show here that if the GIC holds, the problem will exhibit a balanced-growth ‘pseudo-steady-state’ point, by which we mean that there is some ˇm  such that, for all

mt  > mˇ  , 𝔼t [mt+1 ∕mt ] < 𝒢 , and conversely if mt <  ˇm  then 𝔼t [mt+1 ∕mt ] > 𝒢 .

The critical m  will be the ˇm  at which m  growth matches 𝒢 :

                                  𝔼  [m    ]∕m  =  𝔼 [p   ]∕p
                                    t  t+1    t    t  t+1   t
                       𝔼t[mt+1 𝒢 Ψt+1pt]∕(mtpt ) = 𝔼t[pt𝒢Ψt+1 ]∕pt
   ⌊                                     ⌋
   ⌈                                     ⌉
𝔼t  Ψt+1  ((◟mt-−--c(mt)R-∕◝(◜𝒢Ψt+1-)) +-ξξξt+1)◞ ∕mt =  1
                       mt+1
                 ⌊              ℛ            ⌋
                              ◜◞◟◝
               𝔼t⌈ (ˇm − c(mˇ))R ∕𝒢 + Ψt+1ξξξt+1⌉ =  ˇm


                              (ˇm − c(mˇ))ℛ +  1 = ˇm.
(41)

The only difference between (41) and (40) is the substitution of ℛ for  ¯
ℛ .45 ,  46

Theorem 3. For the nondegenerate solution to the problem defined in Section 2.1 when FVAC, WRIC, and GIC all hold, there exists a unique pseudo-steady-state cash-on-hand-to-income ratio ˇm > 0  such that

𝔼t[Ψt+1mt+1  ∕mt ] = 1 if mt = ˇm.
(42)

Moreover, ˇm  is a point of stability in the sense that

 ∀mt  ∈ (0, ˇm ), 𝔼t[mt+1 ]∕mt > 𝒢

∀mt  ∈ (ˇm, ∞ ), 𝔼t[mt+1 ]∕mt  < 𝒢.
(43)

The proofs of these theorems are intuitive, and almost completely parallel; to save space, they are relegated to Appendix H. They involve three steps:

  1. Existence and continuity of 𝔼t[mt+1∕mt ]  or 𝔼t[mt+1 Ψt+1∕mt ]
  2. Existence of the equilibrium point
  3. Monotonicity of 𝔼t[mt+1  − mt]  or 𝔼t [mt+1 Ψt+1 −  mt]

3.4 Example With Balanced-Growth mˇ  But No Target ˆm

Because the equations defining target and pseudo-steady-state m  , (40) and (41), differ only by substitution of ℛ for ℛ¯ = ℛ 𝔼 [Ψ −1]  , if there are no permanent shocks (Ψ  ≡ 1  ), the conditions are identical. For many parameterizations (e.g., under the baseline parameter values used for constructing figure 4), ˆm  and mˇ  will not differ much.

PIC
Figure 7:{FVAC,GIC,¬ GIC-Mod}: No Target Exists But SS Does

An illuminating exception is exhibited in Figure 7, which modifies the baseline parameter values by quadrupling the variance of the permanent shocks, enough to cause failure of the GIC-Mod; now there is no target wealth level mˆ  . But the pseudo-steady-state still exists because it turns off realizations of the permanent shock. It is tempting to conclude that the reason target ˆm  does not exist is that the increase in the size of the shocks induces a precautionary motive that increases the consumer’s effective patience. But that interpretation is not correct when the FHWC holds, because as market resources approach infinity, precautionary saving against noncapital income risk becomes negligible (as the proportion of consumption financed out of such income approaches zero). The correct explanation is more prosaic: The increase in uncertainty boosts the expected uncertainty-modified rate of return factor from ℛ to ¯ℛ >  ℛ which reflects the fact that in the presence of uncertainty the expectation of the inverse of the growth factor increases: 𝒢->  𝒢 . That is, in the limit as m  ↑ ∞ the increase in effective impatience reflected in ÞÞÞ𝒢-< ÞÞÞ𝒢 is entirely due to the certainty-equivalence growth adjustment, not to a (limiting) change in precaution. In fact, the next section will show that an aggregate balanced growth equilibrium will exist even when realizations of the permanent shock are not turned off: The required condition for aggregate balanced growth is the regular GIC, which ignores the magnitude of permanent shocks, not the GIC-Mod.47

Before we get to the formal arguments, the key insight can be understood by considering an economy that starts, at date t  , with the entire population at mt = mˇ  , but then evolves according to the model’s assumed dynamics between t  and t + 1  . Equation (41) will still hold, so for this first period, at least, the economy will exhibit balanced growth: the growth factor for aggregate m  will match the growth factor for permanent income 𝒢 . It is true that there will be people for whom bt+1 = atR∕(𝒢 Ψt+1)  is boosted by a small draw of Ψt+1   . But their contribution to the level of the aggregate variable is given by bt+1 = bt+1Ψt+1   , so their b
 t+1   is reweighted by an amount that exactly unwinds that divisor-boosting. This means that it is possible for the consumption-to-permanent-income ratio for every consumer to be small enough that their market resources ratio is expected to rise, and yet for the economy as a whole to exhibit a balanced growth equilibrium with a finite aggregate balanced growth steady state Mˇ  (this is not numerically the same as the individual pseudo-steady-state ratio ˇm  because the problem’s nonlinearities have consequences when aggregated).48

4 Invariant Relationships

Assume a continuum of ex ante identical buffer-stock households on the unit interval, with constant total mass normalized to one and indexed by i ∈ [0,1 ]  .

Szeidl (2013) proved that such a population will be characterized by invariant distributions of m  , c  , and a  under the condition49

GIC -Szeidl: log ÞÞÞ 𝒢 < 𝔼 [logΨ ]
(44)

which is stronger than our GIC (ÞÞÞ 𝒢 < 1  ), but weaker than our GIC-Mod (ÞÞÞ 𝒢 < 1  ).50

Harmenberg (2021) substitutes a clever change of probability-measure into Szeidl’s proof, with the implication that if the GIC holds, invariant permanent-income-weighted distributions exist. This allows him to prove a conjecture from an earlier draft of this paper (Carroll (2019, Submitted)) that under the GIC, aggregate consumption grows at the same rate 𝒢 as aggregate noncapital income in the long run (with the corollary that aggregate assets and market resources grow at that same rate). Harmenberg (2021) shows that his reformulation of the problem can reduce costs of calculation enormously.51 In confirmation, this notebook finds that the Harmenberg method reduces the simulation size required for a given degree of accuracy by roughly a factor of 100 (!) under the baseline parameter values defined above.

The remainder of this section briefly draws out some implications of these points.

4.1 Individual Balanced Growth of Income, Consumption, and Wealth

Define 𝕄  [∙ ]  to yield the mean of its argument in the population (as distinct from the expectations operator 𝔼[∙]  which represents beliefs about the future). Using boldface capitals for aggregates, the growth factor for aggregate noncapital income is:

pict

because of the independence assumptions we have made about ξξξ  and Ψ  .

Consider an economy that satisfies the Szeidl impatience condition (44) and has existed for long enough by date t  that we can consider it as Szeidl-converged. In such an economy a microeconomist with a population-representative panel dataset could calculate the growth rate of consumption for each individual household, and take the average:

𝕄 [Δ logct+1] = 𝕄 [logct+1pt+1 − logctpt]

             =  𝕄 [logpt+1 − log pt] + 𝕄 [logct+1 − logct].
(45)

Because this economy is Szeidl-converged, distributions of ct  and ct+1   will be identical, so that the second term in (45) disappears; hence, mean cross-sectional growth rates of consumption and permanent income are the same:

𝕄 [Δ logc   ] = 𝕄 [Δ logp    ] = log 𝒢.
         t+1              t+1
(46)

In a Harmenberg-invariant economy (and therefore also any Szeidl-invariant economy), a similar proposition holds in the cross-section as a direct implication of the fact that a constant proportion of total permanent income is accounted for by the successive sets of consumers with any particular m  . This fact is one way of interpreting Harmenberg’s definition of the density of the permanent-income-invariant distribution of m  ; call this density p (m )  .52 Call ct(m )  the total amount of consumption at date t  by persons with market resources m  , and note that in the invariant economy this is given by the converged consumption function c(m )  multiplied by the amount of permanent income accruing to such people p(m )pt  . Since p(m )  is invariant and aggregate permanent income grows according to Pt+1 =  𝒢Pt  , for any m  :

log ct+1 (m ) − logct(m ) = log c(m )p(m )Pt+1 − log c(m )p(m )Pt

                       =  log 𝒢.

4.2 Aggregate Balanced Growth and Idiosyncratic Covariances

Harmenberg shows that the covariance between the individual consumption ratio c  and the idiosyncratic component of permanent income p  does not shrink to zero; thus, covariances are another potential measurement for construction of microfoundations.

Consider a date-t  Harmenberg-converged economy, and define the mean value of the consumption ratio as 𝔠t+n ≡ 𝕄 [ct+n]  . Normalizing period-t  aggregate permanent income to Pt = 1  , total consumption at t + 1  and t + 2  are

Ct+1 =  𝕄 [ct+1pt+1] = 𝔠t+1𝒢1 +  covt+1 (ct+1,pt+1)
                           2
Ct+2 =  𝕄 [ct+2pt+2] = 𝔠t+2𝒢  +  covt+2 (ct+2,pt+2)
(47)

and Harmenberg’s proof that C    − 𝒢C     = 0
 t+2      t+1  allows us to obtain:

(𝔠t+2 − 𝔠t+1) 𝒢2 = 𝒢covt+1 − covt+2.
(48)

In a Szeidl-invariant economy, 𝔠t+2 = 𝔠t+1   , so the economy exhibits balanced growth in the covariance:

covt+2 = 𝒢covt+1.
(49)

The more interesting case is when the economy is Harmenberg- but not Szeidl-invariant. In that case, if the cov  and the 𝔠  terms have constant growth factors Ωcov   and Ω 𝔠   ,53 an equation corresponding to (48) will hold in t + n  :

    𝔠
    ◜t◞+◟n◝               (             )
   (Ωn𝔠𝔠t− Ωn𝔠−1𝔠t)𝒢n = 𝒢 Ωnc−ov1− Ωncov  covt
      n−1                n−1
(Ω 𝔠𝒢)   (Ω 𝔠 − 1)𝔠t𝒢 = Ω cov (𝒢 − Ωcov)covt
(50)

so for the LHS and RHS to grow at the same rates we need

Ωcov = Ω 𝔠𝒢.
(51)

This is intuitive: In the Szeidl-invariant economy, it just reproduces our result above that the covariance exhibits balanced growth because Ω 𝔠 = 1  . The revised result just says that in the Harmenberg case where the mean value 𝔠  of the consumption ratio c  can grow, the covariance must rise in proportion to any ongoing expansion of 𝔠  (as well as in proportion to the growth in p  ).

