CRRA
CRRA
Household Heterogeneity
Dirk Krueger1
Department of Economics
University of Pennsylvania
March 6, 2018
I Introduction 1
1 Overview over the Monograph 3
iii
iv CONTENTS
IV Conclusions 255
Part I
Introduction
1
Chapter 1
3
4 CHAPTER 1. OVERVIEW OVER THE MONOGRAPH
and in which households will be able to partially, but not fully (self-)insure
against random income fluctuations. In part II I will discuss the perhaps most
important workhorse model in this literature. In the standard incomplete
markets model (SIM) households have access only to a single, one period
risk-free bond, so that their period budget constraint reads as
c + qa0 ≤ y(s) + a (1.2)
and the shortsale of bonds a0 might be limited by a constraint of the form
a0 ≥ −Ā. In chapter 5 I will discuss the theoretical properties of various
versions (with alternating assumptions on the household utility function, the
stochastic nature of the income process, and the tightness of the borrowing
constraint) of the model in which the price of the bond q and thus the interest
rate 1 + r = 1/q is exogenously given. (and as a consequence we can analyze
the behavior of one household in isolation). After a brief digression in chapter
6 that discusses the properties of the main driving of this class of models, the
stochastic process for earnings, I will then incorporate the decision problems
from chapter 5 into a dynamic general equilibrium model in which interest
rates and wages are determined endogenously in the labor and capital market.
This is done in chapter 7. There, individual households’ consumption and
saving decisions are aggregated to obtain aggregate labor and capital supply,
firms’ decisions are aggregated to obtain aggregate labor and capital demand,
and wages and interest rates move to clear both markets. Depending on
whether individual uncertainty averages out in the aggregate (no aggregate
uncertainty) wages and interest are constant over time or are themselves
stochastic processes (presence of aggregate uncertainty), leading to severe
computational problems when computing these models. The aggregation
of individual decisions also leads to (possibly time-varying) cross-sectional
consumption and wealth distributions; thus these models are possibly useful
for the study of the effects of redistributive and social insurance policies such
as progressive taxation, unemployment insurance, welfare or social security.
Common to all these models is the assumption of the absence of explicit
insurance arrangements in an environment in which mutual insurance is po-
tentially quite beneficial. In part III I will discuss a strand of the literature
that aims to explain the stylized facts from chapter 3with models that depart
directly from the complete markets model, without a priori ruling out explicit
insurance contracts (as the standard incomplete market model does). Chap-
ter 8 discusses models in which imperfect consumption insurance arises due
to the assumed imperfect enforceability of insurance contracts. If households
6 CHAPTER 1. OVERVIEW OVER THE MONOGRAPH
are given the option to default, with the punishment, say, of being expelled
into financial autarchy forever after (and thus becoming the hand-to-mouth
consumers discussed above), full consumption insurance might not be incen-
tive compatible. I will argue that this results in a model in which households
face a budget constraint of the form (1.1), exactly as in the standard complete
market model, but now in addition face state-contingent shortsale constraints
on the state contingent bonds a0 (s0 ) ≥ Ā(s0 ) that are potentially very tight
and do not permit the implementation of the full consumption insurance al-
location. Finally I will briefly discuss, in chapter 9 models in which perfect
consumption insurance might be impossible to be implemented since individ-
ual incomes are not publicly observable. Consequently a mechanism designer
of consumption insurance contracts will have to respect the incentives of
households to mis-report their incomes and to construct a partial-insurance
consumption insurance contract that induces all households to report their
incomes truthfully.
A short, necessarily subjective assessment of where this area of research
stands and where it might be headed will conclude this monograph.
Chapter 2
7
8CHAPTER 2. WHY MACRO WITH HETEROGENEOUS HOUSEHOLDS (OR
Chapter 3
In this chapter we will discuss the main stylized facts that heterogeneous
agent macro models are designed to explain. We are mainly interested in four
main economic variables, labor earnings, income, wealth and consumption
9
10 CHAPTER 3. SOME STYLIZED FACTS AND SOME PUZZLES
3.1.2 SCF
With respect to income, the PSID as well as the SCF contains data that are
supposedly of higher quality than the income data from the CEX. The SCF is
conducted in three year intervals; the four available surveys are for the years
1989, 1992, 1995 and 1998. It is conducted by the National Opinion Research
center at the University of Chicago and sponsored by the Federal Reserve sys-
tem. It contains rich information about U.S. households’ income and wealth.
In each survey about 4,000 households are asked detailed questions about
their labor earnings, income and wealth. One part of the sample is represen-
tative of the U.S. population, to give an accurate description of the entire
population. The second part over-samples rich households, to get a more
precise idea about the precise composition of this groups’ income and wealth
composition. As we will see, this group accounts for the majority of total
household wealth, and hence it is particularly important to have good infor-
mation about this group. The main advantage of the SCF is the level of detail
of information about income and wealth. The main disadvantage is that it
is not a panel data set, i.e. households are not followed over time. Hence
dynamics of income and wealth accumulation cannot be documented on the
household level with this data set. For further information and some of the
data see http://www.federalreserve.gov/pubs/oss/oss2/98/scf98home.html.
3.1.3 PSID
The Panel Study of Income Dynamics (PSID) is conducted by the Survey
Research Center of the University of Michigan and mainly sponsored by the
National Science Foundation. The PSID is a panel data set that started
with a national sample of 5,000 U.S. households in 1968. The same sam-
ple individuals are followed over the years, barring attrition due to death or
non-response. New households are added to the sample on a consistent basis,
making the total sample size of the PSID about 8700 households. The income
3.1. HOUSEHOLD LEVEL DATA SOURCES 11
and wealth data are not as detailed as for the SCF, but its panel dimension
allows to construct measures of income and wealth dynamics, since the same
households are interviewed year after year. Also the PSID contains data on
consumption expenditures, albeit only food consumption. In addition, in
1990, a representative national sample of 2,000 Latinos, differentially sam-
pled to provide adequate numbers of Puerto Rican, Mexican-American, and
Cuban-Americans, was added to the PSID database. This provides a host of
information for studies on discrimination. For further information and the
complete data set see http://www.isr.umich.edu/src/psid/index.html
3.1.4 CPS
The Current Population Survey (CPS) is conducted by the U.S. Bureau of
the Census and sponsored by the Bureau of Labor Statistics. In its annual
March supplement detailed information about household income is collected.
The survey started to gather information about household income in 1948,
but comprehensive information about household income and income of its
members is available only since 1970’s. The main advantage of the CPS is
its sample size: in each year it contains a representative sample of 40,000 to
60,000 households. However, no information about consumption or wealth
information is collected. Also, this survey, like the SCF is a purely cross-
sectional data set without panel dimension as it does not follow individual
families over time. Fore more details see http://www.bls.census.gov/cps.
• Many other national data sets for other countries. See RED 2011, Vol.
1.
• Luxembourg Income Study as well as the Luxembourg Wealth Study
GDP
10.8 Consumption
Log Real GDP and Consumption p.c.
10.6
10.4
10.2
10
9.8
9.6
9.4
1965 1970 1975 1980 1985 1990 1995 2000 2005 2010
Year
C/GDP
C/Pers. Inc.
0.8
Consumption Share
0.75
0.7
0.65
0.6
1965 1970 1975 1980 1985 1990 1995 2000 2005 2010
Year
0.6
0.5
Consumption Shares
0.4
Nondurable Goods
Durable Goods
0.3 Services
0.2
0.1
0
1965 1970 1975 1980 1985 1990 1995 2000 2005 2010
Year
0.6
Real Consumption Shares
0.5
0.4
Nondurable Goods
Durable Goods
0.3 Services
0.2
0.1
0
1996 1998 2000 2002 2004 2006 2008 2010 2012
Year
10
0
1965 1970 1975 1980 1985 1990 1995 2000 2005 2010
Year
Figure: Labor Income and Net Worth by Age, SCF 2007 ($1,000)
1200.00 120.00
1000.00 100.00
800.00 80.00
600.00 60.00
400.00 40.00
0.00 0.00
20-29 30-39 40-49 50-59 60-69 70 or more
Age Group
Figure 3.6: Income and Net Worth over the Life Cycle
4000
3500
3000
2500
2000
1500
20 30 40 50 60 70 80 90
Age
obtain formal education or training on the job and labor force participation
of women is low because of child bearing and rearing. As more and more
agents finish their education and learn on the job as well as promotions occur,
average wages within the cohort increase. Average personal income at age
45 is almost 2.5 times as high as average personal income at age 25. After
the age of 45 personal income first slowly, then more rapidly declines as more
and more people retire and labor productivity (and thus often wages) fall.
The average household at age 65 has only 60% of the personal income that
the average household at age 45 obtains.
The second main finding is the surprising finding. Not only personal
income, but also consumption follows a hump over the life cycle. In other
words, consumption seems to track income over the life cycle fairly closely.
This is one statement of the so-called excess sensitivity puzzle: consump-
tion appears to be excessively sensitive to predicted changes in income. In
fact, the two standard theories of intertemporal consumption allocation we
will consider in the next section both predict that (under specific assumptions
spelled out explicitly below) consumption follows a martingale and current
income does not help to forecast future consumption. The hump-shaped con-
sumption age profile apparently seems to contradict this hypothesis. Later
in the course we will investigate in detail whether, once we control for house-
hold size (which also happens to follow a hump shape), the hump-shape in
consumption disappears or whether the puzzle persists. Figure ?? (taken
from Fernandez-Villaverde and Krueger, 2007) documents the life cycle pro-
file of consumption, with and without adjustment for family size by household
equivalence scales. The figure is derived using a synthetic cohort analysis, a
technique from Panel data econometrics that allows us to construct average
life cycle profiles for households that we do not observe over their entire life
(we will talk about this technique in detail below). The key observation from
this figure is that consumption displays a hump over the life cycle, and that
this hump persists, even after controlling for family size. The later observa-
tion is not entirely noncontroversial, and we will discuss below the different
positions on this issue.
of log earnings and log consumption. The figure shows that as a cohort ages,
the distribution of earnings and consumption within a cohort fans out.5 Note
that the increase in consumption inequality is substantially less pronounced
than the increase in the dispersion of earnings.
Cross-sectional dispersion (or inequality) has also changed dramatically
over time. Heathcote et al. (2004) and many others have documented a
strong upward trend in wage and even more pronounced in household earn-
ings inequality over the last 30 years. The inequality in hours worked has
remained fairly stable, and consumption inequality has increased by sub-
stantially less than earnings inequality (see Krueger and Perri (2006) and
the discussion in Attanasio et al. (2007)). Finally, it appears that wealth
inequality has increased as well in the 1980’s and since the remained fairly
constant, to the extent the wealth data are available and can be trusted (see
Favilukis, 2007). A substantial body of literature has sprung up in the last
few years trying to explain these trends with structural macro models with
heterogenous agents, of the type studied in these notes.
Heathcote, Perri and Violante (2009)
under Chapter 7 households are discharged of all their debts, are not re-
quired to use any of their future labor income to repay the debt and can
even keep their assets (financial or real estate) below a state-dependent ex-
emption level. Whereas about 1% of all households per year file for personal
bankruptcy, White (1998) computes that currently at least 15% of all US
households would financially benefit from filing for bankruptcy. The fraction
of U.S. households filing for bankruptcy has increased sharply over the last
decade. So has the extent of uncollateralized debt, as a fraction of disposable
household income (the same is very much true for collateralized debt). Also,
charge-off rates of lenders (the fraction of loans the lender does not recover),
have increased substantially (see Livshits et al., 2007). Show pictures on
trends of debt and default (both uncollateralized and collateralized) from
Livshits et al.
Any given model will likely not be able to resolve all these puzzles at
once, and some models will abstract from some of the issues altogether, but
the “stylized facts” of this section should be kept in mind in order to guide
extensions of the models presented next.
22 CHAPTER 3. SOME STYLIZED FACTS AND SOME PUZZLES
Chapter 4
23
24 CHAPTER 4. THE STANDARD COMPLETE MARKETS MODEL
Furthermore, unless otherwise noted, we will assume that the utility function
is additively time-separable and that agents discount the future at common
subjective time discount factor β ∈ (0, 1), so that the period utility function
takes the form Uti (ci , st ) = β t U i (cit (st ), st ) and thus expected lifetime utility
is given by:
X T X
ui (ci ) = β t πt (st )U i (cit (st ), st ) (4.2)
t=0 st ∈S t
where pt (st ) is the period 0 price of one unit of period t consumption, deliv-
ered if event history st has realized.
1. Given {pt (st )}Tt=0,st ∈S t , for each i ∈ I, {cit (st )}Tt=0,st ∈S t maximizes (4.2)
subject to (4.3) and (4.7)
subject
PNto (4.3) and (4.4), for some Pareto weights (αi )N i
i=1 satisfying α ≥ 0
and i=1 αi = 1. First, observe that the assumption is (more than) sufficient
to establish the first welfare theorem, and thus we know that any compet-
itive equilibrium allocation is Pareto efficient. Second, any Pareto efficient
allocation is the solution to the social planner problem in (4.10), for some
Pareto weights (see, e.g. MasColell et. al., chapter 16; this result requires
other parts of assumption 1, especially concavity).
for all dates t and all states st . Hence in any efficient allocation (and thus
in a market economy with a complete set of contingent consumption claims
being traded) the ratio of marginal utilities of consumption of any two agents
is constant across time and states. Also those agents, ceteris paribus (i.e. if
they had the same utility function and the same preference shocks), with
higher relative Pareto weights will consume more in every state of the world
because the utility function is assumed to be strictly concave.
Thus the benevolent social planner whose aim to insure consumption of
all households over time and across states of nature finds it optimal to keep
4.1. THEORETICAL RESULTS 27
This was irrelevant in the social planner problem, but is required for a com-
petitive equilibrium allocation. But in order to compute how much such
2
Negishi’s goal was to prove existence of equilibrium. He did so by first arguing that all
solutions to the social planner problem satisfy the elements of the definition of a competi-
tive equilibrium apart from household optimality (with prices equal to Lagrange multipliers
on the resource constraints of the social planner problem). He then showed that there exist
Pareto weights such that the associated solution to the social planner problem solves the
household maximization problem (again with prices given by the Lagrange multipliers on
the resource constraint). These Pareto weights are given by the solution to the system of
equations (4.18) below.
4.1. THEORETICAL RESULTS 29
an allocation costs we need the appropriate prices. It turns out that the
Lagrange multipliers (that is, the shadow prices) λt (st , α) on the resource
constraints from the social planner problem are appropriate. So for each
agent i define the transfer functions as
XX
ti (α) = λt (st ) cit (st , α) − yti (st ) .
(4.17)
t st ∈S t
Thus ti (α) is the value of lifetime consumption net of the value of lifetime
income, where income and consumption at each history st is valued at its
social value λt (st , α). For our economy, from equation (4.11) the Lagrange
multipliers are given by
Equation (4.17) also makes clear that the Pareto weights that make the
transfer functions ti (α) equal to zero depend on the properties of the indi-
vidual income processes {yti }. As a consequence, the equilibrium level and
thus share θi of consumption of each agent i depends crucially on the value
of her endowment process.
that is, if agents have CRRA utility that is separable in consumption, then
in an efficient (competitive equilibrium) allocation individual consumption
growth is perfectly correlated with and predicted by aggregate consumption
growth. In particular, individual income growth should not help to predict
individual consumption growth once aggregate consumption (income) growth
is accounted for. This result is the starting point of the most basic empiri-
cal tests of perfect consumption insurance, see e.g. Mace (1991), Cochrane
(1991), among others.
To obtain Arrow Debreu prices associated with equilibrium allocations
we obtain from the consumer problem of maximizing (4.2) subject to (4.7)
that
pt+1 (st+1 ) πt+1 (st+1 ) Uci (cit+1 (st+1 ), st+1 )
= β (4.20)
pt (st ) πt (st ) Uci (cit (st ), st )
Under assumption 2 this becomes
−σ
pt+1 (st+1 ) πt+1 (st+1 ) cit+1 (st+1 )
= β (4.21)
pt (st ) πt (st ) cit (st )
−σ
πt+1 (st+1 ) ct+1 (st+1 )
= β (4.22)
πt (st ) ct (st )
Proposition 7 Suppose allocations {(cit (st ))i∈I }Tt=0,st ∈S t and prices {pt (st )}Tt=0,st ∈S t
form an Arrow-Debreu equilibrium. Then under assumption 2 the allocation
{ct (st )}Tt=0,st ∈S t defined by
X
ct (st ) = cit (st ) (4.23)
i∈I
and prices {pt (st )}Tt=0,st ∈S t form an Arrow Debreu equilibrium for the repre-
sentative agent economy in which the representative agent has an endowment
process given by X
yt (st ) = yti (st ) (4.24)
i∈I
and CRRA preferences with parameter σ.
The fact that the equilibrium of the representative agent economy has
consumption allocations given by (4.23) is of course trivial and follows di-
rectly from the market clearing condition. The content of this proposition
lies in the statement that in the representative agent economy (with the rep-
resentative household having the same CRRA utility function as agents in
the heterogeneous agent economy) has the same equilibrium Arrow-Debreu
prices as the economy with I consumers that might differ in their income
processes in an arbitrary way. Thus, in order to derive Arrow-Debreu prices
(and hence all other asset prices) in an economy with complete markets, it
is sufficient to study the corresponding representative agent economy. Most
of the consumption based asset pricing literature since Lucas (1978) (and
indeed most of model-based macroeconomics) has indeed employed models
with a representative consumer. As the previous result suggests, if financial
markets are complete (and sufficiently strong assumptions on agents’ utility
functions are made), then the abstraction from household heterogeneity is
innocuous from the perspective of macroeconomic research. It should also
be noted that for this proposition assumption 2 can be weakened3 , although
3
We have assumed idential CRRA utility functions and time discount factors across all
agents. Koulovatianos (2007), building on the large literature on aggregation in dynamic
models, such as Chatterjee (1994) and Caselli and Ventura (2000), gives necessary and
sufficient conditions on individual utility functions to obtain a representative consumer.
See his Theorem 1 and 2.
Maintaining identical time discount factors the period utility function has to be either
of power or exponential form (with heterogeneity in the CARA, but not in the CRRA
possible). If the time discount factor is allowed to vary across agents, only exponential
utility gives rise to the result stated here.
32 CHAPTER 4. THE STANDARD COMPLETE MARKETS MODEL
the construction of the utility function of the representative agent may more
involved.
N
X N
X
cit (st ) = wti (st )lti (st )
i=1 i=1
and the period utility function is given by U (cit (st ), lti (st )), where lti (st ) ∈ [0, 1]
is the fraction of the time a household works.
The key efficiency conditions now read as
The first condition is the familiar efficient risk sharing condition across house-
holds. The second condition governs the efficient allocation of consumption
4.1. THEORETICAL RESULTS 33
and leisure for an arbitrary household i. We can divide equations (4.26) for
any two agents to arrive at
For given welfare weights α equations (4.25) and (4.27) determine relative
consumption and leisure allocations between households i and j.
Let us consider two important classes of period utility functions U (.) in
the following two examples.
Example 8 Suppose U (cit (st ), lti (st )) is additively separable between consump-
tion and leisure (e.g. U (cit (st ), lti (st )) = v(cit (st )) − g(lti (st )) where v is strictly
concave and g is strictly convex. In this case the implications for efficient
consumption risk sharing are exactly the same as in the case with exogenous
labor supply:
v 0 (cit (st )) αj
= .
v 0 (cjt (st )) αi
In particular, if v is of CRRA form, then as before consumption of each
agent i equals a constant fraction θi of aggregate production. The efficient
allocation of labor supply, in this case, is characterized by
Thus, since g is strictly convex more productive households work harder. But
although labor supply does respond to idiosyncratic productivity shocks wti (st ),
as before the share of consumption of agent i does not.
Example 9 U (cit (st ), lti (st )) is nonseparable. Then in general also consump-
tion shares respond to idiosyncratic wage shocks (which in turn suggests that
one has to be cautious when testing perfect consumption insurance). We
demonstrate this through a second example. Suppose the period utility func-
tion is of familiar Cobb-Douglas form
1−σ
[cit (st )ν (1 − lti (st ))1−ν ] −1
U (cit (st ), lti (st )) = .
1−σ
34 CHAPTER 4. THE STANDARD COMPLETE MARKETS MODEL
Since σ > 0 and ν > 0 the second equation implies that more productive
households (holding Pareto weights constant) consume less leisure and work
more, that is, if wi goes up relative to wj in a particular node, then household
i’s labor supply increases, relative to that of household j. The first equation
implies that more productive households also consume more if and only if σ >
1. If this condition is satisfied, it is easy to verify from the utility function that
Uc,1−l < 0, that is, then the marginal utility of consumption falls with higher
leisure. Recall from equation (4.25) that it is efficient to keep the ratio of
marginal utilities of consumption constant across time and states. If leisure of
household i goes down relative to j and Uc,1−l < 0, then for fixed consumption
Uci /Ucj would rise. Thus to keep Uci /Ucj constant, consumption of household i
has to rise relative to household j in that node: it is efficient to compensate
household i for her higher labor supply with larger consumption. If σ < 1
marginal utility of consumption increases with leisure, and the reverse logic
is true. Finally, if σ → 1 then the utility function is additively separable in
consumption and leisure
U (cit (st ), lti (st )) = ν log cit (st ) + (1 − ν) log 1 − lti (st )
Note that agents purchase Arrow securities {ait+1 (st , st+1 )}st+1∈S for all con-
tingencies st+1 ∈ S that can happen tomorrow, but that, once st+1 is realized,
only the ait+1 (st+1 ) corresponding to the particular realization of st+1 pays off
and thus determines her asset position at the beginning of the next period.
We assume that all agents start their life with an asset position of zero, that
is, ai0 (s0 ) = 0. We can now state the following:
4
Admittedly, such assets could be introduced into the Arrow Debreu market structure
and could be priced in a straightforward manner, once equilibrium prices {pt (st )} for
state-contingent consumption claims have been determined.
5
A full set of one-period Arrow securities is sufficient to make markets “sequen-
tially complete”, in the sense that any (nonnegative) state-contingent consumption al-
location {cit (st )} is attainable with an appropriate sequence of Arrow security holdings
{ait+1 (st , st+1 )} satisfying all sequential markets budget constraints.
36 CHAPTER 4. THE STANDARD COMPLETE MARKETS MODEL
Definition 10 A SM equilibrium is allocations { ĉit (st ), âit+1 (st , st+1 st+1 ∈S }Tt=0,st ∈S t ,
i∈I
and prices for Arrow securities {q̂t (st , st+1 )}Tt=0,st ∈S t ,st+1 ∈S such that
1. For i ∈ I given {q̂t (st , st+1 )}Tt=0,st ∈S t ,st+1 ∈S , for all i, {ĉit (st ), âit+1 (st , st+1 st+1 ∈S }Tt=0,st ∈S t
maximizes (4.2) subject to (4.28) and the constraints cit (st ) ≥ 0 and
ait+1 (st , st+1 ) ≥ −Āi (st+1 , q̂, ci )
2. For all t, st ∈ S t
I
X I
X
ĉit (st ) = yti (st ) (4.29)
i=1 i=1
I
X
âit+1 (st , st+1 ) = 0 for all st+1 ∈ S (4.30)
i=1
would break down. Appendix 4.4 discusses the choice of the No Ponzi con-
dition in greater detail; here we simply present a specification that is easy to
motivate, commonly used and gives rise to the desired equivalence result.
For a given Arrow securities price process q = {qt (st , st+1 )} define date
zero prices of state contingent consumption claims {v0 (st )} implied by q as
follows:
v0 (s0 ) = 1
v0 (st+1 ) = v0 (st )qt (st , st+1 ) = q0 (s0 , s1 ) ∗ . . . ∗ qt (st , st+1 ) ∗ 1. (4.31)
Using these prices derived from q we now state the No Ponzi scheme condition
as
−v0 (sT )aT (sT −1 , sT ) ≤ W (sT ) (4.32)
where ∞ X
X
T
W (s ) = v0 (st )yti (st ) (4.33)
t=T st |sT
We will assume that Āi (st+1 , q) is finite for all t, st+1 , which is a joint re-
striction on the income process {yti } and the Arrow security prices q under
consideration. The no Ponzi condition, stated this way, is sometimes called
the natural debt limit. It is easier to check as it only involves the endow-
ment process, but also prices (either Arrow securities prices or Arrow Debreu
prices).
