294
Macroeconomic Theory Dirk Krueger 1 Department of Economics Stanford University September 25, 2002 1 I am grateful to my teachers in Minnesota, V.V Chari, Timothy Kehoe and Edward Prescott, my colleagues at Stanford, Robert Hall, Beatrix Paal and Tom Sargent, my co-authors Juan Carlos Conesa, Jesus Fernandez-Villaverde and Fabrizio Perri as well as Victor Rios-Rull for helping me to learn modern macroeconomic theory. All remaining errors are mine alone.

Macroeconomics Phd Lectures notes

Embed Size (px)

Citation preview

Page 1: Macroeconomics Phd Lectures notes

Macroeconomic Theory

Dirk Krueger1

Department of EconomicsStanford University

September 25, 2002

1I am grateful to my teachers in Minnesota, V.V Chari, Timothy Kehoe and EdwardPrescott, my colleagues at Stanford, Robert Hall, Beatrix Paal and Tom Sargent,my co-authors Juan Carlos Conesa, Jesus Fernandez-Villaverde and Fabrizio Perri aswell as Victor Rios-Rull for helping me to learn modern macroeconomic theory. Allremaining errors are mine alone.

Page 2: Macroeconomics Phd Lectures notes

ii

Page 3: Macroeconomics Phd Lectures notes

Contents

1 Overview and Summary 1

2 A Simple Dynamic Economy 5

2.1 General Principles for Specifying a Model . . . . . . . . . . . . . 5

2.2 An Example Economy . . . . . . . . . . . . . . . . . . . . . . . . 6

2.2.1 Definition of Competitive Equilibrium . . . . . . . . . . . 7

2.2.2 Solving for the Equilibrium . . . . . . . . . . . . . . . . . 8

2.2.3 Pareto Optimality and the First Welfare Theorem . . . . 11

2.2.4 Negishi’s (1960) Method to Compute Equilibria . . . . . . 14

2.2.5 Sequential Markets Equilibrium . . . . . . . . . . . . . . . 18

2.3 Appendix: Some Facts about Utility Functions . . . . . . . . . . 23

3 The Neoclassical Growth Model in Discrete Time 27

3.1 Setup of the Model . . . . . . . . . . . . . . . . . . . . . . . . . . 27

3.2 Optimal Growth: Pareto Optimal Allocations . . . . . . . . . . . 28

3.2.1 Social Planner Problem in Sequential Formulation . . . . 29

3.2.2 Recursive Formulation of Social Planner Problem . . . . . 31

3.2.3 An Example . . . . . . . . . . . . . . . . . . . . . . . . . 33

3.2.4 The Euler Equation Approach and Transversality Condi-tions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

3.3 Competitive Equilibrium Growth . . . . . . . . . . . . . . . . . . 49

3.3.1 Definition of Competitive Equilibrium . . . . . . . . . . . 50

3.3.2 Characterization of the Competitive Equilibrium and theWelfare Theorems . . . . . . . . . . . . . . . . . . . . . . 52

3.3.3 Sequential Markets Equilibrium . . . . . . . . . . . . . . . 56

3.3.4 Recursive Competitive Equilibrium . . . . . . . . . . . . . 57

4 Mathematical Preliminaries 59

4.1 Complete Metric Spaces . . . . . . . . . . . . . . . . . . . . . . . 60

4.2 Convergence of Sequences . . . . . . . . . . . . . . . . . . . . . . 61

4.3 The Contraction Mapping Theorem . . . . . . . . . . . . . . . . 65

4.4 The Theorem of the Maximum . . . . . . . . . . . . . . . . . . . 71

iii

Page 4: Macroeconomics Phd Lectures notes

iv CONTENTS

5 Dynamic Programming 735.1 The Principle of Optimality . . . . . . . . . . . . . . . . . . . . . 735.2 Dynamic Programming with Bounded Returns . . . . . . . . . . 80

6 Models with Uncertainty 836.1 Basic Representation of Uncertainty . . . . . . . . . . . . . . . . 836.2 Definitions of Equilibrium . . . . . . . . . . . . . . . . . . . . . . 85

6.2.1 Arrow-Debreu Market Structure . . . . . . . . . . . . . . 856.2.2 Sequential Markets Market Structure . . . . . . . . . . . . 876.2.3 Equivalence between Market Structures . . . . . . . . . . 88

6.3 Markov Processes . . . . . . . . . . . . . . . . . . . . . . . . . . . 886.4 Stochastic Neoclassical Growth Model . . . . . . . . . . . . . . . 90

7 The Two Welfare Theorems 937.1 What is an Economy? . . . . . . . . . . . . . . . . . . . . . . . . 937.2 Dual Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 967.3 Definition of Competitive Equilibrium . . . . . . . . . . . . . . . 997.4 The Neoclassical Growth Model in Arrow-Debreu Language . . . 997.5 A Pure Exchange Economy in Arrow-Debreu Language . . . . . 1017.6 The First Welfare Theorem . . . . . . . . . . . . . . . . . . . . . 1037.7 The Second Welfare Theorem . . . . . . . . . . . . . . . . . . . . 1047.8 Type Identical Allocations . . . . . . . . . . . . . . . . . . . . . . 113

8 The Overlapping Generations Model 1158.1 A Simple Pure Exchange Overlapping Generations Model . . . . 116

8.1.1 Basic Setup of the Model . . . . . . . . . . . . . . . . . . 1178.1.2 Analysis of the Model Using Offer Curves . . . . . . . . . 1228.1.3 Inefficient Equilibria . . . . . . . . . . . . . . . . . . . . . 1298.1.4 Positive Valuation of Outside Money . . . . . . . . . . . . 1348.1.5 Productive Outside Assets . . . . . . . . . . . . . . . . . . 1368.1.6 Endogenous Cycles . . . . . . . . . . . . . . . . . . . . . . 1388.1.7 Social Security and Population Growth . . . . . . . . . . 140

8.2 The Ricardian Equivalence Hypothesis . . . . . . . . . . . . . . . 1458.2.1 Infinite Lifetime Horizon and Borrowing Constraints . . . 1468.2.2 Finite Horizon and Operative Bequest Motives . . . . . . 155

8.3 Overlapping Generations Models with Production . . . . . . . . . 1608.3.1 Basic Setup of the Model . . . . . . . . . . . . . . . . . . 1618.3.2 Competitive Equilibrium . . . . . . . . . . . . . . . . . . 1618.3.3 Optimality of Allocations . . . . . . . . . . . . . . . . . . 1688.3.4 The Long-Run Effects of Government Debt . . . . . . . . 172

9 Continuous Time Growth Theory 1779.1 Stylized Growth and Development Facts . . . . . . . . . . . . . . 177

9.1.1 Kaldor’s Growth Facts . . . . . . . . . . . . . . . . . . . . 1789.1.2 Development Facts from the Summers-Heston Data Set . 178

9.2 The Solow Model and its Empirical Evaluation . . . . . . . . . . 183

Page 5: Macroeconomics Phd Lectures notes

CONTENTS v

9.2.1 The Model and its Implications . . . . . . . . . . . . . . . 1869.2.2 Empirical Evaluation of the Model . . . . . . . . . . . . . 189

9.3 The Ramsey-Cass-Koopmans Model . . . . . . . . . . . . . . . . 1999.3.1 Mathematical Preliminaries: Pontryagin’s Maximum Prin-

ciple . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2009.3.2 Setup of the Model . . . . . . . . . . . . . . . . . . . . . . 2009.3.3 Social Planners Problem . . . . . . . . . . . . . . . . . . . 2029.3.4 Decentralization . . . . . . . . . . . . . . . . . . . . . . . 210

9.4 Endogenous Growth Models . . . . . . . . . . . . . . . . . . . . . 2159.4.1 The Basic AK-Model . . . . . . . . . . . . . . . . . . . . 2169.4.2 Models with Externalities . . . . . . . . . . . . . . . . . . 2209.4.3 Models of Technological Progress Based on Monopolistic

Competition: Variant of Romer (1990) . . . . . . . . . . . 232

10 Bewley Models 24510.1 Some Stylized Facts about the Income and Wealth Distribution

in the U.S. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24610.1.1 Data Sources . . . . . . . . . . . . . . . . . . . . . . . . . 24610.1.2 Main Stylized Facts . . . . . . . . . . . . . . . . . . . . . 247

10.2 The Classic Income Fluctuation Problem . . . . . . . . . . . . . 25310.2.1 Deterministic Income . . . . . . . . . . . . . . . . . . . . 25410.2.2 Stochastic Income and Borrowing Limits . . . . . . . . . . 262

10.3 Aggregation: Distributions as State Variables . . . . . . . . . . . 26610.3.1 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26610.3.2 Numerical Results . . . . . . . . . . . . . . . . . . . . . . 273

11 Fiscal Policy 27911.1 Positive Fiscal Policy . . . . . . . . . . . . . . . . . . . . . . . . . 27911.2 Normative Fiscal Policy . . . . . . . . . . . . . . . . . . . . . . . 279

11.2.1 Optimal Policy with Commitment . . . . . . . . . . . . . 27911.2.2 The Time Consistency Problem and Optimal Fiscal Policy

without Commitment . . . . . . . . . . . . . . . . . . . . 279

12 Political Economy and Macroeconomics 281

13 References 283

Page 6: Macroeconomics Phd Lectures notes

vi CONTENTS

Page 7: Macroeconomics Phd Lectures notes

Chapter 1

Overview and Summary

After a quick warm-up for dynamic general equilibrium models in the first partof the course we will discuss the two workhorses of modern macroeconomics, theneoclassical growth model with infinitely lived consumers and the OverlappingGenerations (OLG) model. This first part will focus on techniques rather thanissues; one first has to learn a language before composing poems.I will first present a simple dynamic pure exchange economy with two in-

finitely lived consumers engaging in intertemporal trade. In this model theconnection between competitive equilibria and Pareto optimal equilibria can beeasily demonstrated. Furthermore it will be demonstrated how this connec-tion can exploited to compute equilibria by solving a particular social plannersproblem, an approach developed first by Negishi (1960) and discussed nicely byKehoe (1989).This model with then enriched by production (and simplified by dropping

one of the two agents), to give rise to the neoclassical growth model. Thismodel will first be presented in discrete time to discuss discrete-time dynamicprogramming techniques; both theoretical as well as computational in nature.The main reference will be Stokey et al., chapters 2-4. As a first economicapplication the model will be enriched by technology shocks to develop theReal Business Cycle (RBC) theory of business cycles. Cooley and Prescott(1995) are a good reference for this application. In order to formulate thestochastic neoclassical growth model notation for dealing with uncertainty willbe developed.This discussion will motivate the two welfare theorems, which will then be

presented for quite general economies in which the commodity space may beinfinite-dimensional. We will draw on Stokey et al., chapter 15’s discussion ofDebreu (1954).The next two topics are logical extensions of the preceding material. We will

first discuss the OLG model, due to Samuelson (1958) and Diamond (1965).The first main focus in this module will be the theoretical results that distinguishthe OLG model from the standard Arrow-Debreu model of general equilibrium:in the OLG model equilibria may not be Pareto optimal, fiat money may have