4.3 Implications for Microfoundations

Thus we have microeconomic propositions, for both growth rates and for covariances of observable variables,54 that can be tested in either cross-section or panel microdata to judge (and calibrate) the microfoundations that should hold for any macroeconomic analysis that requires balanced growth for its conclusions.

At first blush, these points are reassuring; one of the most persuasive arguments for the agenda of building microfoundations of macroeconomics is that newly available ‘big data’ allow us to measure cross-sectional covariances with great precision, so that we can use microeconomic natural experiments to disentangle questions that are hopelessly entangled in aggregate time-series data. Knowing that such covariances ought to be a stable feature of a stably growing economy is therefore encouraging.

But this discussion also highlights an uncomfortable point: In the model as specified, permanent income does not have a limiting distribution; it becomes ever more dispersed as the economy with infinite-horizon consumers continues to exist indefinitely.

A few microeconomic data sources attempt direct measurement of ‘permanent income’; Carroll, Slacalek, Tokuoka, and White (2017), for example, show that their assumptions about the magnitude of permanent shocks (and mortality; see below) yield a simulated distribution of permanent income that roughly matches answers in the U.S. Survey of Consumer Finances (‘SCF’) to a question designed to elicit a direct measure of respondents’ permanent income. They use those results to calibrate a model to match empirical facts about the distribution of permanent income and wealth, showing that the model also does fits empirical facts about the marginal propensity to consume. The quantitative credibility of the argument depends on the model’s match to the distribution of permanent income inequality, which would not be possible in a model without a nondegenerate steady-state distribution of permanent income.

For macroeconomists who want to build microfoundations by comparing the microeconomic implications of their models to micro data (directly – not in ratios to difficult-to-meaure ‘permanent income’), it would be something of a challenge to determine how to construct empirical-data-comparable simulated results from a model with no limiting distribution of permanent income.

Death can solve this problem.

4.4 Mortality Yields Invariance

Most heterogeneous-agent models incorporate a constant positive probability of death, following Blanchard (1985) and Yaari (1965). In the Blanchardian model, if the probability of death exceeds a threshold that depends on the size of the permanent shocks, Carroll, Slacalek, Tokuoka, and White (2017) show that the limiting distribution of permanent income has a finite variance. Blanchard (1985) assumes a universal annuitization scheme in which estates of dying consumers are redistributed to survivors in proportion to survivors’ wealth, giving the recipients a higher effective rate of return. This treatment has considerable analytical advantages, most notably that the effect of mortality on the time preference factor is the exact inverse of its effect on the (effective) interest factor. That is, if the ‘pure’ time preference factor is β  and probability of remaining alive (not dead) is ℒ , then the assumption that no utility accrues after death makes the effective discount factor β- = βℒ while the enhancement to the rate of return from the annuity scheme yields an effective interest factor R¯=  R∕ℒ (recall that because of white-noise mortality, the average wealth of the two groups is identical). Combining these, the effective patience factor in the new economy βR¯  is unchanged from its value in the infinite horizon model:

               1∕ρ       1∕ρ
β-¯R = (βℒR ∕ ℒ)   =  (R β)   =  ÞÞÞ.
(52)

The only adjustments this requires to the analysis above are therefore to the few elements that involve a role for the interest factor distinct from its contribution to ÞÞÞ  (principally, the RIC, which becomes Þ  ¯
ÞÞ ∕R  ).

Blanchard (1985)’s innovation was valuable not only for the insight it provided but also because when he wrote, the principal alternative, the Life Cycle model of Modigliani (1966), was computationally challenging given then-available technologies. Despite its (considerable) conceptual value, Blanchard’s analytical solution is now rarely used because essentially all modern modeling incorporates uncertainty, constraints, and other features that rule out analytical solutions anyway.

The simplest alternative to Blanchard is to follow Modigliani in constructing a realistic description of income over the life cycle and assuming that any wealth remaining at death occurs accidentally (not implausible, given the robust finding that for the great majority of households, bequests amount to less than 2 percent of lifetime earnings, Hendricks (20012016)).

Even if bequests are accidental, a macroeconomic model must make some assumption about how they are disposed of: As windfalls to heirs, estate tax proceeds, etc. We again consider the simplest choice, because it represents something of a polar alternative to Blanchard. Without a bequest motive, there are no behavioral effects of a 100 percent estate tax; we assume such a tax is imposed and that the revenues are effectively thrown in the ocean: The estate-related wealth effectively vanishes from the economy.

The chief appeal of this approach is the simplicity of the change it makes in the condition required for the economy to exhibit a balanced growth equilibrium (for consumers without a life cycle income profile). If ℒ is the probability of remaining alive, the condition changes from the plain GIC to a looser mortality-adjusted GIC:

ℒ ÞÞÞ𝒢 < 1.
(53)

With no income growth, what is required to prohibit unbounded growth in aggregate wealth is the condition that prevents the per-capita wealth-to-permanent-income ratio of surviving consumers from growing faster than the rate at which mortality diminishes their collective population. With income growth, the aggregate wealth-to-income ratio will head to infinity only if a cohort of consumers is patient enough to make the desired rate of growth of wealth fast enough to counteract combined erosive forces of mortality and productivity.

5 Conclusions

Numerical solutions to optimal consumption problems, in both life cycle and infinite horizon contexts, have become standard tools since the first reasonably realistic models were constructed in the late 1980s. One contribution of this paper is to show that finite horizon (‘life cycle’) versions of the simplest such models, with assumptions about income shocks (transitory and permanent) dating back to Friedman (1957) and standard specifications of preferences — and without plausible (but computationally and mathematically inconvenient) complications like liquidity constraints — have attractive properties (like continuous differentiability of the consumption function, and analytical limiting MPC’s as resources approach their minimum and maximum possible values).

The main focus of the paper, though, is on the limiting solution of the finite horizon model as the time horizon approaches infinity. This simple model has other appealing features: A ‘Finite Value of Autarky’ condition guarantees convergence of the consumption function, under the mild additional requirement of a ‘Weak Return Impatience Condition’ that will never bind for plausible parameterizations, but provides intuition for the bridge between this model and models with explicit liquidity constraints. The paper also provides a roadmap for the model’s relationships to the perfect foresight model without and with constraints. The constrained perfect foresight model provides an upper bound to the consumption function (and value function) for the model with uncertainty, which explains why the conditions for the model to have a nondegenerate solution closely parallel those required for the perfect foresight constrained model to have a nondegenerate solution.

The main use of infinite horizon versions of such models is in heterogeneous-agent macroeconomics. The paper articulates intuitive ‘Growth Impatience Conditions’ under which populations of such agents, with Blanchardian (tighter) or Modiglianian (looser) mortality will exhibit balanced growth. Finally, the paper provides the analytical basis for many results about buffer-stock saving models that are so well understood that even without analytical foundations researchers uncontroversially use them as explanations of real-world phenomena like the cross-sectional pattern of consumption dynamics in the Great Recession.

A c  Functions Exist, are Concave, and Differentible

To show that (5) defines a sequence of continuously differentiable strictly increasing concave functions {cT,cT− 1,...,cT−k} , we start with a definition. We will say that a function n(z)  is ‘nice’ if it satisfies

  1. n(z)  is well-defined iff z > 0
  2. n(z)  is strictly increasing
  3. n(z)  is strictly concave
  4. n(z)  is C3
  5. n(z) < 0
  6. lim    n (z ) = − ∞
   z↓0 .

(Notice that an implication of niceness is that limz ↓0 n′(z) = ∞ .)

Assume that some vt+1   is nice. Our objective is to show that this implies vt  is also nice; this is sufficient to establish that vt−n  is nice by induction for all n >  0  because vT (m ) = u (m)  and u(m ) = m1 −ρ∕(1 − ρ)  is nice by inspection.

Now define an end-of-period value function 𝔳t(a)  as

            [ 1−ρ                  ]
𝔳t(a ) = β 𝔼t 𝒢t+1vt+1(ℛt+1a +  ξξξt+1 ) .

Since there is a positive probability that ξξξt+1   will attain its minimum of zero and since ℛt+1  > 0  , it is clear that lima ↓0𝔳t(a) = − ∞ and        ′
lima ↓0𝔳t(a) = ∞ . So 𝔳t(a)  is well-defined iff a > 0  ; it is similarly straightforward to show the other properties required for 𝔳t(a)  to be nice. (See Hiraguchi (2003).)

Next define v-(m, c)
  t  as

vt(m, c) = u(c) + 𝔳t(m  − c)
(54)

which is C3   since 𝔳t  and u  are both C3   , and note that our problem’s value function defined in (5) can be written as

vt(m ) = maxc  vt(m, c).
(55)

vt  is well-defined if and only if 0 < c < m  . Furthermore, limc↓0vt(m, c) = limc ↑m vt(m, c) = − ∞ , ∂2vt(m2,c)<  0
   ∂c  , limc ↓0 ∂vt(m,c)= + ∞
         ∂c , and limc↑m ∂vt(m,c)=  − ∞
         ∂c . It follows that the ct(m )  defined by

ct(m ) = argmax  vt(m, c)
          0<c<m
(56)

exists and is unique, and (5) has an internal solution that satisfies

  ′           ′
u (ct(m )) = 𝔳 t(m  − ct(m )).
(57)

Since both u  and 𝔳t  are strictly concave, both ct(m )  and at(m ) = m − ct(m )  are strictly increasing. Since both u  and 𝔳t  are three times continuously differentiable, using (57) we can conclude that c(m )
 t  is continuously differentiable and

 ′             𝔳′′t(at(m ))
ct(m ) = -′′-----------′′-------.
         u (ct(m )) + 𝔳t(at(m))
(58)

Similarly we can easily show that ct(m )  is twice continuously differentiable (as is at(m )  ) (See Appendix B.) This implies that vt(m )  is nice, since vt(m ) = u (ct(m )) + 𝔳t(at(m ))  .