We now have the following
pt+1 (st+1 )
qt (st , st+1 ) = (4.36)
pt (st )
38 CHAPTER 4. THE STANDARD COMPLETE MARKETS MODEL
( ! )
X X
v(a, s) = max U y(s) + a − a0 (s0 )q(s0 |s) +β π(s0 |s)v(a0 (s0 ), s0 )
{a0 (s0 )}s0 ∈S
s0 s0
(4.37)
8
Under assumption 2 and the assumption that {st } is Markov, this assumption is valid.
Using (4.36) and (4.22) we find
pt+1 (st+1 )
q(st , st+1 ) =
pt (st )
−σ
πt+1 (st+1 ) ct+1 (st+1 )
= β
πt (st ) ct (st )
−σ
yt+1 (st+1 )
= βπ(st+1 |st )
yt (st )
= q(st+1 |st )
4.2. EMPIRICAL IMPLICATIONS FOR ASSET PRICING 39
the stream of dividends dj = {djt (st )} where djt (st ) is the amount of consump-
tion goods asset j delivers at node st of the event tree. The period zero (cum
dividend) price of an asset is then simply the value of all the consumption
goods it delivers, that is
∞ X
X
P0j (d) = pt (st )djt (st )
t=0 st
A 0+1 pt (st ) 1
Rt+1 (ŝt+1 ) = t+1 t
= t+1
= t
pt+1 (ŝ )/pt (s ) pt+1 (ŝ ) qt (s , ŝt+1 )
j
and Rt+1 (st+1 ) = 0 for all st+1 6= st+1 .
Example 13 Now consider a one-period risk free bond, that is, a promise
to pay one unit of consumption in every node st+1 tomorrow that can follow
a given node st . The price of such an asset is given by
t+1
P
B t st+1 pt+1 (s ) X
Pt (d; s ) = t
= qt (st , st+1 ).
pt (s ) t+1 s
4.2. EMPIRICAL IMPLICATIONS FOR ASSET PRICING 41
B 1 1
Rt+1 (st+1 ) = =P B
= Rt+1 (st )
PtB (d; st ) s t+1 q t (s t, s
t+1 )
Example 15 An option to buy one share of the Lucas tree at time T (at all
nodes) for a price K has a price Ptcall (st ) at node st given by
X pT (sT )
Ptcall (st ) =
S T
max P T (d; s ) − K, 0
T t
pt (st )
s |s
Such an option is called a call option. A put option is the option to sell the
same asset, and is given by
X pT (sT )
Ptput (st ) = S T
max K − P T (d; s ), 0 .
T t
pt (st )
s |s
The price K is called the strike price (and easily could depend on sT , too).
Attach Lagrange multiplier λt (st ) to this constraint and λt+1 (st+1 ) to the
corresponding constraints (there are as many as there are possible shocks st+1
tomorrow). The first order conditions with respect to ct (st ), ct+1 (st+1 ) and
ajt+1 (st ) read as (remember we assume that asset choices are not constrained
beyond the non-binding No Ponzi conditions)
Substituting the first two equations into the third and re-arranging yields
for all assets j. Thus the unique stochastic discount factor process {mt+1 (st+1 )}
is given by −σ
ct+1 (st+1 )
t+1
mt+1 (s ) = β .
ct (st )
Note that mt+1 (st+1 ) is simply a function of aggregate consumption growth
and the two preference parameters (σ, β). Thus the equation
−σ !
ct+1 (st+1 )
j t+1 t
E β Rt+1 (s )|s = 1
ct (st )
j
is the basic empirical restriction on asset returns {Rt+1 } and consumption
growth {ct+1 /ct } implied by the complete markets model with CRRA prefer-
ences. This equation forms the basis of a large literature that investigates the
asset pricing implications of the complete markets model (or the consumption
capital asset pricing model, CCAPM, as it is sometimes called).
Now let us use and interpret condition (4.40) in the definition of a stochas-
tic discount factor further. Note that (4.40) has to hold for all assets j traded
in this economy, and thus for the risk free bond, in particular. Thus, for the
risk-free bond we have10
B 1
Rt+1 (st ) =
E (mt+1 (st+1 )|st )
1
PtB (d; st ) = t+1 t
B
= E m t+1 (s )|s
Rt+1 (st )
that is, the price of a risk-free bond equals the conditional expectation of
the stochastic discount factor, which perhaps justifies the alternative name
“pricing kernel” more directly.
For ease of notation for what follows, let Et = E(.|st ) denote the condi-
tional expectation. The stochastic discount factor then satisfies, for all assets
j,
j
Et (mt+1 Rt+1 )=1 (4.42)
10 B
Using the fact that Rt+1 (st ) is nonstochastic conditional on st , and employing the
relation between asset prices and returns as the fact that we are considering a one-period
bond that only pays off in period t + 1.
44 CHAPTER 4. THE STANDARD COMPLETE MARKETS MODEL
j j B j B
Et (Rt+1 ) 1 + Et (rt+1 ) 1 + rt+1 + Et (rt+1 ) − rt+1
B
= B
= B
Rt+1 1 + rt+1 1 + rt+1
j B
Et (rt+1 ) − rt+1 j B
= 1+ B
≈ 1 + Et (rt+1 ) − rt+1
1 + rt+1
in (4.43) we obtain
j B j j j
Et (rt+1 ) − rt+1 ≈ −Covt (mt+1 , Rt+1 ) = −ρt (mt+1 , Rt+1 )stdt (mt+1 )stdt (Rt+1 )
(4.44)
for any asset j, the risk-free bond B, and any stochastic discount factor m.
Here ρ is the correlation coefficient and std(.) denotes the standard deviation.
That is, the (conditional) expected excess return of asset j over the risk-free
B
rate rt+1 equals the (negative of) the covariance of the (gross) return of that
asset j with the stochastic discount factor, as long as the real return on the
B
risk free bond rt+1 between period t and t + 1 is close to zero, and thus the
approximation used above is accurate.
Although (4.44) holds true for any asset j and any utility function and
implied stochastic discount factor, the main implications of this equations
j B
have been studied for asset j being stocks, and thus Et (rt+1 ) − rt+1 being
the excess return on equity, and with the utility function of time-separable
CRRA form. In this case equation (4.44) becomes:
" −σ # −σ !
t+1
S B c t+1 S c t+1 (s ) S
Et (rt+1 )−rt+1 = −ρt β , Rt+1 ∗stdt β ∗std t R t+1 .
ct ct (st )
(4.45)
4.2. EMPIRICAL IMPLICATIONS FOR ASSET PRICING 45
One quick way to assess whether the representative agent model has a
chance to rationalize the observed excess returns on equity is the following.
From the data the annual average premium to be explained is roughly 7%.
The annual standard deviation of stock returns is roughly 15.5% in n theodata.
We also observe the time series of aggregate consumption growth ct+1 ct
and
stock
returns Rt+1 in the data. One simple way to ask whether (or, to what
extent) the complete markets model can rationalize the equity premium is
simply to calculate the right hand side for given (σ, β), replacing theoretical
with sample moments. The results of this exercise, which are carried out by
Kocherlakota (1996) and many others since are disastrous for the model, so
let us explain why by interpreting (4.45).
−σ
ct+1 S
First, the is no hope in explaining any premium if ρt ct
, Rt+1 >
0, that is, if (loosely speaking) consumption growth and stock returns are
negatively correlated (since ct+1 /ct is raised to a negative power). Any asset
that pays well in t + 1 (i.e. has high returns between period t and t + 1) in
bad states of the world (low consumption growth ct+1 /ct ) is a good hedge
against consumption risk, and thus does not command any risk premium
(in fact a negative premium). Thus it is the positive correlation between
consumption growth and asset returns11 that generate the equity premium,
rather than the fact that stock returns are very risky. Of course, conditional
S
on a given positive correlation between ct+1 /ct and Rt+1 , the size of the
explained equity premium is increasing in the risk of stock returns and the
volatility of consumption growth. In addition, high values of σ will, for a
given dispersion of consumption growth, help to explain a larger share of the
equity premium, by raising ct+11 /ct to larger powers. Given the empirically
S
weak, but positive correlation between ct+1 /ct and Rt+1 and the low volatility
of ct+1 /ct , the asset pricing literature was led to conclude that the empirically
observed excess returns on stocks can only be rationalized with implausibly
large values of σ. This is a version of Mehra and Prescott’s (1985) statement
of the equity premium puzzle.12
−σ
11 ct+1
Again, the precise statement is: the negative correlation between β ct and
S
Rt+1 .
12
Mehra and Prescott (1985) considered values for σ above 10 as implausible and showed
that for values of σ below that threshold the model-implied equity premium is smaller than
the target in the data by factor of at least 10.
46 CHAPTER 4. THE STANDARD COMPLETE MARKETS MODEL
−σ
ct+1 S
An alternative way to state the puzzle is to realize that −ρt β ct , Rt+1 ≤
1, since correlations are bounded between −1 and 1. Then equation (4.45)
implies that
The entity on the left hand side is the so-called Sharpe ratio, which gives
the excess return a given asset (here stocks) commands per unit of risk (as
measured by the standard deviation of the return). The equation gives an
upper bound for that ratio: it cannot be larger than the volatility of the
stochastic discount factor. This bound, derived in various forms by Shiller
(1982) and Hansen and Jagannathan (1991), is valid for any asset (and any
stochastic discount factor). It indicates that a model that does not generate
a volatile enough stochastic discount factor has no chance in explaining the
equity premium puzzle. For annual data, the Sharpe ratio for stocks is in
the order of about 0.3 − 0.5, whereas the standard deviation of consumption
growth is not larger than 0.035 (the exact value depends somewhat on the
sample period). Thus without large values for σ (and perhaps for all possible
σ) the Hansen-Jagannathan bound is clearly violated by the unique stochastic
discount factor implied by the complete markets model: consumption growth
is simply too smooth in the data to generate a large enough equity premium,
even if it was perfectly correlated with stock returns (which it is not, either).
But was is the problem with a large σ, absent direct empirical evidence13
on its value? First, one may argue that households endowed with such large
risk aversion would make other choices (e.g. buy large amounts of all kinds of
insurance) that are at odds with the data. what is the problem with assuming
a large value for σ? But perhaps more importantly, such high assumed values
would raise a second asset pricing puzzle. To see why, note that the risk-free
13
Work by Barsky et al. (1997) and Kimball, Sahm and Shapiro (2008, 2009) has used
survey responses from the HRS and PSID to shed some light on the empirical distribution
of risk aversion in the U.S. population.
4.2. EMPIRICAL IMPLICATIONS FOR ASSET PRICING 47
rate satisfies
B 1 1 1
Rt+1 = = −σ =
Et (mt+1 ) ct+1 1
Et β ct βEt ct+1 σ
ct
Since in the data consumption is growing over time, thus ct+1 /ct tends to
be positive and on average around 1.02 in U.S. data. Therefore making σ
B
large makes Et (mt+1 ) small, and thus Rt+1 large. Why is this the case?
1
A large σ implies small IES = σ . With a small IES households desire a
smooth consumption profile. But consumption grows in the data. Thus a
large interest rate is needed to persuade them to postpone consumption. In
the data, however, the real risk free rate is small, around 1% on average.
This is Weil’s (1989) risk free rate puzzle.
One possible resolution to this puzzle is simply to make β large (larger
than 1!).14 As long as β(1 + gc )1−σ < 1 (on average), there is no problem for
lifetime utility to converge. But β > 1 and the associated degree of patience
might be considered implausible. Alternatively one might contemplate util-
ity functions in which a households’ attitude towards risk (variation of con-
sumption across states of the world) and towards intertemporal smoothing
(variation of consumption over time) is not governed by the same parameter.
Epstein and Zin (1989. 1991) propose such a class of utility functions:
Remark 17 Epstein and Zin (1989, 1991), extending earlier work by Kreps
and Porteus (1978) propose a class of utility functions that deviate from
expected utility and that can be written as
1
1 1
" # 1− γ
1−σ 1− γ
1 X
u(c, st ) = ct (st )1− γ + β πt (st+1 |st )u(c, st+1 )1−σ
st+1
st+1 . With these preferences the IES is controlled by the parameter γ, and
risk aversion is controlled by the separate parameter σ. These preferences
have become extremely popular in modern consumption-based asset pricing
because they help to jointly explain low risk-free rates as well as high excess
returns on stocks.
4.2.1 An Example
In this section we present an example for which we can solve for the risk free
rate and the equity premium in closed form.15 This is useful since in the ex-
ample it becomes fully transparent which parameters determine the risk-free
rate and the equity premium. The economy is populated by representative
consumers and the income of these households comes from dividends. House-
holds trade risk-free bonds in zero net supply and shares of the stock which
entitles them to the dividends of the stock.16 Also suppose that dividends
follow the process
log(dt+1 ) = log(dt ) + g + ut+1
where ut+1 is iid normally distributed with zero mean and variance σu2 . Since
dividends are the only form of income to the representative household we have
ct = dt for all t and all states of the world. Thus consumption satisfies ct+1
ct
=
g+ut+1 j
e . We can now use this in our asset pricing equation Et (mt+1 Rt+1 ) = 1.
For the risk-free rate we obtain
" −σ #
B c t+1
Rt+1 Et β = 1
ct
B
Et βe−σ(g+ut+1 ) = 1
Rt+1
B
βe−σg Et e−σut+1 = 1
Rt+1
Now we note that because ut+1 is normal with zero mean and variance σu2 we
1 2 2
have Et [e−σut+1 ] = e 2 σ σu and thus the risk-free rate is given by
1 σ[g− 12 σσu2 ]
RB = e
β
15
This is a simplified version of the example presented in Barro (2009) which excludes
disaster risk.
16
Whether or not households also trade a full set of Arrow securities is not crucial in
this representative agent economy since there is no trade in equilibrium.
4.2. EMPIRICAL IMPLICATIONS FOR ASSET PRICING 49
or
1
rB = log(RB ) = ρ + σg − σ 2 σu2 (4.46)
2
1
where we used the definition of the time discount rate ρ as β = 1+ρ and
the approximation ρ ≈ log(1 + ρ). We see that the more patient households
are (the larger is ρ), the higher is the risk free rate. A higher consumption
(dividend) growth rate g and a lower intertemporal elasticity of substitution
raise the risk-free rate as well, for the reason discussed above. Finally, the
last term shows that the risk-free rate falls with the amount of consumption
risk. As we will discuss below, with CRRA preferences households have a
precautionary savings motive (whose size is governed by the parameter σ
as well) that is more potent the larger is consumption risk. To insure that
households are willing to consume their dividends in the presence of the extra
incentive to save, the risk free rate has to fall. This rationalizes the negative
last term in equation (4.46).
Solving for the expected return on equity is harder since we don’t know
the endogenous price of stock PtS . But note that
S
S
S dt+1 + Pt+1 dt+1 Pt+1
Rt+1 = S
= S
+1 .
Pt Pt+1 PtS
t+1 PS
Now conjecture that the price-dividend ratio dt+1 is a time-and state invariant
S
constant so that Pt+1 = κdt+1 . Then
S
S dt+1 Pt+1 1 dt+1 ct+1
Rt+1 = S
+1 S
= +1 =θ
Pt+1 Pt κ dt ct
where θ = κ1 + 1 is a yet to be determined constant. Again using the asset
1 2 2
Again using Et e(1−σ)ut+1 = e 2 (1−σ) σu we have
1 (1−σ)[−g− 12 (1−σ)σu2 ]
θ= e
β
50 CHAPTER 4. THE STANDARD COMPLETE MARKETS MODEL
Thus
and thus is simply an increasing function of risk aversion and the volatility of
consumption (dividends). For similar analyses using Epstein-Zin preferences,
see e.g. Weil (1989), Tallarini (2000) or Barro (2009).
λ(st )
1 X j t 1 X j t 1 X j 1
ln(ct (s )) = b (s ) + ln(α ) − ln (4.51)
N j σN j σN j σ β t πt (st )
1 λ(st )
and using this equation to substitute out ln σ β t πt (st )
in equation (4.50)
yields:
! !
1 1 X 1 1 X
ln(cit (st )) = bi (st ) − bj (st ) + ln(αi ) − ln(αj )
σ N j σ N j
1 X
+ ln(cjt (st ))
N j
1 i t 1
b (s ) − ba (st ) + ln(αi ) − ln(αa ) + ln(cat (st )) (4.52)
=
σ σ
where
1 X j t
ba (st ) ≡ b (s )
N j
1 X
ln(αa ) ≡ ln(αj )
N j
1 X
ln(cat (st )) ≡ ln(cjt (st )) (4.53)
N j
are population averages. Note that we have abused notation somewhat in the
last two expressions, which are sums of logs rather than logs of sums. Finally,
by taking first differences of (4.52) to get rid off the term σ1 (ln(αi ) − ln(αa ))
we arrive at an equation that can be brought to the data (and again sup-
pressing dependence on st ):
1
∆ ln(cit ) = ∆ ln(cat ) + ∆bit − ∆bat
(4.54)
σ
4.3. TESTS OF COMPLETE CONSUMPTION INSURANCE 53
Thus, with this specification of preferences we can test the complete in-
surance hypothesis with the empirical specification
where the error it now captures individual and aggregate changes in pref-
erence shocks as well as potentially measurement error.17 The null hypoth-
esis of complete consumption insurance implies that α1 = 1 and α2 = 0.
Thus the regression equation used to test for complete consumption insur-
ance remains substantially unchanged under the more general utility specifi-
cation with preference shocks, keeping in mind that in the definition above,
ln(cat (st )) 6= ln(ct ). Note that the basic prediction of the theory is that indi-
vidual household i’s consumption should not respond to idiosyncratic shocks
once aggregate consumption is accounted for. Idiosyncratic income shocks
are just one, arguably important, but not the only source of idiosyncratic
risk. Including other measures of idiosyncratic shocks in regression (4.55)
and testing the complete consumption insurance by investigating whether
the corresponding regression coefficient(s) equals zero is equally valid, from
the perspective of the theory presented so far. See especially Cochrane (1991)
on this point.
Finally, we can derive a similar specification under the assumption that
preferences take a Constant Absolute Risk Aversion (CARA) form. Suppose
the period utility function is given by:
1
U i (cit (st ), st ) = − e−γ (ct (s )−b (s ))
i t i t
(4.56)
γ
where γ is the coefficient of absolute risk aversion, then, using the same
manipulations as before one obtains
∆cit = ∆cm i m
t + ∆bt − ∆bt (4.57)
17
Note that, defining
1
it = ∆bit − ∆bat
σ
then the cross-sectional expectation of it across households i equals zero, by definition of
bat .
54 CHAPTER 4. THE STANDARD COMPLETE MARKETS MODEL
where
1 X i
cm
t = c
N j t
1 X i
bm
t = b (4.58)
N j t
Thus, under exponential utility the empirical specification used to test com-
plete consumption insurance is
∆cit = α1 ∆cm i i
t + α2 ∆yt + t (4.59)
The results in Table 1 are, for the most part, inconsistent with the com-
plete insurance hypothesis. For all consumption groups the F-test is deci-
sively rejected at the 5% confidence level (in fact it would be rejected at the
1% confidence level). Loosely speaking, this is mostly due to the fact that in-
dividual income growth does help to explain individual consumption growth,
20
The test statistic for the F-test is distributed as an F (2, N − 3), where N is the
number of usable household observations in the sample. For the difference specification
N = 10, 695 whereas for the growth rate specification it varies by consumption category
because households with nonpositive consumption in either interview have to be excluded.
A ∗ after the test statistic indicates that the null hypothesis can be rejected at the 95%
confidence level.
56 CHAPTER 4. THE STANDARD COMPLETE MARKETS MODEL
For this specification the results are more supportive of perfect consump-
tion smoothing. In particular, the F-test is rejected only for nondurable
consumption expenditures, again primarily due to the fact that individual
consumption growth is sensitive to individual income growth. Note that here
the intercept is significant for almost all measures of consumption (in theory
it should be zero). It is these results, in addition to the findings that the es-
timated value of α2 is quantitatively small, that lead Mace to the conclusion
4.3. TESTS OF COMPLETE CONSUMPTION INSURANCE 57
that perfect risk sharing is a good first approximation of the data, at least
for exponential utility.
However, as Nelson (1994) argues, if one excludes households classified as
“incomplete income reporters” by the CEX (households that either did not
respond to certain income questions or gave inconsistent answers) and uses
expenditures for the entire quarter as the relevant measure of consumption
(as opposed to Mace who used only the expenditure in the month preceding
the interview) one finds a strong rejection of complete consumption insurance
even for the CARA utility specification.
i.e. for the moment we ignore the presence of aggregate consumption growth.
Note that if we had a single cross-section of first differenced household obser-
vations, one would exactly run a regression of the from (4.60).21 For simplicity
also assume that in the cross-section E (∆yti ) = 0, that is, individual income
changes are measured in deviation from aggregate income changes.22 The
null hypothesis of perfect risk sharing implies α̂2 = 0. For ease of exposition
we abstract from a constant in the regression and assume α0 = 0.
Suppose we have N household observations for one period t, and suppose
that the income variable is measured with error
where ∆zti is the measured income change, ∆yti is the true income change
and ∆vti is an additive measurement error satisfying E (∆vti ) = 0, where E(.)
is the cross-sectional (across households) expectation, recalling that we hold
t fixed. Furthermore assume that V ar (∆vti ) = σv2 , that V ar (∆yti ) = σy2 and
that ∆vti , it and ∆yti are all mutually independent.23
The equation we estimate is then
N −1 N j j
P
j=1 ∆ct ∆zt
plimN →∞ α̂2 = plimN →∞ j 2
N −1 N
P
j=1 ∆zt
N −1 N j j
P i i
j=1 (α2 ∆yt + t ) ∆yt + ∆vt
= plimN →∞ j 2
N −1 N j
P
j=1 ∆yt + ∆vt
i 2
plimN →∞ N −1 N
P
j=1 (∆yt )
= α2 j 2 j 2
plimN →∞ N −1 N
P
−1
PN
j=1 ∆y t + plim N →∞ N j=1 ∆vt
σy2 1
= α2 · 2 2
= α2 · 2
σy + σv 1 + σσv2
y
be correlated with the error term and the standard consumption insurance
regressions will yield inconsistent result. Future versions of these notes will
discuss the papers by Schulhofer-Wohl (2011) and Mazzocco and Saini (2012)
in greater detail.
Consider the following example based on Mazzocco and Saini (2012):
∞ X
2 X
X
max αi β t πt (st )U i (cit (st ))
{cit (st ),lti (st )}
i=1 t=0 st ∈S t
s.t.
N
X
cit (st ) = Ȳt (st )
i=1
cit (st ) ≥ 0
• Plotting αi Uci (cit (st )) against cit (st ), the curves for i = 1 and i = 2
intersect once (and only once). If σ1 = σ2 , they are parallel to each
other (and on top of each other if α1 = α2 ).
• Plotting the efficient cit (st ) against Ȳt (st ), the curves (expenditure func-
tions) intersect once (and only once) if σ1 < σ2 . Let intersection point
be denoted by Ȳ ∗ .
where
2
1X
log cat+1 = log cit+1
2 i=1
• Now suppose Ȳt < Ȳ ∗ < Ȳt+1 , then with efficient risk sharing and
σ1 < σ2 we have
• In bad times ∆Ȳt+1 < 0, ∆ log cit+1 is particularly low for low σ house-
holds (thus εit+1 < 0) and ∆ log yt+1
i
is low for these households. Re-
versely for high σ households
s.t.
N
X XN
cit (st ) = wti (st )lti (st ) + Yt (st )
i=1 i=1
cit (st ) ≥ 0, lti (st ) ∈ [0, 1]
• Step 3: Conditional on given ρit (st ), what is the efficient split between
consumption and leisure
together with constraint determines indirect utility function V i (ρit (st ), wti (st )).
• Note: if labor supply is exogenous, this step is trivial and V i (ρit (st ), wti (st )) =
U i (cit (st ))
V i,j (ρi,j t i t j t
t (s ), wt (s ), wt (s ))
= maxj αi V i (ρit (st ), wti (st ))
ρit (st ),ρt (st )
• Assume that number of households is even (if not, need to group three
households into one group).