1

Page 8: Macroeconomics Phd Lectures notes

2 CHAPTER 1. OVERVIEW AND SUMMARY

positive value, for a given economy there may be a continuum of equilibria(and the core of the economy may be empty). All this could not happen inthe standard Arrow-Debreu model. References that explain these differences indetail include Geanakoplos (1989) and Kehoe (1989). Our discussion of theseissues will largely consist of examples. One reason to develop the OLG modelwas the uncomfortable assumption of infinitely lived agents in the standardneoclassical growth model. Barro (1974) demonstrated under which conditions(operative bequest motives) an OLG economy will be equivalent to an economywith infinitely lived consumers. One main contribution of Barro was to providea formal justification for the assumption of infinite lives. As we will see thismethodological contribution has profound consequences for the macroeconomiceffects of government debt, reviving the Ricardian Equivalence proposition. Asa prelude we will briefly discuss Diamond’s (1965) analysis of government debtin an OLG model.In the next module we will discuss the neoclassical growth model in con-

tinuous time to develop continuous time optimization techniques. After havinglearned the technique we will review the main developments in growth the-ory and see how the various growth models fare when being contrasted withthe main empirical findings from the Summers-Heston panel data set. We willbriefly discuss the Solow model and its empirical implications (using the arti-cle by Mankiw et al. (1992) and Romer, chapter 2), then continue with theRamsey model (Intriligator, chapter 14 and 16, Blanchard and Fischer, chapter2). In this model growth comes about by introducing exogenous technologicalprogress. We will then review the main contributions of endogenous growth the-ory, first by discussing the early models based on externalities (Romer (1986),Lucas (1988)), then models that explicitly try to model technological progress(Romer (1990).All the models discussed up to this point usually assumed that individuals

are identical within each generation (or that markets are complete), so thatwithout loss of generality we could assume a single representative consumer(within each generation). This obviously makes life easy, but abstracts from alot of interesting questions involving distributional aspects of government policy.In the next section we will discuss a model that is capable of addressing theseissues. There is a continuum of individuals. Individuals are ex-ante identical(have the same stochastic income process), but receive different income realiza-tions ex post. These income shocks are assumed to be uninsurable (we thereforedepart from the Arrow-Debreu world), but people are allowed to self-insure byborrowing and lending at a risk-free rate, subject to a borrowing limit. Deaton(1991) discusses the optimal consumption-saving decision of a single individualin this environment and Aiyagari (1994) incorporates Deaton’s analysis into afull-blown dynamic general equilibrium model. The state variable for this econ-omy turns out to be a cross-sectional distribution of wealth across individuals.This feature makes the model interesting as distributional aspects of all kindsof government policies can be analyzed, but it also makes the state space verybig. A cross-sectional distribution as state variable requires new concepts (de-veloped in measure theory) for defining and new computational techniques for

Page 9: Macroeconomics Phd Lectures notes

3

computing equilibria. The early papers therefore restricted attention to steadystate equilibria (in which the cross-sectional wealth distribution remained con-stant). Very recently techniques have been developed to handle economies withdistributions as state variables that feature aggregate shocks, so that the cross-sectional wealth distribution itself varies over time. Krusell and Smith (1998)is the key reference. Applications of their techniques to interesting policy ques-tions could be very rewarding in the future. If time permits I will discuss suchan application due to Heathcote (1999).

For the next two topics we will likely not have time; and thus the corre-sponding lecture notes are work in progress. So far we have not consideredhow government policies affect equilibrium allocations and prices. In the nextmodules this question is taken up. First we discuss fiscal policy and we startwith positive questions: how does the governments’ decision to finance a givenstream of expenditures (debt vs. taxes) affect macroeconomic aggregates (Barro(1974), Ohanian (1997))?; how does government spending affect output (Baxterand King (1993))? In this discussion government policy is taken as exogenouslygiven. The next question is of normative nature: how should a benevolent gov-ernment carry out fiscal policy? The answer to this question depends cruciallyon the assumption of whether the government can commit to its policy. A gov-ernment that can commit to its future policies solves a classical Ramsey problem(not to be confused with the Ramsey model); the main results on optimal fiscalpolicy are reviewed in Chari and Kehoe (1999). Kydland and Prescott (1977)pointed out the dilemma a government faces if it cannot commit to its policy-this is the famous time consistency problem. How a benevolent governmentthat cannot commit should carry out fiscal policy is still very much an openquestion. Klein and Rios-Rull (1999) have made substantial progress in an-swering this question. Note that we throughout our discussion assume that thegovernment acts in the best interest of its citizens. What happens if policies areinstead chosen by votes of selfish individuals is discussed in the last part of thecourse.

As discussed before we assumed so far that government policies were eitherfixed exogenously or set by a benevolent government (that can or can’t commit).Now we relax this assumption and discuss political-economic equilibria in whichpeople not only act rationally with respect to their economic decisions, but alsorationally with respect to their voting decisions that determine macroeconomicpolicy. Obviously we first had to discuss models with heterogeneous agents sincewith homogeneous agents there is no political conflict and hence no interestingdifferences between the Ramsey problem and a political-economic equilibrium.This area of research is not very far developed and we will only present twoexamples (Krusell et al. (1997), Alesina and Rodrik (1994)) that deal with thequestion of capital taxation in a dynamic general equilibrium model in whichthe capital tax rate is decided upon by repeated voting.

Page 10: Macroeconomics Phd Lectures notes

4 CHAPTER 1. OVERVIEW AND SUMMARY

Page 11: Macroeconomics Phd Lectures notes

Chapter 2

A Simple DynamicEconomy

2.1 General Principles for Specifying a Model

An economic model consists of different types of entities that take decisionssubject to constraints. When writing down a model it is therefore crucial toclearly state what the agents of the model are, which decisions they take, whatconstraints they have and what information they possess when making theirdecisions. Typically a model has (at most) three types of decision-makers

1. Households: We have to specify households preferences over commodi-ties and their endowments of these commodities. Households are as-sumed to maximize their preferences, subject to a constraint set thatspecifies which combination of commodities a household can choose from.This set usually depends on the initial endowments and on market prices.

2. Firms: We have to specify the technology available to firms, describ-ing how commodities (inputs) can be transformed into other commodities(outputs). Firms are assumed to maximize (expected) profits, subject totheir production plans being technologically feasible.

3. Government: We have to specify what policy instruments (taxes, moneysupply etc.) the government controls. When discussing government policyfrom a positive point of view we will take government polices as given(of course requiring the government budget constraint(s) to be satisfied),when discussing government policy from a normative point of view wewill endow the government, as households and firms, with an objectivefunction. The government will then maximize this objective function bychoosing policy, subject to the policies satisfying the government budgetconstraint(s)).

5

Page 12: Macroeconomics Phd Lectures notes

6 CHAPTER 2. A SIMPLE DYNAMIC ECONOMY

In addition to specifying preferences, endowments, technology and policy, wehave to specify what information agents possess when making decisions. Thiswill become clearer once we discuss models with uncertainty. Finally we haveto be precise about how agents interact with each other. Most of economicsfocuses on market interaction between agents; this will be also the case in thiscourse. Therefore we have to specify our equilibrium concept, by makingassumptions about how agents perceive their power to affect market prices.In this course we will focus on competitive equilibria, by assuming that allagents in the model (apart from possibly the government) take market pricesas given and beyond their control when making their decisions. An alternativeassumption would be to allow for market power of firms or households, whichinduces strategic interactions between agents in the model. Equilibria involvingstrategic interaction have to be analyzed using methods from modern gametheory, which you will be taught in the second quarter of the micro sequence.To summarize, a description of any model in this course should always con-

tain the specification of the elements in bold letters: what commodities aretraded, preferences over and endowments of these commodities, technology, gov-ernment policies, the information structure and the equilibrium concept.

2.2 An Example Economy

Time is discrete and indexed by t = 0, 1, 2, . . . There are 2 individuals that liveforever in this pure exchange economy. There are no firms or any government inthis economy. In each period the two agents trade a nonstorable consumptiongood. Hence there are (countably) infinite number of commodities, namelyconsumption in periods t = 0, 1, 2, . . .

Definition 1 An allocation is a sequence (c1, c2) = (c1t , c2t )∞t=0 of consump-tion in each period for each individual.

Individuals have preferences over consumption allocations that can be rep-resented by the utility function

u(ci) =∞Xt=0

βt ln(cit) (2.1)

with β ∈ (0, 1).This utility function satisfies some assumptions that we will often require in

this course. These are further discussed in the appendix to this chapter. Notethat both agents are assumed to have the same time discount factor β.Agents have deterministic endowment streams ei = eit∞t=0 of the consump-

tion goods given by

e1t =

½20if t is evenif t is odd

e2t =

½02if t is evenif t is odd

Page 13: Macroeconomics Phd Lectures notes

2.2. AN EXAMPLE ECONOMY 7

There is no uncertainty in this model and both agents know their endowmentpattern perfectly in advance. All information is public, i.e. all agents knoweverything. At period 0, before endowments are received and consumption takesplace, the two agents meet at a central market place and trade all commodities,i.e. trade consumption for all future dates. Let pt denote the price, in period 0,of one unit of consumption to be delivered in period t, in terms of an abstractunit of account. We will see later that prices are only determined up to aconstant, so we can always normalize the price of one commodity to 1 and makeit the numeraire. Both agents are assumed to behave competitively in thatthey take the sequence of prices pt∞t=0 as given and beyond their control whenmaking their consumption decisions.After trade has occurred agents possess pieces of paper (one may call them

contracts) stating

in period 212 I, agent 1, will deliver 0.25 units of the consumptiongood to agent 2 (and will eat the remaining 1.75 units)in period 2525 I, agent 1, will receive one unit of the consumption

good from agent 2 (and eat it).

and so forth. In all future periods the only thing that happens is that agentsmeet (at the market place again) and deliveries of the consumption goods theyagreed upon in period 0 takes place. Again, all trade takes place in period 0and agents are committed in future periods to what they have agreed upon inperiod 0. There is perfect enforcement of these contracts signed in period 0.1

2.2.1 Definition of Competitive Equilibrium

Given a sequence of prices pt∞t=0 households solve the following optimizationproblem

maxcit∞t=0

∞Xt=0

βt ln(cit)

s.t.∞Xt=0

ptcit ≤

∞Xt=0

pteit

cit ≥ 0 for all t

Note that the budget constraint can be rewritten as

∞Xt=0

pt(eit − cit) ≥ 0

1A market structure in which agents trade only at period 0 will be called an Arrow-Debreumarket structure. We will show below that this market structure is equivalent to a marketstructure in which trade in consumption and a particular asset takes place in each period, amarket structure that we will call sequential markets.

Page 14: Macroeconomics Phd Lectures notes

8 CHAPTER 2. A SIMPLE DYNAMIC ECONOMY

The quantity eit−cit is the net trade of consumption of agent i for period t whichmay be positive or negative.

For arbitrary prices pt∞t=0 it may be the case that total consumption inthe economy desired by both agents, c1t + c

2t at these prices does not equal total

endowments e1t + e2t ≡ 2. We will call equilibrium a situation in which prices

are “right” in the sense that they induce agents to choose consumption so thattotal consumption equals total endowment in each period. More precisely, wehave the following definition

Definition 2 A (competitive) Arrow-Debreu equilibrium are prices pt∞t=0 andallocations (cit∞t=0)i=1,2 such that

1. Given pt∞t=0, for i = 1, 2, cit∞t=0 solves

maxcit∞t=0

∞Xt=0

βt ln(cit) (2.2)

s.t.∞Xt=0

ptcit ≤

∞Xt=0

pteit (2.3)

cit ≥ 0 for all t (2.4)

2.

c1t + c2t = e

1t + e

2t for all t (2.5)

The elements of an equilibrium are allocations and prices. Note that wedo not allow free disposal of goods, as the market clearing condition is statedas an equality.2 Also note the ˆ’s in the appropriate places: the consumptionallocation has to satisfy the budget constraint (2.3) only at equilibrium pricesand it is the equilibrium consumption allocation that satisfies the goods marketclearing condition (2.5). Since in this course we will usually talk about com-petitive equilibria, we will henceforth take the adjective “competitive” as beingunderstood.