B ct(m )  is Twice Continuously Differentiable

First we show that ct(m )  is C1   . Define y  as y ≡  m + dm  . Since u′(c (y)) − u′(c(m )) = 𝔳′(a(y)) − 𝔳′(a(m ))
    t           t        t  t       t  t  and at(y)−-at(m) = 1 − ct(y)−ct(m)
   dm              dm  ,

pict

Since ct  and at  are continuous and increasing,       u′(c (y))−u′(c(m))
  lim  ---tct(y)−ct(mt)---<  0
dm →+0  and       𝔳′(at(y))−𝔳′(at(m))
dmlim→+0-t-at(y)−att(m)--- < 0  are satisfied. Then u′(ct(y))− u′(ct(m))   𝔳′(at(y))−𝔳′(at(m ))
---ct(y)−-ct(m)---+  t-at(y)−att(m-)---< 0  for sufficiently small dm  . Hence we obtain a well-defined equation:

                          ′      ′
ct(y) − ct(m )           𝔳t(ata(yt()y)−)−-𝔳ta(t(amt()m))
------------- = -u′(ct(y))−u′(ct(m))---𝔳′(at(y))−𝔳′(at(m-)).
     dm          --ct(y)−ct(m)---+  -t-at(y)−att(m-)---

This implies that the right-derivative,  ′+
ct (m )  is well-defined and

               𝔳′′(at(m ))
c′+t (m ) =-′′----t------′′-------.
         u (ct(m )) + 𝔳 t(at(m ))

Similarly we can show that c′+(m ) = c′−(m )
 t        t  , which means c′(m )
 t  exists. Since 𝔳t  is C3   ,  ′
ct(m )  exists and is continuous.  ′
ct(m )  is differentiable because  ′′
𝔳t  is   1
C   , ct(m)  is   1
C   and  ′′           ′′
u (ct(m )) + 𝔳t (at(m )) < 0  .  ′′
ct(m )  is given by

 ′′       a′t(m )𝔳′t′′(at) [u ′′(ct) + 𝔳′′t(at)] − 𝔳′′t(at)[c′tu′′′(ct) + a′t𝔳′′t′ (at)]
ct(m ) =  ---------------------′′-------′′----2--------------------.
                            [u (ct) + 𝔳 t(at)]
(59)

Since  ′′
𝔳t(at(m ))  is continuous,  ′′
ct(m )  is also continuous.

C 𝒯  Is a Contraction Mapping

We must show that our operator 𝒯  satisfies all of Boyd’s conditions.

Boyd’s operator T  maps from 𝒞г(𝒜, ℬ )  to 𝒞 (𝒜,ℬ )  . A preliminary requirement is therefore that { 𝒯z} be continuous for any г − bounded z  , {𝒯z } ∈ 𝒞 (ℝ++,  ℝ)  . This is not difficult to show; see Hiraguchi (2003).

Consider condition (1). For this problem,

pict

so x(∙) ≤ y(∙)  implies {𝒯x }(mt) ≤ { 𝒯y}(mt )  by inspection.55

Condition (2) requires that {𝒯0 } ∈ 𝒞г(𝒜, ℬ )  . By definition,

                       { (  1−ρ )      }
{ 𝒯0}(m  ) =    max        ct---- + β0
        t    ct∈[κmt,¯κmt]    1 − ρ

the solution to which is patently u(¯κmt )  . Thus, condition (2) will hold if       1−ρ
(¯κmt )  is г  -bounded, which it is if we use the bounding function

              1−ρ
г(m ) = η + m    ,
(60)

defined in the main text.

Finally, we turn to condition (3), {𝒯(z + ζг )}(mt) ≤ {𝒯z }(mt ) + ζα г (mt )  . The proof will be more compact if we define ˘c  and ˘a  as the consumption and assets functions56 associated with 𝒯z  and ˆc  and ˆa  as the functions associated with 𝒯(z + ζг )  ; using this notation, condition (3) can be rewritten

pict

Now note that if we force the ⌣  consumer to consume the amount that is optimal for the ∧ consumer, value for the ⌣  consumer must decline (at least weakly). That is,

pict

Thus, condition (3) will certainly hold under the stronger condition

pict

where the last line follows because 0 < α < 1  by assumption.57

Using               1−ρ
г (m ) = η + m  and defining ˆat = ˆa(mt )  , this condition is

pict

which by imposing PF-FVAC (equation (19), which says ℶ < 1  ) can be rewritten as:

         [ 1−ρ               1−ρ]     1− ρ
η > β-𝔼t-𝒢-t+1-(ˆatℛt+1-+-ξξξt+1)----−--m-t--.
                    1 − ℶ
(61)

But since η  is an arbitrary constant that we can pick, the proof thus reduces to showing that the numerator of (61) is bounded from above:

             [                              ]         [               ]
 (1 − ℘ )β 𝔼t  𝒢1t−+ρ1(ˆatℛt+1 + 𝜃𝜃𝜃t+1∕(1 − ℘ ))1−ρ  + ℘β 𝔼t  𝒢1t−+ρ1(ˆatℛt+1 )1− ρ − m1t−ρ
             [ 1−ρ                               1−ρ]
≤(1 − ℘ )β 𝔼t  𝒢t+1((1 − ¯κ)mt ℛt+1 + 𝜃𝜃𝜃t+1∕(1 − ℘))
  + ℘ βR1−ρ((1 − ¯κ)mt )1− ρ − m1t−ρ
             [ 1−ρ                               1−ρ]
=(1 − ℘ )β 𝔼t  𝒢t+1((1 − ¯κ)mt ℛt+1 + 𝜃𝜃𝜃t+1∕(1 − ℘))
         (         (         1∕ρ)1 −ρ    )
      1−ρ(      1− ρ   1∕ρ(Rβ)----        )
  + m t    ℘ βR     ℘       R        −  1

                                                             (                )
             [                                      ]        |     (R β)1∕ρ    |
=(1 − ℘ )β 𝔼t  𝒢1t−+ρ1((1 − ¯κ)mt ℛt+1 + 𝜃𝜃𝜃t+1∕(1 − ℘))1−ρ +  m1−t ρ| ℘1∕ρ--------− 1|
                                                             ( ◟----◝◜R---◞   )
                                                                <1 by WRIC
             [ 1−ρ           1−ρ]            ρ 1−ρ
<(1 − ℘ )β 𝔼t  𝒢t+1(𝜃𝜃𝜃∕(1 − ℘ ))     = ℶ (1 − ℘) 𝜃𝜃𝜃-  .

We can thus conclude that equation (61) will certainly hold for any:

                  ρ 1− ρ
η >  η=  ℶ(1-−-℘-)𝜃𝜃𝜃----
     --      1 − ℶ
(62)

which is a positive finite number under our assumptions.

The proof that 𝒯  defines a contraction mapping under the conditions (31) and (28) is now complete.

C.1 𝒯  and v

In defining our operator 𝒯  we made the restriction κmt  ≤ ct ≤ ¯κmt  . However, in the discussion of the consumption function bounds, we showed only (in (32)) that κ-mt ≤  ct(mt ) ≤ ¯κtmt
  t  . (The difference is in the presence or absence of time subscripts on the MPC’s.) We have therefore not proven (yet) that the sequence of value functions (5) defines a contraction mapping.

Fortunately, the proof of that proposition is identical to the proof above, except that we must replace ¯κ  with ¯κT− 1   and the WRIC must be replaced by a slightly stronger (but still quite weak) condition. The place where these conditions have force is in the step at (62). Consideration of the prior two equations reveals that a sufficient stronger condition is

pict

where we have used (30) for ¯κ
 T− 1   (and in the second step the reversal of the inequality occurs because we have assumed ρ > 1  so that we are exponentiating both sides by the negative number 1 − ρ  ). To see that this is a weak condition, note that for small values of ℘  this expression can be further simplified using             − 1
(1 + ℘1∕ρÞÞÞR)   ≈  1 − ℘1 ∕ρÞÞÞR   so that it becomes

pict

Calling the weak return patience factor ÞÞÞ℘ = ℘1 ∕ρÞÞÞ
  R         R   and recalling that the WRIC was   ℘
ÞÞÞ R < 1  , the expression on the LHS above is    −ρ
βÞÞÞR   times the WRPF. Since we usually assume β  not far below 1 and parameter values such that ÞÞÞR  ≈ 1  , this condition is clearly not very different from the WRIC.

The upshot is that under these slightly stronger conditions the value functions for the original problem define a contraction mapping in г − bounded space with a unique v(m )  . But since limn →∞  κT−n = κ-  and limn →∞  ¯κT−n = ¯κ  , it must be the case that the v(m )  toward which these vT−n  ’s are converging is the same v(m )  that was the endpoint of the contraction defined by our operator 𝒯  . Thus, under our slightly stronger (but still quite weak) conditions, not only do the value functions defined by (5) converge, they converge to the same unique v  defined by 𝒯  .58

D Convergence in Euclidian Space

D.1 Convergence of vt

Boyd’s theorem shows that 𝒯  defines a contraction mapping in an г  -bounded space. We now show that 𝒯  also defines a contraction mapping in Euclidian space.

Calling  ∗
v the unique fixed point of the operator 𝒯  , since  ∗          ∗
v (m ) = 𝒯v (m )  ,

∥vT− n+1 − v∗∥г ≤ αn −1∥vT −  v∗∥г .
(63)

On the other hand, vT − v∗ ∈ 𝒞г (𝒜,ℬ )  and κ =  ∥vT − v∗∥г <  ∞ because vT  and v ∗ are in 𝒞г (𝒜,ℬ )  . It follows that

               ∗          n−1
|vT −n+1(m ) − v (m )| ≤ κ α  |г (m )|.
(64)

Then we obtain

                   ∗
nl→im∞vT −n+1(m ) = v (m ).
(65)

Since           m1−ρ-
vT (m ) = 1−ρ  ,            (¯κm)1−ρ
vT−1(m ) ≤   1−ρ  <  vT(m )  . On the other hand, vT −1 ≤ vT  means 𝒯vT −1 ≤ 𝒯vT  , in other words, vT−2(m ) ≤ vT−1(m )  . Inductively one gets vT −n(m ) ≥ vT−n− 1(m )  . This means that {vT −n+1(m )}∞n=1   is a decreasing sequence, bounded below by v∗ .

D.2 Convergence of ct

Given the proof that the value functions converge, we now show the pointwise convergence of consumption functions              ∞
{cT −n+1(m )}n=1   .