N/2
X
max V 2i−1,2i (ρt2i−1,2i (st ), wt2i−1 (st ), wt2i (st ))
{ρt2i−1,2i (st )} i=1
N/2
X
s.t. ρ2i−1,2i
t (st ) = Ȳt (st ).
i=1
wi (st ) = wi (ŝτ )
wj (st ) = wj (ŝτ )
ρi (ρi,j t i t j t j i,j t i t j t
t (s ), wt (s ), wt (s )) > ρ (ρt (s ), wt (s ), wt (s ))
ρi (ρi,j t i t j t j i,j t i t j t
t (ŝ ), wt (s ), wt (s )) < ρ (ρt (ŝ ), wt (s ), wt (s ))
• Key: weather is very important for these villages, and has lots of annual
and seasonal variation. Life is risky there.
• Monthly data from 1975 - 1985. About 120 observations for about 30
households in each village.
• Efficient risk sharing tests: results fairly uniform across the two tests.
For few (less than 5%) of pairs efficient risk sharing rejected. But
enough to still formally reject efficient risk sharing on village level.
where
N X
X
T
v0 (st ) yti (st ) − cit (st )
SN (s ) = (4.65)
t=T st |sT
is the time zero value of future income net of consumption (that is, the time
zero value of period savings) from node sT up to date N. Note that for a
fixed sT , the sequence {SN (sT )} need not converge24 , but if it does, then
∞ X
X
T
v0 (st ) yti (st ) − cit (st )
lim inf SN (s ) = (4.66)
N →∞
t=T st |sT
x∗ = lim inf xn
n→∞
x∗ = lim xn
n→∞
x∗ = −∞
68 CHAPTER 4. THE STANDARD COMPLETE MARKETS MODEL
rules out a sequence of asset holdings and consumption for which, at any
node st+1 , the beginning of the period debt ait+1 (st , st+1 ) exceeds the node
SN (st+1 )
st+1 of future savings, − lim inf
v0 (st+1 )
.
Let us interpret the No Ponzi scheme condition a bit further. Writing out
(4.64) explicitly using (4.65) we obtain:
N X
X
−v0 (sT )aiT (sT −1 , sT ) ≤ lim inf v0 (st ) yti (st ) − cit (st )
(4.68)
N →∞
t=T st |sT
Now suppose that the sequential budget constraint holds with equality (which
it will as an optimum as long as the utility function is strictly increasing).
Then
X
yti (st ) − cit (st ) = qt (st , st+1 )ait+1 (st , st+1 ) − ait (st ) (4.69)
st+1
as the value of the asset portfolio bought in node st . Using equation (4.70),
equation (4.68) becomes
N X
X
−v0 (sT )aiT (sT −1 , sT ) ≤ v0 (st ) qt (st ) · ait+1 (st ) − ait (st )
lim inf
N →∞
t=T st |sT
lim inf v0 (sT )qT (sT ) · aiT +1 (sT ) − v0 (sT )aiT (sT )
=
N →∞
X
+ v0 (sT +1 )qT +1 (sT +1 ) · aiT +2 (sT +1 ) − v0 (sT +1 )aiT +1 (sT +1 )
sT +1 |sT
X
+... + v0 (sN )qN (sN ) · aiN +1 (sN ) − v0 (sN )aiN (sN )
sN |sT
X
= lim inf −v0 (sT )aiT (sT ) + v0 (sN )qN (sN ) · aiN +1 (sN )
N →∞
sN |s T
(4.71)
4.5. APPENDIX B: PROOFS 69
The last equality is due to the fact that all intermediate terms cancel out.
We demonstrate this for the first terms
X
v0 (sT )qT (sT ) · aiT +1 (sT ) = v0 (sT ) qT (sT , sT +1 )aiT +1 (sT , sT +1 )
sT +1
(4.72)
X X
v0 (sT +1 )aiT +1 (sT +1 ) = v0 (sT )qT (sT , sT +1 )aiT +1 (sT +1 )
sT +1 |sT sT +1 |sT
X
= v0 (sT ) qT (sT , sT +1 )aiT +1 (sT , sT +1 )
sT +1
(4.73)
Simplifying equation (4.71) we can therefore conclude that the no Ponzi
scheme condition, as stated in (4.64), can equivalently be written as
X
lim inf v0 (sN )qN (sN ) · aiN +1 (sN ) ≥ 0 (4.74)
N →∞ N T
s |s
and thus rules out asset sequences for which the present value of future debt
is bounded above zero. Using this No Ponzi condition Wright (1987) then
states and proves proposition 11 in the main text.
N X
X
pt (st ) yti (st ) − cit (st )
0 ≤ lim inf
N →∞
t=T st |s0
N X
X
pt (st ) yti (st ) − cit (st )
= lim inf
N →∞
t=0 st
∞ X
X
pt (st ) yti (st ) − cit (st )
= (4.77)
t=0 st
and recursively
X
aN (st ) = cit (st ) − yti (st ) + qt (st , st+1 )aN (st , st+1 ) (4.79)
st+1
25
The last equality is valid whenever the series converges. One can circumvent the prob-
lem of possible nonconvergence by adjusting the definition of Arrow Debreu equilibrium:
one defines the Arrow debreu budget constraint as
p0 (ci − y i ) ≤ 0
where
T X
X
p0 c = lim inf pt (st )ct (st )
T →∞
t=0 st
Then all arguments go through even if, for a given price process p, the infinite sum does
not converge for some process c. For details see Wright (1987), in particular footnote 6.
4.6. APPENDIX C: PROPERTIES OF CRRA UTILITY 71
st+1 |st
(4.80)
With aN (sN ) = 0 for all sN we have, via continuing substitution
N X
X
T N T
v0 (st ) cit (st ) − yti (st )
v0 (s )a (s ) = (4.81)
t=T st |sT
and hence the sequential budget constraint. Taking lim inf 0 s on both sides of
(4.81) yields
N X
X
T T −1
v0 (st ) yti (st ) − cit (st )
−v0 (s )aT (s , sT ) = lim inf (4.84)
N →∞
t=T st |sT
and hence (ci , a) as constructed above satisfies the no Ponzi scheme condition.
1−σ
with σ > 0. Note that limσ→1 c 1−σ−1 = ln(c) which justifies the second line
in the definition of the CRRA utility function above. First we note that
U satisfies the “usual” properties: U is continuous, three times continu-
ously differentiable, strictly increasing (i.e. U 0 (c) > 0), strictly concave (i.e.
U 00 (c) < 0) and satisfies the Inada conditions
lim U 0 (c) = +∞
c&0
lim U 0 (c) = 0
c%+∞
that is, as the inverse of the percentage change in the marginal rate of substi-
tution MRS between consumption at t and t + 1 in response to a percentage
change in the consumption ratio ct+1 ct
. For the CRRA utility function note
that
∂u(c) −σ
∂ct+1 ct+1
∂u(c)
= M RS(ct+1 , ct ) = β
ct
∂ct
and thus
−1
−σ−1
−σβ ct+1
ct
1
iest (ct+1 , ct ) = − =
c −σ
σ
t+1
β c
t
ct+1
ct
where rt+1 is the real interest rate between periods t and t + 1. Thus, for any
model in which the intertemporal Euler equation holds with equality the IES
can alternatively be written as:
ct+1 ct+1
d ct
d ct
ct+1 ct+1
ct ct
iest (ct+1 , ct ) = −
∂u(c)
= −
∂ct+1 d 1+r1
d t+1
∂u(c) 1
∂ct
1+rt+1
∂u(c)
∂ct+1
∂u(c)
∂ct
that is, the IES measures the percentage change in the consumption growth
rate in response to a percentage change in the gross real interest rate, the
intertemporal price of consumption.
Note that for the CRRA utility function the Euler equation reads as
−σ
ct+1
(1 + rt+1 )β = 1.
ct
Taking logs on both sides and rearranging one obtains
or
1 1
ln(ct+1 ) − ln(ct ) =
ln(β) + ln(1 + rt+1 ). (4.86)
σ σ
This equation can then be used to obtain empirical estimates of the IES.
With time series data on consumption growth and real interest rates the IES
1
σ
can be estimated from a regression of the former on the later.27
27
Note that in order to interpret (4.86) as a regression one needs a theory where the
error term comes from. In models with risk this error term can be linked to expectational
errors, and (4.86) with error term arises as a first order approximation to the stochastic
version of the Euler equation. We will discuss the stochastic Euler equation at length in
part II of this monograph.
4.6. APPENDIX C: PROPERTIES OF CRRA UTILITY 75
29
Note that the class of (non-time separable) Epstein-Zin utility functions discussed in
remark 17 above inherit the homotheticity property from the time separable lifetime utility
with CRRA period utility.
Part II
77
79
Permanent income type models assume that agents do not have access to a
complete set of contingent consumption claims. The main difference between
this type of models and the complete markets model thus manifests itself in
the budget constraints. Whereas the complete markets model in sequential
formulation has budget constraints of the form (4.28), for the SIM model we
have1
ct (st ) + qt (st )at+1 (st ) = yt (st ) + at (st−1 ) (5.1)
Here qt (st ) is the price at date t, event history st , of one unit of consumption
delivered in period t + 1 regardless of what event st+1 is realized. Notation
is crucial here: at+1 is the quantity of risk-free one period bonds being pur-
chased in period t and paying off in period t + 1. Thus at+1 (st ) is a function
only of st and not of st+1 whereas in the complete markets model with a full
set of Arrow securities these assets are indexed by (st , st+1 ) = st+1 and each
asset pays off only at a particular event history st+1 . Obviously each of these
Arrow securities has a potentially different price qt (st , st+1 ), as discussed in
the previous chapter.
The basic income fluctuation problem is to maximize
T X
X
u(c) = β t πt (st )U (ct (st ), st ) (5.2)
t=0 st ∈S t
1
In principle the budget constraints have to hold only as inequalities. At the optimal
allocations they will always hold as strict equality as long as household utility is strictly
increasing in consumption, an assumption we maintain throughout.
81
82 CHAPTER 5. THE SIM IN PARTIAL EQUILIBRIUM
where 1 + r = 1q is the gross real interest rate. Equation (5.7) together with
the Euler equation
U 0 (c0 ) = β(1 + r)U 0 (c1 ) (5.8)
uniquely determine the optimal consumption allocation (c0 , c1 ) over the life
cycle of the household, as a function of the model parameters (y0 , y1 , r, β) and
any parameters characterizing the period utility function (e.g. the parameter
σ in the case of CRRA utility).
3. Provided that the household has income in the second period, an in-
crease in r reduces the present value of lifetime income W0 (y0 , y1 , r, a0 )
and hence reduces current and future consumption. This is the human
capital or wealth effect.
84 CHAPTER 5. THE SIM IN PARTIAL EQUILIBRIUM
Which effect dominates depends on the particular form of the utility func-
tion and the income profile. We can at least partially rank the magnitudes
of these effects for the case in which households have CRRA period utility,
discussed in the previous chapter. Recall that the parameter 1/σ measures
the intertemporal elasticity of substitution as discussed in the appendix of
the previous chapter and hence the potency of the substitution effect.3
In the case of CRRA utility the explicit solution of the model is given by:
W0 (y0 , y1 , r, a0 )
c0 = 1 1 (5.10)
β σ (1+r) σ
1+ 1+r
1 1
β σ (1 + r) σ W0 (y0 , y1 , r, a0 )
c1 = 1 1 (5.11)
β σ (1+r) σ
1+ 1+r
1
The dependence of W0 on 1 + r captures the wealth effect, the term (1 + r) σ
captures the substitution effect (absent if 1/σ = 0) and the term 1 + r in the
second denominator encompasses the income effect. Since
∂W0 (y0 , y1 , r, a0 ) y1
=− ≤0
∂(1 + r) (1 + r)2
the human wealth effect is negative for consumption in both periods, and is
arbitrarily small or large depending on the size of future income y1 , and thus
can either dominate, or is being dominated by the other two effects.
Now suppose the wealth effect is absent (e.g. because y1 = 0). Then
−2 1 !
∂c0 1 1
−1 βσ 1 1 1
= W0 (y0 , y1 , r, a0 ) ∗ 1 + β (1 + r)
σ σ ∗ (1 + r) − (1 + r) σ
σ
∂(1 + r) (1 + r)2 σ
1 1 !
1 1
−2 β σ (1 + r) σ 1
−1
= W0 (y0 , y1 , r, a0 ) ∗ 1 + β (1 + r)
σ σ ∗ 2
1− (5.12)
(1 + r) σ
The first term in the last bracket is the positive income effect, the second term
the negative intertemporal substitution effect. expression the human capital
effect, the second term is the combination between income and substitution
effect. For CRRA utility the relative magnitude of these two effects only
3
In the absence of risk, risk aversion (as measured by σ) of households is irrelevant for
the interpretation of the solution.
5.2. THE GENERAL MODEL WITH CERTAINTY EQUIVALENCE 85
would make in the absence of risk. The following section will then relax the
assumptions leading to certainty equivalence and study versions of the model
in which households engage in precautionary saving behavior.4
For the rest of this chapter we impose the following assumption on the
exogenous prices (interest rates) of the one period risk-free bond.
Assumption 18 The price of the one period bond is nonstochastic and con-
stant over time,
1
qt (st ) = q = .
1+r
the present value of all future labor income, including initial wealth. We
assume that W0 is finite (otherwise the household maximization problem does
not have a solution).5 Forming the Lagrangian, taking first order conditions
and combining yields the standard Euler equation
and assuming that the sum is finite for all t, we require that for all t bond holdings adhere
to the “natural borrowing limit”
at+1 ≥ −Wt+1 (5.14)
Note that Wt+1 is finite if r > 0 and the sequence {yt } is bounded from above.
5.2. THE GENERAL MODEL WITH CERTAINTY EQUIVALENCE 87
Specific Case ρ = r
If ρ = r then the right hand side of (5.16) is constant over time and thus
Uc (ct , st ) is time-invariant. In periods in which the preference shifter st makes
marginal utility high, consumption also has to be high, since marginal utility
is decreasing in consumption. This may provide us with a nontrivial theory
of life-cycle consumption profiles even in the absence of risk, and even under
the assumption ρ = r.
Example 19 If U is separable between consumption and preference shifters
st , then we immediately obtain that household consumption is constant over
time (over the life cycle)
ct = ct+1 (5.18)
something that seems counterfactual6 in light of the discussion in chapter
3. We can combine () with the intertemporal budget constraint () to solve
explicitly for the level of consumption. This yields, for all t,
−1 rW0
θ 1+r if T < ∞
c0 = ct = ct = r (5.19)
1+r
W0 if T = ∞
6
Note that under the separability assumption, even if ρ 6= r, consumption does not
display a life-cycle hump shape, but rather monotonically trends upwards (if r > ρ, i.e.
if incentives to postpone consumption dominate impatience) or monotonically downwards
(if ρ > r, i.e. if impatience dominates incentives to postpone consumption).
88 CHAPTER 5. THE SIM IN PARTIAL EQUILIBRIUM
where θ = 1 − (1+r)1 T +1 is an adjustment factor needed if the lifetime hori-
zon T of the household is finite. Thus consumption at each date is equal to
“permanent income”
rW0
ct =
θ (1 + r)
At each date households consume annuity value of their lifetime wealth W0 ,
composed of initial financial wealth and the present discounted value of life-
time labor income. This is the purest version of the permanent income hy-
pothesis: consumption at each date should equal permanent income.
General Case ρ 6= r
For a fixed sequence of preference shifters {st }, a decline in the interest rate
r makes the life cycle consumption profile less steep, as future consumption
is substituted in favor of current consumption, ceteris paribus. The reverse
is true for an increase in the interest rate. Note that this is not a statement
about the level of the consumption profile but rather its slope, because the
effect of interest rate changes on consumption levels also depends on the
size of the income and human wealth effects, as discussed in the previous
section. Thus the Euler equations only capture the substitution effect and
make predictions about the shape of the consumption profile over the life
8
See Heckman (1974) for the original source of this point. A hump-shaped profile
of hours worked in turn would be part of the optimal household labor supply choice in
the presence of a hump-shaped life cycle wage profile (at least under the appropriate
assumptions on household preferences).
9
Aguiar and Hurst (2005, 2007) argue that older households (especially those in retire-
ment) pay lower prices for the same consumption goods because they spend more time
shopping for good deals. This results in a decline in life cycle consumption expenditures to-
wards the end of life, without necessarily implying a fall in utility-generating consumption
services that enter the utility function.
90 CHAPTER 5. THE SIM IN PARTIAL EQUILIBRIUM
1
∆ ln ct+1 = [ln(1 + r) − ln(1 + ρ)] (5.24)
σ
Using the Euler equations and the intertemporal budget constraint one can
solve for the optimal consumption allocation, which, for all t ≥ 0, is given
by:
t
1+r σ 1−γ
ct = W0 (5.25)
1+ρ 1 − γ T +1
≡ M P C(t, T ) ∗ W0 (5.26)
σ1
10 [1+r]1−σ
where γ is defined as γ = . Holding lifetime income W0 fixed we
1+ρ
find that an increase in the lifetime horizon T reduces the marginal propen-
sities to consume out of lifetime wealth M P C(t, T ) at all ages t.
the event history st budget constraint and taking first order conditions with
respect to ct (st ) and ct+1 (st+1 ) yields
β t πt (st )Uc (ct (st ), st ) = λt (st ) (5.30)
β t+1 πt+1 (st+1 )Uc (ct+1 (st+1 ), st+1 ) = λt+1 (st+1 ) (5.31)
Taking first order conditions with respect to at+1 (st ) yields
X
λt (st )q = λt+1 (st+1 ) (5.32)
st+1 |st
Combining yields
1 X t+1
β t πt (st )Uc (ct (st ), st ) = β πt+1 (st+1 )Uc (ct+1 (st+1 ), st+1 ) (5.33)
q t+1 t
s |s
for all sT .
For the infinite horizon case, T = ∞, we impose
∞
X yτ +1 (sτ +1 )
at+1 (st ) ≥ − inf (5.28)
{sτ +1 }|st
τ =t+1
(1 + r)τ −(t+1)
≡ − inf Wt+1 (st ) ≡ −Āt+1 (st ) (5.29)
{sτ +1 }|st
where the right hand side is the minimum (infimum, if the minimum does not exist)
realized present discount value of future income over all infinite histories {sτ +1 } that can
follow node st . We assume that the process {Āt+1 (st )} is bounded above, which is true
as long as r > 0 and current income only depends on the current shock: yτ (sτ ) = y(sτ ).
In that case define
ymin = min y(st )
st ∈S
and
∞
X yτ +1 (sτ +1 )
Āt+1 (st ) = inf
{s τ +1
}|st
τ =t+1
(1 + r)τ −(t+1)
∞
X 1
= ymin τ −(t+1)
τ =t+1
(1 + r)
ymin (1 + r)ymin
= 1 =
1 − 1+r r
This No Ponzi constraint, the natural extension of the concept of the “natural borrowing
constraint” to a stochastic income process, insures that households having incurred the
maximal amount of debt can repay it with probability 1 (by setting consumption to zero
in all future periods). As long as the utility function satisfies the Inada conditions this
constraint will never be binding.
92 CHAPTER 5. THE SIM IN PARTIAL EQUILIBRIUM
or
X
t t 1+r
Uc (ct (s ), s ) = πt+1 (st+1 |st )Uc (ct+1 (st+1 ), st+1 )
1+ρ
st+1 |st
1+r
E Uc (ct+1 (st+1 ), st+1 )|st
= (5.34)
1+ρ
or, more compactly,
t 1+r
Et Uc (ct+1 , st+1 )
Uc (ct , s ) = (5.35)
1+ρ
where Et is the expectation of st+1 conditional on st . This is the standard
stochastic Euler equation for the standard incomplete markets (SIM) model.
We immediately see that if ρ = r, then the marginal utility of consumption
process {Uc (ct , st )} follows a martingale, i.e. its period t expectation of the
t + 1 variable equals the t variable.12
Remark 23 Recall that for the standard complete markets (SCM) model,
t+1 |st )
under the assumption13 that Arrow securities prices satisfy qt (st+1 ) = πt+1 (s
1+r
12
A martingale is a stochastic process {xt } that satisfies, with probability 1,
Et (xt+1 ) = xt .
Et (xt+1 ) ≥ xt
Et (xt+1 ) ≤ xt
Et ct+1 = α1 + α2 ct (5.38)
1+ρ 1+ρ
with α1 = c̄ 1 − 1+r
and α2 = 1+r
. If furthermore ρ = r, then equation
(5.38) turns into:
Et ct+1 = ct (5.39)
i.e. not only marginal utility but consumption itself follows a martingale.
We summarize the most important implications of (5.38) or (5.39) as:
3. Note, however, that the income realization in period yt+1 can affect
the choice ct+1 , but only through that part that constitutes a deviation
16
Of course, the assumption ρ = r is not required for the result, as it is straightforward
to show that, with quadratic utility, the consumption dynamics in the no-risk case is given
by
ct+1 = α1 + α2 ct
compared to equation (5.38) in the main text.
5.2. THE GENERAL MODEL WITH CERTAINTY EQUIVALENCE 95
T −t X T −t X
X πt+τ (st+τ |st )ct+τ (st+τ ) t−1
X πt+τ (st+τ |st )yt+τ (st+τ )
≤ a t (s ) +
τ =0 t+τ t
(1 + r)τ τ =0 t+τ t
(1 + r)τ
s |s s |s
(5.43)
T −t
! T −t
!
X ct+τ t X yt+τ t
E s ≤ at (st−1 ) + E s ≡ Wt (st )
τ =0
(1 + r)τ τ =0
(1 + r)τ
(5.44)
or, in short,
T −t T −t
X ct+τ X yt+τ
Et ≤ Et + at ≡ W t (5.45)
τ =0
(1 + r)τ τ =0
(1 + r)τ
Using this result in the intertemporal budget constraint (5.45), which at the
optimal allocation holds with equality, we can solve for consumption in closed
22
Since
Et ct+1 = ct
Et ct+2 = Et Et+1 ct+2 = Et ct+1 = ct
98 CHAPTER 5. THE SIM IN PARTIAL EQUILIBRIUM
form23 :
θt−1 rW
1+r
t
if T < ∞
ct = r (5.46)
1+r
W t if T = ∞
1
where θt = 1 − (1+r)T −t+1
.
Note that this equation holds for the fixed initial node s0 as well, so that (say,
for T = ∞),
∞
!
r X yt+τ
c0 (s0 ) = a0 + E 0
1+r τ =0
(1 + r)τ
where a0 is the exogenous initial asset position the household starts her life
with (whereas for future periods Wt (st ) is a function of the endogenous choice
at (st−1 )). Compare this to the consumption function under certain income
in (5.19): in period 0 households make identical consumption (and thus sav-
ing) choices in the presence and absence of income risk: the SIM model
with quadratic utility (and nonbinding borrowing constraints) exhibits cer-
tainty equivalence. Of course, in the stochastic case future consumption and
asset holdings respond to future income shocks (which are absent in the de-
terministic case). However, for a given amount of assets carried into period
t, households at that time period would make identical consumption-savings
choices for period t in the deterministic and the stochastic income case.
23
More explicitly,
( t
t θt−1 rW1+r
t (s )
if T < ∞
ct (s ) = r t
1+r Wt (s ) if T = ∞
Note that this equation holds for the fixed initial node s0 as well, so that (say, for T = ∞),
∞
!
r X yt+τ
c0 (s0 ) = a0 + E0
1+r τ =0
(1 + r)τ
where a0 is the exogenous initial asset position the household starts her life with (whereas
for future periods Wt (st ) is a function of the endogenous choice at (st−1 )).