2.2.2 Solving for the Equilibrium

For arbitrary prices pt∞t=0 let’s first solve the consumer problem. Attachthe Lagrange multiplier λi to the budget constraint. The first order necessary

2Different people have different tastes as to whether one should allow free disposal or not.Personally I think that if one wishes to allow free disposal, one should specify this as part oftechnology (i.e. introduce a firm that has available a technology that uses positive inputs toproduce zero output; obviously for such a firm to be operative in equilibrium it has to be thecase that the price of the inputs are non-positive -think about goods that are actually badssuch as pollution).

Page 15: Macroeconomics Phd Lectures notes

2.2. AN EXAMPLE ECONOMY 9

conditions for cit and cit+1 are then

βt

cit= λipt (2.6)

βt+1

cit+1= λipt+1 (2.7)

and hence

pt+1cit+1 = βptc

it for all t (2.8)

for i = 1, 2.Equations (2.8), together with the budget constraint can be solved for the

optimal sequence of consumption of household i as a function of the infinitesequence of prices (and of the endowments, of course)

cit = cit (pt∞t=0)

In order to solve for the equilibrium prices pt∞t=0 one then uses the goodsmarket clearing conditions (2.5)

c1t (pt∞t=0) + c2t (pt∞t=0) = e1t + e2t for all tThis is a system of infinite equations (for each t one) in an infinite numberof unknowns pt∞t=0 which is in general hard to solve. Below we will discussNegishi’s method that often proves helpful in solving for equilibria by reducingthe number of equations and unknowns to a smaller number.For our particular simple example economy, however, we can solve for the

equilibrium directly. Sum (2.8) across agents to obtain

pt+1¡c1t+1 + c

2t+1

¢= βpt(c

1t + c

2t )

Using the goods market clearing condition we find that

pt+1¡e1t+1 + e

2t+1

¢= βpt(e

1t + e

2t )

and hence

pt+1 = βpt

and therefore equilibrium prices are of the form

pt = βtp0

Without loss of generality we can set p0 = 1, i.e. make consumption at period0 the numeraire.3 Then equilibrium prices have to satisfy

pt = βt

3Note that multiplying all prices by µ > 0 does not change the budget constraints of agents,so that if prices pt∞t=0 and allocations (cit∞t=0)i∈1,2 are an AD equilibrium, so are pricesµpt∞t=0 and allocations (cit∞t=0)i=1,2

Page 16: Macroeconomics Phd Lectures notes

10 CHAPTER 2. A SIMPLE DYNAMIC ECONOMY

so that, since β < 1, the period 0 price for period t consumption is lower than theperiod 0 price for period 0 consumption. This fact just reflects the impatienceof both agents.

Using (2.8) we have that cit+1 = cit = ci0 for all t, i.e. consumption isconstant across time for both agents. This reflects the agent’s desire to smoothconsumption over time, a consequence of the strict concavity of the period utilityfunction. Now observe that the budget constraint of both agents will hold withequality since agents’ period utility function is strictly increasing. The left handside of the budget constraint becomes

∞Xt=0

ptcit = c

i0

∞Xt=0

βt =ci01− β

for i = 1, 2.

The two agents differ only along one dimension: agent 1 is rich first, which,given that prices are declining over time, is an advantage. For agent 1 the righthand side of the budget constraint becomes

∞Xt=0

pte1t = 2

∞Xt=0

β2t =2

1− β2

and for agent 2 it becomes

∞Xt=0

pte2t = 2β

∞Xt=0

β2t =2β

1− β2

The equilibrium allocation is then given by

c1t = c10 = (1− β)2

1− β2=

2

1 + β> 1

c2t = c20 = (1− β)2β

1− β2=

1 + β< 1

which obviously satisfies

c1t + c2t = 2 = e

1t + e

2t for all t

Therefore the mere fact that the first agent is rich first makes her consumemore in every period. Note that there is substantial trade going on; in eacheven period the first agent delivers 2 − 2

1+β =2β1+β to the second agent and in

all odd periods the second agent delivers 2 − 2β1+β to the first agent. Also note

that this trade is mutually beneficial, because without trade both agents receivelifetime utility

u(eit) = −∞

Page 17: Macroeconomics Phd Lectures notes

2.2. AN EXAMPLE ECONOMY 11

whereas with trade they obtain

u(c1) =∞Xt=0

βt ln

µ2

1 + β

¶=ln³

21+β

´1− β

> 0

u(c2) =∞Xt=0

βt ln

µ2β

1 + β

¶=ln³2β1+β

´1− β

< 0

In the next section we will show that not only are both agents better off inthe competitive equilibrium than by just eating their endowment, but that, ina sense to be made precise, the equilibrium consumption allocation is sociallyoptimal.

2.2.3 Pareto Optimality and the First Welfare Theorem

In this section we will demonstrate that for this economy a competitive equi-librium is socially optimal. To do this we first have to define what sociallyoptimal means. Our notion of optimality will be Pareto efficiency (also some-times referred to as Pareto optimality). Loosely speaking, an allocation is Paretoefficient if it is feasible and if there is no other feasible allocation that makes nohousehold worse off and at least one household strictly better off. Let us nowmake this precise.

Definition 3 An allocation (c1t , c2t )∞t=0 is feasible if

1.

cit ≥ 0 for all t, for i = 1, 2

2.

c1t + c2t = e

1t + e

2t for all t

Feasibility requires that consumption is nonnegative and satisfies the re-source constraint for all periods t = 0, 1, . . .

Definition 4 An allocation (c1t , c2t )∞t=0 is Pareto efficient if it is feasible andif there is no other feasible allocation (c1t , c2t )∞t=0 such that

u(ci) ≥ u(ci) for both i = 1, 2

u(ci) > u(ci) for at least one i = 1, 2

Note that Pareto efficiency has nothing to do with fairness in any sense: anallocation in which agent 1 consumes everything in every period and agent 2starves is Pareto efficient, since we can only make agent 2 better off by makingagent 1 worse off.

Page 18: Macroeconomics Phd Lectures notes

12 CHAPTER 2. A SIMPLE DYNAMIC ECONOMY

We now prove that every competitive equilibrium allocation for the economydescribed above is Pareto efficient. Note that we have solved for one equilibriumabove; this does not rule out that there is more than one equilibrium. One can,in fact, show that for this economy the competitive equilibrium is unique, butwe will not pursue this here.

Proposition 5 Let (cit∞t=0)i=1,2 be a competitive equilibrium allocation. Then(cit∞t=0)i=1,2 is Pareto efficient.Proof. The proof will be by contradiction; we will assume that (cit∞t=0)i=1,2

is not Pareto efficient and derive a contradiction to this assumption.So suppose that (cit∞t=0)i=1,2 is not Pareto efficient. Then by the definition

of Pareto efficiency there exists another feasible allocation (cit∞t=0)i=1,2 suchthat

u(ci) ≥ u(ci) for both i = 1, 2

u(ci) > u(ci) for at least one i = 1, 2

Without loss of generality assume that the strict inequality holds for i = 1.Step 1: Show that

∞Xt=0

ptc1t >

∞Xt=0

ptc1t

where pt∞t=0 are the equilibrium prices associated with (cit∞t=0)i=1,2. If not,i.e. if

∞Xt=0

ptc1t ≤

∞Xt=0

ptc1t

then for agent 1 the ˜-allocation is better (remember u(c1) > u(c1) is assumed)and not more expensive, which cannot be the case since c1t∞t=0 is part ofa competitive equilibrium, i.e. maximizes agent 1’s utility given equilibriumprices. Hence

∞Xt=0

ptc1t >

∞Xt=0

ptc1t (2.9)

Step 2: Show that

∞Xt=0

ptc2t ≥

∞Xt=0

ptc2t

If not, then

∞Xt=0

ptc2t <

∞Xt=0

ptc2t

Page 19: Macroeconomics Phd Lectures notes

2.2. AN EXAMPLE ECONOMY 13

But then there exists a δ > 0 such that

∞Xt=0

ptc2t + δ ≤

∞Xt=0

ptc2t

Remember that we normalized p0 = 1. Now define a new allocation for agent 2,by

c2t = c2t for all t ≥ 1c20 = c20 + δ for t = 0

Obviously

∞Xt=0

ptc2t =

∞Xt=0

ptc2t + δ ≤

∞Xt=0

ptc2t

and

u(c2) > u(c2) ≥ u(c2)

which can’t be the case since c2t∞t=0 is part of a competitive equilibrium, i.e.maximizes agent 2’s utility given equilibrium prices. Hence

∞Xt=0

ptc2t ≥

∞Xt=0

ptc2t (2.10)

Step 3: Now sum equations (2.9) and (2.10) to obtain

∞Xt=0

pt(c1t + c

2t ) >

∞Xt=0

pt(c1t + c

2t )

But since both allocations are feasible (the allocation (cit∞t=0)i=1,2 because it isan equilibrium allocation, the allocation (cit∞t=0)i=1,2 by assumption) we havethat

c1t + c2t = e

1t + e

2t = c

1t + c

2t for all t

and thus

∞Xt=0

pt(e1t + e

2t ) >

∞Xt=0

pt(e1t + e

2t ),

our desired contradiction.

Page 20: Macroeconomics Phd Lectures notes

14 CHAPTER 2. A SIMPLE DYNAMIC ECONOMY

2.2.4 Negishi’s (1960) Method to Compute Equilibria

In the example economy considered in this section it was straightforward tocompute the competitive equilibrium by hand. This is usually not the case fordynamic general equilibrium models. Now we describe a method to computeequilibria for economies in which the welfare theorem(s) hold. The main idea isto compute Pareto-optimal allocations by solving an appropriate social plannersproblem. This social planner problem is a simple optimization problem whichdoes not involve any prices (still infinite-dimensional, though) and hence mucheasier to tackle in general than a full-blown equilibrium analysis which consistsof several optimization problems (one for each consumer) plus market clearingand involves allocations and prices. If the first welfare theorem holds then weknow that competitive equilibrium allocations are Pareto optimal; by solvingfor all Pareto optimal allocations we have then solved for all potential equilib-rium allocations. Negishi’s method provides an algorithm to compute all Paretooptimal allocations and to isolate those who are in fact competitive equilibriumallocations.We will repeatedly apply this trick in this course: solve a simple social

planners problem and use the welfare theorems to argue that we have solvedfor the allocations of competitive equilibria. Then find equilibrium prices thatsupport these allocations. The news is even better: usually we can read offthe prices as Lagrange multipliers from the appropriate constraints of the socialplanners problem. In later parts of the course we will discuss economies in whichthe welfare theorems do not hold. We will see that these economies are muchharder to analyze exactly because there is no simple optimization problem thatcompletely characterizes the (set of) equilibria of these economies.Consider the following social planners problem

max(c1t ,c2t )∞t=0

αu(c1) + (1− α)u(c2) (2.11)

= max(c1t ,c2t )∞t=0

∞Xt=0

βt£α ln(c1t ) + (1− α) ln(c2t )

¤s.t.

cit ≥ 0 for all i, all t

c1t + c2t = e1t + e

2t ≡ 2 for all t

for a Pareto weight α ∈ [0, 1]. The social planner maximizes the weighted sum ofutilities of the two agents, subject to the allocation being feasible. The weight αindicates how important agent 1’s utility is to the planner, relative to agent 2’sutility. Note that the solution to this problem depends on the Pareto weights,i.e. the optimal consumption choices are functions of α

(c1t , c2t )∞t=0 = (c1t (α), c2t (α))∞t=0We have the following

Proposition 6 An allocation (c1t , c2t )∞t=0 is Pareto efficient if and only if itsolves the social planners problem (2.11) for some α ∈ [0, 1]

Page 21: Macroeconomics Phd Lectures notes

2.2. AN EXAMPLE ECONOMY 15

Proof. Omitted (but a good exercise)This proposition states that we can characterize the set of all Pareto effi-

cient allocations by varying α between 0 and 1 and solving the social plannersproblem for all α’s. As we will demonstrate, by choosing a particular α, the asso-ciated efficient allocation for that α turns out to be the competitive equilibriumallocation.Now let us solve the planners problem for arbitrary α ∈ (0, 1).4 Attach La-

grange multipliers µt2 to the resource constraints (and ignore the non-negativityconstraints on cit since they never bind, due to the period utility function satisfy-ing the Inada conditions). The reason why we divide by 2 will become apparentin a moment.The first order necessary conditions are

αβt

c1t=

µt2

(1− α)βt

c2t=

µt2

Combining yields

c1tc2t

1− α(2.12)

c1t =α

1− αc2t (2.13)

i.e. the ratio of consumption between the two agents equals the ratio of thePareto weights in every period t. A higher Pareto weight for agent 1 resultsin this agent receiving more consumption in every period, relative to agent 2.Using the resource constraint in conjunction with (2.13) yields

c1t + c2t = 2

α

1− αc2t + c

2t = 2

c2t = 2(1− α) = c2t (α)

c1t = 2α = c1t (α)

i.e. the social planner divides the total resources in every period according to thePareto weights. Note that the division is the same in every period, independentof the agents endowments in that particular period. The Lagrange multipliersare given by

µt =2αβt

c1t= βt

(if we wouldn’t have done the initial division by 2 we would have to carry the12 around from now on; the results below wouldn’t change at all).