Consider any convergent subsequence {           }
  cT−n(i)(m ) of {cT− n+1(m )} ∞n=1   converging to c∗ . By the definition of c   (m )
 T−n  , we have

                         1−ρ
u(cT−n(i)(m ))+ β 𝔼T −n(i)[𝒢T− n(i)+1vT −n(i)+1(m )]
                                        1−ρ
               ≥  u(cT−n(i)) + β 𝔼T −n(i)[𝒢T− n(i)+1vT−n(i)+1(m )],
(66)

for any cT−n(i) ∈ [κm, ¯κm ]  . Now letting n(i)  go to infinity, it follows that the left hand side converges to u(c∗) + β 𝔼t[𝒢1t−ρv (m )]  , and the right hand side converges to u(c      ) + β 𝔼 [𝒢1−ρv (m )]
   T− n(i)       t t  . So the limit of the preceding inequality as n (i)  approaches infinity implies

u (c∗) + β 𝔼 [𝒢1−ρv(m )] ≥ u(c     ) + β 𝔼 [𝒢1− ρv(m )].
           t  t+1             T−n(i)       t  t+1
(67)

Hence,                  {                           }
c∗ ∈    argmax     u (cT−n(i)) + β 𝔼t[𝒢1− ρv(m )]
    cT−n(i)∈[κm,¯κm]                   t+1 . By the uniqueness of c(m )  , c∗ = c(m )  .

E Perfect Foresight Liquidity Constrained Solution

Under perfect foresight in the presence of a liquidity constraint requiring b ≥ 0  , this appendix taxonomizes the varieties of the limiting consumption function `c(m )  that arise under various parametric conditions.

Table 5:Appendix: Perfect Foresight Liquidity Constrained Taxonomy
|--------For-constrained-`c and-unconstrained-¯c-consumption--functions---------|
|Main  Condition  |                      |                                    |
|   Subcondition  |        Math          |  Outcome,  Comments    or Results  |
|-----------------|----------------------|------------------------------------|
|¬GIC             |          1 <    ÞÞÞ ∕𝒢 | Constraint never binds for m  ≥ 1   |
|    and RIC      |ÞÞÞ ∕R    < 1           |   FHWC    holds (R >  𝒢);           |
|                 |                      |   `c(m ) = ¯c(m ) for m ≥ 1          |
|    and ¬RIC     |          1 <    ÞÞÞ ∕R |   `c(m ) is degenerate: `c(m ) = 0   |
|GIC              |ÞÞÞ ∕𝒢    < 1           | Constraint binds in finite time  ∀ m  |
|                 |                      |                                    |
|    and RIC      |ÞÞÞ ∕R    < 1           |    FHWC   may  or may  not hold    |
|                 |                      |      limm ↑∞ ¯c(m ) − `c(m ) = 0      |
|                 |                      |      limm ↑∞ `κκκ(m ) = κ-             |
|    and ¬RIC     |          1 <    ÞÞÞ ∕R | ¬FHWC                              |
|                 |                      |                                    |
----------------------------------------------limm-↑∞-`κκκ(m-) =-0----------------

Conditions are applied from left to right; for example, the second row indicates conclusions in the case where ¬ GIC and RIC both hold, while the third row indicates that when the GIC and the RIC both fail, the consumption function is degenerate; the next row indicates that whenever the GICholds, the constraint will bind in finite time.

Results are summarized in table 5.

E.1 If GIC Fails

A consumer is ‘growth patient’ if the perfect foresight growth impatience condition fails (¬ GIC, 1 < ÞÞÞ∕𝒢 ). Under ¬ GIC the constraint does not bind at the lowest feasible value of mt  = 1  because 1 < (Rβ )1∕ρ∕𝒢 implies that spending everything today (setting ct = mt = 1  ) produces lower marginal utility than is obtainable by reallocating a marginal unit of resources to the next period at return R  :59

    1 < (Rβ )1∕ρ𝒢 −1
            − ρ
    1 < Rβ 𝒢
u ′(1) < Rβu ′(𝒢 ).

Similar logic shows that under these circumstances the constraint will never bind at m  = 1  for a constrained consumer with a finite horizon of n  periods, so for m  ≥ 1  such a consumer’s consumption function will be the same as for the unconstrained case examined in the main text.

RIC fails, FHWC holds. If the RIC fails (1 < ÞÞÞR   ) while the finite human wealth condition holds, the limiting value of this consumption function as n ↑ ∞ is the degenerate function

`cT −n(m ) = 0(bt + h).
(68)

(that is, consumption is zero for any level of human or nonhuman wealth).

RIC fails, FHWC fails. ¬ FHWC implies that human wealth limits to h = ∞ so the consumption function limits to either `cT −n(m ) = 0  or `cT−n(m ) = ∞ depending on the relative speeds with which the MPC approaches zero and human wealth approaches ∞ .60

Thus, the requirement that the consumption function be nondegenerate implies that for a consumer satisfying ¬ GIC we must impose the RIC (and the FHWC can be shown to be a consequence of ¬ GIC and RIC). In this case, the consumer’s optimal behavior is easy to describe. We can calculate the point at which the unconstrained consumer would choose c = m  from Equation (16):

       m#  = (m#  − 1 + h)κ-
m  (1 − κ) = (h − 1)κ
  #     --          (-      )
                      --κ---
       m#  = (h − 1)  1 − κ-
(69)

which (under these assumptions) satisfies 0 < m#  <  1  .61 For m  < m#   the unconstrained consumer would choose to consume more than m  ; for such m  , the constrained consumer is obliged to choose `c(m ) = m  .62 For any m  > m#   the constraint will never bind and the consumer will choose to spend the same amount as the unconstrained consumer, ¯c(m )  .

(Stachurski and Toda (2019) obtain a similar lower bound on consumption and use it to study the tail behavior of the wealth distribution.)

E.2 If GIC Holds

Imposition of the GIC reverses the inequality in (68), and thus reverses the conclusion: A consumer who starts with m  =  1
  t  will desire to consume more than 1. Such a consumer will be constrained, not only in period t  , but perpetually thereafter.

Now define bn#   as the bt  such that an unconstrained consumer holding bt = bn#   would behave so as to arrive in period t + n  with bt+n = 0  (with b0#   trivially equal to 0); for example, a consumer with bt−1 = b1
        #   was on the ‘cusp’ of being constrained in period t − 1  : Had bt−1   been infinitesimally smaller, the constraint would have been binding (because the consumer would have desired, but been unable, to enter period t  with negative, not zero, b  ). Given the GIC, the constraint certainly binds in period t  (and thereafter) with resources of mt = m0# =  1 + b0#  = 1  : The consumer cannot spend more (because constrained), and will not choose to spend less (because impatient), than       0
ct = c# = 1  .

We can construct the entire ‘prehistory’ of this consumer leading up to t  as follows. Maintaining the assumption that the constraint has never bound in the past, c  must have been growing according to ÞÞÞ𝒢 , so consumption n  periods in the past must have been

 n     − n      −n
c# =  ÞÞÞ𝒢  ct = ÞÞÞ 𝒢 .
(70)

The PDV of consumption from t − n  until t  can thus be computed as

ℂtt−n = ct−n(1 + ÞÞÞ∕R + ⋅⋅ ⋅ + (ÞÞÞ∕R )n)
         n                  n
      = c#(1(+ ÞÞÞR +  ⋅⋅⋅) + ÞÞÞ R)
          −n  1 − ÞÞÞn+R1
      = ÞÞÞ 𝒢   ---------
        (      1 − ÞÞÞR)
          ÞÞÞ−𝒢-n−-ÞÞÞR-
      =    1 − ÞÞÞ
                 R

and note that the consumer’s human wealth between t − n  and t  (the relevant time horizon, because from t  onward the consumer will be constrained and unable to access post-t  income) is

hn# =  1 + ⋅⋅⋅ + ℛ −n
(71)

while the intertemporal budget constraint says

  t       n     n
ℂ t− n  = b# + h #

from which we can solve for the  n
b#   such that the consumer with         n
bt− n = b#   would unconstrainedly plan (in period t − n  ) to arrive in period t  with bt = 0  :
                   hn#
             ◜(-----◞◟ ----)◝
 n     t       1 − ℛ −(n+1 )
b# = ℂ t−n −   -1-−-ℛ-−-1--  .
(72)

Defining   n     n
m # =  b# + 1  , consider the function `c(m )  defined by linearly connecting the points {mn#, cn# } for integer values of n ≥  0  (and setting `c(m ) = m  for m <  1  ). This function will return, for any value of m  , the optimal value of c  for a liquidity constrained consumer with an infinite horizon. The function is piecewise linear with ‘kink points’ where the slope discretely changes; for infinitesimal 𝜖  the MPC of a consumer with assets        n
m  = m # − 𝜖  is discretely higher than for a consumer with assets        n
m  = m # + 𝜖  because the latter consumer will spread a marginal dollar over more periods before exhausting it.

In order for a unique consumption function to be defined by this sequence (72) for the entire domain of positive real values of b  , we need  n
b#   to become arbitrarily large with n  . That is, we need

      n
nli→m∞ b# = ∞.
(73)

E.2.1 If FHWC Holds

The FHWC requires ℛ − 1 < 1  , in which case the second term in (72) limits to a constant as n ↑ ∞ , and (73) reduces to a requirement that

     ( Þ −n   Þ   Þ   nÞ  )
 lim    ÞÞ-𝒢-−--(ÞÞR∕ÞÞ-𝒢)-ÞÞR-    = ∞
n→ ∞         1 − ÞÞÞR
          ( ÞÞÞ −n−  ℛ −nÞÞÞ  )
      lim    --𝒢----------R    = ∞
     n→ ∞       1 − ÞÞÞR
                 (  ÞÞÞ −𝒢n  )
             lim    -------   =  ∞.
             n→∞   1 − ÞÞÞR

Given the GIC   −1
ÞÞÞ 𝒢  > 1  , this will hold iff the RIC holds, ÞÞÞR < 1  . But given that the FHWC R >  𝒢 holds, the GIC is stronger (harder to satisfy) than the RIC; thus, the FHWC and the GIC together imply the RIC, and so a well-defined solution exists. Furthermore, in the limit as n  approaches infinity, the difference between the limiting constrained consumption function and the unconstrained consumption function becomes vanishingly small, because the date at which the constraint binds becomes arbitrarily distant, so the effect of that constraint on current behavior shrinks to nothing. That is,
lmi→m∞ `c(m ) − ¯c(m ) = 0.
(74)

E.2.2 If FHWC Fails

If the FHWC fails, matters are a bit more complex.