5.2. THE GENERAL MODEL WITH CERTAINTY EQUIVALENCE 99
Here (Et − Et−1 )yt+s = Et yt+s − Et−1 yt+s is the revision of the expectation
about period t + s income between period t − 1 and period t, and thus ηt is
the annuity value of the revisions of expectations, between
t − 1 and
t, of all
1
future incomes. Again, the adjustment factor θt = 1 − (1+r)T −t+1 controls
for the length of remaining lifetime if T is finite. Thus the realized change
in consumption between two periods equals the annuity value of the revision
in expectations about the present discounted value of future income, ηt . Its
exact value depends on the specifics of the stochastic income process that
households face. We now consider several concrete examples.
yt = ytp + ut (5.49)
ytp = yt−1
p
+ vt (5.50)
where ytp is the “permanent” part of current income, ut is the transitory part
and vt is the innovation to the permanent part of income.25 We assume that
ut and vt are uncorrelated iid random variables with Et ut+s = Et vt+s = 0 for
s > 0, where we adopt the timing convention that ut , vt are known when Et
is taken. This process can be rewritten as
and thus
t+s
X
yt+s = yt−1 + ut+s − ut−1 + vτ (5.52)
τ =t
We now demonstrate that agents behaving according to the SIM model with
quadratic utility react quite differently to permanent income shocks vt and
transitory income shocks ut . We then have that
ut if s = 0
Et yt+s = yt−1 − ut−1 + vt + (5.53)
0 if s > 0
Et−1 yt+s = yt−1 − ut−1 + 0 + 0 (5.54)
and hence
ut + vt if s = 0
(Et − Et−1 ) yt+s = (5.55)
vt if s > 0
Thus we can calculate ηt in equation (5.48) as
r
ηt = ut + θt vt
1+r
and thus we find that, for the specific income process with permanent and
transitory shocks, the realized change in consumption is given by
r
θt ∆ct = ut + θt vt (5.56)
1+r
rθt−1
∆ct = ut + vt . (5.57)
1+r
Thus households adjust their consumption one for one with a permanent in-
rθt−1
come shock vt , but only change their consumption mildly, by 1+r in response
to a purely temporary shock ut .
Example 27 The last example assumes that income shocks are either per-
manent or transitory. Now suppose that households are only subject to one
income shock, and that a fraction γ of the shock is mean-reverting whereas a
fraction 1 − γ is permanent. Thus the income process takes the form
with γ ∈ [0, 1]. Comparing (5.58) to (5.51) we see that the case of γ = 0 cor-
responds to a process with only permanent shocks, whereas γ = 1 corresponds
to a process with only transitory shocks. Under such a process we find that
εt if s = 0
(Et − Et−1 ) yt+s = (5.59)
(1 − γ)εt if s > 0
and thus
rγ
θt ∆ct = εt + (1 − γ)θt εt (5.60)
1+r
Therefore the household’s consumption response to the εt shock is a convex
combination (with weights γ and 1 − γ) of the response to a transitory and
to a permanent shock.
Example 28 Finally, for a simple AR(1) income process of the form
yt = δyt−1 + εt (5.61)
with 0 < δ < 1 we find that
T −t s
rεt X δ
θt ∆ct = . (5.62)
1 + r s=0 1+r
Comparing this result to example 26 above we see that the change in con-
sumption in reaction to a persistent (but not permanent) shock εt falls quan-
titatively in between that induced by a purely transitory shock (which equals
rθt−1 P −t δ s
1+r
), since Ts=0 1+r
> 1 and that of a permanent shock (which equals
1), since δ < 1.
5.3 Prudence
So far we have analyzed the SIM model in partial equilibrium, under the
assumptions that household have quadratic utility and that constraints on
borrowing were not binding. We showed that household behavior exhibit
certainty equivalence: ex ante, prior to the realizations of risk, consumption
and savings choices are identical without and with risk. We also studied how
consumption and thus saving responds to income shocks ex post.
In the next two sections we will in turn relax both key assumptions that
gave risk to certainty equivalence, and show that the relaxation of either as-
sumption can give rise to precautionary saving behavior such that an increase
102 CHAPTER 5. THE SIM IN PARTIAL EQUILIBRIUM
in income risk induces households to save more and consume less. In this
section26 we demonstrate that households engage in precautionary saving if
they have convex marginal utility, U 000 (c) > 0. In section 5.4 it is shown that
in the presence of liquidity constraints agents may exhibit the same behavior
even under quadratic utility.27
To set the stage for this section, recall that with quadratic utility and
absent binding liquidity constraints the consumption function was given by
(5.46) and exhibited certainty equivalence. Specifically, the optimal con-
r
sumption rule was only a function of an annuity factor 1+r (adjusted by an
additional term if T is finite) and otherwise only depends on the expected
present discounted value of lifetime income Wt . From the definition of Wt
we see that only the conditional (on period t information) first moment of
future incomes Et yt+s matters for the consumption choice, but not the extent
of income risk (the conditional variance of future labor income) or higher mo-
ments of the random variables {yt+s }s≥1 . Thus a change in the riskiness of
future labor income (with unchanged mean) or any other higher moment of
the distribution of future labor income leaves the current consumption (and
thus saving) choice completely unaltered.
The key steps in deriving this result were to exploit quadratic utility
to obtain equation (5.38) as a special case of the general stochastic Euler
equation (5.35) for the SIM model, reproduced here for completeness, but
with preference shocks suppressed:
1+r
Et Uc (ct+1 , st+1 )
Uc (ct ) = (5.63)
1+ρ
and then to combine this Euler equation with the intertemporal budget con-
straint (5.45), which was derived from the sequential budget constraints, a
derivation that in turn required the absence of binding borrowing constraints.
In this section we will continue to use the intertemporal budget constraint,
but will investigate the implications of the intertemporal Euler equation of
utility is not quadratic (and thus marginal utility is not linear in consump-
tion, which was required for obtaining (5.38).
26
For this part the key references include Kimball (1990), Barsky, Mankiw and Zeldes
(1986), Deaton (1991) and Carroll (1997).
27
See Zeldes (1989a, 1989b) and again Deaton (1991) for the classic references.
5.3. PRUDENCE 103
W0 = y0 + ȳ1 (5.65)
c 0 + a1 = y 0 (5.66)
c1 = a1 + ȳ1 + ỹ1 (5.67)
c1 = y0 − c0 + ȳ1 + ỹ1
= W0 − c0 + ỹ1 (5.68)
Note that c1 is a random variable that varies with the realization of the
stochastic component of period 1 income ỹ1 .
Finally define
s = W0 − c0 (5.69)
= a1 + ȳ1 = E0 (c1 )
useful later on when stating the general precautionary result. With these
definitions the stochastic Euler equation (5.63), ignoring preference shocks,
becomes
Uc (c0 ) = E0 Uc (W0 − c0 + ỹ1 ) (5.70)
Note that the only endogenous choice in this equation is the deterministic
number c0 , the consumption choice for period 0, since W0 is an exogenous
constant and ỹ1 is an exogenous random variable. We now want to analyze
how the optimal choice of c0 depends on the parameters of the model, and
especially, how it varies with the characteristics of income risk, measured by
the variance of labor income in period 1, i.e. with σy2 = V ar0 (ỹ1 ).
−ε with prob. 21
ỹ1 = (5.71)
ε with prob. 12
with 0 < ε < ȳ1 . Therefore σy2 = ε2 and ε measures the amount of income
risk the household faces. We wish to determine under what condition c0 (ε)
is a strictly decreasing function. Writing out equation (5.70) and using the
assumption that ỹ1 follows a two-point distribution yields:
1
Uc (c0 ) = [Uc (W0 − c0 + ε) + Uc (W0 − c0 − ε)] . (5.72)
2
Totally differentiating (5.72) with respect to ε delivers:
dc0 1 dc0 dc0
Ucc (c0 ) = Ucc (W0 − c0 + ε)) − + 1 + Ucc (W0 − c0 − ε)) − −1
dε 2 dε dε
(5.73a)
and thus
1
dc0 (ε) 2
[Ucc (W0 − c0 + ε) − Ucc (W0 − c0 − ε)]
= (5.74)
dε Ucc (c0 ) + 12 [Ucc (W0 − c0 + ε)) + Ucc (W0 − c0 − ε))]
5.3. PRUDENCE 105
is positive.
But this is true for arbitrary ε > 0 if and only if Uccc (c) > 0. Hence
consumption in period 0 strictly declines in reaction to a marginal increase in
period 1 income risk σy2 if and only if the third derivative of the utility function
is positive. Using equation (5.66) and (5.69) it then follows immediately that
a1 and s are strictly increasing functions of σy2 for all levels of σy2 if and only
if Uccc (c) > 0 for all c. Thus a sufficient (and necessary) condition for the
household to exhibit precautionary saving behavior (i.e. to increase saving
in response to increased income risk) in the absence of binding borrowing
constraints is strictly convex marginal utility Uc . Of course it is a simple
corollary of this result that under the assumption Uccc (c) > 0 households
deviate from certainty equivalence behavior; simply realize that c0 (ε = 0) >
c0 (ε > 0).
An Application
The model above has been used by Barsky, Mankiw and Zeldes (1986) to
show that the presence of income risk alone can invalidate the Ricardian
equivalence hypothesis. Recall that this hypothesis states that for a given
process of government spending a change in the timing of taxation does not
affect household behavior and macroeconomic aggregates. Thus if Ricardian
equivalence holds it should not matter whether current government expendi-
tures are financed via current taxes of by government debt that is redeemed
later (using future taxes).
A simple example is sufficient to make this argument. Suppose that
government spending satisfies G0 = G1 = 0 and that the benchmark policy
is one of no taxation in either period (and thus the situation analyzed in the
previous subsection). Now consider the Ricardian experiment of lowering
taxes in period 0, say, in order to stimulate consumption in the economy
in a Keynesian-style fiscal expansion. Thus in period 0 households receive
a lump-sum subsidy of size t, and in period 1 have to pay positive taxes in
order to finance the repayment of the debt the government incurred for paying
the transfers. Let τ denote the tax rate on labor income in period 1. We
assume that the government deals with a continuum population of measure
106 CHAPTER 5. THE SIM IN PARTIAL EQUILIBRIUM
1, each member of which faces the income process specified in (5.71). Thus
government tax revenues from the income tax defined by τ are given by30
τ E0 (y1 ) = τ ȳ1
and the size of the period 0 transfers thus has to satisfy (recall that we
assumed r = 0)
t = τ ȳ1
in order for the intertemporal government budget constraint to hold.
Therefore we note that, after the policy innovation, the expected present
discounted value of lifetime labor income for each household equals
and the fiscal policy reform leaves the expected present discounted value of
lifetime income unchanged. If households were certainty equivalence con-
sumers, their consumption behavior would therefore be unaffected by the
Ricardian tax policy experiment. We note, however, that after-tax income
in the second period now equals (1 − τ )(ȳ1 + ỹ1 ), with associated variance
(1 − τ )2 σy2 < σy2 . Therefore, from the result in the previous section (as along
as the size of τ, t is small), the household consumes more and saves less in re-
sponse to the small temporary, debt-financed tax cut. Ricardian equivalence
fails, despite the fact that the household behaves fully rational and forward
looking, faces no borrowing constraints, and the tax is (or at least looks
like) a lump-sum tax. The tax change does affect the extent of income risk,
however, and thus induces a decline in precautionary saving. With Barsky,
Mankiw and Zeldes’ (1986) words, the household is a “Ricardian consumer
with Keynesian propensities” to consume out of current income.
31
The compensating risk premium θ∗ (c, ỹ1 ) is defined as
equation32
E0 U (c + ỹ1 ) = U (c − θ(c, ỹ1 )) (5.77)
is explicitly characterized by
1 Ucc (c)
θ(c, ỹ1 ) = − σy2 + o(σy2 ), (5.78)
2 Uc
where o(σy2 ) is a term that converges to zero faster than does σy2 . This result
justifies the coefficient of absolute risk aversion
Ucc (c)
r(c) = − (5.79)
Uc (c)
as an appropriate measure to quantify the willingness of a household to pay
to avoid a small risk. Correspondingly, the coefficient of relative risk aversion
cUcc (c)
σ(c) = − (5.80)
Uc (c)
measures the willingness to pay to avoid small relative gambles in which
percentages of c are at risk. Of course, for the special cases of CRRA utility
1−σ
U (c) = c1−σ and CARA utility U (c) = − γ1 e−γc these entities are given,
correspondingly, by
σ
r(c) = and σ(c) = σ
c
r(c) = γ and σ(c) = γc.
Now define the equivalent precautionary saving premium ψ(s, ỹ1 ) implicitly
by the equation
E0 Uc (s + ỹ1 ) = Uc (s − ψ(s, ỹ1 )) (5.81)
Comparing equations (5.77) and (5.81) it follows immediately that all results
(and especially the result summarized in equation (5.78) above) derived by
Pratt (1964) for the equivalent risk premium θ(c, ỹ1 ) immediately apply to
the equivalent precautionary saving premium ψ(s, ỹ1 ), with all expressions
in the original Pratt (1964) results replaced by one higher derivative.
Similarly, Kimball (1990) defines the compensating precautionary saving
premium by ψ ∗ (s, ỹ1 ) as
that is, the amount of extra saving (starting from the saving level s without
risk) the household would find optimal in the presence of risk, relative to
the risk-free case. Since the compensating precautionary saving premium is
easier to interpret (and, at least locally, equal to the equivalent premium)
we will focus on this measure of precautionary saving from now on, referring
the reader to Kimball (1990) for a complete treatment. Again comparing
equations (5.82) and (5.76) allows for a straightforward application of the
results in Pratt (1964) to the precautionary saving premium. Consequently,
we have the following
1 Uccc (s)
ψ ∗ (s, ỹ1 ) = − σy2 + o(σy2 )
2 Ucc (s)
o(σy2 )
lim = 0. (5.83)
σy →0 σy2
2
Uccc (s)
p(s) = −
Ucc (s)
and in both cases risk aversion and prudence is controlled by a single param-
eter (σ and γ, respectively).33
Now that we have fully characterized the precautionary saving premium
through proposition 29 we want to further interpret what this entity ψ ∗ (s, ỹ1 )
actually measures, in terms of observable behavior. To do so, we first recall
that the Euler equation for the case without risk reads as
For future reference define c0 (W0 , ỹ1 ) as the optimal consumption choice as-
sociated with lifetime wealth W0 and income in the second period determined
33
As for risk aversion and the risk premium, one can also define the index of relative
prudence as
cUccc (c)
pr(c) = − (5.86)
Ucc (c)
and establish the same results as above if income risk is proportional to wealth W0 . For
CRRA and CARA utility the coefficient of relative prudence is given, respectively, by
pr(c) = σ+1
pr(c) = γc.
5.3. PRUDENCE 111
by the random variable ỹ1 , and as s(W0 , ỹ1 ) the optimal saving (out of life-
time wealth) function. The corresponding optimal policy functions in the
absence of income risk are defined as c0 (W0 , 0) and s(W0 , 0), which are of
course related by
c0 (W0 , 0) + s(W0 , 0) = W0 .
Furthermore define the lifetime wealth W0 needed to make a given consump-
tion c0 optimal as W0 (c0 , ỹ1 ). Note that W0 (c, ỹ1 ) is defined as the quantity
x that satisfies
W0 (c0 (x, ỹ1 ), ỹ1 ) = x.
It is then easy to show34 that ψ ∗ (s, ỹ1 ) equals the additional wealth required
to keep consumption at a given level c0 if a small income risk is introduced,
that is
ψ ∗ (W0 (c0 , 0) − c0 , ỹ1 ) = W0 (c0 , ỹ1 ) − W0 (c0 , 0) (5.89)
In other words, ψ ∗ (s, ỹ1 ) can also be interpreted as the magnitude of the
rightward shift of the consumption function c0 (W0 , 0) with wealth W0 on the
x-axis, at a particular level of consumption c0 , as income risk is turned on.
We now collect further important results about precautionary saving con-
tained in Kimball (1990).
Remark 30 The previous results were local in the sense that we considered
a specific value for c0 or s, and small risks. However, Kimball (1990) also
34
By the definition of W0 (c0 , ỹ1 ) we have
where the last equality follows from equation (5.88). By definition of ψ ∗ (W0 (c0 , 0) − c0 , ỹ1 )
Equating the left hand sides of the previous two equations yields
But since Ucc < 0 for all c, this last equation implies that, since both W0 (c0 , 0) +
ψ ∗ (W0 (c0 , 0) − c0 , ỹ1 ) and W0 (c0 , ỹ1 ) are deterministic numbers:
provides a global when he shows that, for any two utility functions U, V, if
Uccc (s) Vccc (s)
− >−
Ucc (s) Vcc (s)
for all s then
ψU∗ (s, ỹ1 ) > ψV∗ (s, ỹ1 )
for all s and all nondegenerate random variables ỹ1 . This result is particularly
useful for two utility functions U, V that yield the same consumption and
savings functions in the absence of risk (e.g. if both U and V are of CRRA
form with σU > σV ; recall that ρ = r = 0). Then this result implies that
for the same introduction of risk ỹ1 households with utility function U shift
the consumption function to the right uniformly more than households with
V when income risk is introduced.
Remark 31 Often the situation of interest is one where the initial choice
situation already contains some risk, and then additional risk is added, and
the question arises what happens to consumption and saving in response to
this additional risk. As long as the new risk is independently distributed to
the already present risk, the results go through unchanged. However, if the
additional risk is, for example, a mean-preserving spread of the old risk (and
thus the additional risk is not independent of the initial risk), then the results
stated above do not carry over.
Example 33 Assume that U (c) = log(c), which has positive and strictly
decreasing absolute prudence, and also assume that ỹ1 follows the process
from subsection 5.3.1. We can then determine the consumption function in
closed form as:
3 1 1
c0 (W0 ; ε) = w − 8ε2 + (W0 )2 2
4 4
where c0 (W0 ; ε) is well-defined for wealth levels W0 ≥ ε. We note that
1
c0 (W0 ; ε) < c0 (W0 ; ε = 0) = W0 for all W0 ≥ ε
2
and that for all ε > 0 the function c0 (W0 ; ε) is strictly increasing and strictly
concave, with
lim (c0 (W0 ; ε) − c0 (W0 ; ε = 0)) = 0
W0 →∞
and
∂c0 (W0 ; ε) 1 ∂c0 (W0 ; ε = 0)
lim = = .
w→∞ ∂W0 2 ∂W0
114 CHAPTER 5. THE SIM IN PARTIAL EQUILIBRIUM
Figure ?? shows the consumption function (for the case ε = 0.2) and the pre-
cautionary saving premium at c0 = 0.2. As predicted by the previous remark
the slope of the consumption function with risk is larger than the slope of the
consumption function without risk, for all wealth levels W0 .
We derived equation (5.91) from (5.92) since it will be useful in the analysis
below. We note that since ct is known when expectations Et are taken, the
5.3. PRUDENCE 115
−σ −σ
1 1
term Et ct
= ct
does not create any problem for obtaining a
closed-form solution. However, the term Et (ct+1 )−σ does, since it is the
to obtain estimates of α̂1 = bσ1 and to test whether β̂ = 0, where the variables
Xt often include household income in period t.
Although estimating the approximated stochastic Euler equation has been
the predominant approach to determine the parameters and test the main
implication of the partial equilibrium SIM model, critiques of this approach
(see e.g. Carroll, 2001) question whether the approximation (5.93) used to
derive equation (5.95) is accurate. Related to this point, by using this ap-
proximation equation (5.95) ignores the importance of precautionary saving
(despite the fact that with CRRA utility the household is prudent and has
preference-induced precautionary saving motive).
To see this point, now assume that ln(ct+1 ) is normally distributed36 with
mean µ = Et ln(ct+1 ) and variance σc2 . Although it is in general not desirable
to make distributional assumptions on endogenous variables (such as ct+1 ),
we do it here to demonstrate the potential problem with the Euler equation
estimation approach, in an analytically tractable way. With this assumption
we find that
Et (c−σ −σ ln(ct+1 )
t+1 ) = Et e
(u−µ)2
Z ∞ − 2
−σu e
2σc
= e √ du
−∞ 2πσc
2
−
[u−(µ−σσc2 )]
2 )2 −µ2 Z ∞ 2
(µ−σσc
e 2σc
= e 2σc2
√ du
−∞ 2πσc
2 )2 −µ2
(µ−σσc
2
= e 2σc
1 2 2
= e 2 σ σc −µσ
1 2 2
= e 2 σ σc −σEt ln(ct+1 ) (5.96)
Z = ln(κ + ỹ1 )
and thus
∆at+1 = ut (5.104)
i.e. assets follow a random walk. Permanent income shocks vt are fully ab-
sorbed by consumption and thus do not induce changes in the asset position
of the household. Transitory income shocks in contrast trigger a consump-
rut ut
tion response of only 1+r , and thus the remaining part of the shock 1+r is
at+1
absorbed by 1+r and thus assets change by the whole transitory shock ut .
This discussion shows that a certainty equivalent consumer will violate
any fixed borrowing limit at+1 ≥ −Ā with probability 1 (since assets follow
a random walk). The asset process might not may not violate the no Ponzi
condition if we specify it carefully37 , but calls into question the plausibility
of the assumption that liquidity constraints are never binding, a maintained
assumption so far.
3. The previous argument also implies that any change in the environ-
ment that affects the incidence of future binding borrowing constraints
affects current consumption. Importantly, suppose the variance of fu-
ture income increases (say, for yt+1 ), making lower realizations of yt+1
possible or more likely. If consequently the set of yt+1 values for which
the borrowing constraint binds becomes larger, then
consumption growth for the low-wealth group, but not for the high-wealth
group. Zeldes (broadly) finds such evidence and concludes that borrowing
constraints are important in shaping consumption choices, at least for the
currently wealth-poor. In evaluating his result, however, you should keep in
mind that we argued above that with preferences that exhibit prudence, cur-
rent income may help predict consumption growth, if it contains information
about the extent of future income uncertainty. While it is not entirely obvi-
ous why this should be the case for low-wealth households and not for high-
wealth households, note that wealth accumulation is an endogenous choice,
so the low wealth-group sample is not simply a random sample. Therefore it
is not inconceivable that low-wealth households have income processes with
very different stochastic properties that high-wealth households, including
predictability of future income risk by current income. [TBC]
Case ρ < r
In this case both the (im-)patience motive and the precautionary saving
motive point towards postponing consumption in favor of saving. In fact, it
is optimal to accumulate assets without bounds as we will demonstrate next.
Proposition 34 [Sotomayor (1984), Chamberlain and Wilson (200)]: Let
r > ρ > 0 and assume that either U is bounded or that the income process is
iid with support yt ∈ [a, A] where a ≥ 0 and A < ∞. Then
lim ct = ∞ almost surely (5.115)
t→∞
126 CHAPTER 5. THE SIM IN PARTIAL EQUILIBRIUM
and
lim at = ∞ almost surely (5.116)
t→∞
The basic intuition for this result can be derived from the Euler equation
with liquidity constraints, equation (5.108). Under the assumption that ρ < r
we obtain that
1+r
Uc (ct ) ≥ Et Uc (ct+1 ) (5.117)
1+ρ
> Et Uc (ct+1 ).
Hence the stochastic process for marginal utility {Uc (ct )}, which is strictly
positive, follows a supermartingale.42 We now can invoke the martingale
convergence theorem (which is reviewed in remark 35 below) it follows that
the sequence of random variables {Uc (ct )} converges almost surely to some
limit random variable Uc (c). Iterating on the Euler equation, for all t
t+1
1+ρ
E0 Uc (ct+1 ) ≤ Uc (c0 ) (5.118)
1+r
and thus E0 Uc (ct+1 ) → 0. But since Uc (.) is strictly positive, it must be the
case that not only {Uc (ct )} converges almost surely to some limit random
variable Uc (c), which is what the martingale convergence theorem guaran-
tees, but it converges almost surely to Uc (c) = 0. From the strict concavity
and the Inada conditions of the utility function it follows that consumption
converges to ∞ with probability one, that is, equation (5.115) is satisfied.
With a bounded labor income process, to finance diverging consumption re-
quires diverging asset income and thus equation (5.116) follows. will exist.
Although we refer the reader to Sotomayor (1984) and Chamberlain and Wil-
son (2000) for proofs of this result, note that the assumptions in the previous
proposition are used to insure that lifetime utility of the consumer remains
finite under the optimal consumption allocation.
42
In fact, for the seqence of random variables {Uc (ct )} to be a supermartingale it is
required that
E |Uc (ct )| < ∞
which is assured if the endowment process {yt } is such that positive consumption is is
possible with probability 1 (or alternatively, if Uc (0) < ∞, but this would violate the
Inada condition).
5.5. PRUDENCE AND LIQUIDITY CONSTRAINTS: THEORY 127
{s ∈ F : X(s) ≤ a} ∈ F.