4Note that for α = 0 and α = 1 the solution to the problem is trivial. For α = 0 we havec1t = 0 and c

2t = 2 and for α = 1 we have the reverse.

Page 22: Macroeconomics Phd Lectures notes

16 CHAPTER 2. A SIMPLE DYNAMIC ECONOMY

Hence for this economy the set of Pareto efficient allocations is given by

PO = (c1t , c2t )∞t=0 : c1t = 2α and c2t = 2(1− α) for some α ∈ [0, 1]

How does this help us in finding the competitive equilibrium for this economy?Compare the first order condition of the social planners problem for agent 1

αβt

c1t=µt2

or

βt

c1t=µt2α

with the first order condition from the competitive equilibrium above (see equation (2.6)):

βt

c1t= λ1pt

By picking λ1 =12α and pt = βt these first order conditions are identical. Sim-

ilarly, pick λ2 =1

2(1−α) and one sees that the same is true for agent 2. So forappropriate choices of the individual Lagrange multipliers λi and prices pt theoptimality conditions for the social planners’ problem and for the householdmaximization problems coincide. Resource feasibility is required in the com-petitive equilibrium as well as in the planners problem. Given that we founda unique equilibrium above but a lot of Pareto efficient allocations (for each αone), there must be an additional requirement that a competitive equilibriumimposes which the planners problem does not require.In a competitive equilibrium households’ choices are constrained by the bud-

get constraint; the planner is only concerned with resource balance. The laststep to single out competitive equilibrium allocations from the set of Paretoefficient allocations is to ask which Pareto efficient allocations would be afford-able for all households if these holds were to face as market prices the Lagrangemultipliers from the planners problem (that the Lagrange multipliers are the ap-propriate prices is harder to establish, so let’s proceed on faith for now). Definethe transfer functions ti(α), i = 1, 2 by

ti(α) =Xt

µt£cit(α)− eit

¤The number ti(α) is the amount of the numeraire good (we pick the period 0consumption good) that agent i would need as transfer in order to be able toafford the Pareto efficient allocation indexed by α. One can show that the ti asfunctions of α are homogeneous of degree one5 and sum to 0 (see HW 1).

5In the sense that if one gives weight xα to agent 1 and x(1 − α) to agent 2, then thecorresponding required transfers are xt1 and xt2.

Page 23: Macroeconomics Phd Lectures notes

2.2. AN EXAMPLE ECONOMY 17

Computing ti(α) for the current economy yields

t1(α) =Xt

µt£c1t (α)− e1t

¤=

Xt

βt£2α− e1t

¤=

1− β− 2

1− β2

t2(α) =2(1− α)

1− β− 2β

1− β2

To find the competitive equilibrium allocation we now need to find the Paretoweight α such that t1(α) = t2(α) = 0, i.e. the Pareto optimal allocation thatboth agents can afford with zero transfers. This yields

0 =2α

1− β− 2

1− β2

α =1

1 + β∈ (0, 0.5)

and the corresponding allocations are

c1t

µ1

1 + β

¶=

2

1 + β

c2t

µ1

1 + β

¶=

1 + β

Hence we have solved for the equilibrium allocations; equilibrium prices aregiven by the Lagrange multipliers µt = βt (note that without the normalizationby 1

2 at the beginning we would have found the same allocations and equilibrium

prices pt =βt

2 which, given that equilibrium prices are homogeneous of degree0, is perfectly fine, too).To summarize, to compute competitive equilibria using Negishi’s method

one does the following

1. Solve the social planners problem for Pareto efficient allocations indexedby Pareto weight α

2. Compute transfers, indexed by α, necessary to make the efficient allocationaffordable. As prices use Lagrange multipliers on the resource constraintsin the planners’ problem.

3. Find the Pareto weight(s) α that makes the transfer functions 0.

4. The Pareto efficient allocations corresponding to α are equilibrium allo-cations; the supporting equilibrium prices are (multiples of) the Lagrangemultipliers from the planning problem

Page 24: Macroeconomics Phd Lectures notes

18 CHAPTER 2. A SIMPLE DYNAMIC ECONOMY

Remember from above that to solve for the equilibrium directly in generalinvolves solving an infinite number of equations in an infinite number of un-knowns. The Negishi method reduces the computation of equilibrium to a finitenumber of equations in a finite number of unknowns in step 3 above. For aneconomy with two agents, it is just one equation in one unknown, for an economywith N agents it is a system of N−1 equations in N−1 unknowns. This is whythe Negishi method (and methods relying on solving appropriate social plan-ners problems in general) often significantly simplifies solving for competitiveequilibria.

2.2.5 Sequential Markets Equilibrium

The market structure of Arrow-Debreu equilibrium in which all agents meet onlyonce, at the beginning of time, to trade claims to future consumption may seemempirically implausible. In this section we show that the same allocations asin an Arrow-Debreu equilibrium would arise if we let agents trade consumptionand one-period bonds in each period. We will call a market structure in whichmarkets for consumption and assets open in each period Sequential Markets andthe corresponding equilibrium Sequential Markets (SM) equilibrium.6

Let rt+1 denote the interest rate on one period bonds from period t to periodt+1. A one period bond is a promise (contract) to pay 1 unit of the consumptiongood in period t + 1 in exchange for 1

1+rt+1units of the consumption good in

period t. We can interpret qt ≡ 11+rt+1

as the relative price of one unit of the

consumption good in period t + 1 in terms of the period t consumption good.Let ait+1 denote the amount of such bonds purchased by agent i in period t andcarried over to period t+1. If ait+1 < 0 we can interpret this as the agent takingout a one-period loan at interest rate rt+1. Household i’s budget constraint inperiod t reads as

cit +ait+1

(1 + rt+1)≤ eit + ait (2.14)

or

cit + qtait+1 ≤ eit + ait

Agents start out their life with initial bond holdings ai0 (remember that period0 bonds are claims to period 0 consumption). Mostly we will focus on thesituation in which ai0 = 0 for all i, but sometimes we want to start an agent offwith initial wealth (ai0 > 0) or initial debt (a

i0 < 0). We then have the following

definition

Definition 7 A Sequential Markets equilibrium is allocations ¡cit, ait+1¢i=1,2∞t=1,interest rates rt+1∞t=0 such that

6In the simple model we consider in this section the restriction of assets traded to one-period riskless bonds is without loss of generality. In more complicated economies (withuncertainty, say) it would not be. We will come back to this issue in later chapters.

Page 25: Macroeconomics Phd Lectures notes

2.2. AN EXAMPLE ECONOMY 19

1. For i = 1, 2, given interest rates rt+1∞t=0 cit, ait+1∞t=0 solves

maxcit,ait+1∞t=0

∞Xt=0

βt ln(cit) (2.15)

s.t.

cit +ait+1

(1 + rt+1)≤ eit + a

it (2.16)

cit ≥ 0 for all t (2.17)

ait+1 ≥ −Ai (2.18)

2. For all t ≥ 02Xi=1

cit =2Xi=1

eit

2Xi=1

ait+1 = 0

The constraint (2.18) on borrowing is necessary to guarantee existence ofequilibrium. Suppose that agents would not face any constraint as to howmuch they can borrow, i.e. suppose the constraint (2.18) were absent. Supposethere would exist a SM-equilibrium ¡cit, ait+1¢i=1,2∞t=1, rt+1∞t=0.Without con-straint on borrowing agent i could always do better by setting

ci0 = ci0 +ε

1 + r1

ai1 = ai1 − ε

ai2 = ai2 − (1 + r2)ε

ait+1 = ait+1 −tYt=1

(1 + rt+1)ε

i.e. by borrowing ε > 0 more in period 0, consuming it and then rolling over theadditional debt forever, by borrowing more and more. Such a scheme is oftencalled a Ponzi scheme. Hence without a limit on borrowing no SM equilibriumcan exist because agents would run Ponzi schemes.In this section we are interested in specifying a borrowing limit that prevents

Ponzi schemes, yet is high enough so that households are never constrainedin the amount they can borrow (by this we mean that a household, knowingthat it can not run a Ponzi scheme, would always find it optimal to chooseait+1 > −Ai). In later chapters we will analyze economies in which agents faceborrowing constraints that are binding in certain situations. Not only are SMequilibria for these economies quite different from the ones to be studied here,but also the equivalence between SM equilibria and AD equilibria will breakdown.

Page 26: Macroeconomics Phd Lectures notes

20 CHAPTER 2. A SIMPLE DYNAMIC ECONOMY

We are now ready to state the equivalence theorem relating AD equilibriaand SM equilibria. Assume that ai0 = 0 for all i = 1, 2.