Given failure of FHWC, (73) requires

        (   − n       −n)    (      − (n+1))
    lim   ℛ----ÞÞÞR-−-ÞÞÞ-𝒢-  +    1 −-ℛ-------  = ∞
    n→∞       ÞÞÞR − 1            ℛ − 1 − 1
    (               −1  )         (   − n )
lim   --ÞÞÞR--- − --ℛ------ ℛ − n −  --ÞÞÞ𝒢---   = ∞
n→∞   ÞÞÞR  − 1   ℛ −1 − 1           ÞÞÞR  − 1

If RIC Holds. When the RIC holds, rearranging (75) gives

     (  ÞÞÞ −n  )        (   Þ          − 1  )
 lim    ---𝒢---  − ℛ −n   -ÞÞR---+ --ℛ------   =  ∞
n→ ∞   1 − ÞÞÞR            1 − ÞÞÞR  ℛ − 1 − 1

and for this to be true we need
 ÞÞÞ− 1 >  ℛ −1
  𝒢
𝒢 ∕ÞÞÞ  >  𝒢∕R
   1  >  ÞÞÞ∕R

which is merely the RIC again. So the problem has a solution if the RIC holds. Indeed, we can even calculate the limiting MPC from
               (    )
                 cn
lim κn# =  lim    -#n-
n→∞       n→ ∞   b#
(75)

which with a bit of algebra63 can be shown to asymptote to the MPC in the perfect foresight model:64

 lim  κκ`κ(m ) = 1 − ÞÞÞR.
m → ∞
(77)

If RIC Fails. Consider now the ¬ RIC case, ÞÞÞ  > 1
 R  . We can rearrange (75)as

     (         −1                −1           )        (    − n )
 lim    --ÞÞÞR-(ℛ----−-1)---−  ---ℛ---(ÞÞÞR-−-1)---- ℛ − n −  --ÞÞÞ𝒢---    = ∞.    (78)
n→ ∞   (ℛ− 1 − 1 )(ÞÞÞR − 1)   (ℛ −1 − 1)(ÞÞÞR − 1)           ÞÞÞR  − 1

which makes clear that with ¬FHWC    ⇒  ℛ −1 > 1  and ¬RIC  ⇒  ÞÞÞR > 1  the numerators and denominators of both terms multiplying ℛ − n  can be seen transparently to be positive. So, the terms multiplying   −n
ℛ  in (75) will be positive if
ÞÞÞRℛ − 1 − ÞÞÞR  >   ℛ −1ÞÞÞR − ℛ −1
          −1
        ℛ     >   ÞÞÞR
           𝒢  >   ÞÞÞ

which is merely the GIC which we are maintaining. So the first term’s limit is + ∞ . The combined limit will be + ∞ if the term involving ℛ −n  goes to + ∞ faster than the term involving −  ÞÞÞ−n
    𝒢 goes to − ∞ ; that is, if
  −1       −1
ℛ     >  ÞÞÞ 𝒢
𝒢∕R   >  𝒢 ∕ÞÞÞ

ÞÞÞ∕R   >  1

which merely confirms the starting assumption that the RIC fails.

What is happening here is that the cn#   term is increasing backward in time at rate dominated in the limit by 𝒢 ∕ÞÞÞ  while the b#   term is increasing at a rate dominated by 𝒢 ∕R  term and

𝒢∕R   >  𝒢 ∕ÞÞÞ                                (79)

because ¬RIC  ⇒  ÞÞÞ > R  .

Consequently, while limn ↑∞ bn# =  ∞ , the limit of the ratio cn#∕bn#   in (75) is zero. Thus, surprisingly, the problem has a well defined solution with infinite human wealth if the RIC fails. It remains true that ¬ RIC implies a limiting MPC of zero,

 lim  `κκκ (m ) = 0,
m →∞
(80)

but that limit is approached gradually, starting from a positive value, and consequently the consumption function is not the degenerate `c(m ) = 0  . (Figure 8 presents an example for ρ = 2  , R =  0.98  , β = 1.00  , 𝒢 = 0.99  ; note that the horizontal axis is bank balances b = m  − 1  ; the part of the consumption function below the depicted points is uninteresting — c = m  — so not worth plotting).

PIC
Figure 8:Appendix: Nondegenerate c  Function with ¬ FHWC and ¬ RIC

We can summarize as follows. Given that the GIC holds, the interesting question is whether the FHWC holds. If so, the RIC automatically holds, and the solution limits into the solution to the unconstrained problem as m  ↑ ∞ . But even if the FHWC fails, the problem has a well-defined and nondegenerate solution, whether or not the RIC holds.

Although these results were derived for the perfect foresight case, we know from work elsewhere in this paper and in other places that the perfect foresight case is an upper bound for the case with uncertainty. If the upper bound of the MPC in the perfect foresight case is zero, it is not possible for the upper bound in the model with uncertainty to be greater than zero, because for any κ >  0  the level of consumption in the model with uncertainty would eventually exceed the level of consumption in the absence of uncertainty.

Ma and Toda (2020) characterize the limits of the MPC in a more general framework that allows for capital and labor income risks in a Markovian setting with liquidity constraints, and find that in that much more general framework the limiting MPC is also zero.

F The Perfect Foresight Liquidity Constrained Solution as a Limit

Formally, suppose we change the description of the problem by making the following two assumptions:

℘    = 0
ct  ≤ mt,

and we designate the solution to this consumer’s problem `ct(m )  . We will henceforth refer to this as the problem of the ‘restrained’ consumer (and, to avoid a common confusion, we will refer to the consumer as ‘constrained’ only in circumstances when the constraint is actually binding).

Redesignate the consumption function that emerges from our original problem for a given fixed ℘  as ct(m; ℘ )  where we separate the arguments by a semicolon to distinguish between m  , which is a state variable, and ℘  , which is not. The proposition we wish to demonstrate is

lim c (m; ℘ ) = `c(m ).
 ℘↓0 t          t
(81)

We will first examine the problem in period T − 1  , then argue that the desired result propagates to earlier periods. For simplicity, suppose that the interest, growth, and time-preference factors are β =  R = 𝒢 = 1  , and there are no permanent shocks, Ψ  = 1  ; the results below are easily generalized to the full-fledged version of the problem.

The solution to the restrained consumer’s optimization problem can be obtained as follows. Assuming that the consumer’s behavior in period T  is given by cT(m )  (in practice, this will be cT (m ) = m  ), consider the unrestrained optimization problem

                   {            ∫  ¯             }
 ∗                                 𝜃𝜃𝜃
`aT−1(m ) = argmaax    u(m − a ) +    vT(a + 𝜃𝜃𝜃)dℱ 𝜃𝜃𝜃  .
                                 𝜃𝜃𝜃
(82)

As usual, the envelope theorem tells us that v′T (m ) = u′(cT(m ))  so the expected marginal value of ending period T − 1  with assets a  can be defined as

           ∫ ¯𝜃𝜃𝜃
`𝔳′   (a) ≡    u′(cT(a + 𝜃𝜃𝜃))dℱ𝜃𝜃𝜃,
 T− 1       𝜃𝜃𝜃

and the solution to (82) will satisfy

u ′(m −  a) = `𝔳′  (a).
              T−1
(83)

`a ∗  (m )
  T−1  therefore answers the question “With what level of assets would the restrained consumer like to end period T − 1  if the constraint cT−1 ≤ mT  −1   did not exist?” (Note that the restrained consumer’s income process remains different from the process for the unrestrained consumer so long as ℘ >  0  .) The restrained consumer’s actual asset position will be

`aT−1(m ) = max [0,`a∗T−1(m )],

reflecting the inability of the restrained consumer to spend more than current resources, and note (as pointed out by Deaton (1991)) that

m1  = (`𝔳′   (0))−1∕ρ
  #     T −1

is the cusp value of m  at which the constraint makes the transition between binding and non-binding in period T −  1  .

Analogously to (83), defining

             [                                           ]
 ′              − ρ          ∫ ¯𝜃𝜃𝜃                   − ρ
𝔳T−1(a;℘ ) ≡  ℘a   + (1 − ℘ )   (cT (a + 𝜃𝜃𝜃∕ (1 − ℘ )))   dℱ𝜃𝜃𝜃  ,
                              𝜃𝜃𝜃
(84)

the Euler equation for the original consumer’s problem implies

        −ρ    ′
(m  − a)   = 𝔳T− 1(a; ℘)
(85)

with solution a∗  (m; ℘)
 T−1  . Now note that for any fixed a > 0  , lim℘ ↓0 𝔳′  (a;℘ ) = `𝔳 ′ (a)
        T−1          T−1  . Since the LHS of (83) and (85) are identical, this means that lim    a∗  (m; ℘ ) = `a ∗ (m )
   ℘↓0 T−1           T−1  . That is, for any fixed value of        1
m >  m #   such that the consumer subject to the restraint would voluntarily choose to end the period with positive assets, the level of end-of-period assets for the unrestrained consumer approaches the level for the restrained consumer as ℘ ↓ 0  . With the same a  and the same m  , the consumers must have the same c  , so the consumption functions are identical in the limit.

Now consider values        1
m  ≤ m #   for which the restrained consumer is constrained. It is obvious that the baseline consumer will never choose a ≤  0  because the first term in (84) is lima ↓0 ℘a− ρ = ∞ , while lima ↓0(m −  a)−ρ  is finite (the marginal value of end-of-period assets approaches infinity as assets approach zero, but the marginal utility of consumption has a finite limit for m > 0  ). The subtler question is whether it is possible to rule out strictly positive a  for the unrestrained consumer.

The answer is yes. Suppose, for some m <  m1#   , that the unrestrained consumer is considering ending the period with any positive amount of assets a = δ > 0  . For any such δ  we have that lim    𝔳′  (a;℘ ) = `𝔳′  (a)
   ℘↓0 T− 1         T−1  . But by assumption we are considering a set of circumstances in which  ∗
`aT −1(m ) < 0  , and we showed earlier that        ∗             ∗
lim ℘↓0aT−1(m; ℘ ) = `a T−1(m )  . So, having assumed a =  δ > 0  , we have proven that the consumer would optimally choose a < 0  , which is a contradiction. A similar argument holds for m =  m1#   .