Ft ⊂ Ft+1
for all t. For our purposes we can interpret Ft as capturing the information
the household has at time t. Imagine that at time zero nature draws an
infinite sequence s ∈ F, but which s was drawn is unknown to the household
(she just knows the probability measure π over all possible sequences). As
time unfolds the household learns more and more about what infinite history
s was drawn. For example, suppose that st ∈ {1, 2} for all t. Then s ∈ F is
an infinite history of 1’s and 2’s. Then
F0 = {, F, F01 , F02 }
where
F0i = {ŝ ∈ F : ŝ0 = i},
that is, at time zero the household can distinguish between infinite histories
whose first entry differs (but not among histories with identical first entries).
Similarly
F1 = {, F, F111 , F112 , F121 , F122 , . . .}
where the set
F1ij = {ŝ ∈ F : ŝ0 = i, ŝ1 = j}.
Now households can distinguish between infinite histories that differ in one of
the first two entries. A stochastic process is a sequence {Xt , Ft }∞
t=0 such that
43
For ease of notation we treat s0 as random here too, rather than as the fixed intial
node (as we did so far).
128 CHAPTER 5. THE SIM IN PARTIAL EQUILIBRIUM
{Ft }∞ ∞
t=0 is a filtration and {Xt }t=0 is a sequence of random variables such that
Xt is measurable with respect to Ft for all t. We say that {Xt }∞ t=0 is adapted
∞ t t
to the filtration {Ft }t=0 . Our notation {ct (s ), πt (s )} can now be related to
this rigorous way of describing the stochastic structure of the economy. First,
we can construct the probability πt (st ) of event history st from the probability
measure π over infinite histories as
Z
t
πt (s ) = 1A(st ) dπ
where
A(st ) = {ŝ ∈ F : ŝt = st }.
Note that formally πt (.) is a probability measure on the measurable space
(S t , Ft ). Finally, although consumption, defined as a stochastic process, is a
sequence of functions ct : F → R mapping infinite histories into the real
numbers, the measurability of ct with respect to Ft requires that for any two
infinite histories s, ŝ ∈ F with st = ŝt we have
ct (s) = ct (ŝ).
Thus we can write, in short, ct (st ) for all infinite histories for which the
finite history until period t is given by st . Finally, with our definition of what
exactly a stochastic process is we can now define a martingale as a stochastic
process {Xt , Ft }∞
t=0 such that for all t
Xt = E [Xt+1 |Ft ]
E (|Xt |) < ∞
for all t.
5.5. PRUDENCE AND LIQUIDITY CONSTRAINTS: THEORY 129
almost surely. The martingale convergence theorem then states that a super-
martingale {Xt , Ft }∞
t=0 that satisfies
converges to a limit random variable X ∗ almost surely and that the limit
random variable X ∗ satisfies E(|X ∗ |) ≤ K. Note that X ∗ in general is a
nondegenerate random variable, rather than a fixed number.
Case ρ = r
Sotomayor (1984) and Chamberlain and Wilson (2000) show that the pre-
vious result (that consumption and assets diverge to ∞ almost surely) goes
through even for ρ = r, if the income process is sufficiently stochastic, in a
sense to be made precise below.
Proposition 36 Define
∞
r X yτ
xt = (5.119)
1 + r τ =t (1 + r)τ −t
Then
c̄ := lim ct = sup xt =: x̄ (5.120)
t→∞ t
Proof. See Sargent and Ljungquist (2nd edition), chapter 16.3.1 Note
that for this result U need not be bounded, a maintained assumption of
Chamberlain and Wilson (2000) who of course are mainly interested in the
stochastic labor income case.
The number xt is the period t annuity value of future labor income. Since
the income process {yt } is bounded from above (because yt can take only
finitely many values) the sequence {xt } is bounded from above and thus x̄ is
a finite number which gives the maximal (over time) annuity value of income.
The intuition for this result is easiest to see if there is a finite date T at
which the borrowing constraint binds for the last time. In that case aT = 0
and aT +τ > 0 for τ > 0. Since ρ = r and there is no income risk the Euler
equation then implies (see equation (5.108)) that:
cT = cT +1 = cT +τ = ĉ for all τ > 0 (5.121)
ct ≤ cT for all t < T (5.122)
The budget constraints from period T onwards read as
aT +1
cT + = yT
1+r
aT +τ +1
cT +τ + = yT +τ + aT +τ
1+r
Consolidating these into a lifetime budget constraint (from T onwards) yields46
∞ ∞
X cT +τ X yT +τ
=
τ =0
(1 + r)τ τ =0
(1 + r)τ
46
This follows as long as
aT +τ +1
lim = 0.
τ 7→∞ (1 + r)τ +1
5.5. PRUDENCE AND LIQUIDITY CONSTRAINTS: THEORY 131
and thus
∞ ∞ ∞
X cT +τ X 1 (1 + r)ĉ X yT +τ
= ĉ = =
τ =0
(1 + r)τ τ =0
(1 + r)τ r τ =0
(1 + r)τ
cannot be optimal, intuitively, because the household “leaves resources on the table” rather
than consuming it. The condition
aT +τ +1
lim ≤0
τ 7→∞ (1 + r)τ +1
is called the transversality condition. It is an optimality condition (like the Euler equa-
tion), that, under appropriate conditions on the utility function, the endowment process
and the interest process, is jointly sufficient, together with the Euler equations, for an
optimal consumption-savings plan. Under alternative conditions on the fundamentals the
transversality condition is a necessary condition for an optimal allocation.
Note the fundamental distinction between a No-Ponzi condition and the transversality
condition. The No-Ponzi condition is required for the household maximization problem to
have a solution and in the current context would read as
aT +τ +1
lim ≥ 0.
τ 7→∞ (1 + r)τ +1
Taking the No Ponzi condition and the transversality condition (and asserting its neces-
sity) together, any candidate for an optimal consumption-savings plan of a well defined
deterministic consumption-savings problem has to satisfy:
aT +τ +1
lim = 0.
τ 7→∞ (1 + r)τ +1
132 CHAPTER 5. THE SIM IN PARTIAL EQUILIBRIUM
That is, in this case consumption is weakly rising over time until the date the
borrowing constraint binds for the last time, and then constant and equal
to the annuity value of income from that date onwards. The proposition
generalizes this result to the case in which the borrowing constraint binds
infinitely often and thus consumption ceases to rise only in the time limit.
Stochastic Labor Income Now let’s turn to the more interesting case in
which the endowment process is stochastic. Here the result that consumption
and asset holdings diverge as time extends to infinity is restored, provided
that, for any event history st the present discounted value of future income is
sufficiently stochastic. The next proposition, again proved under various as-
sumptions by Sotomayor (1984) and Chamberlain and Wilson (2000), states
this formally.
lim ct = ∞ (5.124)
t→∞
lim at = ∞ (5.125)
t→∞
Note from (b) that Sotomayor (1984) needs no assumptions on the bound-
edness of the utility and the level of variability of the income process (other
that it has to be nondegenerate) to prove the result, but needs the iid assump-
tion. Dispensing with the iid comes at the cost of having to make U bounded
and the income process “sufficiently” stochastic so that present discounted
value of future income leaves set [α, α + ε] with probability of at least ε. So
unfortunately the theorem does not apply to an economy with both serially
correlated shocks and standard utility functions (CRRA or CARA).
Note that one can show that condition (5.123) holds for income following a
finite-state Markov chain with Π(y) > 0 for all y and π(y 0 |y) > 0 for all y, y 0 .
The previous theorem implies that, under the appropriate conditions, the
5.5. PRUDENCE AND LIQUIDITY CONSTRAINTS: THEORY 133
Case ρ > r
At this stage it will prove helpful to formulate the income fluctuation problem
with borrowing constraints recursively. For the general problem in which
income follows a stationary, finite state Markov chain47 the Bellman equation
is
( )
1 X
v(a, y) = max U (c) + π(y 0 |y)v(a0 , y 0 ) (5.126)
a0 ,c≥0 1 + ρ y0
a0
s.t. c + = y+a (5.127)
1+r
As first order condition we obtain
1+r X
Uc (c) ≥ π(y 0 |y)v(a0 , y 0 ) (5.128)
1 + ρ y0
= if a0 > 0
or ( )
1+r X
Uc (c) = max Uc (y + a), π(y 0 |y)Uc (c0 ) (5.132)
1 + ρ y0
47
We will deal with the case of nonstationary income below.
134 CHAPTER 5. THE SIM IN PARTIAL EQUILIBRIUM
Both equations (5.126) and (5.132) can be used to compute optimal policy
functions a0 (a, y) and c(a, y) as well as the value function v(a, y) using exactly
the same iterative procedures as described in the last section, just with the
modification that the Bellman equation has the additional constraint a0 ≥ 0
and the Euler equation has two parts now.
Income Process IID If the income process is iid we can reduce the state
space from two to one dimension by introducing the variable cash at hand
x = a + y. The Bellman equation becomes
( )
a0
1 X
v(x) = max u x− + π(y 0 )v(a0 + y 0 ) (5.133)
0≤a0 ≤(1+r)x 1+r 1 + ρ y0
dc(x)
>0 (5.136)
dx
There exists an x̄ > y such that a0 (x) = 0 for all x ≤ x̄ and a0 (x) > 0 for all
0
x > x̄. Finally dc(x)
dx
≤ 1 and dadx(x) < 1 + r.
5.5. PRUDENCE AND LIQUIDITY CONSTRAINTS: THEORY 135
x0 = x (5.139)
xt = a0 (xt−1 ) + y ≥ y (5.140)
If there exists a smallest T such that xT = y then we found a contradiction,
since then a0 (xT −1 ) = 0 and xT −1 > 0. So suppose that xt > y for all t. But
then a0 (xt ) > 0 by assumption. Hence
1+r 0
v 0 (x0 ) = v (x1 )
1+ρ
t
1+r
= v 0 (xt )
1+ρ
t
1+r
< v 0 (y)
1+ρ
t
1+r
= u0 (y) (5.141)
1+ρ
where the inequality follows from the fact that xt > y and the strict concavity
of v. the last equality follows from the envelope theorem and the fact that
a0 (y) = 0 so that c(y) = y.
136 CHAPTER 5. THE SIM IN PARTIAL EQUILIBRIUM
c0 (x) ∈ (0, 1]. Optimal asset holdings are either constant at the borrowing
0
limit or strictly increasing in cash at hand, i.e. a0 (x) = 0 or dadx(x) ∈ (0, 1 + r)
Proposition 40 There exists x̄ > y1 such that for all x ≤ x̄ we have c(x) =
x and a0 (x) = 0
Proof. Suppose, to the contrary, that a0 (x) > 0 for all x ≥ y1 . Then,
using the first order condition and the envelope condition we have for all
x ≥ y1
1+r 0 0 1+r 0
v(x) = Ev (x ) ≤ v (y1 ) < v 0 (y1 ) (5.143)
1+ρ 1+ρ
Picking x = y1 yields a contradiction.
Hence there is a cutoff level for cash at hand below which the consumer
consumes all cash at hand and above which he consumes less than cash at
hand and saves a0 (x) > 0. So far the results are strikingly similar to the
deterministic case. Unfortunately here it basically ends, and therefore our
analytical ability to characterize the optimal policies. In particular, the very
important proposition showing that there exists x̃ such that if x ≥ x̃ then
x0 < x̃ does not go through anymore, which is obviously quite problematic
for computational considerations. In fact we state, without a proof, a result
due to Schechtman and Escudero (1977).
Proof. See Schechtman and Escudero (1977), Lemma 3.6 and Theorem
3.7
Fortunately there are fairly general conditions under which one can, in
fact, prove the existence of an upper bound for the state space. Again we
will refer to Schechtman and Escudero for the proof of the following results.
Intuitively why would cash at hand go off to infinity even if the agents are
impatient relative to the market interest rate, i.e. even if β(1 + r) < 1? If
138 CHAPTER 5. THE SIM IN PARTIAL EQUILIBRIUM
agents are very risk averse, face borrowing constraints and a positive proba-
bility of having very low income for a long time, they may find it optimal to
accumulated unbounded funds over time to self-insure against the eventual-
ity of this unlikely, but very bad event to happen. It turns out that if one
assumes that the risk aversion of the agent is sufficiently bounded, then one
can rule this out.
Proposition 42 Suppose that the marginal utility function has the property
that there exist finite eu0 such that
Proof. See Schechtman and Escudero (1977), Theorems 3.8 and 3.9
The number eu0 is called the asymptotic exponent of u0 . Note that if the
utility function is of CRRA form with risk aversion parameter σ, then since
we have eu0 = −σ and hence for these utility function the previous proposition
applies. Also note that for CARA utility function
c
logc e−c = −c logc e = − (5.146)
ln(c)
c
− lim = −∞ (5.147)
c→∞ ln(c)
a0 a00
X
0 0 0 0 0
U y+a− = β(1 + r) π(y |y)U y + a − (5.153)
1+r y 0
1+r
Defining
a0
ct = y + a − (5.154)
1+r
a00
ct+1 = y 0 + a0 − (5.155)
1+r
we obtain back our stochastic Euler equation (5.35). There are several im-
portant special cases that merit a brief discussion.
cT (a, y) = a + y (5.158)
a0T (a, y) = 0 (5.159)
1. Value function iteration: now we look for a time invariant value func-
tion v(a, y) and associated policy functions a0 (a, y) and c(a, y). We need
to find a fixed point to Bellman’s equation, since there is no final period
to start from. Thus we make an initial guess for the value function,
5.6. PRUDENCE AND LIQUIDITY CONSTRAINTS: COMPUTATION143
since the income distribution is nonstationary and thus in general the wealth
distribution will be for any fixed interest rate r. Hence when dealing with
general equilibrium models with infinitely lived agents, one of the maintained
assumptions is stationarity of the individual income processes.49
The partial equilibrium problem with nonstationary income process and
infinite horizon can be tackled, however, through an appropriate change of
variables. For this trick to work it is crucial to assume that households
have CRRA period utility. First assume that, following Deaton (1991) that
the labor income process of the household features iid growth rates, that is,
assume that
yt+1
zt+1 = (5.165)
yt
is equal to an iid random variable, and thus implies that the natural loga-
rithm of income follows a random walk, potentially with drift
where {εt } is a sequence of iid random variables with E(εt ) = 0. Note that
this specification implies positive income in all periods, with probability one.
In order to overcome the problem that the state space for assets is unbounded,
which makes the dynamic programming problem of the household difficult
to tackle, normalize all variables by the current level of income
ct
θt = (5.168)
yt
xt at + y t
wt = = (5.169)
yt yt
49
Note that the same comment does not apply for economies in which agents die with
probability 1 in finite time in which case typically a stationary general equilibrium can be
found.
5.6. PRUDENCE AND LIQUIDITY CONSTRAINTS: COMPUTATION145
Note that Deaton (1991) is not able to develop a condition under which the
state space for w is bounded above. It is obviously bounded below by 1
since assets at are restricted to be nonnegative. In accordance to the model
with stationary income he establishes a threshold w̄ such that θ(w) = w and
w0 (w) = 1 for all w ≤ w̄; i.e. for low levels of normalized cash at hand it is
optimal for the consumer to eat all cash at hand today and save nothing for
tomorrow.
As discussed in the next chapter, a very popular specification of the house-
hold labor income process postulates that log-labor income is composed of
a permanent component that follows a random walk and a transitory shock.
That is, suppose labor income follows the process51
Here ηt+1 and εt+1 are permanent and transitory income shocks that are
independently (over time and of each other) and normally distributed with
2 σ2
variances (σε2 , ση2 ) and means (− σ2ε , − 2η ). Denote the cumulative distribution
function of the Normal distribution by N.
As before one can make the problem stationary by expressing all variables
relative to the permanent component of income, pt , see e.g. Carroll (1997).
Now the state space consists of assets a, the permanent component of log-
income p and the transitory shock e. The dynamic programming problem
can be written as
1−σ Z Z
c 0 0 0 0 0
v(a, p, ε) = max +β v(a , p + η , ε )dN (ε )dN (η )
c,a0 ≥0 1−σ ε0 η 0
s.t.
a0
c+ = exp(p + ε) + a
1+r
With the usual definition of cash at hand
x = y+a
= exp(p + ε) + a
51
In example 26 we studied a similar process, but there the level of labor income (rather
than its log) was subject to permanent and transitory shocks (and the period utility
function was quadratic rather than of CRRA form).
5.6. PRUDENCE AND LIQUIDITY CONSTRAINTS: COMPUTATION147
we can reduce the state space and re-write the recursive problem as
1−σ Z Z
c 0 0 0 0 0
v(x, p) = max +β v(exp(p + ε ) + (1 + r)(x − c), p + η )dN (ε )dN (η )
0≤c≤x 1−σ ε0 η 0
(5.175)
0
where we have used the fact that by definition of x and the budget constraint
x0 = exp(p0 + ε0 ) + a0
= exp(p0 + ε0 ) + (1 + r)(x − c)
Note, however, that since p follows a random walk, the problem above is not
stationary. Therefore we attempt to make it stationary by defining variables
relative to the permanent level of income:
c x
c̃ = and w =
exp(p) exp(p)
with
x0 exp(p0 + ε0 ) + (1 + r)(x − c)
w0 = =
exp(p0 ) exp(p0 )
(1 + r)(x − c) (1 + r)(w − c̃)
= exp(ε0 ) + 0
= exp(ε0 ) +
exp(p + η ) exp(η 0 )
and thus the Bellman equation becomes
1−σ
(1 + r)(w − c̃)
Z Z
c̃ 0 1−σ 0 0 0
ṽ(w) = max +β exp(η ) ṽ exp(ε ) + dN (ε )dN (η )
0≤c̃≤w 1−σ ε0 η 0 exp(η 0 )
148 CHAPTER 5. THE SIM IN PARTIAL EQUILIBRIUM
which we need to solve numerically for the value function ṽ(w) and the as-
sociated policy function c̃(w). Once we have determined those, consumption
and asset levels are given by
x
c(x, p) = exp(p)c̃
exp(p)
0
a (x, p) = (x − c(x, p))(1 + r).
Alternatively one can use the Euler equation (and policy function iter-
ation) to determine the optimal consumption policy. In original form, the
Euler equation associated with Bellman equation (5.175) reads as
Z Z
−σ −σ 0 0 0 0 0
c(a, p, ε) = max (exp(p + ε) + a) , β(1 + r) c(a , p + η , ε )dN (ε )dN (η )
ε0 η0
mapping these numbers into the appropriate labor income realizations. Now
one can generate a time series for consumption and asset holdings (for the
infinite horizon case) by
c0 = c(a0 , y0 ) (5.176)
a1 = a0 (a0 , y0 ) (5.177)
and recursively
ci = c(ai , yi ) (5.178)
ai+1 = a0 (ai , yi ) (5.179)
The same applies to the finite time horizon case, but here the policy functions
being applied also vary with time.
52
In the canonical consumption-savings problem the sate space is typically small. For
problems with multiple continuous state variables and smooth functions to be approx-
imated Krueger and Kubler (2004) propose a sparse grid method for approximation of
multidimensional (but smooth) functions. Recently, Judd, Maliar and Maliar (2012) de-
veloped a related algorithm that endogenizes the choice of the sparse grid.
53
Recently Carroll (2006) proposed an endogenous grid method that in practice speeds
up computation of the class of problems discussed in this section tremendously.
150 CHAPTER 5. THE SIM IN PARTIAL EQUILIBRIUM
Example 43 One popular stochastic earnings process (see section 5.6.3 and
also the next chapter) models the logarithm of earnings as the sum of a
permanent and transitory shock, as did equations (5.49) and (5.50) for the
level of earnings:
where εit is the transitory shock and ηit is the permanent shock. This process
fits into the general structure of equation (5.180) by letting N = 2, (x1it , x2it ) =
(εit , ηit ) and α01 = 1, ατ1 = 0 for all τ > 0, and ατ2 = 1 for all τ ≥ 0. Recall
that this process also implies
where the remaining error term ξit is uncorrelated with the permanent and
transitory income shock at all leads and lags. One can loosely interpret πtη
as the (potentially time- or age-varying) marginal propensity to consume out
of a permanent income shock, with πtε possessing the same interpretation
for a transitory shock, but should keep in mind that ∆ log(cit ) is the change
in log-consumption (i.e. consumption growth) rather than the change in
consumption levels. In fact, ignoring this distinction (both for income and
consumption), the canonical PIH model with certainty equivalence and in-
finite horizon has the implication that πtη = 1 and πtε = 1+rr
, as shown in
55
That is, after consumption growth has been purged from its predictable components
(that are due to, e.g. life cycle effects and family size and composition as well as other
observable household characteristics).
56
Blundell et al. (2008) provide an appendix where they argue that this consumption
growth process is consistent with a first order approximation of the Euler equation (of
course, with non-binding borrowing constraints) in a model where income follows the
permanent-transitory process stipulated and households have CRRA period utility.
5.7. PRUDENCE AND LIQUIDITY CONSTRAINTS: EMPIRICAL IMPLICATIONS153
section 5.6.3.
Note that with this postulated consumption process
Covi (πtη ηit + πtε εit + ξit , εit )
φεt = 1− = 1 − πtε
V ari (εit )
Covi (πtη ηit + πtε εit + ξit , ηit )
φηt = 1− = 1 − πtη
V ari (ηit )
and thus the insurance coefficients are just one minus the marginal propen-
sities to consume out of the various income shocks. But again, as long as
(εit , ηit ) are not directly observable, neither the φxt nor the πtx can be empir-
ically estimated.
However, given the assumed income and consumption growth processes,
recognize that, using equation (5.183),
Covi (∆ log(yit ), ∆ log(yit+1 )) = Covi (ηit + εit − εit−1 , ηit+1 + εit+1 − εit ) = −V ari (εit )
Covi (∆ log(cit ), ∆ log(yit+1 )) = Covi (πtη ηit + πtε εit + ξit , ηit+1 + εit+1 − εit )
= −πtε V ari (εit ) = −Covi (∆ log(cit ), εit ).
and thus
Covi (∆ log(cit ), ηit )
φηt = 1 −
V ari (ηit )
Covi (∆ log(cit ), ∆ log(yit−1 ) + ∆ log(yit ) + ∆ log(yit+1 ))
= 1−
Covi (∆ log(yit ), ∆ log(yit−1 ) + ∆ log(yit ) + ∆ log(yit+1 ))
154 CHAPTER 5. THE SIM IN PARTIAL EQUILIBRIUM
at+2 (st+1 )
ct+1 (st+1 ) + = yt+1 (st+1 ) + at+1 (st ) (5.186)
1+r
5.8. APPENDIX A: THE INTERTEMPORAL BUDGET CONSTRAINT155
at+2 (st+1 )
!