Proposition 8 Let allocations ¡cit¢i=1,2∞t=0 and prices pt∞t=0 form an Arrow-Debreu equilibrium. Then there exist

¡Ai¢i=1,2

and a corresponding sequen-

tial markets equilibrium with allocations ¡cit, ait+1¢i=1,2∞t=0 and interest ratesrt+1∞t=0 such that

cit = cit for all i, all t

Reversely, let allocations ¡cit, ait+1¢i=1,2∞t=0 and interest rates rt+1∞t=0 forma sequential markets equilibrium. Suppose that it satisfies

ait+1 > −Ai for all i, all trt+1 > 0 for all t

Then there exists a corresponding Arrow-Debreu equilibrium ¡cit¢i=1,2∞t=0, pt∞t=0such that

cit = cit for all i, all t

Proof. Step 1: The key to the proof is to show the equivalence of the budgetsets for the Arrow-Debreu and the sequential markets structure. Normalizep0 = 1 and relate equilibrium prices and interest rates by

1 + rt+1 =ptpt+1

(2.19)

Now look at the sequence of sequential markets budget constraints and assumethat they hold with equality (which they do in equilibrium, due to the nonsa-tiation assumption)

ci0 +ai1

1 + r1= ei0 (2.20)

ci1 +ai2

1 + r2= ei1 + a

i1 (2.21)

...

cit +ait+1

1 + rt+1= eit + a

it (2.22)

Substituting for ai1 from (2.21) in (2.20) one gets

ci0 +ci1

1 + r1+

ai2(1 + r1) (1 + r2)

= ei0 +ei1

(1 + r1)

Page 27: Macroeconomics Phd Lectures notes

2.2. AN EXAMPLE ECONOMY 21

and, repeating this exercise, one gets7

TXt=0

citQtj=1(1 + rj)

+aiT+1QT+1

j=1 (1 + rj)=

TXt=0

eitQtj=1(1 + rj)

Now note that (using the normalization p0 = 1)

tYj=1

(1 + rj) =p0p1∗ p1p2· · · ∗ pt−1

pt=1

pt(2.23)

Taking limits with respect to t on both sides gives, using (2.23)

∞Xt=0

ptcit + lim

T→∞aiT+1QT+1

j=1 (1 + rj)=∞Xt=0

pteit

Given our assumptions on the equilibrium interest rates we have

limT→∞

aiT+1QT+1j=1 (1 + rj)

≥ limT→∞

−AiQT+1j=1 (1 + rj)

= 0

and hence

∞Xt=0

ptcit ≤

∞Xt=0

pteit

Step 2: Now suppose we have an AD-equilibrium ¡cit¢i=1,2∞t=0, pt∞t=0.We want to show that there exist a SM equilibrium with same consumptionallocation, i.e.

cit = cit for all i, all t

Obviously ¡cit¢i=1,2∞t=0 satisfies market clearing. Defining as asset holdingsait+1 =

∞Xτ=1

pt+τ¡cit+τ − eit+τ

¢pt+1

we see that the allocation satisfies the SM budget constraints (remember 1 +rt+1 =

ptpt+1

) Also note that

ait+1 > −∞Xτ=1

pt+τeit+τ

pt+1≥ −

∞Xt=0

pteit > −∞

7We define

0Yj=1

(1 + rj) = 1

Page 28: Macroeconomics Phd Lectures notes

22 CHAPTER 2. A SIMPLE DYNAMIC ECONOMY

so that we can take

Ai =∞Xt=0

pteit

This borrowing constraint, equalling the value of the endowment of agent i atAD-equilibrium prices is also called the natural debt limit. This borrowing limitis so high that agent i, knowing that she can’t run a Ponzi scheme, will neverreach it.It remains to argue that ¡cit¢i=1,2∞t=0 maximizes utility, subject to the

sequential markets budget constraints and the borrowing constraints. Takeany other allocation satisfying these constraints. In step 1. we showed thatthis allocation satisfies the AD budget constraint. If it would be better thancit = cit∞t=0 it would have been chosen as part of an AD-equilibrium, which itwasn’t. Hence cit∞t=0 is optimal within the set of allocations satisfying the SMbudget constraints at interest rates 1 + rt+1 =

ptpt+1

.

Step 3: Now suppose ¡cit, ait+1¢i∈I∞t=1 and rt+1∞t=0 form a sequentialmarkets equilibrium satisfying

ait+1 > −Ai for all i, all trt+1 > 0 for all t

We want to show that there exists a corresponding Arrow-Debreu equilibrium¡cit¢i∈I∞t=0, pt∞t=0 with

cit = cit for all i, all t

Again obviously ¡cit¢i∈I∞t=0 satisfies market clearing and, as shown in step1, the AD budget constraint. It remains to be shown that it maximizes utilitywithin the set of allocations satisfying the AD budget constraint. For p0 = 1and pt+1 =

pt1+rt+1

the set of allocations satisfying the AD budget constraint

coincides with the set of allocations satisfying the SM-budget constraint (forappropriate choices of asset holdings). Since in the SM equilibrium we have theadditional borrowing constraints, the set over which we maximize in the AD caseis larger, since the borrowing constraints are absent in the AD formulation. Butby assumption these additional constraints are never binding (ait+1 > −Ai).Then from a basic theorem of constrained optimization we know that if theadditional constraints are never binding, then the maximizer of the constrainedproblem is also the maximizer of the unconstrained problem, and hence cit∞t=0is optimal for household i within the set of allocations satisfying her AD budgetconstraint.This proposition shows that the sequential markets and the Arrow-Debreu

market structures lead to identical equilibria, provided that we choose the noPonzi conditions appropriately (equal to the natural debt limits, for example)and that the equilibrium interest rates are sufficiently high.8 Usually the anal-

8This assumption can be sufficiently weakened if one introduces borrowing constraints ofslightly different form in the SM equilibrium to prevent Ponzi schemes. We may come backto this later.

Page 29: Macroeconomics Phd Lectures notes

2.3. APPENDIX: SOME FACTS ABOUT UTILITY FUNCTIONS 23

ysis of our economies is easier to carry out using AD language, but the SMformulation has more empirical appeal. The preceding theorem shows that wecan have the best of both worlds.For our example economy we find that the equilibrium interest rates in the

SM formulation are given by

1 + rt+1 =ptpt+1

=1

β

or

rt+1 = r =1

β− 1 = ρ

i.e. the interest rate is constant and equal to the subjective time discount rateρ = 1

β − 1.

2.3 Appendix: Some Facts about Utility Func-tions

The utility function

u(ci) =∞Xt=0

βt ln(cit) (2.24)

described in the main text satisfies the following assumptions that we will oftenrequire in our models:

1. Time separability: total utility from a consumption allocation ci equalsthe discounted sum of period (or instantaneous) utility U(cit) = ln(c

it). In

particular, the period utility at time t only depends on consumption inperiod t and not on consumption in other periods. This formulation rulesout, among other things, habit persistence.

2. Time discounting: the fact that β < 1 indicates that agents are impatient.The same amount of consumption yields less utility if it comes at a latertime in an agents’ life. The parameter β is often referred to as (subjective)time discount factor. The subjective time discount rate ρ is defined byβ = 1

1+ρ and is often, as we will see, intimately related to the equilibrium

interest rate in the economy (because the interest rate is nothing else butthe market time discount rate).

3. Homotheticity: Define the marginal rate of substitution between consump-tion at any two dates t and t+ s as

MRS(ct+s, ct) =

∂u(c)∂ct+s

∂u(c)∂ct

Page 30: Macroeconomics Phd Lectures notes

24 CHAPTER 2. A SIMPLE DYNAMIC ECONOMY

The function u is said to be homothetic ifMRS(ct+s, ct) =MRS(λct+s,λct)for all λ > 0 and c. It is easy to verify that for u defined above we have

MRS(ct+s, ct) =

βt+s

ct+s

βt

ct

=

λβt+s

ct+s

λβt

ct

=MRS(λct+s,λct)

and hence u is homothetic. This, in particular, implies that if an agent’slifetime income doubles, optimal consumption choices will double in eachperiod (income expansion paths are linear).9 It also means that consump-tion allocations are independent of the units of measurement employed.

4. The instantaneous utility function or felicity function U(c) = ln(c) is con-tinuous, twice continuously differentiable, strictly increasing (i.e. U 0(c) >0) and strictly concave (i.e. U 00(c) < 0) and satisfies the Inada conditions

limc&0

U 0(c) = +∞lim

c%+∞U 0(c) = 0

These assumptions imply that more consumption is always better, but anadditional unit of consumption yields less and less additional utility. TheInada conditions indicate that the first unit of consumption yields a lot ofadditional utility but that as consumption goes to infinity, an additionalunit is (almost) worthless. The Inada conditions will guarantee that anagent always chooses ct ∈ (0,∞) for all t

5. The felicity function U is a member of the class of Constant Relative RiskAversion (CRRA) utility functions. These functions have the following

important properties. First, define as σ(c) = −U 00(c)cU 0(c) the (Arrow-Pratt)

coefficient of relative risk aversion. Hence σ(c) indicates a household’sattitude towards risk, with higher σ(c) representing higher risk aversion.For CRRA utility functions σ(c) is constant for all levels of consumption,and for U(c) = ln(c) it is not only constant, but equal to σ(c) = σ = 1.Second, the intertemporal elasticity of substitution ist(ct+1, ct) measuresby how many percent the relative demand for consumption in period t+1,relative to demand for consumption in period t, ct+1ct declines as the relative

price of consumption in t+1 to consumption in t, qt =1

1+rt+1changes by

one percent. Formally

ist(ct+1, ct) = −

·d(

ct+1ct)

ct+1ct

¸·d 11+rt+1

11+rt+1

¸ = −·d(

ct+1ct)

d 11+rt+1

¸·

ct+1ct1

1+rt+1

¸9In the absense of borrowing constraints and other frictions which we will discuss later.

Page 31: Macroeconomics Phd Lectures notes

2.3. APPENDIX: SOME FACTS ABOUT UTILITY FUNCTIONS 25

But combining (2.6) and (2.7) we see that

βU 0(ct+1)U 0(ct)

=pt+1pt

=1

1 + rt+1

which, for U(c) = ln(c) becomes

ct+1ct

=1

β

µ1

1 + rt+1

¶−1

and thus

d³ct+1ct

´d 11+rt+1

= − 1β

µ1

1 + rt+1

¶−2

and therefore

ist(ct+1, ct) = −

·d(

ct+1ct)

d 11+rt+1

¸·

ct+1ct1

1+rt+1

¸ = −− 1

β

³1

1+rt+1

´−21β

³1

1+rt+1

´−2 = 1

Therefore logarithmic period utility is sometimes also called isoelastic util-ity.10 Hence for logarithmic period utility the intertemporal elasticity sub-stitution is equal to (the inverse of) the coefficient of relative risk aversion.

10In general CRRA utility functions are of the form

U(c) =c1−σ − 11− σ

and one can easily compute that the coefficient of relative risk aversion for this utility functionis σ and the intertemporal elasticity of substitution equals σ−1.In a homework you will show that

ln(c) = limσ→1

c1−σ − 11− σ

i.e. that logarithmic utility is a special case of this general class of utility functions.

Page 32: Macroeconomics Phd Lectures notes

26 CHAPTER 2. A SIMPLE DYNAMIC ECONOMY

Page 33: Macroeconomics Phd Lectures notes

Chapter 3

The Neoclassical GrowthModel in Discrete Time

3.1 Setup of the Model

The neoclassical growth model is arguably the single most important workhorsein modern macroeconomics. It is widely used in growth theory, business cycletheory and quantitative applications in public finance.Time is discrete and indexed by t = 0, 1, 2, . . . In each period there are three

goods that are traded, labor services nt, capital services kt and a final outputgood yt that can be either consumed, ct or invested, it. As usual for a completedescription of the economy we have to specify technology, preferences, endow-ments and the information structure. Later, when looking at an equilibrium ofthis economy we have to specify the equilibrium concept that we intend to use.

1. Technology: The final output good is produced using as inputs labor andcapital services, according to the aggregate production function F

yt = F (kt, nt)

Note that I do not allow free disposal. If I want to allow free disposal, Iwill specify this explicitly by defining an separate free disposal technology.Output can be consumed or invested

yt = it + ct

Investment augments the capital stock which depreciates at a constantrate δ over time

kt+1 = (1− δ)kt + it

We can rewrite this equation as

it = kt+1 − kt + δkt

27

Page 34: Macroeconomics Phd Lectures notes

28CHAPTER 3. THE NEOCLASSICALGROWTHMODEL IN DISCRETE TIME

i.e. gross investment it equals net investment kt+1 − kt plus depreciationδkt.We will require that kt+1 ≥ 0, but not that it ≥ 0. This assumes that,since the existing capital stock can be disinvested to be eaten, capital isputty-putty. Note that I have been a bit sloppy: strictly speaking thecapital stock and capital services generated from this stock are differentthings. We will assume (once we define the ownership structure of thiseconomy in order to define an equilibrium) that households own the capitalstock and make the investment decision. They will rent out capital to thefirms. We denote both the capital stock and the flow of capital services bykt. Implicitly this assumes that there is some technology that transformsone unit of the capital stock at period t into one unit of capital servicesat period t. We will ignore this subtlety for the moment.