These arguments demonstrate that for any m  > 0  , lim    c   (m; ℘) = `c    (m )
   ℘↓0 T−1          T− 1  which is the period T − 1  version of (81). But given equality of the period T − 1  consumption functions, backwards recursion of the same arguments demonstrates that the limiting consumption functions in previous periods are also identical to the constrained function.

Note finally that another intuitive confirmation of the equivalence between the two problems is that our formula (87) for the maximal marginal propensity to consume satisfies

lim  ¯κ  = 1,
℘↓0

which makes sense because the marginal propensity to consume for a constrained restrained consumer is 1 by our definitions of ‘constrained’ and ‘restrained.’

G The Limiting MPC’s

For mt  > 0  we can define et(mt) = ct(mt)∕mt  and at(mt ) = mt − ct(mt)  and the Euler equation (6) can be rewritten

                 ⌊(            (                   ) ) − ρ⌋
                                 ◜----=mt+◞1◟𝒢t+1----◝
                 ||            | Ra (m  ) + 𝒢  ξξξ   | |    |
et(mt )−ρ = βR 𝔼t |||| et+1(mt+1) || --t---t-----t+1-t+1|| ||    ||
                 ⌈(            (         mt        ) )    ⌉

                        ρ   [                     −ρ         ]
        =    (1 − ℘ )βRm t 𝔼t (et+1(mt+1)mt+1 𝒢t+1)  | ξξξt+1 > 0
                     [(                             ) −ρ         ]
        +  ℘βR1 −ρ𝔼t    et+1(ℛt+1at (mt))mt-−--ct(mt-)    | ξξξt+1 = 0 .
                                             mt

Consider the first conditional expectation in (6), recalling that if ξξξt+1 > 0  then ξξξt+1 ≡ 𝜃𝜃𝜃t+1∕(1 − ℘)  . Since limm ↓0at(m ) = 0  , 𝔼t[(et+1(mt+1 )mt+1𝒢t+1)− ρ | ξξξt+1 > 0 ]  is contained within bounds defined by (e  (𝜃𝜃𝜃∕(1 − ℘ ))𝒢 Ψ 𝜃𝜃𝜃∕(1 − ℘))−ρ
  t+1 --          ----  and       ¯           ¯¯          −ρ
(et+1(𝜃𝜃𝜃∕(1 − ℘))𝒢Ψ 𝜃𝜃𝜃 ∕(1 − ℘ ))  both of which are finite numbers, implying that the whole term multiplied by (1 − ℘)  goes to zero as   ρ
m t  goes to zero. As mt ↓ 0  the expectation in the other term goes to ¯κ−tρ+1(1 − ¯κt)− ρ  . (This follows from the strict concavity and differentiability of the consumption function.) It follows that the limiting ¯κt satisfies  −ρ       1− ρ −ρ        −ρ
¯κt  = β ℘R   κ¯t+1 (1 − ¯κt)  . Exponentiating by ρ  , we can conclude that

               κ¯ = ℘ −1∕ρ(βR)−1∕ρR(1 − ¯κ )¯κ
                 t                        t t+1
    ◜----ÞÞÞ◞R◟---◝
 1∕ρ  −1     1∕ρ
℘◟---R-◝ (◜βR-)-◞κ¯t = (1 − ¯κt)¯κt+1
    ≡ ℘1∕ρÞÞÞR

which yields a useful recursive formula for the maximal marginal propensity to consume:

(℘1∕ρÞÞÞR ¯κt)−1 = (1 − ¯κt)− 1¯κ−1
   −1            1∕ρ    −1 t+1
  ¯κt (1 − ¯κt) = ℘   ÞÞÞR¯κ t+1
           −1        1∕ρÞ   −1
         ¯κ t =  1 + ℘  ÞÞR¯κ t+1.

As noted in the main text, we need the WRIC (31) for this to be a convergent sequence:

0 ≤ ℘1∕ρÞÞÞR <  1,
(86)

Since ¯κT =  1  , iterating (86) backward to infinity (because we are interested in the limiting consumption function) we obtain:

                      1∕ρ
lni→m∞ ¯κT−n = ¯κ ≡  1 − ℘  ÞÞÞR
(87)

and we will therefore call ¯κ  the ‘limiting maximal MPC.’

The minimal MPC’s are obtained by considering the case where mt  ↑ ∞ . If the FHWC holds, then as mt ↑ ∞ the proportion of current and future consumption that will be financed out of capital approaches 1. Thus, the terms involving ξξξt+1   in (86) can be neglected, leading to a revised limiting Euler equation

                       [                              ]
(mtet(mt ))−ρ  = βR 𝔼t  (et+1 (at(mt )ℛt+1) (Rat (mt)))−ρ

and using L’Hôpital’s rule limmt → ∞ et(mt) = κt  , and limmt →∞  et+1(at(mt)ℛt+1 ) = κt+1   so a further limit of the Euler equation is
         −ρ        (                )−ρ
  (mt κt)    =  βR  κt+1R (1 − κt)mt
  R◟− ◝1◜ÞÞÞ◞  κt  =  (1 − κt)κt+1
≡ÞÞÞR=(1−κ)

and the same sequence of derivations used above yields the conclusion that if the RIC 0 ≤ ÞÞÞR < 1  holds, then a recursive formula for the minimal marginal propensity to consume is given by
 − 1        −1Þ
κt  =  1 + κ-t+1ÞÞR
(88)

so that ({κ−1  })∞
  -T− n  n=0   is also an increasing convergent sequence, and we define

κ− 1 ≡ lim κ −T1−n
      n↑∞
(89)

as the limiting (inverse) marginal MPC. If the RIC does not hold, then limn →∞ κ-−T1−n = ∞ and so the limiting MPC is κ-= 0  .

For the purpose of constructing the limiting perfect foresight consumption function, it is useful further to note that the PDV of consumption is given by

  (           2      )       −1
ct◟1-+-ÞÞÞR-+◝ÞÞÞ◜-R-+-⋅⋅⋅◞  = ctκT− n.
    =1+ ÞÞÞ (1+ÞÞÞ κ−1 )...
         R   R-t+2

which, combined with the intertemporal budget constraint, yields the usual formula for the perfect foresight consumption function:
ct = (bt + ht)κt
(90)

H Unique and Stable Target and Steady State Points

This appendix proves Theorems 4-5 and:

Lemma 1. If mˇ  and mˆ  both exist, then mˇ ≤ mˆ  .

H.1 Proof of Theorem 4

Theorem 4. For the nondegenerate solution to the problem defined in Section 2.1 when FVAC, WRIC, and GIC-Mod all hold, there exists a unique cash-on-hand-to-permanent-income ratio mˆ > 0  such that

𝔼t[mt+1 ∕mt] = 1 if mt = ˆm.
(91)

Moreover, ˆm  is a point of ‘stability’ in the sense that

 ∀m  ∈  (0, ˆm), 𝔼 [m    ] > m
    t            t  t+1      t
∀mt ∈ (mˆ, ∞ ), 𝔼t[mt+1 ] < mt.
(92)

The elements of the proof of Theorem 4 are:

H.2 Existence and Continuity of 𝔼t[mt+1 ∕mt ]

The consumption function exists because we have imposed sufficient conditions (the WRIC  and FVAC  ; Theorem 1).

Section 2.8 shows that for all t  , at− 1 = mt−1 − ct−1 > 0  . Since mt  = at−1ℛt + ξξξt  , even if ξξξ
 t takes on its minimum value of 0, a   ℛ  >  0
 t−1  t  , since both a
 t−1   and ℛ
  t  are strictly positive. With mt  and mt+1   both strictly positive, the ratio 𝔼t[mt+1 ∕mt ]  inherits continuity (and, for that matter, continuous differentiability) from the consumption function.

H.3 Existence of a point where 𝔼t[mt+1∕mt ] = 1  .

This follows from:

  1. Existence and continuity of 𝔼t[mt+1∕mt ]  (just proven)
  2. Existence a point where 𝔼  [m    ∕m  ] < 1
  t  t+1   t
  3. Existence a point where 𝔼t [mt+1 ∕mt ] > 1
  4. The Intermediate Value Theorem

H.3.1 Existence of m  where 𝔼t[mt+1 ∕mt] < 1

If RIC holds. Logic exactly parallel to that of Section 3.1 leading to equation (38), but dropping the 𝒢t+1   from the RHS, establishes that

                           [                        ]
                            ℛt+1-(mt-−-c(mt-)) +-ξξξt+1
lmitm↑∞ 𝔼t [mt+1 ∕mt ] = mltim↑∞ 𝔼t            m
                                         t
                 = 𝔼t [(R ∕𝒢t+1)ÞÞÞR]
                 = 𝔼t [ÞÞÞ ∕𝒢t+1]

                 < 1

where the inequality reflects imposition of the GIC-Mod (25).

If RIC fails. When the RIC fails, the fact that          ′
limm ↑∞ c(m ) = 0  (see equation (29)) means that the limit of the RHS of (93) as m  ↑ ∞ is ℛ¯ = 𝔼t [ℛt+1 ]  . In the next step of this proof, we will prove that the combination GIC-Mod and ¬ RIC implies ℛ¯ < 1  .

So we have limm ↑∞ 𝔼t[mt+1∕mt ] < 1  whether the RIC holds or fails.

H.3.2 Existence of m  > 1 where 𝔼  [m    ∕m  ] > 1
  t  t+1   t

Paralleling the logic for c  in Section 3.2: the ratio of 𝔼t[mt+1 ]  to mt  is unbounded above as mt ↓ 0  because limmt ↓0𝔼t [mt+1 ] > 0  .

Intermediate Value Theorem. If 𝔼t[mt+1 ∕mt]  is continuous, and takes on values above and below 1, there must be at least one point at which it is equal to one.

H.3.3 𝔼t[mt+1] − mt  is monotonically decreasing.

Now define ζζζ(m ) ≡ 𝔼 [m    ] − m
     t     t  t+1      t  and note that

ζζζ(mt) < 0 ↔  𝔼t[mt+1∕mt ] < 1
ζζζ(m ) = 0 ↔  𝔼 [m   ∕m  ] = 1
    t         t   t+1   t
ζζζ(mt) > 0 ↔  𝔼t[mt+1∕mt ] > 1,

so that ζζζ(ˆm ) = 0  . Our goal is to prove that ζζζ(∙)  is strictly decreasing on (0,∞  )  using the fact that

          (     )            [(     )                                ]
  ′         -d--                -d--  (                             )
ζζζ(mt ) ≡   dm    ζζζ(mt) = 𝔼t    dm     ℛt+1 (mt − c(mt )) + ξξξt+1 − mt
               t                  ′t
                        = ℛ¯ (1 − c (mt)) − 1.