πt+1 (st+1 |st )at+1 (st ) ct+1 (st+1 ) + 1+r
− yt+1 (st+1 )
= πt+1 (st+1 |st )
1+r 1+r
(5.187)
Since at+1 (st ) is only a function of st (but not st+1 ), equation (5.187) holds
for every node st+1 following st . Thus we can sum (5.187) over all nodes
st+1 |st to obtain
at+2 (st+1 )
!
at+1 (st ) X πt+1 (st+1 |st ) X ct+1 (st+1 ) + 1+r
− yt+1 (st+1 )
= at+1 (st ) = πt+1 (st+1 |st )
1+r t+1 t
1 + r t+1 t
1+r
s |s s |s
(5.188)
Substituting (5.188) back into (5.185) yields
X πt+1 (st+1 |st )ct+1 (st+1 ) X πt+1 (st+1 |st )at+2 (st+1 )
ct (st ) + +
t+1 t
1+r t+1 t
(1 + r)2
s |s s |s
t+1 t t+1
X πt+1 (s |s )yt+1 (s )
= yt (st ) + + at (st−1 ) (5.189)
1+r
st+1 |st
we obtain equation (5.43). But equation (5.191) follows from the No-Ponzi
conditions imposed in (5.27) and (5.28). For T < ∞ this is trivial since
aT +1 (sT ) ≥ 0 for all sT immediately implies (5.191). For T = ∞ we note
that the constraint
156 CHAPTER 5. THE SIM IN PARTIAL EQUILIBRIUM
rWt
θt ct = (5.192)
1+r
rWt−1
θt−1 ct−1 = . (5.193)
1+r
T −t
rWt rat r X yt+s
θt ct = = + Et (5.194)
1+r 1 + r 1 + r s=0 (1 + r)s
yields
T −t
r X yt+s
θt ct = r (at−1 + yt−1 − ct−1 ) + Et (5.196)
1 + r s=0 (1 + r)s
Bringing the term rct−1 to the left and dividing both sides by (1 + r) delivers
T −t
θt ct + rct−1 r r X yt+s
= (at−1 + yt−1 ) + E t s (5.197)
1+r 1+r (1 + r)2 s=0 (1 + r)
5.9. APPENDIX B: DERIVATION OF CONSUMPTION RESPONSE TO INCOME SHOCKS 157
Subtracting θt−1 ct−1 from both sides, using (5.193) and the definition of Wt−1
one gets
T −t T −t+1
θt ct + rct−1 r X yt+s r r X yt−1+s
−θt−1 ct−1 = E t s + yt−1 − E t−1
1+r (1 + r)2 s=0 (1 + r) 1 + r 1+r s=0
(1 + r)s
(5.198)
But now we note that the left hand side of equation (5.198) equals:
θt ct + rct−1 θt ct 1 r
− θt−1 ct−1 = − 1− − ct−1
1+r 1+r (1 + r)T −t+2 1 + r
θt ct 1 1
= − 1+r− − r ct−1
1+r 1+r (1 + r)T −t+1
θt ∆ct
= (5.199)
1+r
Plugging (5.199) back into equation (5.198) and multiplying both sides by
(1 + r) finally yields
T −t −t+1
TX
r X yt+s yt−1+s
θt ∆ct = Et s + ry t−1 − rE t−1
1 + r s=0 (1 + r) s=0
(1 + r)s
T −t −t+1
TX
r X yt+s yt−1+s
= Et s − rE t−1
1 + r s=0 (1 + r) s=1
(1 + r)s
T −t −t+1
TX
r X yt+s r yt−1+s
= Et s − Et−1
1 + r s=0 (1 + r) 1+r s=1
(1 + r)s−1
T −t T −t
r X yt+s r X yt+s
= Et s − Et−1
1 + r s=0 (1 + r) 1+r s=0
(1 + r)s
T −t
r X (Et − Et−1 )yt+s
= = ηt
1 + r s=0 (1 + r)s
A Digression: Stochastic
Earnings or Wage Processes
159
160CHAPTER 6. A DIGRESSION: STOCHASTIC EARNINGS OR WAGE PROCESSES
3. The stochastic part of the earnings process log(yit ): this is the part
we will approximate as a Markov chain below and take as input into
our quantitative models. An infinite horizon model abstracts from the
life cycle aspects of earnings as specified in part 1. and 2. Thus the
key model input in this class of models is the stochastic component of
earnings. Even in quantitative life cycle models where parts 1. and 2.
are modelled as well, the stochastic part of the earnings process is a key
ingredient of the model. Typically, the stochastic part of log-earnings
is modeled as the sum of a transitory and a persistent shock of the form
where
zit = ρzit−1 + πt ηit (6.4)
and the innovations are uncorrelated standard normal random vari-
ables, that are also uncorrelated with the (αi , βi ). Note that the persis-
tent part has to be initialized with some zi0 , often this initial condition
is simply set to zero. Thus the stochastic part of log-earnings is mod-
eled as a sum of a transitory shock with (potentially time varying)
variance φ2t and a persistent shock whose conditional variance is given
by πt2 and the persistence is given by the (time-invariant) parameter
ρ. Sometimes a low-order MA term is allowed for the transitory shock,
and some authors a priori impose ρ = 1, estimating the stochastic part
as a combination of a transitory and a fully permanent shock. The key
question for this part of the process is how persistent income shocks are,
that is, how big is ρ and how large is πt2 relative to φ2t . Since we know
162CHAPTER 6. A DIGRESSION: STOCHASTIC EARNINGS OR WAGE PROCESSES
from our previous discussion that more persistent shocks are harder to
self-insure against, the empirical answer to this question is evidently of
great importance for the quantitative properties of our models.
6.1 Estimation
The parameters to be estimated are the θt ’s, the variances and the covariance
of (αi , βi ) and the (φ2t , πt2 ) as well as ρ. In a first stage the θt ’s are estimated
by regressing individual or household log-earnings on observable household
characteristics, in Guvenen’s (2007) case a cubic polynomial in potential
experience hit . In a second stage the stochastic part of the earnings equation
is estimated (if αi , βi are random coefficients their variances are typically
estimated jointly with the remaining stochastic part of the process). While
there are various methods to do this, the most popular method is to use a
minimum distance estimator that, by choice of the parameters, minimizes
the weighted distance between specific cross-sectional moments measured in
the data and implied by the model.
The moments used in estimation are cross-sectional variances and auto-
covariances of log(eiht ), that is,
E(log(eiht )2 )
E(log(eiht ) log(eih+n,t+n ))
for n ≥ 1. For the data, E(.) denotes cross-sectional sample averages, for
the statistical model given jointly by (6.2)-(6.4) we can compute these cross-
sectional moments explicitly (alternatively one could simulate them easily).
Both in the data and in the model these moments could be computed sepa-
rately for different experience levels h, in practice small sample sizes in the
data leads most researchers to combine (i.e. take some average of) the data
for different experience groups. Guvenen (2007) provides the formulas for
these moments from the model and argues that from these moments the pa-
rameters of the statistical model are identified. In particular, the moments
from the model are given by
6.2 Results
Guvenen (2007) estimates the process above on US PSID data from 1968 to
1993. The following table summarizes his results.
The first three rows display the results if the stochastic earnings process is
estimated under the restriction that σβ2 = 0 (and consequently σαβ = 0). We
observe that with this restriction the persistent shock is very persistent as
the estimate of ρ is close to, and statistically indistinguishable from 1. This
result is very much consistent with the previous literature that has imposed
σβ2 = 0. There is substantial heterogeneity in the intercept of life cycle earn-
ings profiles (σ̂α2 is significantly different from zero) and this heterogeneity
is decreasing with educational attainment. Both the estimates of the size of
transitory and persistent shocks is consistent with the previous literature as
well (the table displays time averages of πt2 and φ2t ).
Once slope heterogeneity is allowed (the last three rows of the table), three
findings stand out. First, the estimated heterogeneity across households in
the slope of life cycle earnings profile is significantly different from zero in a
statistical sense, and is substantial from an economic point of view. Guvenen
(2007) calculated that of the entire cross-sectional earnings dispersion of a
cohort that has reached age 55, 75% of that variance for individuals with
a college degree and 55% of those with high school degree is accounted for
164CHAPTER 6. A DIGRESSION: STOCHASTIC EARNINGS OR WAGE PROCESSES
Thus the results in the table essentially contrast two views of how the
stochastic earnings process of individuals or households look like. According
to what Guvenen (2007) calls the Restricted Income Profile (RIP) view, life
cycle earnings are characterized by substantial heterogeneity in initial levels,
no heterogeneity in slopes and highly persistent shocks over the life cycle. In
contrast, according to the Heterogeneous Income Profiles (HIP) view, the key
heterogeneity across households is in the slopes of life cycle earnings profiles
which are realized at labor market entry. Subsequent earnings shocks are
not very persistent (and thus, as we have seen above, should be quite easy to
self-insure against). We will now discuss further where the initial evidence
(mainly by MaCurdy, 1982) in favor of RIP came from and why σβ2 = 0 and
ρ ≈ 1 seem to go hand in hand, and σβ2 0 and ρ 1 seem to go hand in
hand.
2
The variances of the transitory component declines somewhat in size, while that of
the persistent component increases (but remember, the persistent component is not nearly
as persistent anymore). Also note that the correlation between αi and βi is estimated to
be negative (albeit insignificantly so). This may suggest that households face trade-offs
between careers with low initial earnings levels but subsequently higher earnings growth,
and those with higher levels and lower growth.
6.3. INTERPRETATION 165
6.3 Interpretation
according to the model. Let us also assume that the process started at
t = −∞ so that
∞
X
zit = π ρt−τ ηiτ
τ =0
and hence
3
Simplicity means that the basic argument does not depend on these assumptions (but
the exact algebra does).
166CHAPTER 6. A DIGRESSION: STOCHASTIC EARNINGS OR WAGE PROCESSES
6.3.2 Size of ρ
The last point we want to discuss is that why, if the truth is HIP but we
estimate the earnings process under the restriction σβ2 = 0, we may end up
with an estimate of ρ that is biased upward. Take two households with the
same initial income and deterministic but different earnings growth
log(eit ) = α + βi t (6.6)
log(e11 ) = α + β1
log(e21 ) = α + β2
and has to interpret this, from the perspective of (6.7), as a negative shock
for household 1 and a positive shock for household 2. Since under the truth
(6.6) earnings of both households deviate more and more from the common
average over time, the econometrician has to interpret these larger and larger
deviations as repeated and very persistent negative shocks for household 1
and positive shocks for household 2. Guvenen’s (2007) figure 1 provides an
instructive visualization of this point. Thus when one estimates the earnings
process without allowing for slope heterogeneity when in fact it is present one
tends to find significantly higher values for ρ than when estimating it with σβ2
permitted. Note that the difference between an estimate (on annual basis)
of ρ = 0.98 (RIP) and ρ = 0.82 (HIP) is huge: under the RIP estimate after
20 years still 2/3’s of a shock is present in current earnings, whereas under
the HIP estimate only 2% of the shock 20 years ago is present. Obviously
self-insurance will do very well dealing with the latter shock, and not well at
all dealing with the former shock.
168CHAPTER 6. A DIGRESSION: STOCHASTIC EARNINGS OR WAGE PROCESSES
Chapter 7
169
170 CHAPTER 7. THE SIM IN GENERAL EQUILIBRIUM
by keeping interest rates and wages fixed when analyzing the change in
household behavior due to changes in particular policies (tax reform,
social security reform, welfare reform), may over- or understate the full
effects of such a reform. A claim, e.g., that a reform of the social se-
curity system towards a system that invests more funds in the stock
market is welfare improving may ignore the fact that, once large ad-
ditional funds are invested in the stock market, the excess return of
stocks over bonds (or the population growth rate plus the growth rate
of productivity), may diminish or vanish altogether. Only in cases in
which one can reasonably expect relative prices to remain uninfluenced
by policy reforms (for a small open economy, say) or one can convinc-
ingly argue that price effects are quantitatively unimportant1 would a
partial equilibrium analysis yield unbiased results.
labor endowment and hence labor income in the economy is constant over
time, i.e. there is no aggregate uncertainty. Without this assumption there
would be no hope for the existence of a stationary equilibrium in which wages
w and interest rates r are constant over time. In the next section we will
consider a model that is similar to this one, but will include one source of ag-
gregate uncertainty, in addition to the idiosyncratic labor income uncertainty
present in this model.
Each agent’s preferences over stochastic consumption processes are given
by
X ∞
u(c) = E0 β t U (ct ) (7.4)
t=0
1
with β = 1+ρ and ρ > 0. As before the agent can self-insure against idiosyn-
cratic labor endowment shocks by purchasing at period t uncontingent claims
to consumption at period t + 1 at a price qt = 1+r1t+1 . Again, in a stationary
equilibrium rt+1 will be constant across time. The agents budget constraint
is given by
ct + at+1 = wt yt + (1 + rt )at (7.5)
Note that we slightly change the way assets are traded: instead of zero
1
coupon bonds being traded at a discount q = 1+r we now consider a bond
that trades at price 1 today and earns gross interest rate (1+rt+1 ) tomorrow.
We do this for the following reason: the asset being traded will be physical
capital, with the interest rate being determined by the marginal product of
capital. As long as the interest rate is constant as in Aiyagari (1994) both for-
mulations are equivalent (if you derive the corresponding Euler equations for
both formulations, you’ll find that they’re identical). With aggregate uncer-
tainty as in Krusell and Smith (1998), however, it would make a substantial
difference whether agents can trade a risk free bond or, as Krusell and Smith
assume, risky capital. Thus, in order to be consistent with their formulation,
I opted for changing the budget constraint. Evidently all theoretical results
from the last section about the income fluctuation problem still apply (as
long as the interest rate is nonstochastic), because the optimality conditions
for the households are identical across both formulations.
We impose an exogenous borrowing constraint on asset holdings: at+1 ≥
0. Aiyagari (1994) considers several alternative borrowing constraints, but
uses the no-borrowing constraint in his applications. The agent starts out
with initial conditions (a0 , y0 ). His consumption at period t after endowment
7.1. A MODEL WITHOUT AGGREGATE UNCERTAINTY 173
shock history y t has been realized is denoted by ct (a0 , y t ) and his asset hold-
ings by at+1 (a0 , y t ). Let Φ0 (a0 , y0 ) denote the initial measure over (a0 , y0 )
across households. In accordance with our previous assumption the marginal
distribution of Φ0 with respect to y0 is assumed to be Π. At each point of
time an agent is characterized by her current asset position at and her cur-
rent income yt . These are her individual state variables. What describes the
aggregate state of the economy is the cross-sectional distribution over individ-
ual characteristics Φt (at , yt ). This concludes the description of the household
side of the economy.
On the production side we assume that competitive firms, taking as given
wages wt and interest rates rt , have access to a standard neoclassical produc-
tion technology3
Yt = F (Kt , Lt ) (7.6)
2.
rt = Fk (Kt , Lt ) − δ (7.8)
wt = FL (Kt , Lt ) (7.9)
3. For all t
Z X
Kt+1 = at+1 (a0 , y t )π(y t |y0 )dΦ0 (a0 , y0 ) (7.10)
y t ∈Y t
Z X
Lt = L̄ = yt π(y t |y0 )dΦ0 (a0 , y0 ) (7.11)
y t ∈Y t
Z X
ct (a0 , y t )π(y t |y0 )dΦ0 (a0 , y0 ) + Kt+1
y t ∈Y t
= F (Kt , Lt ) + (1 − δ)Kt (7.12)
The last three conditions are the asset market clearing, the labor market
clearing and the goods market clearing condition, respectively
Now let us define a recursive competitive equilibrium. We have already
conjectured what the correct state space is for our economy, with (a, y) being
the individual state variables and Φ(a, y) being the aggregate state variable.
First we need to define an appropriate measurable space on which the mea-
sures Φ are defined. Define the set A = [0, ∞) of possible asset holdings
and the set Y of possible labor endowment realizations. Define by P(Y ) the
power set of Y (i.e. the set of all subsets of Y ) and by B(A) the Borel σ-
algebra of A. Let Z = A × Y and B(Z) = P(Y ) × B(A). Finally define by M
the set of all probability measures on the measurable space M = (Z, B(Z)).
Why all this? Because our measures Φ will be required to be elements of M.
Now we are ready to define a recursive competitive equilibrium. At the heart
of any RCE is the recursive formulation of the household problem. Note that
7.1. A MODEL WITHOUT AGGREGATE UNCERTAINTY 175
3. For all Φ ∈ M
Z
K (Φ ) = K(H(Φ)) = a0 (a, y; Φ)dΦ
0 0
(7.18)
Z
L(Φ) = ydΦ (7.19)
Z Z
c(a, y; Φ)dΦ + a0 (a, y; Φ)dΦ (7.20)
= F (K(Φ), L(Φ)) + (1 − δ)K(Φ)
for all (a, y) ∈ Z and all (A, Y) ∈ B(Z). QΦ ((a, y), (A, Y)) is the probability
that an agent with current assets a and current income y ends up with
assets a0 in A tomorrow and income y 0 in Y tomorrow. Suppose that Y is a
singleton, say Y = {y1 }. The probability that tomorrow’s income is y 0 = y1 ,
given today’s income is π(y 0 |y). The transition of assets is non-stochastic
as tomorrows assets are chosen today according to the function a0 (a, y). So
either a0 (a, y) falls into A or it does not. Hence the probability of transition
from (a, y) to {y1 } × A is π(y 0 |y) if a0 (a, y) falls into A and zero if it does
not fall into A. If Y contains more than one element, then one has to sum
over the appropriate π(y 0 |y).
How does the function QΦ help us to determine tomorrow’s measure over
(a, y) from today’s measure? Suppose QΦ were a Markov transition matrix
for a finite state Markov chain and Φt would be the distribution today. Then
to figure out the distribution Φt tomorrow we would just multiply Q by Φt ,
or
Φt+1 = QTΦt Φt (7.22)
where here T stands for the transpose of a matrix. But a transition function
is just a generalization of a Markov transition matrix to uncountable state
spaces. Hence we need integrals:
Z
0
Φ (A, Y) = (H(Φ)) (A, Y) = QΦ ((a, y), (A, Y))Φ(da × dy) (7.23)
r = FK (K, L) − δ (7.24)
w = KL (K, L) (7.25)
3.
Z
K = a0 (a, y)dΦ (7.26)
Z
L = ydΦ (7.27)
Z Z
c(a, y)dΦ + a0 (a, y)dΦ = F (K, L) + (1 − δ)K (7.28)
Note the big simplification: value functions, policy functions and prices
are not any longer indexed by measures Φ, all conditions have to be satisfied
only for the equilibrium measure Φ. The last requirement states that the
measure Φ reproduces itself: starting with measure over incomes and assets
Φ today generates the same measure tomorrow. In this sense a stationary
RCE is the equivalent of a steady state, only that the entity characterizing
the steady state is not longer a number (the aggregate capital stock, say) but
a rather complicated infinite-dimensional object, namely a measure.5
where Ea(r) are the average asset holdings in the economy. This condition
requires equality between the demand for capital by firms and the supply of
capital by households (last period’s demand for assets, with physical capital
being the only asset in the economy).
5
If we restrict attention to a finite set of capital stocks, A = {a1 , . . . , aM } then Φ is
an M ∗ N × 1 column vector and the Markov transition function Q is an M ∗ N × M ∗ N
matrix with generic element qij,kl giving the probability of going from (a, y) = (ai , yl )
to (a0 , y 0 ) = (ak , yl ) tomorrow. Using the convention that rows index states today and
columns index states tomorrow, an invariant measure Φ has to satisfy the matrix equation
Φ = QT Φ.
That is, Φ is the eigenvector (rescaled to have the sum of its rows equal to 1) associated
with an eigenvalue λ = 1 of the matrix QT . Since QT is a stochastic matrix (every row
sums to 1 and all entires are nonnegative) it has at least one unit eigenvalue and thus at
least one stationary measure Φ. But if QT has multiple unit eigenvalues there are multiply
statioanry measures (in fact, a continuum of them, since convex combinations of stationary
measures are themselves stationary measures.
7.1. A MODEL WITHOUT AGGREGATE UNCERTAINTY 179
¿From (7.24) it is clear that the capital demand of the firm K(r) is a
function of r alone, defined implicitly as
Proposition 47 (Huggett 1993) For ρ > 0, r > −1 and y1 > 0 and CRRA
utility with σ > 1 the functional equation has a unique solution v which is
strictly increasing, strictly concave and continuously differentiable in its first
argument. The optimal policies are continuous functions that are strictly
increasing (for c(a, y)) or increasing or constant at zero (for a0 (a, y))
Similar results can be proved for the iid case and arbitrary bounded U
with ρ > r and ρ > 0, see Aiyagari (1994).
With respect to the boundedness of the state space, as seen in the pre-
vious section we require ρ > r and additional assumptions. Under these
assumptions there exists an ā s.t. a0 (ā, yN ) = ā and a0 (a, y) ≤ ā for all
y ∈ Y and all a ∈ [0, ā] = A. From now on we will restrict the state space to
Y × A and it will be understood that Z = A × Y. Thus, under the maintained
assumptions by Huggett or Aiyagari there exists an optimal policy a0r (a, y)
indexed by r. The next step is to ask what more is needed to make aggregate
asset demand Z
Ea(r) = a0r (a, y)dΦr (7.37)
has a unique fixed point (that Tr∗ maps M into itself follows from SLP,
Theorem 8.2). To show this Aiyagari (in the working paper version, and
quite loosely described) draws on a theorem in SLP and Huggett on a similar
theorem due to Hopenhayn and Prescott (1992). In both theorems the key
condition is a monotone mixing condition that requires a positive probability
7.1. A MODEL WITHOUT AGGREGATE UNCERTAINTY 181
Proposition 48 If
1. Qr is a transition function
2. Qr is increasing
Then the operator Tr∗ has a unique fixed point Φr and for all Φ0 ∈ M
the sequence of measures defined by
Φn = (T ∗ )n Φ0 (7.43)
converges weakly to Φr
tablished that both K(r) and Ea(r) are continuous functions on r ∈ (−δ, ρ),
that for low r we have that Ea(r) < K(r) and for high r < ρ we have
Ea(r) > K(r). These arguments together then guarantee the existence of
r∗ such that
K(r∗ ) = Ea(r∗ ) (7.46)
and hence the existence of a stationary recursive competitive equilibrium.
With respect to uniqueness the verdict is negative, as we can’t prove the
monotonicity of the function Ea(r) which is due to offsetting income and
substitution effects of saving with respect to the interest rate directly, as
well as the indirect effect that the interest rate has on the wage rate. Even
harder is the question about the stability of the stationary equilibrium. We
know that, provided the economy starts with initial distribution Φ0 = Φr∗ ,
then the economy remains there forever. The question arises whether, for
an arbitrary initial distribution Φ0 6= Φr∗ it is the case that Φt → Φr∗ and
rt → r∗ (either locally or globally). Note that to assess this question one has
to examine either the sequential equilibrium or the law of motion H in the
full-blown RCE. To the best of my knowledge no stability result has been
established for these types of economies as of today.
1. Fix an r ∈ (−δ, ρ). For a fixed r we can solve the household’s recursive
problem (e.g. by value function iteration or policy function iteration).
This yields a value function vr and decision rules a0r , cr , which obviously
depend on the r we picked.
S CM = δK CM (7.48)
S ∗ = δK ∗ > δK CM = S CM (7.49)
For this income process log(yt ) can take any value in (−∞, ∞). Our theory
developed so far has assumed a finite state space for yt ∈ Y or equiva-
lently Y log for log(yt ). The transformation of processes with continuous state
space into finite state Markov chains was pioneered in economics by George
186 CHAPTER 7. THE SIM IN GENERAL EQUILIBRIUM
Tauchen (1986) and roughly goes like this. First we pick the finite set Y log .
Aiyagari picks the number of states to be 7. Since log(yt ) can take any value
between (−∞, ∞), first subdivide the real line into 7 intervals
5
I1 = (−∞, − σε )
2
5 3
I2 = [− σε , − σε )
2 2
3 1
I3 = [− σε , − σε )
2 2
1 1
I4 = [− σε , σε )
2 2
1 3
I5 = [ σε , σε )
2 2
3 5
I6 = [ σε , σε )
2 2
5
I7 = [ σε , ∞) (7.54)
2
and take as state space for log-income the set of “midpoints” of the intervals
(x−θsi )2
Z − 2
0 e 2σy
Π = πT Π (7.57)
7.1. A MODEL WITHOUT AGGREGATE UNCERTAINTY 187
for Π. Given that the states for log(y) are given by Y log , we find the state
space for levels of income as
Y = {y1 , . . . , y7 } (7.60)
e−3σε e−2σε e−σε 1 eσε e2σε e3σε
{ , , , , , , }
ȳ ȳ ȳ ȳ ȳ ȳ ȳ
and now the average (and aggregate) labor endowment equals
X
yΠ(y) = 1. (7.61)
y∈Y
Quite a few labor economists these days believe, however, that the persistence
parameter should be higher than θ = 0.9, closer to 1 (what happens if θ = 1?).
But then there is substantial debate whether the simple AR(1) process for
the log of labor earnings is the appropriate stochastic process to model this
variable. Instead let us turn to the results.
188 CHAPTER 7. THE SIM IN GENERAL EQUILIBRIUM
Results
First note that for the Cobb-Douglas production function and L̄ = 1 we have
Y = K α and
r + δ = αK α−1 (7.64)
Thus
αY αδ
r+δ = = (7.65)
K s
where s is the aggregate saving rate. Thus
αδ
s= (7.66)
r+δ
The other implications of the model that Aiyagari discusses, the welfare
cost of idiosyncratic income fluctuations (as well as the welfare benefit from
access to self-insurance), as well as the predictions of the model with respect
to the wealth and consumption distribution will be addressed once the model
is enriched by aggregate fluctuations.
• Consequence of h = 1:
– Can’t borrow.
– Fraction γ ∈ (0, 1) of income lost.
State Transitions
• Assets a ∈ A = R.