2. Preferences: There is a large number of identical, infinitely lived house-holds. Since all households are identical and we will restrict ourselvesto type-identical allocations1 we can, without loss of generality assumethat there is a single representative household. Preferences of each house-hold are assumed to be representable by a time-separable utility function(Debreu’s theorem discusses under which conditions preferences admit acontinuous utility function representation)

u (ct∞t=0) =∞Xt=0

βtU(ct)

3. Endowments: Each household has two types of endowments. At period 0each household is born with endowments k0 of initial capital. Furthermoreeach household is endowed with one unit of productive time in each period,to be devoted either to leisure or to work.

4. Information: There is no uncertainty in this economy and we assume thathouseholds and firms have perfect foresight.

5. Equilibrium: We postpone the discussion of the equilibrium concept to alater point as we will first be concerned with an optimal growth problem,where we solve for Pareto optimal allocations.

3.2 Optimal Growth: Pareto Optimal Alloca-tions

Consider the problem of a social planner that wants to maximize the utility ofthe representative agent, subject to the technological constraints of the economy.Note that, as long as we restrict our attention to type-identical allocations, an

1Identical households receive the same allocation by assumption. In the next quarter I)or somebody else) may come back to the issue under which conditions this is an innocuousassumption,

Page 35: Macroeconomics Phd Lectures notes

3.2. OPTIMAL GROWTH: PARETO OPTIMAL ALLOCATIONS 29

allocation that maximizes the utility of the representative agent, subject to thetechnology constraint is a Pareto efficient allocation and every Pareto efficientallocation solves the social planner problem below. Just as a reference we havethe following definitions

Definition 9 An allocation ct, kt, nt∞t=0 is feasible if for all t ≥ 0

F (kt, nt) = ct + kt+1 − (1− δ)kt

ct ≥ 0, kt ≥ 0, 0 ≤ nt ≤ 1k0 ≤ k0

Definition 10 An allocation ct, kt, nt∞t=0 is Pareto efficient if it is feasibleand there is no other feasible allocation ct, kt, nt∞t=0 such that

∞Xt=0

βtU(ct) >∞Xt=0

βtU(ct)

3.2.1 Social Planner Problem in Sequential Formulation

The problem of the planner is

w(k0) = maxct,kt,nt∞t=0

∞Xt=0

βtU(ct)

s.t. F (kt, nt) = ct + kt+1 − (1− δ)kt

ct ≥ 0, kt ≥ 0, 0 ≤ nt ≤ 1k0 ≤ k0

The function w(k0) has the following interpretation: it gives the total lifetimeutility of the representative household if the social planner chooses ct, kt, nt∞t=0optimally and the initial capital stock in the economy is k0. Under the assump-tions made below the function w is strictly increasing, since a higher initialcapital stock yields higher production in the initial period and hence enablesmore consumption or capital accumulation (or both) in the initial period.We now make the following assumptions on preferences and technology.Assumption 1: U is continuously differentiable, strictly increasing, strictly

concave and bounded. It satisfies the Inada conditions limc&0 U0(c) = ∞ and

limc→∞ U 0(c) = 0. The discount factor β satisfies β ∈ (0, 1)Assumption 2: F is continuously differentiable and homogenous of de-

gree 1, strictly increasing in both arguments and strictly concave. FurthermoreF (0, n) = F (k, 0) = 0 for all k, n > 0. Also F satisfies the Inada conditionslimk&0 Fk(k, 1) =∞ and limk→∞ Fk(k, 1) = 0. Also δ ∈ [0, 1]From these assumptions two immediate consequences for optimal allocations

are that nt = 1 for all t since households do not value leisure in their utilityfunction. Also, since the production function is strictly increasing in capital,k0 = k0. To simplify notation we define f(k) = F (k, 1)+ (1− δ)k, for all k. The

Page 36: Macroeconomics Phd Lectures notes

30CHAPTER 3. THE NEOCLASSICALGROWTHMODEL IN DISCRETE TIME

function f gives the total amount of the final good available for consumptionor investment (again remember that the capital stock can be eaten). Fromassumption 2 the following properties of f follow more or less directly: f iscontinuously differentiable, strictly increasing and strictly concave, f(0) = 0,f 0(k) > 0 for all k, limk&0 f

0(k) =∞ and limk→∞ f 0(k) = 1− δ.

Using the implications of the assumptions, and substituting for ct = f(kt)−kt+1 we can rewrite the social planner’s problem as

w(k0) = maxkt+1∞t=0

∞Xt=0

βtU(f(kt)− kt+1) (3.1)

0 ≤ kt+1 ≤ f(kt)k0 = k0 > 0 given

The only choice that the planner faces is the choice between letting the consumereat today versus investing in the capital stock so that the consumer can eat moretomorrow. Let the optimal sequence of capital stocks be denoted by k∗t+1∞t=0.The two questions that we face when looking at this problem are

1. Why do we want to solve such a hypothetical problem of an even more hy-pothetical social planner. The answer to this questions is that, by solvingthis problem, we will have solved for competitive equilibrium allocationsof our model (of course we first have to define what a competitive equilib-rium is). The theoretical justification underlying this result are the twowelfare theorems, which hold in this model and in many others, too. Wewill give a loose justification of the theorems a bit later, and postponea rigorous treatment of the two welfare theorems in infinite dimensionalspaces until the next quarter.

2. How do we solve this problem?2 The answer is: dynamic programming.The problem above is an infinite-dimensional optimization problem, i.e.we have to find an optimal infinite sequence (k1, k2, . . . ) solving the prob-lem above. The idea of dynamic programing is to find a simpler maximiza-tion problem by exploiting the stationarity of the economic environmentand then to demonstrate that the solution to the simpler maximizationproblem solves the original maximization problem.

To make the second point more concrete, note that we can rewrite the prob-

2Just a caveat: infinite-dimensional maximization problems may not have a solution evenif the u and f are well-behaved. So the function w may not always be well-defined. In ourexamples, with the assumptions that we made, everything is fine, however.

Page 37: Macroeconomics Phd Lectures notes

3.2. OPTIMAL GROWTH: PARETO OPTIMAL ALLOCATIONS 31

lem above as

w(k0) = maxkt+1∞t=0 s.t.

0≤kt+1≤f(kt), k0 given

∞Xt=0

βtU(f(kt)− kt+1)

= maxkt+1∞t=0 s.t.

0≤kt+1≤f(kt), k0 given

(U(f(k0)− k1) + β

∞Xt=1

βt−1U(f(kt)− kt+1))

= maxk1 s.t.

0≤k1≤f(k0), k0 given

U(f(k0)− k1) + β

maxkt+1∞t=1

0≤kt+1≤f(kt), k1 given

∞Xt=1

βt−1U(f(kt)− kt+1)

= maxk1 s.t.

0≤k1≤f(k0), k0 given

U(f(k0)− k1) + β

maxkt+2∞t=0

0≤kt+2≤f(kt+1), k1 given

∞Xt=0

βtU(f(kt+1)− kt+2)

Looking at the maximization problem inside the [ ]-brackets and comparingto the original problem (3.1) we see that the [ ]-problem is that of a socialplanner that, given initial capital stock k1, maximizes lifetime utility of therepresentative agent from period 1 onwards. But agents don’t age in our model,the technology or the utility functions doesn’t change over time; this suggeststhat the optimal value of the problem in [ ]-brackets is equal to w(k1) and hencethe problem can be rewritten as

w(k0) = max0≤k1≤f(k0)k0 given

U(f(k0)− k1) + βw(k1)

Again two questions arise:

2.1 Under which conditions is this suggestive discussion formally correct? Wewill come back to this in a little while.

2.2 Is this progress? Of course, the maximization problem is much easiersince, instead of maximizing over infinite sequences we maximize overjust one number, k1. But we can’t really solve the maximization problem,because the function w(.) appears on the right side, and we don’t knowthis function. The next section shows ways to overcome this problem.

3.2.2 Recursive Formulation of Social Planner Problem

The above formulation of the social planners problem with a function on theleft and right side of the maximization problem is called recursive formulation.Now we want to study this recursive formulation of the planners problem. Sincethe function w(.) is associated with the sequential formulation, let us changenotation and denote by v(.) the corresponding function for the recursive formu-lation of the problem. Remember the interpretation of v(k): it is the discountedlifetime utility of the representative agent from the current period onwards if the

Page 38: Macroeconomics Phd Lectures notes

32CHAPTER 3. THE NEOCLASSICALGROWTHMODEL IN DISCRETE TIME

social planner is given capital stock k at the beginning of the current period andallocates consumption across time optimally for the household. This function v(the so-called value function) solves the following recursion

v(k) = max0≤k0≤f(k)

U(f(k)− k0) + βv(k0) (3.2)

Note again that v and w are two very different functions; v is the valuefunction for the recursive formulation of the planners problem and w is thecorresponding function for the sequential problem. Of course below we want toestablish that v = w, but this is something that we have to prove rather thansomething that we can assume to hold! The capital stock k that the plannerbrings into the current period, result of past decisions, completely determineswhat allocations are feasible from today onwards. Therefore it is called the“state variable”: it completely summarizes the state of the economy today (i.e.all future options that the planner has). The variable k0 is decided (or controlled)today by the social planner; it is therefore called the “control variable”, becauseit can be controlled today by the planner.3

Equation (3.2) is a functional equation (the so-called Bellman equation): itssolution is a function, rather than a number or a vector. Fortunately the math-ematical theory of functional equations is well-developed, so we can draw onsome fairly general results. The functional equation posits that the discountedlifetime utility of the representative agent is given by the utility that this agentreceives today, U(f(k)− k0), plus the discounted lifetime utility from tomorrowonwards, βv(k0). So this formulation makes clear the planners trade-off: con-sumption (and hence utility) today, versus a higher capital stock to work with(and hence higher discounted future utility) from tomorrow onwards. Hence, fora given k this maximization problem is much easier to solve than the problemof picking an infinite sequence of capital stocks kt+1∞t=0 from before. The onlyproblem is that we have to do this maximization for every possible capital stockk, and this posits theoretical as well as computational problems. However, it willturn out that the functional equation is much easier to solve than the sequentialproblem (3.1) (apart from some very special cases). By solving the functionalequation we mean finding a value function v solving (3.2) and an optimal policyfunction k0 = g(k) that describes the optimal k0 for the maximization part in(3.2), as a function of k, i.e. for each possible value that k can take. Again weface several questions associated with equation (3.2):

1. Under what condition does a solution to the functional equation (3.2) existand, if it exist, is unique?

2. Is there a reliable algorithm that computes the solution (by reliable wemean that it always converges to the correct solution, independent of theinitial guess for v

3These terms come from control theory, a field in applied mathematics. Control theory isused in many technical applications such as astronautics.

Page 39: Macroeconomics Phd Lectures notes

3.2. OPTIMAL GROWTH: PARETO OPTIMAL ALLOCATIONS 33

3. Under what conditions can we solve (3.2) and be sure to have solved (3.1),i.e. under what conditions do we have v = w and equivalence between theoptimal sequential allocation kt+1∞t=0 and allocations generated by theoptimal recursive policy g(k)

4. Can we say something about the qualitative features of v and g?

The answers to these questions will be given in the next two sections: theanswers to 1. and 2. will come from the Contraction Mapping Theorem, tobe discussed in Section 4.3. The answer to the third question makes up whatRichard Bellman called the Principle of Optimality and is discussed in Section5.1. Finally, under more restrictive assumptions we can characterize the solutionto the functional equation (v, g) more precisely. This will be done in Section 5.2.In the remaining parts of this section we will look at specific examples where wecan solve the functional equation by hand. Then we will talk about competitiveequilibria and the way we can construct prices so that Pareto optimal alloca-tions, together with these prices, form a competitive equilibrium. This will beour versions of the first and second welfare theorem for the neoclassical growthmodel.