Now, we show that (given our other assumptions)  ′
ζζζ (m )  is decreasing (but for different reasons) whether the RIC holds or fails.

If RIC holds. Equation (15) indicates that if the RIC holds, then κ-> 0  . We show at the bottom of Section 2.9.1 that if the RIC holds then 0 < κ-< c′(mt) < 1  so that

pict

which is negative because the GIC-Mod says ÞÞÞ 𝒢 < 1  .

If RIC fails. Under ¬ RIC, recall that limm ↑∞ c′(m ) = 0  . Concavity of the consumption function means that c′ is a decreasing function, so everywhere

pict

which means that   ′
ζζζ(mt )  from (93) is guaranteed to be negative if

        [    ]
 ¯       -R--
ℛ  ≡ 𝔼t  𝒢 Ψ   < 1.
(93)

But the combination of the GIC-Mod holding and the RIC failing can be written:

pict

and multiplying all three elements by R∕ÞÞÞ  gives

pict

which satisfies our requirement in (93).

H.4 Proof of Theorem 5

Theorem 5. For the nondegenerate solution to the problem defined in Section 2.1 when FVAC, WRIC, and GIC all hold, there exists a unique pseudo-steady-state cash-on-hand-to-income ratio ˇm > 0  such that

𝔼 [Ψ    m    ∕m  ] = 1 if m = ˇm.
 t  t+1  t+1   t         t
(94)

Moreover, ˇm  is a point of stability in the sense that

 ∀mt  ∈ (0, ˇm ), 𝔼t[mt+1 ]∕mt > 𝒢

∀mt  ∈ (ˇm, ∞ ), 𝔼t[mt+1 ]∕mt  < 𝒢.
(95)

The elements of the proof are:

H.4.1 Existence and Continuity of the Ratio

Since by assumption                  ¯
0 < Ψ--≤ Ψt+1 ≤  Ψ  < ∞ , our proof in H.2 that demonstrated existence and continuity of 𝔼t[mt+1 ∕mt]  implies existence and continuity of 𝔼t[Ψt+1mt+1 ∕mt ]  .

H.4.2 Existence of a stable point

Since by assumption 0 < Ψ--≤ Ψt+1 ≤  ¯Ψ  < ∞ , our proof in Subsection H.2 that the ratio of 𝔼t[mt+1 ]  to mt  is unbounded as mt ↓ 0  implies that the ratio 𝔼t [Ψt+1mt+1  ]  to mt  is unbounded as mt ↓ 0  .

The limit of the expected ratio as mt  goes to infinity is most easily calculated by modifying the steps for the prior theorem explicitly:

                                [     (                     )   ]
 lim  𝔼 [Ψ   m    ∕m  ] = lim  𝔼    𝒢t+1--(R∕𝒢t+1-)a-(mt-) +-ξξξt+1--∕𝒢--
mt↑∞  t   t+1  t+1   t   mt ↑∞   t               mt
                                [                       ]
                      =  lim  𝔼   (R-∕𝒢)a(mt-) +-Ψt+1-ξξξt+1
                        mt ↑∞   t           mt
                             [ (R∕𝒢 )a(m ) + 1]
                      =  lim    ----------t-----
                        mt ↑∞         mt
                      = (R ∕𝒢)ÞÞÞR

                      = ÞÞÞ 𝒢
                      < 1

where the last two lines are merely a restatement of the GIC (18).

The Intermediate Value Theorem says that if 𝔼t[Ψt+1mt+1  ∕mt ]  is continuous, and takes on values above and below 1, there must be at least one point at which it is equal to one.

H.4.3 𝔼t[Ψt+1mt+1 ] − mt  is monotonically decreasing.

Define ζζζ(mt ) ≡ 𝔼t[Ψt+1mt+1  ] − mt  and note that

ζζζ(mt ) < 0 ↔ 𝔼t [Ψt+1mt+1  ∕mt] < 1

ζζζ(mt ) = 0 ↔ 𝔼t [Ψt+1mt+1  ∕mt] = 1
ζζζ(mt ) > 0 ↔ 𝔼t [Ψt+1mt+1  ∕mt] > 1,

so that ζζζ(ˆm ) = 0  . Our goal is to prove that ζζζ(∙)  is strictly decreasing on (0,∞  )  using the fact that

          (     )            [(     )                                  ]
  ′         -d--                -d--  (                               )
ζζζ(mt ) ≡   dm    ζζζ(mt) = 𝔼t    dm     ℛ (mt − c(mt )) + Ψt+1 ξξξt+1 − mt
               t                   t  ′
                        = (R ∕𝒢)(1 − c (mt)) − 1.

Now, we show that (given our other assumptions)  ′
ζζζ (m )  is decreasing (but for different reasons) whether the RIC holds or fails (¬ RIC).

If RIC holds. Equation (15) indicates that if the RIC holds, then κ-> 0  . We show at the bottom of Section 2.9.1 that if the RIC holds then 0 < κ-< c′(mt) < 1  so that

pict

which is negative because the GIC says ÞÞÞ  < 1
 𝒢  .

I Relational Diagrams for the Inequality Conditions

This appendix explains in detail the paper’s ‘inequalities’ diagrams (Figures 13).

PIC

Figure 9:Appendix: Inequality Conditions for Perfect Foresight Model
(Start at a node and follow arrows)

I.1 The Unconstrained Perfect Foresight Model

A simple illustration is presented in Figure 9, whose three nodes represent values of the absolute patience factor ÞÞÞ  , the permanent-income growth factor 𝒢 , and the riskfree interest factor R  . The arrows represent imposition of the labeled inequality condition (like, the uppermost arrow, pointing from ÞÞÞ  to 𝒢 , reflects imposition of the GIC condition (clicking GIC should take you to its definition; definitions of other conditions are also linked below)).65 Annotations inside parenthetical expressions containing ≡ are there to make the diagram readable for someone who may not immediately remember terms and definitions from the main text. (Such a reader might also want to be reminded that R, β,  and Γ  are all in ℝ++   , and that ρ >  1  ).

Navigation of the diagram is simple: Start at any node, and deduce a chain of inequalities by following any arrow that exits that node, and any arrows that exit from successive nodes. Traversal must stop upon arrival at a node with no exiting arrows. So, for example, we can start at the ÞÞÞ  node and impose the GIC and then the FHWC, and see that imposition of these conditions allows us to conclude that ÞÞÞ <  R  .

One could also impose ÞÞÞ  < R  directly (without imposing GIC  and FHWC  ) by following the downward-sloping diagonal arrow exiting ÞÞÞ  . Although alternate routes from one node to another all justify the same core conclusion (ÞÞÞ <  R  , in this case), ⁄=  symbol in the center is meant to convey that these routes are not identical in other respects. This notational convention is used in category theory diagrams,66 to indicate that the diagram is not commutative.67

Negation of a condition is indicated by the reversal of the corresponding arrow. For example, negation of the RIC, ¬RIC  ≡ ÞÞÞ >  R  , would be represented by moving the arrowhead from the bottom right to the top left of the line segment connecting ÞÞÞ  and R  .

If we were to start at R  and then impose ¬FHWC  , that would reverse the arrow connecting R  and 𝒢 , but the 𝒢 node would then have no exiting arrows so no further deductions could be made. However, if we also reversed GIC  (that is, if we imposed ¬GIC  ), that would take us to the ÞÞÞ  node, and we could deduce R > ÞÞÞ  . However, we would have to stop traversing the diagram at this point, because the arrow exiting from the ÞÞÞ node points back to our starting point, which (if valid) would lead us to the conclusion that R > R  . Thus, the reversal of the two earlier conditions (imposition of ¬FHWC  and ¬GIC  ) requires us also to reverse the final condition, giving us ¬RIC  .68

Under these conventions, Figure 1 in the main text presents a modified version of the diagram extended to incorporate the PF-FVAC (reproduced here for convenient reference).

PIC
Figure 10:Appendix: Relation of GIC, FHWC, RIC, and PFVAC

An arrowhead points to the larger of the two quantities being compared. For example, the diagonal arrow indicates that ÞÞÞ < R1∕ρ𝒢1−1∕ρ  , which is an alternative way of writing the PF-FVAC, (19)

This diagram can be interpreted, for example, as saying that, starting at the ÞÞÞ  node, it is possible to derive the PF -FVAC  69 by imposing both the GIC and the FHWC; or by imposing RIC and ¬ FHWC. Or, starting at the 𝒢 node, we can follow the imposition of the FHWC (twice — reversing the arrow labeled ¬FHWC  ) and then ¬RIC  to reach the conclusion that ÞÞÞ < 𝒢 . Algebraically,

FHWC    :  𝒢 <  R

 ¬RIC   :  R < ÞÞÞ
           𝒢 < ÞÞÞ
(96)

which leads to the negation of both of the conditions leading into ÞÞÞ  . ¬ GIC is obtained directly as the last line in (96) and ¬ PF-FVAC follows if we start by multiplying the Return Patience Factor (RPF=ÞÞÞ ∕R  ) by the FHWF (=𝒢∕R  ) raised to the power 1∕ ρ − 1  , which is negative since we imposed ρ >  1  . FHWC implies FHWF < 1  so when FHWF is raised to a negative power the result is greater than one. Multiplying the RPF (which exceeds 1 because ¬ RIC) by another number greater than one yields a product that must be greater than one:

                          >1-from ¬RIC
                          ◜(    ◞◟1∕ρ)◝  >◜1-fro◞m◟-FHW◝C
                            (Rβ-)---         1∕ρ− 1
                      1 <      R       (𝒢∕R )
                          (               )
                               (R β)1∕ρ
                      1 <   ------1∕ρ------
                            (R∕𝒢 )  R 𝒢∕R
R1∕ρ𝒢1− 1∕ρ = (R∕ 𝒢)1∕ρ𝒢  < ÞÞÞ

which is one way of writing ¬PF -FVAC  .

The complexity of this algebraic calculation illustrates the usefulness of the diagram, in which one merely needs to follow arrows to reach the same result.

After the warmup of constructing these conditions for the perfect foresight case, we can represent the relationships between all the conditions in both the perfect foresight case and the case with uncertainty as shown in Figure 3 in the paper (reproduced here).

PIC
Figure 11:Appendix: Relation of All Inequality Conditions

Finally, the next diagram substitutes the values of the various objects in the diagram under the baseline parameter values and verifies that all of the asserted inequality conditions hold true.