• If h = 0 and d = 0:
B(a, y, h = 0; d = 0) = {c ≥ 0, a0 ∈ A :
c + q(a, y, h; a0 )a0 ≤ wy + a}.
Note that for some (a, y) ∈ A × Y this set might be empty (and the
household will need to default).
• If h = 0 and d = 1:
B(a, y, h = 0; d = 1) = {c ≥ 0, a0 = 0 : c ≤ wy}.
• If h = 1:
B(a, y, h = 1) = {c ≥ 0, a0 ≥ 0 :
c + q(a, y, h; a0 )a0 ≤ (1 − γ)wy + a}
• Maximization problem
Z
max0 n(a, y, h; a0 ) ∗
n(a,y,h;a )≥0 (a,y,h;a0 )
[1 − p(a, y, h; a0 )] a0
0 0
q(a, y, h; a )a −
1+r
1. Given (w, q), v solves the household Bellman equation and c, a0 , d are
the optimal policy functions.
2. Given (w, r), (K, L) satisfy the production firms’ first order condition.
[1 − p(a, y, h; a0 )] a0
0 0
q(a, y, h; a )a −
1+r
X
= max0 n(a, y, h; a0 )a0 ∗
n(a,y,h;a )≥0
(a,y,h;a0 )
[1 − p(a, y, h; a0 )]
0
q(a, y, h; a ) −
1+r
7.2. AN INCOMPLETE MARKETS MODEL WITH UNSECURED DEBT AND EQUILIBRIUM DEF
• Calibration to
government policy and all other exogenous elements that define preferences,
endowments and technology. Now, unexpectedly, either government policy
or some exogenous elements of the economy (such as the labor productivity
process) change. This change was completely unexpected by all agents of the
economy (a zero probability event), so that no anticipation actions were taken
by any agent. The exogenous change may be either transitory or permanent;
for the general discussion to follow this does not make a difference. We
want to study the transition path induced by the exogenous change, from
the old stationary equilibrium to a new stationary equilibrium (which may
coincide with the old stationary equilibrium in case the exogenous change is
of transitory nature, or may differ from it in case the exogenous change is
permanent.
For concreteness, suppose that the economy to start with is the standard
Aiyagari economy we studied in the previous section. As an example, we
consider as exogenous unexpected change the sudden permanent introduc-
tion of a capital income tax at rate τ. The receipts are rebated lump-sum
to households as government transfers T. The initial policy is characterized
by τ = T = 0. Obviously, since a tax on capital income changes households’
savings decisions, we expect that, due to the policy change, individual be-
havior and thus aggregate variables such as the interest rate, wage rate and
the capital stock change. We would also hope that, over time, the economy
settles down to its new stationary equilibrium associated with the capital
income tax. But since the economy starts, pre-reform, with an aggregate
state (aggregate capital, wealth distribution) not equal to the final station-
ary equilibrium, one would expect that it requires time for the economy to
settle down to its new stationary equilibrium. In other words, there will be
a nontrivial transition path induced by the reform.
Note that value functions are now functions of time also, since aggregate
prices and policies may change over time.
wt = FL (Kt , Lt ) (7.67)
rt = FK (Kt , Lt ) − δ. (7.68)
Tt = τt rt Kt (7.69)
for all t ≥ 0.
4. (Market Clearing):
Z
ct (at , yt )dΦt + Kt+1 = F (Kt , Lt ) + (1 − δ)Kt (7.70)
7.3. UNEXPECTED AGGREGATE SHOCKS AND TRANSITION DYNAMICS199
Z
Lt = yt dΦt (7.71)
Z
Kt+1 = at+1 (at , yt )dΦt (7.72)
had enough time to settle down to the new stationary equilibrium. The key
insight then is to realize that if vT = v∞ , then for a given sequence of prices
{rt , wt }Tt=1 the household problem can be solved backwards (note that this
is true independent of wether people live forever or not). This suggests the
following algorithm.
Algorithm 53 1. Fix T
2. Compute the stationary equilibrium Φ0 , v0 , c0 , a0 , r0 , w0 , K0 associated
with τ = τ0 = 0
3. Compute the stationary equilibrium Φ∞ , v∞ , c∞ , a∞ , r∞ , w∞ , K∞ asso-
ciated with τ∞ = τ. Assume that
ΦT , vT , cT , aT , rT , wT , KT = Φ∞ , v∞ , c∞ , a∞ , r∞ , w∞ , K∞
−1
4. Guess a sequence of capital stocks {K̂t }Tt=1 Note that since the capital
stock at time t = 1 is determined by decisions at time 0, K̂1 = K0 . Also
note that Lt = L0 = L̄ is fixed. Thus with the guesses on the capital
−1
stock we also obtain {r̂t , ŵt }Tt=1 determined by
ŵt = FL (K̂t , L̄)
r̂t = FK (K̂t , L̄) − δ
T̂t = τt r̂t K̂t .
−1 −1
5. Since we know vT (a, y) and {r̂t , ŵt , T̂t }Tt=1 we can solve for {v̂t , ĉt , ât+1 }Tt=1
backwards.
−1
6. With the policy functions {ât+1 } we can define the transition laws {Γ̂t }Tt=1 .
But since we know Φ0 = Φ1 from the initial stationary equilibrium, we
can iterate the distributions forward
Φ̂t+1 = Γ̂t (Φ̂t )
for t = 1, . . . , T − 1.
7. With {Φ̂t }Tt=1 we can compute
Z
Ât = adΦ̂t
for t = 1, . . . , T.
7.3. UNEXPECTED AGGREGATE SHOCKS AND TRANSITION DYNAMICS201
8. Check whether
max Ât − K̂t < ε
1≤t<T
−1
If yes, go to 9. If not, adjust your guesses for {K̂t }Tt=1 in 4.
9. Check whether
Φ̂T − ΦT
< ε. If yes, the transition converges smoothly
into the new steady state and we are done and should save {v̂t , ât+1 , ĉt , Φ̂t , r̂t , ŵt , K̂t }.
If not, go to 1. and increase T.
c1−σ
U (c) =
1−σ
202 CHAPTER 7. THE SIM IN GENERAL EQUILIBRIUM
v0 (a, y; g) = v1 (a, y)
or
Yt = st F (Kt , Lt ) (7.76)
where {st } is a sequence of random variables that follows a finite state Markov
chain with transition matrix π. We will follow Krusell and Smith (1998) and
assume that the aggregate productivity shock can take only two values
st ∈ {sb , sg } = S (7.77)
with sb < sg and denote by π(s0 |s) the conditional probability of the aggre-
gate state transiting from s today to s0 tomorrow. Krusell and Smith (1998)
interpret sb as an economic recession and sg as an expansion.
The idiosyncratic labor productivity yt is assumed to take only two values
yt ∈ {yu , ye } = Y (7.78)
where yu < ye stands for the agent being unemployed (having low labor pro-
ductivity) and ye stands for the agent being employed. Krusell and Smith
attach the employed-unemployed interpretation to the idiosyncratic labor
productivity variable, an interpretation which will be important in the cal-
ibration section. Obviously the probability of being unemployed is higher
during recessions than during expansions, and the Markov chain governing
idiosyncratic labor productivity should reflect this dependence of idiosyn-
cratic uncertainty on the aggregate state of the economy.
7.4. AGGREGATE UNCERTAINTY AND DISTRIBUTIONS AS STATE VARIABLES205
2. For the recursive equilibrium defined above Krusell and Smith assert
that the current cross-sectional asset distribution and the current shock
are sufficient aggregate state variables. This is not formally proved,
and such a proof would be very hard. What one has to prove is that a
recursive equilibrium generates a sequential equilibrium in the following
sense: Cross-sectional distributions are generated, starting from the
initial condition (s0 , Φ0 ) as follows
Φ1 (s1 ) = H(s0 , Φ0 , s1 ) (7.94)
Φt+1 (st+1 ) = H(st , Φt (st ), st+1 ) (7.95)
prices are generated by
rt (st ) = r(st , Φt (st )) (7.96)
wt (st ) = w(st , Φt (st )) (7.97)
and, starting from initial conditions (a0 , y0 ), an individual households’
allocation is generated by
c0 (a0 , y0 , s0 ) = c(a0 , y0 , s0 , Φ0 ) (7.98)
a1 (a0 , y0 , s0 ) = a0 (a0 , y0 , s0 , Φ0 ) (7.99)
and in general recursively
ct (a0 , y t , st ) = c(at (a0 , y t−1 , st−1 ), yt , st , Φt ) (7.100)
at+1 (a0 , y t , st ) = a0 (at (a0 , y t−1 , st−1 ), yt , st , Φt ) (7.101)
Similarly optimal choices of the firm can be generated. Thus, a given
recursive equilibrium generates sequential allocations and prices; it re-
mains to be verified that these prices are in fact a sequential equilib-
rium.
3. Finally, the issue of existence of a recursive equilibrium arises. We
know that, since a sequential equilibrium in general exists, there is a
state space large enough such that a recursive equilibrium (recursive
in that state space) exists. So the issue is whether a recursive equilib-
rium in which the aggregate state only contains the current shock and
the current wealth distribution does exist. Although this state space
seems “natural” in some sense, there is no guarantee of existence of
such a recursive equilibrium. The analysis of this economy is purely
computational in spirit as neither the existence, uniqueness, stability or
qualitative features of the equilibrium can be theoretically established.
7.4. AGGREGATE UNCERTAINTY AND DISTRIBUTIONS AS STATE VARIABLES209
Another way of putting it, the average capital stock tomorrow is not equal to
the saving function of some average, representative consumer, evaluated at
today’s average capital stock. If the optimal policy function for tomorrow’s
assets, would feature a constant propensity to save out of current assets and
income (a0 being linear in a, with same slope for all y ∈ Y ), then exact
aggregation would occur and in fact the average capital stock today would
be a sufficient statistic for the average capital stock tomorrow. This insight
is important for the quantitative results to come.
The computational strategy that Krusell and Smith (and many others
since) follow is to approximate the distribution Φ with a finite set of mo-
ments. Remember that Φ is the distribution over (a, y). Obviously, since y
can only take two values, the second dimension is not the problem, so in what
follows we focus on the discussion of the distribution over assets a. Let the
n-dimensional vector m denote the first n moments of the asset distribution
(i.e. the marginal distribution of Φ with respect to its first argument).
We now posit that the agents use an approximate law of motion
m0 = Hn (s, m) (7.103)
mapping the first n moments of the asset distribution today, m, into the first
n moments of the asset distribution tomorrow, m0 . Note that by doing so
agents are boundedly rational in the sense that moments of higher order than
n of the current wealth distribution may help to more accurately forecast the
210 CHAPTER 7. THE SIM IN GENERAL EQUILIBRIUM
for s ∈ {sb , sg }. Here (as , bs ) are parameters that need to be determined. The
recursive problem of the household then becomes
( )
XX
v(a, y, s, K) = max 0
U (c) + β π(y 0 , s0 |y, s)v(a0 , y 0 , s0 , K 0 )
c,a ≥0
y 0 ∈Y s0 ∈S
s.t. (7.105)
0
c + a = w(s, K)y + (1 + r(s, K))a (7.106)
log(K 0 ) = as + bs log(K) (7.107)
Note that we have reduced the state space from something that includes the
infinite-dimensional space of measures to a four dimensional space (a, y, s, K) ∈
R × Y × S × R. The algorithm for computing an approximate recursive equi-
librium is then as follows:
1. Guess (as , bs )
7.4.3 Calibration
In this economy no analytical results can be proved and one has to resort
to numerical analysis. Krusell and Smith take the model period to be one
quarter (note that this is a business cycle model). As preferences they as-
sume CRRA utility with rather low risk aversion of σ = 1 (i.e. log-utility).
The time discount factor is chosen to be β = 0.994 = 0.96 on an annual
basis, which implies a yearly subjective time discount rate of ρ = 4.1%
as in Aiyagari. Also similar to Aiyagari they take the capital share to be
α = 0.36. As annual depreciation rate they choose δ = (1−0.025)4 −1 = 9.6%,
slightly higher than Aiyagari, but within the range of values commonly used
in real business cycle studies. The remaining parameters describe the joint
aggregate-idiosyncratic labor productivity process. Unfortunately the paper
itself does not contain a precise discussion of the parameterization, so that
the discussion here relies partly on Krusell and Smith’s information from
the paper, partly on Imrohoroglu’s (1989) paper to which they refer to and
partly on the FORTRAN code posted on Tony Smith’s web site.
The calibration strategy is to first calibrate the aggregate component of
the productivity process, i.e. the set S and the 2×2 matrix π(s0 |s). Remember
212 CHAPTER 7. THE SIM IN GENERAL EQUILIBRIUM
that the aggregate state represents recessions and expansions. With respect
to S they choose
S = {0.99, 1.01} (7.110)
I would think that the standard deviation of the technology shock is a bit on
the small side with 0.01. In fact, using Cooley and Prescott’s (1995) contin-
uous state process and discretizing into a two state chain yields a standard
deviation of the shock of about 0.02. Krusell and Smith claim that with their
aggregate process they are able to generate aggregate fluctuations of output
similar to US data, which, given the information in the paper I wasn’t able
to verify. Given that Cooley and Prescott need sufficiently more variance
in the technology shock to generate business cycles of reasonable size, this
must mean that the economy with heterogenous agents and uninsurable id-
iosyncratic risk, for a given aggregate shock variance, is more volatile than
its representative agent counterpart.
As for the transition matrix for the aggregate shock the first assumption
made is symmetry, so that π(sg |sg ) = π(sb |sb ). Krusell and Smith choose
π(sg |sg ) such that, conditional on being in the good state today, the expected
time in the good state are 8 quarters, or
or
1
8 =
1 − π(sg |sg )
7
π(sg |sg ) = (7.112)
8
so that
7 1
0
π(s |s) = 8
1
8
7 (7.113)
8 8
It follows that, conditional on being in the bad state today, the expected
time of staying there is also 8 quarters.
With respect to idiosyncratic labor productivity, the state space is chosen
as
Y = {0.25, 1} (7.114)
Remember that the idiosyncratic states are meant to represent employment
and unemployment, so that it is assumed that an unemployed person makes
7.4. AGGREGATE UNCERTAINTY AND DISTRIBUTIONS AS STATE VARIABLES213
In short, the best times for finding a job are when the economy moves from
a recession to an expansion, the worst chances are when the economy moves
from a boom into a recession. Combining the aggregate transition prob-
abilities with the idiosyncratic probabilities, conditional on the aggregate
transitions, finally yields as transition matrix (7.82)
0.525 0.035 0.09375 0.0099
0.35 0.84 0.03125 0.1151
π= 0.03125 0.0025 0.292 0.0245
(7.119)
0.09375 0.1225 0.583 0.8505
where τj is the set of time indices for which st = sj and Tj is the cardinality
of that set. Note that, since the regression is on log-income, σj can be inter-
preted as the average percentage error for the prediction of the capital stock
216 CHAPTER 7. THE SIM IN GENERAL EQUILIBRIUM
2
ε̂jt
P
t∈τj
Rj2 = 1 − P 2 (7.127)
t∈τj log Kt+1 − log K
where G(y) is a function of the stochastic income process. Thus under cer-
tainty equivalence
at+1 = at + H(y) (7.132)
and thus the saving function a0 has slope 1 under certainty equivalence (and
ρ = r). In Krusell and Smith’s economy agents are prudent and face liquidity
constraints, but almost act as if they are certainty equivalence consumers.
Why? A few reasons come to my mind:
1. Agents are prudent, but not all that much. A σ = 1 is at the lower end
of the empirical estimates for risk aversion. As we saw from Aiyagari’s
(1994) paper, the amount of precautionary saving increases significantly
with increases in σ, and so should the deviation of agents decision rules
from certainty equivalence.
Krusell and Smith consider various sensitivity tests with respect to these
and other dimensions (change in time discount rates, endogenous labor-
leisure choice etc.), and claim that the result of quasi-aggregation (only the
mean capital stock matters for forecasting tomorrows mean capital stock) is
robust to changes in the model parameterization.
s.t.
0
c + a = w(s, K)y + (1 + r(s, K))a (7.134)
log(K 0 ) = as + bs log(K) (7.135)
They pick B = {0.9858, 0.9894, 0.993}. Hence annual discount rates differ
between 2.8% for the most patient agents and 5.6% for the most impatient
agents. As transition matrix Krusell and Smith propose (remember that the
model period is one quarter)
0.995 0.005 0
γ = 0.000625 0.99875 0.000625 (7.136)
0 0.005 0.995
which implies that 80% of the population has discount factor in the middle
and 10% are on either of the two extremes (in the stationary distribution).
220 CHAPTER 7. THE SIM IN GENERAL EQUILIBRIUM
It also implies that patient and impatient agents remain so for an expected
period of 50 years. De facto this parameterization creates three deterministic
types of agents. The patient agents are the ones that will accumulate most
of the wealth in this economy. With this trick of stochastic discount factors
Krusell and Smith are able to approximate the US wealth distribution to a
reasonably accurate degree (see their Figure 3): in the modified economy the
richest 1% of the population holds 24% of all wealth, the richest 5% hold 55%
percent of all wealth. Also note that the quasi-aggregation result extends to
the economy with stochastic discount factors.
Part III
221
223
In this chapter we will consider models that, in spirit, build on the com-
plete markets model considered in Chapter 3, in the same fashion the models
in Chapter 4 and 5 built on the simple Pilch model in Chapter 3. Most em-
pirical studies reject the complete insurance hypothesis and thus cast doubt
upon the complete markets model as a reasonable description of reality.
Therefore models are desired that predict some, but not perfect risk shar-
ing. The Pilch model (in its general equilibrium form) generates some “risk
sharing” via self-insurance: agents smooth part of their income fluctuation
by asset accumulation and decumulation, with the part being determined
by preferences and the nature of the income shocks. But remember that
there are no formal risk-sharing arrangements in the Pilch model, as explicit
contingent insurance contracts, which agents in the model would have an
incentive to trade and financial intermediaries would have an incentive to
offer, are ruled out by assumption, without any good reason from within the
model.9
The models we study in this chapter will allow a full set of contingent
consumption claims being traded (in the decentralized version) or being allo-
cated by the social planner (in the centralized version of the model). That full
insurance does not arise as optimal and equilibrium allocation is due to infor-
mational and/or enforcement frictions, which are explicitly modeled. These
models adhere to the principle, most forcefully articulated by Townsend that
the world is constrained efficient, and that is up to the researcher/modeler
to find the right set of constraints that give rise to model allocations which
are in line with empirical consumption allocations.
We will study two such sets of constraints: the first stemming from the
fact that in actual economies financial contracts are only imperfectly enforce-
able (because there exist explicit bankruptcy provisions in the legal code or
it is too costly to always enforce repayment), the second deriving from the
fact that individual incomes and/or actions are imperfectly observable, so
that contingent claims payments cannot be directly conditioned on these.
Both types of models will deliver allocations characterized by some, but (de-
pending on parameterizations) imperfect insurance, which is due explicitly
to these frictions.
9
Townsend (1985) and Cole and Kocherlakota (1997) study environments with private
information and show that the optimal contract between agents and financial intermedi-
aries is a simple uncontingent debt contract, closely resembling the one-period uncontin-
gent bonds that we let agents trade in the Pilch model.
224
Chapter 8
Limited Enforceability of
Contracts
225
226 CHAPTER 8. LIMITED ENFORCEABILITY OF CONTRACTS
For given initial s0 , the constrained Pareto Frontier Ws0 : [U 1 (s0 ), Ū 1 (s0 )] →
[U 2 (s0 ), Ū 2 (s0 )] is defined as
The state space is (w, s) ∈ [U 1 (s0 ), Ū 1 (s0 )]×S. This is a standard dynamic
programming problem that can be solved with standard techniques. Note
that first the values of autarky have to be determined, which is done by
2
I deviate from the timing convention in Kocherlakota, who lets the economy start after
an initial s0 has been drawn, and analyzes allocations from period 1 onward. I formulate
the problem so that consumption also takes place in period 0, after the period 0 shock has
been realized. This is more consistent with my discussion in earlier chapters and will help
to simplify the discussion of the continuum economy.
230 CHAPTER 8. LIMITED ENFORCEABILITY OF CONTRACTS
8.4 Decentralization
Since both the papers by Kocherlakota and by Kehoe and Levine will be
presented in class, I will skip the discussion of how constrained efficient al-
locations can be decentralized within a subgame perfect equilibrium of a
3
Strictly speaking one also needs the upper bound on utilities Ū (s). One guesses these
bounds and then iterates between guesses and solutions of the Bellman equation, until one
attains V (Ū (s), s) = U 2,Aut (s) for all s.
8.4. DECENTRALIZATION 231
For any allocation (c1 , c2 ), the implied interest rate are said to be high if
XX
Q(st |s0 ) ∗ c1t (st ) + c2t (st ) < ∞
(8.23)
t≥0 st
The main implication of this result is that one can solve for the entire set
of potential equilibrium allocations by solving the recursive planning problem
and then find Arrow securities prices and borrowing constraints from (8.20)
and (8.21) that, together with the consumption allocation, form a sequential
markets equilibrium.
the consumption distribution. This example follows the one used in Krueger
and Perri’s (2006) application of limited commitment models to income and
consumption inequality.
Let us assume that the aggregate state of the world can take two values
st ∈ S = {1, 2}. Let individual endowments be given by the functions
1 1 + ε if st = 1
e (st ) = (8.25)
1 − ε if st = 2
2 1 − ε if st = 1
e (st ) = (8.26)
1 + ε if st = 2
That is, if st = 1 then agent 1 has currently high endowment 1 + ε, if st = 2
then agent 2 has currently high endowment. Here ε measures the variability
of individual endowment and also turns out be the (unconditional) standard
deviation of the cross-sectional income distribution.
Which agent is rich follows a stochastic process that is assumed to be
Markov with transition probabilities
δ 1−δ
π= (8.27)
1−δ δ
where δ ∈ (0, 1) denotes the conditional probability of agent 1 being rich
tomorrow, conditional on being rich today (and symmetrically for agent 2).
Thus δ is a measure of persistence of the income process, with δ = 21 denoting
the iid case and δ = 1 reflecting a deterministic income process (that is,
permanent shocks). The stationary distribution associated with π, for any
given δ ∈ (0, 1) is given by Π(s) = 21 for all s ∈ S, and we assume that
Π(s0 ) = 21 for all s0 ∈ S. This assumption makes agents ex ante identical and
allows us to restrict attention to symmetric equilibrium allocations.
The good thing about this model is that we can solve for the continuation
expected discounted utility from autarky analytically. Since there are only
two aggregate states at each point of time, the continuation value from the
autarkic allocation can take two values, one for the currently rich agent,
denoted by U (1+ε), and one for the currently poor agent, denoted by U (1−ε).
These two values solve the following recursion
U (1 + ε) = (1 − β)u(1 + ε) + δU (1 + ε) + (1 − δ)U (1 − ε)
U (1 − ε) = (1 − β)u(1 − ε) + δU (1 − ε) + (1 − δ)U (1 + ε) (8.28)
234 CHAPTER 8. LIMITED ENFORCEABILITY OF CONTRACTS
The proof of this lemma is straightforward and hence omitted. The in-
terpretation is also fairly simple: without income fluctuations the autarkic
allocation is first-best; since an increase in income variability reduces current
consumption and increases future consumption risk, it reduces the continua-
tion utility for the currently poor agent; for the rich agent initially the direct
effect of currently higher consumption dominates the risk effect, but as ε be-
comes large the risk effect becomes dominant; the last two properties follow
from strict concavity of U (1 + ε) and the signs of the derivatives at ε = 0
and ε → 1. The first figure at the end of this chapter graphically summarizes
the lemma.
We now want to characterize constrained-efficient consumption alloca-
tions in this model and analyze how they change with changes in ε. For this
we note that in any insurance mechanism as the one described in this model it
is efficient to transfer resources from the currently rich to the currently poor
agent. Therefore the constrained-efficient consumption distribution features
maximal insurance, subject to delivering at least the continuation utility of
autarky to the currently rich agent. This argument motivates the following
proposition, which is due to Kehoe and Levine (2001):4
ε)) = U (1 + ε). Obviously one solution for εc (ε) = ε, i.e. the autarkic
consumption allocation. We note from the previous lemma that if ε ≤ ε1 ,
then this is in fact the resulting consumption allocation. In this case, for
small ε, there is no risk sharing possible whatsoever, since at any level of
risk sharing the rich agent would have an incentive to default. However, if
ε ∈ (ε1 , ε2 ) then there exists an εc (ε) < ε1 such that U (1 + ε) = U (1 + εc (ε)).