3.2.3 An Example

Consider the following example. Let the period utility function be given byU(c) = ln(c) and the aggregate production function be given by F (k, n) =kαn1−α and assume full depreciation, i.e. δ = 1. Then f(k) = kα and thefunctional equation becomes

v(k) = max0≤k0≤kα

ln (kα − k0) + βv(k0)

Remember that the solution to this functional equation is an entire functionv(.). Now we will apply several methods to solve this functional equation.

Guess and Verify

We will guess a particular functional form of a solution and then verify that thesolution has in fact this form (note that this does not rule out that the functionalequation has other solutions). This method works well for the example at hand,but not so well for most other examples that we are concerned with. Let usguess

v(k) = A+B ln(k)

where A and B are coefficients that are to be determined. The method consistsof three steps:

1. Solve the maximization problem on the right hand side, given the guessfor v, i.e. solve

max0≤k0≤kα

ln (ka − k0) + β (A+B ln(k0))

Page 40: Macroeconomics Phd Lectures notes

34CHAPTER 3. THE NEOCLASSICALGROWTHMODEL IN DISCRETE TIME

Obviously the constraints on k0 never bind and the objective function isstrictly concave and the constraint set is compact, for any given k. Thefirst order condition is sufficient for the unique solution. The FOC yields

1

kα − k0 =βB

k0

k0 =βBkα

1 + βB

2. Evaluate the right hand side at the optimum k0 = βBkα

1+βB . This yields

RHS = ln (ka − k0) + β (A+B ln(k0))

= ln

µkα

1 + βB

¶+ βA+ βB ln

µβBkα

1 + βB

¶= − ln(1 + βB) + α ln(k) + βA+ βB ln

µβB

1 + βB

¶+ αβB ln (k)

3. In order for our guess to solve the functional equation, the left hand side ofthe functional equation, which we have guessed to equal LHS= A+B ln(k)must equal the right hand side, which we just found. If we can findcoefficients A,B for which this is true, we have found a solution to thefunctional equation. Equating LHS and RHS yields

A+B ln(k) = − ln(1 + βB) + α ln(k) + βA+ βB ln

µβB

1 + βB

¶+ αβB ln (k)

(B − α(1 + βB)) ln(k) = −A− ln(1 + βB) + βA+ βB ln

µβB

1 + βB

¶(3.3)

But this equation has to hold for every capital stock k. The right handside of (3.3) does not depend on k but the left hand side does. Hencethe right hand side is a constant, and the only way to make the left handside a constant is to make B − α(1 + βB) = 0. Solving this for B yieldsB = α

1−αβ . Since the left hand side of (3.3) is 0, the right hand side betteris, too, for B = α

1−αβ . Therefore the constant A has to satisfy

0 = −A− ln(1 + βB) + βA+ βB ln

µβB

1 + βB

¶= −A− ln

µ1

1− αβ

¶+ βA+

αβ

1− αβln(αβ)

Solving this mess for A yields

A =1

1− β

·αβ

1− αβln(αβ) + ln(1− αβ)

¸

Page 41: Macroeconomics Phd Lectures notes

3.2. OPTIMAL GROWTH: PARETO OPTIMAL ALLOCATIONS 35

We can also determine the optimal policy function k0 = g(k) as

g(k) =βBkα

1 + βB

= αβkα

Hence our guess was correct: the function v∗(k) = A+B ln(k), with A,B asdetermined above, solves the functional equation, with associated policy func-tion g(k) = αβkα. Note that for this specific example the optimal policy of thesocial planner is to save a constant fraction αβ of total output kα as capital stockfor tomorrow and and let the household consume a constant fraction (1 − αβ)of total output today. The fact that these fractions do not depend on the levelof k is very unique to this example and not a property of the model in general.Also note that there may be other solutions to the functional equation; we havejust constructed one (actually, for the specific example there are no others, butthis needs some proving). Finally, it is straightforward to construct a sequencekt+1∞t=0 from our policy function g that will turn out to solve the sequentialproblem (3.1) (of course for the specific functional forms used in the example):

start from k0 = k0, k1 = g(k0) = αβkα0 , k2 = g(k1) = αβkα1 = (αβ)1+αkα

2

0 and

in general kt = (αβ)P t−1

j=0 αj

kαt

0 . Obviously, since 0 < α < 1 we have that

limt→∞ kt = (αβ)

11−α

for all initial conditions k0 > 0 (which, not surprisingly, is the unique solutionto g(k) = k).

Value Function Iteration: Analytical Approach

In the last section we started with a clever guess, parameterized it and used themethod of undetermined coefficients (guess and verify) to solve for the solutionv∗ of the functional equation. For just about any other than the log-utility,Cobb-Douglas production function case this method would not work; even yourmost ingenious guesses would fail when trying to be verified.Consider the following iterative procedure for our previous example

1. Guess an arbitrary function v0(k). For concreteness let’s take v0(k) = 0for all

2. Proceed recursively by solving

v1(k) = max0≤k0≤kα

ln (kα − k0) + βv0(k0)

Note that we can solve the maximization problem on the right hand sidesince we know v0 (since we have guessed it). In particular, since v0(k

0) = 0for all k0 we have as optimal solution to this problem

k0 = g1(k) = 0 for all k

Page 42: Macroeconomics Phd Lectures notes

36CHAPTER 3. THE NEOCLASSICALGROWTHMODEL IN DISCRETE TIME

Plugging this back in we get

v1(k) = ln (kα − 0) + βv0(0) = ln k

α = α ln k

3. Now we can solve

v2(k) = max0≤k0≤kα

ln (kα − k0) + βv1(k0)

since we know v1 and so forth.

4. By iterating on the recursion

vn+1(k) = max0≤k0≤kα

ln (kα − k0) + βvn(k0)

we obtain a sequence of value functions vn∞n=0 and policy functionsgn∞n=1. Hopefully these sequences will converge to the solution v∗ andassociated policy g∗ of the functional equation. In fact, below we willstate and prove a very important theorem asserting exactly that (undercertain conditions) this iterative procedure converges for any initial guessand converges to the correct solution, namely v∗.

In the first homework I let you carry out the first few iterations in thisprocedure. Note however, that, in order to find the solution v∗ exactly youwould have to carry out step 2. above a lot of times (in fact, infinitely manytimes), which is, of course, infeasible. Therefore one has to implement thisprocedure numerically on a computer.

Value Function Iteration: Numerical Approach

Even a computer can carry out only a finite number of calculation and canonly store finite-dimensional objects. Hence the best we can hope for is anumerical approximation of the true value function. The functional equa-tion above is defined for all k ≥ 0 (in fact there is an upper bound, butlet’s ignore this for now). Because computer storage space is finite, we willapproximate the value function for a finite number of points only.4 For thesake of the argument suppose that k and k0 can only take values in K =0.04, 0.08, 0.12, 0.16, 0.2. Note that the value functions vn then consists of5 numbers, (vn(0.04), vn(0.08), vn(0.12), vn(0.16), vn(0.2))Now let us implement the above algorithm numerically. First we have to pick

concrete values for the parameters α and β. Let us pick α = 0.3 and β = 0.6.

1. Make an initial guess v0(k) = 0 for all k ∈ K4In this course I will only discuss so-called finite state-space methods, i.e. methods in

which the state variable (and the control variable) can take only a finite number of values.Ken Judd, one of the world leaders in numerical methods in economics teaches an exellentsecond year class in computational methods, in which much more sophisticated methods forsolving similar problems are discussed. I strongly encourage you to take this course at somepoint of your career here in Stanford.

Page 43: Macroeconomics Phd Lectures notes

3.2. OPTIMAL GROWTH: PARETO OPTIMAL ALLOCATIONS 37

2. Solve

v1(k) = max0≤k0≤k0.3k0∈K

©ln¡k0.3 − k0¢+ 0.6 ∗ 0ª

This obviously yields as optimal policy k0(k) = g1(k) = 0.04 for all k ∈ K(note that since k0 ∈ K is required, k0 = 0 is not allowed). Plugging thisback in yields

v1(0.04) = ln(0.040.3 − 0.04) = −1.077v1(0.08) = ln(0.080.3 − 0.04) = −0.847v1(0.12) = ln(0.120.3 − 0.04) = −0.715v1(0.16) = ln(0.160.3 − 0.04) = −0.622v1(0.2) = ln(0.20.3 − 0.04) = −0.55

3. Let’s do one more step by hand

v2(k) =

max0≤k0≤k0.3k0∈K

ln¡k0.3 − k0¢+ 0.6v1(k0)

Start with k = 0.04 :

v2(0.04) = max0≤k0≤0.040.3

k0∈K

©ln¡0.040.3 − k0¢+ 0.6v1(k0)ª

Since 0.040.3 = 0.381 all k0 ∈ K are possible. If the planner choosesk0 = 0.04, then

v2(0.04) = ln¡0.040.3 − 0.04¢+ 0.6 ∗ (−1.077) = −1.723

If he chooses k0 = 0.08, then

v2(0.04) = ln¡0.040.3 − 0.08¢+ 0.6 ∗ (−0.847) = −1.710

If he chooses k0 = 0.12, then

v2(0.04) = ln¡0.040.3 − 0.12¢+ 0.6 ∗ (−0.715) = −1.773

If k0 = 0.16, then

v2(0.04) = ln¡0.040.3 − 0.16¢+ 0.6 ∗ (−0.622) = −1.884

Finally, if k0 = 0.2, then

v2(0.04) = ln¡0.040.3 − 0.2¢+ 0.6 ∗ (−0.55) = −2.041

Page 44: Macroeconomics Phd Lectures notes

38CHAPTER 3. THE NEOCLASSICALGROWTHMODEL IN DISCRETE TIME

Hence for k = 0.04 the optimal choice is k0(0.04) = g2(0.04) = 0.08 andv2(0.04) = −1.710. This we have to do for all k ∈ K. One can already seethat this is quite tedious by hand, but also that a computer can do thisquite rapidly. Table 1 below shows the value of¡

k0.3 − k0¢+ 0.6v1(k0)for different values of k and k0. A ∗ in the column for k0 that this k0 isthe optimal choice for capital tomorrow, for the particular capital stock ktoday

Table 1

k0

k0.04 0.08 0.12 0.16 0.2

0.04 −1.7227 −1.7097∗ −1.7731 −1.8838 −2.04070.08 −1.4929 −1.4530∗ −1.4822 −1.5482 −1.64390.12 −1.3606 −1.3081∗ −1.3219 −1.3689 −1.44050.16 −1.2676 −1.2072∗ −1.2117 −1.2474 −1.30520.2 −1.1959 −1.1298 −1.1279∗ −1.1560 −1.2045

Hence the value function v2 and policy function g2 are given by

Table 2

k v2(k) g2(k)

0.04 −1.7097 0.080.08 −1.4530 0.080.12 −1.3081 0.080.16 −1.2072 0.080.2 −1.1279 0.12

In Figure 3.1 we plot the true value function v∗ (remember that for this ex-ample we know to find v∗ analytically) and selected iterations from the numeri-cal value function iteration procedure. In Figure 3.2 we have the correspondingpolicy functions.We see from Figure 3.1 that the numerical approximations of the value func-

tion converge rapidly to the true value function. After 20 iterations the approx-imation and the truth are nearly indistinguishable with the naked eye. Lookingat the policy functions we see from Figure 2 that the approximating policy func-tion do not converge to the truth (more iterations don’t help). This is due to thefact that the analytically correct value function was found by allowing k0 = g(k)to take any value in the real line, whereas for the approximations we restrictedk0 = gn(k) to lie in K. The function g10 approximates the true policy function asgood as possible, subject to this restriction. Therefore the approximating value

Page 45: Macroeconomics Phd Lectures notes

3.2. OPTIMAL GROWTH: PARETO OPTIMAL ALLOCATIONS 39

0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2-3

-2.5

-2

-1.5

-1

-0.5

0

Capital Stock k Today

Val

ue F

unct

ion

Value Function: True and Approximated

V0

V1

V2

V10 True Value Function

Figure 3.1:

function will not converge exactly to the truth, either. The fact that the valuefunction approximations come much closer is due to the fact that the utility andproduction function induce “curvature” into the value function, something thatwe may make more precise later. Also note that we we plot the true value andpolicy function only on K, with MATLAB interpolating between the points inK, so that the true value and policy functions in the plots look piecewise linear.