PIC
Figure 12:Appendix: Numerical Relation of All Inequality Conditions

J Apparent Balanced Growth in 𝔠  and cov (c,p )

Section 4.2 demonstrates some propositions under the assumption that, when an economy satisfies the GIC, there will be constant growth factors Ω 𝔠   and Ωcov   respectively for 𝔠  (the average value of the consumption ratio) and cov(c,p )  . In the case of a Szeidl-invariant economy, the main text shows that these are Ω  = 1
 𝔠  and Ω   =  𝒢
 cov . If the economy is Harmenberg- but not Szeidl-invariant, no proof is offered that these growth factors will be constant.

J.1 log c  and log (cov(c,p))  Grow Linearly

Figures 13 and 14 plot the results of simulations of an economy that satisfies Harmenberg- but not Szeidl-invariance with a population of 4 million agents over the last 1000 periods (of a 2000 period simulation).70 The first figure shows that log𝔠  increases apparently linearly. The second figure shows that log (− cov(c,p ))  also increases apparently linearly. (These results are produced by the notebook ApndxBalancedGrowthcNrmAndCov.ipynb).

PIC
Figure 13:Appendix: log  𝔠  Appears to Grow Linearly
PIC
Figure 14:Appendix: log (− cov(c,p ))  Appears to Grow Linearly

References

   Abowd, John M., and David Card (1989): “On the Covariance Structure of Earnings and Hours Changes,” Econometrica, 57, 411–445.

   Aiyagari, S. Rao (1994): “Uninsured Idiosyncratic Risk and Aggregate Saving,” Quarterly Journal of Economics, 109, 659–684.

   Alvarez, Fernando, and Nancy L Stokey (1998): “Dynamic programming with homogeneous functions,” Journal of economic theory, 82(1), 167–189.

   Bellman, Richard (1957): Dynamic Programming. Princeton University Press, Princeton, NJ, USA, 1 edn.

   Benhabib, Jess, Alberto Bisin, and Shenghao Zhu (2015): “The wealth distribution in Bewley economies with capital income risk,” Journal of Economic Theory, 159, 489–515, Available at https://www.nber.org/papers/w20157.pdf.

   Bewley, Truman (1977): “The Permanent Income Hypothesis: A Theoretical Formulation,” Journal of Economic Theory, 16, 252–292.

   Blanchard, Olivier J. (1985): “Debt, Deficits, and Finite Horizons,” Journal of Political Economy, 93(2), 223–247.

   Boyd, John H. (1990): “Recursive Utility and the Ramsey Problem,” Journal of Economic Theory, 50(2), 326–345.

   Cagetti, Marco (2003): “Wealth Accumulation Over the Life Cycle and Precautionary Savings,” Journal of Business and Economic Statistics, 21(3), 339–353.

   Carroll, Christopher D. (1992): “The Buffer-Stock Theory of Saving: Some Macroeconomic Evidence,” Brookings Papers on Economic Activity, 1992(2), 61–156, https://www.econ2.jhu.edu/people/ccarroll/BufferStockBPEA.pdf.

   __________  (1997): “Buffer Stock Saving and the Life Cycle/Permanent Income Hypothesis,” Quarterly Journal of Economics, CXII(1), 1–56.

   __________  (2019, Submitted): “Theoretical Foundations of Buffer Stock Saving,” Quantitative Economics.

   Carroll, Christopher D., Martin Holm, and Miles S. Kimball (2019): “Liquidity Constraints and Precautionary Saving,” Manuscript, Johns Hopkins University, https://www.econ2.jhu.edu/people/ccarroll/papers/LiqConstr.

   Carroll, Christopher D., and Miles S. Kimball (1996): “On the Concavity of the Consumption Function,” Econometrica, 64(4), 981–992, https://www.econ2.jhu.edu/people/ccarroll/concavity.pdf.

   Carroll, Christopher D., and Andrew A. Samwick (1997): “The Nature of Precautionary Wealth,” Journal of Monetary Economics, 40(1), 41–71.

   Carroll, Christopher D., Jiri Slacalek, Kiichi Tokuoka, and Matthew N. White (2017): “The Distribution of Wealth and the Marginal Propensity to Consume,” Quantitative Economics, 8, 977–1020, At https://www.econ2.jhu.edu/people/ccarroll/papers/cstwMPC.

   Chamberlain, Gary, and Charles A. Wilson (2000): “Optimal Intertemporal Consumption Under Uncertainty,” Review of Economic Dynamics, 3(3), 365–395.

   Christopher D. Carroll, Alexander M. Kaufman, Jacqueline L. Kazil, Nathan M. Palmer, and Matthew N. White (2018): “The Econ-ARK and HARK: Open Source Tools for Computational Economics,” in Proceedings of the 17th Python in Science Conference, ed. by Fatih Akici, David Lippa, Dillon Niederhut, and M Pacer, pp. 25 – 30.

   Clarida, Richard H. (1987): “Consumption, Liquidity Constraints, and Asset Accumulation in the Face of Random Fluctuations in Income,” International Economic Review, XXVIII, 339–351.

   Daly, Moira, Dmytro Hryshko, and Iourii Manovskii (2016): “Improving the measurement of earnings dynamics,” Discussion paper, National Bureau of Economic Research.

   Deaton, Angus S. (1991): “Saving and Liquidity Constraints,” Econometrica, 59, 1221–1248, https://www.jstor.org/stable/2938366.

   Durán, Jorge (2003): “Discounting long run average growth in stochastic dynamic programs,” Economic Theory, 22(2), 395–413.

   Friedman, Milton A. (1957): A Theory of the Consumption Function. Princeton University Press.

   Gourinchas, Pierre-Olivier, and Jonathan Parker (2002): “Consumption Over the Life Cycle,” Econometrica, 70(1), 47–89.

   Harmenberg, Karl (2021): “Aggregating heterogeneous-agent models with permanent income shocks,” Journal of Economic Dynamics and Control, 129, 104185.

   Hendricks, Lutz (2001): Bequests and Retirement Wealth in the United States. University of Arizona.

   __________  (2016): “Wealth Distribution and Bequests,” Lecture Notes, Economics 821, University of North Carolina.

   Hiraguchi, Ryoji (2003): “On the Convergence of Consumption Rules,” Manuscript, Johns Hopkins University.

   Hryshko, Dmytro, and Iourii Manovskii (2020): “How much consumption insurance in the US?,” Manuscript, University of Alerta.

   Jappelli, Tullio, and Luigi Pistaferri (2000): “Intertemporal Choice and Consumption Mobility,” Econometric Society World Congress 2000 Contributed Paper Number 0118.

   Krueger, Dirk, Kurt Mitman, and Fabrizio Perri (2016): “Macroeconomics and Household Heterogeneity,” Handbook of Macroeconomics, 2, 843–921.

   Li, Huiyu, and John Stachurski (2014): “Solving the income fluctuation problem with unbounded rewards,” Journal of Economic Dynamics and Control, 45, 353–365.

   Ma, Qingyin, John Stachurski, and Alexis Akira Toda (2020): “The income fluctuation problem and the evolution of wealth,” Journal of Economic Theory, 187.

   Ma, Qingyin, and Alexis Akira Toda (2020): “A Theory of the Saving Rate of the Rich,” .

   MaCurdy, Thomas (1982): “The Use of Time Series Processes to Model the Error Structure of Earnings in a Longitudinal Data Analysis,” Journal of Econometrics, 18(1), 83–114.

   Martins-da Rocha, V Filipe, and Yiannis Vailakis (2010): “Existence and uniqueness of a fixed point for local contractions,” Econometrica, 78(3), 1127–1141.

   Matkowski, Janusz, and Andrzej S. Nowak (2011): “On Discounted Dynamic Programming With Unbounded Returns,” Economic Theory, 46, 455–474.

   Modigliani, Franco (1966): “The Life Cycle Hypothesis, the Demand for Wealth, and the Supply of Capital,” Social Research, 33, 160–217.

   Muth, John F. (1960): “Optimal Properties of Exponentially Weighted Forecasts,” Journal of the American Statistical Association, 55(290), 299–306.

   Rabault, Guillaume (2002): “When do borrowing constraints bind? Some new results on the income fluctuation problem,” Journal of Economic Dynamics and Control, 26(2), 217–245.

   Ramsey, Frank (1928): “A Mathematical Theory of Saving,” Economic Journal, 38(152), 543–559.

   Riehl, Emily (2017): Category theory in context. Courier Dover Publications.

   Rincón-Zapatero, Juan Pablo, and Carlos Rodríguez-Palmero (2003): “Existence and uniqueness of solutions to the Bellman equation in the unbounded case,” Econometrica, 71(5), 1519–1555.

   Schechtman, Jack, and Vera Escudero (1977): “Some results on ‘An Income Fluctuation Problem’,” Journal of Economic Theory, 16, 151–166.

   Scheinkman, José, and Laurence Weiss (1986): “Borrowing Constraints and Aggregate Economic Activity,” Econometrica, 54(1), 23–46.

   Schmitt-Grohé, Stephanie, and Martın Uribe (2003): “Closing small open economy models,” Journal of international Economics, 61(1), 163–185.

   Stachurski, John, and Alexis Akira Toda (2019): “An Impossibility Theorem for Wealth in Heterogeneous-Agent Models with Limited Heterogeneity,” Journal of Economic Theory, 182, 1–24.

   Stokey, Nancy L., Robert E. Lucas, and Edward C. Prescott (1989): Recursive Methods in Economic Dynamics. Harvard University Press.

   Szeidl, Adam (2013): “Stable Invariant Distribution in Buffer-Stock Saving and Stochastic Growth Models,” Manuscript, Central European University, Available at http://www.personal.ceu.hu/staff/Adam_Szeidl/papers/invariant_revision.pdf.

   Toche, Patrick (2005): “A Tractable Model of Precautionary Saving in Continuous Time,” Economics Letters, 87(2), 267–272.

   Yaari, Menahem E (1965): “Uncertain lifetime, life insurance, and the theory of the consumer,” The Review of Economic Studies, 32(2), 137–150.

   Yao, Jiaxiong (2012): “The Theoretical Foundations of Buffer Stock Saving: A Note,” Manuscript, Johns Hopkins University.

   Zeldes, Stephen P. (1989): “Optimal Consumption with Stochastic Income: Deviations from Certainty Equivalence,” Quarterly Journal of Economics, 104(2), 275–298.