In this case the resulting consumption allocation features some risk sharing,
but not complete risk sharing. The second figure at the end of this chapter
shows a sample path for the income process and the resulting constrained-
efficient sample path for the consumption allocation.
How does the dispersion of the consumption distribution εc (ε) vary with
an increase in income dispersion. From the previous discussion we immedi-
ately obtain the following proposition (see Krueger and Perri, 2006a):
This discussion is summarized in the third figure at the end of this chap-
ter. Similarly one can establish how the consumption distribution varies
with changes in the persistence of the income process. See Kehoe and Levine
(2001) or Krueger and Perri (2006a) for a proof.
Here V (w, y) is the total (normalized) resource cost the planner has to
minimally spend in order to fulfill his utility promises w, without violating
the individual rationality constraint of the agent: he can’t promise less from
tomorrow onwards, in any state of the world, than the agent would obtain
from the autarkic allocation. As before, multiplying the current cost by the
5
Of course this is not their idea; it goes back at least to Abreu (1986) and Spear and
Srivastava (1987).
8.6. A CONTINUUM ECONOMY 239
Since V is strictly convex, this implies that either g(w; y 0 ) < w or g(w; y 0 ) =
U Aut (y 0 ). That is, for states in which the individual rationality constraints are
not binding utility promises for tomorrow are lower than for today; for states
in which the constraint is binding utility promises are raised sufficiently in
order to prevent the agent from reneging. Define w = miny∈Y U Aut (y) and
w̄ = maxy∈Y U Aut (y). It is clear that for all w we must have g(w; y 0 ) ≥
w. On the other hand we have that for ymax = arg maxy∈Y U Aut (y) we
have g(U Aut (ymax ), ymax ) = U Aut (ymax ) and for all w > U Aut (ymax ) we have
g(w, y 0 ) ≤ g(w, ymax ) ≤ U Aut (y 0 ). Thus for iid income shocks the state space
for promised utility w, denoted by W, is bounded: W = [w, w̄]. For 2 income
shocks the fourth figure at the end of this chapter demonstrates the situation;
for a formal proof of the assertions above see Krueger and Perri (2006b). You
may want it instructive to follow an agent with given initial utility promise
w0 through his life with a sequence of income shocks: one positive income
shock brings utility to w = w, with a sequence of bad income shock making
the agent move down in promised utility, until he hits w = w, with a single
good shock putting him back at w = w̄.
The remaining discussion of this model has strong similarities with the
discussion in Aiyagari (1994) and Huggett (1993) for standard incomplete
markets models. First, for a given interest rate R, the dynamic programming
problem (??) delivers value function V (w, y) and policy functions h(w, y) and
g(w, y; y 0 ). In order to find the associated stationary distribution over utility
promises and income shocks ΨR we first determine the Markov transition
function QR induced by π and g(w, y; y 0 ). First one has to prove that there is
an upper bound for utility promises w, denoted by w̄ (this is straightforward
for π being iid, but not for the general case). Then denote the state space for
utility promises by W = [w, w̄], define Z = W ×Y and B(Z) = B(W )×P(Y ).
8.6. A CONTINUUM ECONOMY 241
Of course one has to prove that such invariant measure exists and is unique,
the proof of which for the iid case is contained in Krueger (1999). For the
general, non-iid case I’m not aware of any such result, which is similar to the
status of theoretical results available for the standard incomplete markets
model discussed in the previous chapter.
The last figure at the end of this chapter shows the stationary consump-
tion distribution associated with ΨR (w, y), again for the iid case with only
two possible income shocks. Since consumption C(h(w, y)) is strictly increas-
ing in utility promises w, the consumption distribution mimics the utility dis-
tribution ΨR . In particular, all agents with currently high shock consume the
same, and the maximum in the consumption distribution, with agents with
a sequence of bad shocks working themselves down through the consumption
distribution until they have hit rock bottom.
So for a given intertemporal shadow interest rate R we now know how de-
termine the associated stationary utility promise distribution. So far nothing
assures that the social planner can satisfy this distribution of utility promises
with the aggregate resources available to society. Total resources available
are Z X
ydΨR = yΠ(y) (8.43)
y
How about total resources needed by the planner? An agent that enters
this period with utility promises w and current income y is awarded current
utility h(w, y). Thus requires, in terms of resources, C(h(w, y)). Thus total
resources required to satisfy utility distribution ΨR in the current period (and
by stationarity in each period) are
Z
C(h(w, y))dΨR (8.44)
242 CHAPTER 8. LIMITED ENFORCEABILITY OF CONTRACTS
π(y 0 |y)
q(y 0 |y) = qπ(y 0 |y) = (8.46)
R
which justifies that R is in fact called a shadow interest rate: it turns out
to be the equilibrium risk free interest rate in the corresponding equilibrium
with borrowing constraints that are not too tight.
Finally we want to discuss the relationship between initial promised utili-
ties w0 and initial assets a0 . So suppose we have found a stationary constrained-
efficient utility distribution Ψ(w, y), a corresponding R∗ and associated value
and policy functions V (w, y), h(w, y) and g(w, y; y 0 ). First, analogously to
(8.15) and (8.16) from the recursive policy functions we can construct sequen-
tial constrained-efficient consumption allocations {ct (w0 , y0 )} for an agent
8.6. A CONTINUUM ECONOMY 243
Q(y0 ) = Π(y0 )
Q(y t ) = q(yt |yt−1 ) ∗ . . . q(y1 |y0 )Π(y0 ) (8.47)
ct (a0 , y t ) = ct (a−1 t
0 (a0 , y0 ), y ) (8.49)
where w0 = a−1 0 (a0 , y0 ) is the inverse function of (8.48) with respect to the
first argument (note that this function is well-defined because a0 (w0 , y0 ) is
strictly increasing in w0 ). The distribution Φ(a0 , y0 ) is then determined as
Φ(a0 , y0 ) = Ψ(a−1
0 (a0 , y0 ), y0 ) (8.50)
What one then can prove (see Krueger (1999) is that for an initial dis-
tribution Φ(a0 , y0 ) given by (8.50), the allocation determined by (8.49) and
prices (8.46) are a competitive equilibrium with borrowing constraints that
are not too tight (or alternatively, form a competitive equilibrium in the spirit
of Kehoe and Levine (1993), where equilibrium prices are given by (8.47)).
A few final comments: the method of using utility promises as state
variables to make dynamic contracting problems recursive has been used
in a number of applications, for example in the study of optimal unemploy-
ment insurance (Shavell and Weiss 1979, Hopenhayn and Nicolini 1997, Zhao
2000), asset pricing and the equity premium (Alvarez and Jermann 2000,
Lustig 2001), sovereign debt (Atkeson 1991, Kletzer and Wright 2000), redis-
tributive taxation (Krueger and Perri 2006b, Attanasio and Rios-Rull 2001),
time-consistent monetary policy (Chang 1998), time-consistent fiscal policy
(Phelan and Stacchetti 1999, Sleet 1998) and risk sharing in village economies
(Udry 1994, Ligon, Thomas and Worrall 2001).
Also, instead of using utility promises Marcet and Marimon (1999) have
developed a recursive method that use (cumulative) Lagrange multipliers as
244 CHAPTER 8. LIMITED ENFORCEABILITY OF CONTRACTS
state variables. As before, the same problems with the curse of dimensionality
when the number of heterogeneous agents becomes large arises.
So what next: on the methodological side one would like to figure out
how to handle models with a large number of agents and aggregate uncer-
tainty, both theoretically and numerically. Lustig (2001) takes a big step in
that direction. On the substantial side a careful study of redistribution and
insurance over the business cycle, of optimal insurance of large income risks
in the future, both by government policies and private arrangements, and of
the dynamics of the income, consumption and wealth distribution seems to
be fruitful avenues for future research.
Chapter 9
Private Information
In this section we consider another friction that may prevent perfect risk
sharing from occurring as a result of efficient or equilibrium consumption
allocations. As in standard Arrow Debreu theory now households and fi-
nancial intermediaries can write legally binding and enforceable contracts.
However, we now assume that individual income realizations (and individual
consumption) is private information of the agent. Financial intermediaries or
the social planner have to rely on reports of income by agents. Consumption
allocations therefore have to structured in such a way that households find
it optimal to tell the truth about their income realization, rather than to
lie about it. We will first consider the problem of a financial intermediary
dealing with a single agent in isolation, before discussing a model with many
agents, an aggregate resource constraint and an endogenous interest rate
245
246 CHAPTER 9. PRIVATE INFORMATION
from the risk-neutral principal. Both parties can commit to long-term con-
tracts, so that in the absence of private information the optimal consumption
allocation for the agent, subject to the financial intermediary breaking even,
is
ct (y t ) = E(yt ) = E(y) = 1
where the last equality is by assumption (that is, we normalize mean income
to 1). Thus the agent hands over his realized income in every period and
receives constant consumption equal to mean income back from the financial
intermediary. His lifetime utility equals
∞ X
X
U (c) = (1 − β) πt (y t )β t u(ct (y t )) = u(1)
t=0 yt
The big problem with this consumption allocation is that the agent’s
income realizations are private information. If the agent is promised constant
consumption independent of his income report, then he would always report
yt = y1 , keep the difference between his true income and the report, receive
1 from the financial intermediary and do strictly better. Then, however, the
principal would lose money, since his profits would be
∞ X
X
W = (1 − β) πt (y t )β t (y1 − 1)
t=0 yt
= y1 − 1 < 0.
above that without providing these incentives the principal is going to lose
money).
What we want to construct in the following is the efficient long-term
insurance contract between the two parties, explicitly taking into account
the informational frictions in this environment. We will again immediately
proceed to the recursive formulation of the problem, understanding that in
principle one should first write down the sequential problem, then the re-
cursive problem and then prove the principle of optimality.1 The recursive
formulation of the contracting problem again makes use of promised lifetime
utility w as a state variable. Let us first pose the dynamic programming
problem and then explain it:
X
V (w) = min πs [(1 − β)ts + βV (ws )] (9.1)
{ts ,ws }N
s=1
s.t.
X
w = πs [(1 − β)u(ts + ys ) + βws ] (9.2)
(1 − β)u(ts + ys ) + βws ≥ (1 − β)u(tk + ys ) + βwk ∀s, k (9.3)
ts ∈ [−ys , ∞) (9.4)
ws ∈ [w, w̄] (9.5)
The constraints (9.3) say the following: suppose state s with associated
income realization ys is realized. If the agent reports the truth, he receives
transfers ts and continuation utility ws , for total lifetime utility
u(ts + ys ) + βws .
u(tk + ys ) + βwk .
Note that even when mis-reporting income, his true lifetime utility depends
on his true, rather than his reported income ys . The constraints (9.3) simply
state that it has to be in the agents’ own interest to truthfully reveal his
income. The domain restriction on transfers is self-explanatory: the principal
can make arbitrarily large transfers, but cannot force the agent to make
payments bigger than their current (truthfully reported) income ys . With
respect to the bounds on promised utility we note the following. Let w̄ =
supc u(c) and w = inf c u(c); evidently it is not possible to deliver more lifetime
utility that w̄ even with infinite resources and it is not possible to deliver less
lifetime utility than w even without giving the agent any consumption even.
We explicitly allow w = −∞ and w̄ = +∞, but will make more restrictive
assumptions later. For now we assume that
Assumption: The utility function u : [0, ∞) → R is strictly increasing,
strictly concave, at least twice differentiable and satisfies the Inada condi-
tions.
So far is has been left unclear what determines the w the agent will start
the contract with. A higher initial w means more lifetime utility for the
agent, but less lifetime utility for the principal. The expected lifetime utility
for the principal, conditional on delivering w to the agent, is given by
W (w) = 1 − V (w)
since V (w) measures the lifetime expected costs from the contract and 1
measures the expected revenues from the agent, equal to his expected per
period and (by normalization of 1 − β) lifetime revenue. Thus varying the w
traces out the constrained utility possibility frontier between principal and
agent. One particularly important w is that w that solves V (w) = 1, that is,
that lifetime utility of the agent that yields 0 profits for the principal. If V
is strictly increasing in w, such w is necessarily unique.
9.1. PARTIAL EQUILIBRIUM 249
u(c) = w
or cf b (w) = u−1 (w). The value of the transfers needed to deliver the constant
consumption stream cf b (w) are given by
V (w) = cf b (w) − 1.
Obviously
V (w) ≤ V (w)
with strict inequality if any of the incentive constraints is binding. On the
other hand, the principal can decide to give constant transfers ts = t (and
constant continuation utility ws = w) Surely, since transfers are independent
of reported income, this induces truth-telling. In order to obtain the utility
promise w, the transfer has to satisfy
X
πs u(ys + t) = w
s
V̄ (w) = t̄(w)
and evidently
V (w) ≤ V̄ (w)
with strict inequality whenever the distribution of income shocks is not de-
generate.
Observe the following facts. Since cf b (w) is strictly increasing and strictly
convex, so is V (w). Furthermore, as long as u satisfies the Inada conditions we
have V 0 (w) = 0 and V 0 (w̄) = ∞, since u(.) and cf b (.) are inverse functions
of each other. Finally we have V (w) = −1 and V (w̄) = ∞. Similarly,
the function t̄(w) and thus V̄ (w) are strictly increasing and strictly convex,
250 CHAPTER 9. PRIVATE INFORMATION
with V̄ (w̄) = ∞. While this does not necessarily mean that V (w) is strictly
increasing (it is straightforward to show that it is) and strictly convex (it is
harder to show that it is), this discussion gives us a fairly tight bound on
V (w).
The fact that V (w) is strictly increasing in w is intuitive: delivering
higher lifetime utility to the agent requires higher transfers on average by
the principal. Since marginal utility is declining, a given additional unit of
transfers increases utility by less and less, thus one would expect the cost
function to be strictly convex. We will go ahead and assert this without
proof.
Now we discuss further properties of the optimal contract. Consider the
constraints
or
u(ts + ys ) − u(ts−1 + ys ) ≥ u(ts + ys−1 ) − u(ts−1 + ys−1 )
Since u is strictly concave and ys > ys−1 we have ts−1 ≥ ts (it helps to draw a
picture to convince you of this). From (9.6) it then easily follows that ws ≥
ws−1 . That is, it is efficient to give households with lower income realizations
higher transfers. But in order to provide the incentives of not always claiming
to have had low income realizations, these households are “punished” with
lower continuation utilities. We summarize this in the following
Second we claim that the principal only has to worry about the agent
lying locally, that is, if income is ys , the principal has to only worry about
the agent reporting ys−1 or ys+1 . As long as the agent does not have an
incentive to lie locally, he does not want to lie and report ys−2 or ys+2 or
even more extreme values. Formally
9.1. PARTIAL EQUILIBRIUM 251
Proposition 66 Suppose that the true income state is s and that all local
constraints of the form
Proof. We will show that, if these constraints are satisfied, then a house-
hold would not want to report s − 1 when the true state is s + 1. A simple
repetition of the argument will then show that the household would not want
to report s − 1 for any state k ≥ s. The case of mis-reporting upwards is, of
course, symmetric and hence omitted.
Since s > s − 1, we have, from the previous result, ts ≤ ts−1 and thus
(again by concavity of u)
But
(1 − β)u(ts+1 + ys+1 ) + βws+1 ≥ (1 − β)u(ts + ys+1 ) + βws
(use (9.9), but for s + 1) and thus
That is, if the household does not want to lie one state down from s to s − 1,
he does not want to lie two states down, from s + 1 to s − 1 either.
This result reduces the set of incentive constraints substantially. It turns
out that we can reduce it even further. Without proof we state the following
proposition (if you are interested in the proof, consult Thomas and Worrall
(1990) or the discussion of the same paper in Sargent and Ljungqvist’s book;
note that the proof requires strict convexity of the cost function V ).
Proof. Omitted
Intuitively, this result makes a lot of sense. Remember, the purpose of the
contract is for the principal to insure the agent against bad income shocks,
with high transfers in states with bad income realizations. This, of course,
triggers the incentive to report lower income than actually realized, rather
than higher income. Thus the lying-down constraints are the crucial con-
straints. With this result the dynamic programming problem becomes more
manageable. In particular, if N = 2, there is only one incentive constraint
and the promise keeping constraint. It is then quite feasible to characterize
the qualitative properties of the optimal contract in more detail, which we
leave to the discussion of Atkeson and Lucas (1992) in class.
Let us consider instead a simple example with N = 2 income states. From
our characterization of the binding pattern of the constraints the dynamic
programming problem becomes
V (w) = min π [(1 − β)t1 + βV (w1 )] + (1 − π) [(1 − β)t2 + βV (w2 )]
{t1 ,t2 ,w1 ,w2 }
s.t.
w = π [(1 − β)u(t1 + y1 ) + βw1 ] + (1 − π) [(1 − β)u(t2 + y2 ) + βw2 ]
(1 − β)u(t2 + y2 ) + βw2 ≥ (1 − β)u(t1 + y2 ) + βw1
where π is the probability of low income y1 . Let us proceed under the as-
sumption that V is differentiable (which is not straightforward to prove).
The the first order and envelope conditions read as
π(1 − β) − λπ(1 − β)u0 (t1 + y1 ) + µ(1 − β)u0 (t1 + y2 ) = 0
(9.11)
0 0
(1 − π)(1 − β) − λ(1 − π)(1 − β)u (t2 + y2 ) − µ(1 − β)u (t2 + y2 ) = 0
(9.12)
0
πβV (w1 ) − λπβ + µβ = 0
(9.13)
0
(1 − π)βV (w2 ) − λ(1 − π)β − µβ = 0
(9.14)
0
V (w) = λ
(9.15)
Rewriting equations (9.13) and (9.14) and using (9.15) yields
µ µ
λ = V 0 (w) = V 0 (w1 ) + = V 0 (w2 ) −
π 1−π
9.2. ENDOGENOUS INTEREST RATES IN GENERAL EQUILIBRIUM253
Thus, since µ > 0, we have (by strict convexity of the cost function V ) that
w1 < w < w2 , that is, utility promises spread out over time. Rewriting (9.11)
and (9.12) yields
µ
1 = λu0 (t1 + y1 ) − u0 (t1 + y2 )
π
µ 0 µ µ 0
= λ− u (t1 + y1 ) + [u0 (t1 + y1 ) − u0 (t1 + y2 )] > λ − u (t1 + y1 )
π π π
µ
1 = λ+ u0 (t2 + y2 )
1−π
and thus
µ 0
1 = V 0 (w2 )u0 (c2 ) = V 0 (w1 )u0 (c1 ) + [u (c1 ) − u0 (t1 + y2 )]
π
> V 0 (w1 )u0 (c1 )
The first equality (and again using strict convexity of the cost function) show
that if future promises w2 are increased (in response to, say, an increase in w),
then it is necessarily optimal to also increase c2 . That is, the principal should
spread costs over time, increasing both current costs and well as future costs.
[To be completed]
9.3 Applications
9.3.1 New Dynamic Public Finance
Paper that makes the closest connection to the traditional Ramsey tax lit-
erature is Werning (2007). Key paper starting this literature is Golosov,
Kocherlakota and Tsyvinski (2003) and accessible summaries are Golosov,
Tsyvinski and Werning (2006) as well as Kocherlakota (20xx).
254 CHAPTER 9. PRIVATE INFORMATION
Part IV
Conclusions
255
Bibliography
257
258 BIBLIOGRAPHY
[17] Attanasio, O., J. Banks, C. Meghir and G. Weber, (1999), “Humps and
Bumps in Lifetime Consumption”, Journal of Business and Economic
Statistics.
[19] Barro, R. (2009), “Rare Disasters, Asset Prices, and Welfare Costs,”
American Economic Review, 99, 243–264.
[51] Conesa, J. and Krueger, D. (1999), “Social Security Reform with Het-
erogeneous Agents,” Review of Economic Dynamics, 2, 757-795.
[61] Deaton, A. and Paxson, C. (1997), “The Effects of Economic and Pop-
ulation Growth on National Saving and Inequality,” Demography, 34,
97-114.
[65] Engen, E., Gale, W. and Scholz, J. (1996), “The Illusory Effect of
Saving Incentives on Saving”, Journal of Economic Perspectives, 10,
113-38.
[71] Galı̀, J. (1990), “Finite Horizon, Life-Cycle Savings and Time Series
Evidence on Consumption,” Journal of Monetary Economics, 26, 433-
52.
[72] Gervais, M. and P. Klein (2010), “Measuring Consumption Smoothing
in CEX Data,” Journal of Monetary Economics, 57, 988-999.
[73] Gollier, C. (2002), “What Does the Classical Theory Have to Say About
Household Portfolios?,” in Household Portfolios, L. Guiso, M. Halias-
sos, T. Jappelli (eds.), MIT Press, Cabridge, MA.
[74] Golosov, M., N. Kocherlakota and A. Tsyvinski (2003), “Optimal Indi-
rect and Capital Taxation,” Review of Economic Studies, 70, 569-587.
[75] Golosov, M., A. Tsyvinski and I. Werning (2006), “New Dynamic Pub-
lic Finance: A User’s Guide,” NBER Macroeconomics Annual 2006,
317-363.
[76] Gourinchas, P. and Parker, J. (2002), “Consumption over the Life Cy-
cle,” Econometrica 70, 47-89.
[77] Green, E. (1987), “Lending and the Smoothing of Uninsurable In-
come,” in: Prescott, E. and Wallace, N. (eds) Contractual Arrange-
ments for Intertemporal Trade, University of Minnesota Press, 3-25.
[78] Guiso, L., Jappelli, T. and Terlizzese, D. (1992), “Earnings Uncertainty
and Precautionary Saving,” Journal of Monetary Economics 30, 307-
37.
[79] Guvenen, F. (2007a), “An Empirical Investigation of Labor Income
Processes,” Working Paper, University of Texas
[80] Guvenen, F. (2007b), “Learning your Earning: Are Labor Income
Shocks Really Very Persistent?” American Economic Review, 97, 687-
712.
[81] Haliassos, M. and Bertaut, C. (1995), “Why Do so Few Hold Stocks?,”
Economic Journal, 105, 1110-29.
[82] Hall, R. (1978), “Stochastic Implications of the Life Cycle-Permanent
Income Hypothesis: Theory and Evidence,” Journal of Political Econ-
omy, 86, 971-987.
264 BIBLIOGRAPHY
[89] Hayashi, F., Altonji, J. and Kotlikoff, L. (1996), “Risk Sharing Between
and Within Families,” Econometrica, 64, 261-94.
[106] Hurd, M. (1987), “Savings of the Elderly and Desired Bequests,” Amer-
ican Economic Review, 77, 298-312.
[109] Judd, K., L. Maliar and S. Maliar (2012), “Merging Simulation and
Projection Approaches to Solve High-Dimensional Problems,” NBER
Working Paper 18501.
[113] King, M. and Dycks Mireaux, L. (1982), “Asset Holdings and the Life-
Cycle,” Economic Journal 92, 247-67.
[135] Li, Wenli and R. Yao (2007), “The Life-Cycle Effects of House Price
Changes,” Journal of Money, Credit, and Banking, 1375-1409.
[136] Livshits, I., J. MacGee and M. Tertilt (2007), “Accounting for the Rise
in Consumer Bankruptcies,” Working Paper, Stanford University.
[138] MaCurdy, T. (1982), “The Use of Time Series Processes to Model the
Error Structure of Earnings in a Longitudinal Data Analysis,” Journal
of Econometrics, 18, 82-114.
[157] Shea, J. (1995), “Union Contracts and the Life-Cycle Permanent In-
come Hypothesis,” American Economic Review, 85, 186-200.
[163] Storesletten, K., C. Telmer and A. Yaron (2007), “Asset Pricing with
Idiosyncratic Risk and Overlapping Generations,” Review of Economic
Dynamics, 10, 519-548.
[168] Weil, P. (1989), “The Equity Premium Puzzle and the Risk-Free Rate
Puzzle,” Journal of Monetary Economics, 24, 401-421.
[171] Yaari, M. (1965), “Uncertain Lifetime, Life Insurance, and the Theory
of the Consumer,” Review of Economic Studies, 32, 137–150.