3.2.4 The Euler Equation Approach and TransversalityConditions

We now relate our example to the traditional approach of solving optimizationproblems. Note that this approach also, as the guess and verify method, willonly work in very simple examples, but not in general, whereas the numericalapproach works for a wide range of parameterizations of the neoclassical growth

Page 46: Macroeconomics Phd Lectures notes

40CHAPTER 3. THE NEOCLASSICALGROWTHMODEL IN DISCRETE TIME

0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.20.04

0.05

0.06

0.07

0.08

0.09

0.1

0.11

0.12

Capital Stock k Today

Pol

icy

Func

tion

Policy Function: True and Approximated

g1

g2

g10

True Policy Function

Figure 3.2:

model. First let us look at a finite horizon social planners problem and then atthe related infinite-dimensional problem

The Finite Horizon Case

Let us consider the social planner problem for a situation in which the repre-sentative consumer lives for T < ∞ periods, after which she dies for sure andthe economy is over. The social planner problem for this case is given by

wT (k0) = maxkt+1Tt=0

TXt=0

βtU(f(kt)− kt+1)

0 ≤ kt+1 ≤ f(kt)k0 = k0 > 0 given

Obviously, since the world goes under after period T, kT+1 = 0. Also, givenour Inada assumptions on the utility function the constraints on kt+1 will never

Page 47: Macroeconomics Phd Lectures notes

3.2. OPTIMAL GROWTH: PARETO OPTIMAL ALLOCATIONS 41

be binding and we will disregard them henceforth. The first thing we note isthat, since we have a finite-dimensional maximization problem and since the setconstraining the choices of kt+1Tt=0 is closed and bounded, by the Bolzano-Weierstrass theorem a solution to the maximization problem exists, so thatwT (k0) is well-defined. Furthermore, since the constraint set is convex andwe assumed that U is strictly concave (and the finite sum of strictly concavefunctions is strictly concave), the solution to the maximization problem is uniqueand the first order conditions are not only necessary, but also sufficient.Forming the Lagrangian yields

L = U(f(k0)− k1) + . . .+ βtU(f(kt)− kt+1) + βt+1U(f(kt+1)− kt+2) + . . .+ βTU(f(kT )− kT+1)

and hence we can find the first order conditions as

∂L

∂kt+1= −βtU 0(f(kt)− kt+1) + βt+1U 0(f(kt+1)− kt+2)f 0(kt+1) = 0 for all t = 0, . . . , T − 1

or

U 0(f(kt)− kt+1)| z = βU 0(f(kt+1)− kt+2)| z f 0(kt+1)| z for all t = 0, . . . , T − 1

Cost in utilityfor saving1 unit more

capital for t+ 1

=

Discountedadd. utilityfrom one moreunit of cons.

Add. productionpossible withone more unit

of capital in t+ 1

(3.4)

The interpretation of the optimality condition is easiest with a variational argu-ment. Suppose the social planner in period t contemplates whether to save onemore unit of capital for tomorrow. One more unit saved reduces consumptionby one unit, at utility cost of U 0(f(kt)− kt+1). On the other hand, there is onemore unit of capital to produce with tomorrow, yielding additional productionf 0(kt+1). Each additional unit of production, when used for consumption, isworth U 0(f(kt+1)−kt+2) utiles tomorrow, and hence βU 0(f(kt+1)−kt+2) utilestoday. At the optimum the net benefit of such a variation in allocations must bezero, and the result is the first order condition above. This first order conditionsome times is called an Euler equation (supposedly because it is loosely linkedto optimality conditions in continuous time calculus of variations, developed byEuler). Equations (3.4) is second order difference equation, a system of T equa-tions in the T + 1 unknowns kt+1Tt=0 (with k0 predetermined). However, wehave the terminal condition kT+1 = 0 and hence, under appropriate conditions,can solve for the optimal kt+1Tt=0 uniquely. We can demonstrate this for ourexample from above.Again let U(c) = ln(c) and f(k) = kα. Then (3.4) becomes

1

kαt − kt+1=

βαkα−1t+1

kαt+1 − kt+12kαt+1 − kt+2 = αβkα−1t+1 (k

αt − kt+1) (3.5)

Page 48: Macroeconomics Phd Lectures notes

42CHAPTER 3. THE NEOCLASSICALGROWTHMODEL IN DISCRETE TIME

with k0 > 0 given and kT+1 = 0. A little trick will make our life easier. Define

zt =kt+1kαt. The variable zt is the fraction of output in period t that is saved

as capital for tomorrow, so we can interpret zt as the saving rate of the socialplanner. Dividing both sides of (3.5) by kαt+1 we get

1− zt+1 =αβ(kαt − kt+1)

kt+1= αβ

µ1

zt− 1¶

zt+1 = 1 + αβ − αβ

zt

This is a first order difference equation. Since we have the boundary condi-tion kT+1 = 0, this implies zT = 0, so we can solve this equation backwards.Rewriting yields

zt =αβ

1 + αβ − zt+1We can now recursively solve backwards for the entire sequence ztTt=0, giventhat we know zT = 0. We obtain as general formula (verify this by plugging itinto the first order difference equation)

zt = αβ1− (αβ)T−t1− (αβ)T−t+1

and hence

kt+1 = αβ1− (αβ)T−t1− (αβ)T−t+1

kαt

ct =1− αβ

1− (αβ)T−t+1kαt

One can also solve for the discounted future utility at time zero from the optimalallocation to obtain

wT (k0) = α ln(k0)TXj=0

(αβ)j −TXj=1

βT−j ln

ÃjXi=0

(αβ)i!

+αβTXj=1

βT−j"j−1Xi=0

(αβ)i(ln(αβ) + ln

ÃPj−1i=0 (αβ)

iPji=0 (αβ)

i

!)#

Note that the optimal policies and the discounted future utility are functions ofthe time horizon that the social planner faces. Also note that for this specificexample

limT→∞

αβ1− (αβ)T−t1− (αβ)T−t+1 k

αt

= αβkαt

Page 49: Macroeconomics Phd Lectures notes

3.2. OPTIMAL GROWTH: PARETO OPTIMAL ALLOCATIONS 43

and

limt→∞w

T (k0) =1

1− β

·αβ

1− αβln(αβ) + ln(1− αβ)

¸+

α

1− αβln(k0)

So is this the case that the optimal policy for the social planners problem withinfinite time horizon is the limit of the optimal policies for the T−horizon plan-ning problem (and the same is true for the value of the planning problem)?Our results from the guess and verify method seem to indicate this, and for thisexample this is indeed true, but a) this needs some proof and b) it is not atall true in general, but very specific to the example we considered.5. We can’tin general interchange maximization and limit-taking: the limit of the finitemaximization problems is not necessary equal to maximization of the problemin which time goes to infinity.In order to prepare for the discussion of the infinite horizon case let us

analyze the first order difference equation

zt+1 = 1 + αβ − αβ

zt

graphically. On the y-axis of Figure 3.3 we draw zt+1 against zt on the x-axis.Since kt+1 ≥ 0, we have that zt ≥ 0 for all t. Furthermore, as zt approaches 0from above, zt+1 approaches −∞. As zt approaches +∞, zt+1 approaches 1+αβfrom below asymptotically. The graph intersects the x-axis at z0 = αβ

1+αβ . Thedifference equation has two steady states where zt+1 = zt = z. This can be seenby

z = 1 + αβ − αβ

z

z2 − (1 + αβ)z + αβ = 0

(z − 1)(z − αβ) = 0

z = 1 or z = αβ

From Figure 3.3 we can also determine graphically the sequence of optimalpolicies ztTt=0.We start with zT = 0 on the y-axis, go to the zt+1 = 1+αβ− αβ

ztcurve to determine zT−1 and mirror it against the 45-degree line to obtain zT−1on the y-axis. Repeating the argument one obtains the entire ztTt=0 sequence,

5An easy counterexample is the cake-eating problem without discounting

maxctTt=0

∞Xt=0

u(ct)

s.t. 1 =TXt=0

ct

with u bounded and strictly concave, whose solution for the finite time horizon is obviouslyct =

1T+1

, for all t. The limit as T →∞ would be ct = 0, which obviously can’t be optimal.

For example ct = (1− a)at, for any a ∈ (0, 1) beats that policy.

Page 50: Macroeconomics Phd Lectures notes

44CHAPTER 3. THE NEOCLASSICALGROWTHMODEL IN DISCRETE TIME

and hence the entire kt+1Tt=0 sequence. Note that going with t backwards tozero, the zt approach αβ. Hence for large T and for small t (the optimal policiesfor a finite time horizon problem with long horizon, for the early periods) comeclose to the optimal infinite time horizon policies solved for with the guess andverify method.

z = zt+1 t

zt+1

αβ 1 zt

z =1+αβ-αβ/z t+1 t

zT

zT-1

Figure 3.3:

Page 51: Macroeconomics Phd Lectures notes

3.2. OPTIMAL GROWTH: PARETO OPTIMAL ALLOCATIONS 45

The Infinite Horizon Case

Now let us turn to the infinite horizon problem and let’s see how far we can getwith the Euler equation approach. Remember that the problem was to solve

w(k0) = maxkt+1∞t=0

∞Xt=0

βtU(f(kt)− kt+1)

0 ≤ kt+1 ≤ f(kt)k0 = k0 > 0 given

Since the period utility function is strictly concave and the constraint set isconvex, the first order conditions constitute necessary conditions for an optimalsequence k∗t+1∞t=0 (a proof of this is a formalization of the variational argu-ment I spelled out when discussing the intuition for the Euler equation). As areminder, the Euler equations were

βU 0(f(kt+1)− kt+2)f 0(kt+1) = U 0(f(kt)− kt+1) for all t = 0, . . . , t, . . .

Again this is a second order difference equation, but now we only have an initialcondition for k0, but no terminal condition since there is no terminal time period.In a lot of applications, the transversality condition substitutes for the missing

Page 52: Macroeconomics Phd Lectures notes

46CHAPTER 3. THE NEOCLASSICALGROWTHMODEL IN DISCRETE TIME

terminal condition. Let us first state and then interpret the TVC6

limt→∞βtU 0(f(kt)− kt+1)f 0(kt)| z kt|z = 0

value in discountedutility terms of onemore unit of capital

TotalCapitalStock

= 0

The transversality condition states that the value of the capital stock kt, whenmeasured in terms of discounted utility, goes to zero as time goes to infinity.Note that this condition does not require that the capital stock itself convergesto zero in the limit, only the (shadow) value of the capital stock has to convergeto zero.The transversality condition is a tricky beast, and you may spend some more

time on it next quarter. For now we just state the following theorem.

Theorem 11 Let U,β and F (and hence f) satisfy assumptions 1. and 2. Thenan allocation kt+1∞t=0 that satisfies the Euler equations and the transversalitycondition solves the sequential social planners problem, for a given k0.

This theorem states that under certain assumptions the Euler equations andthe transversality condition are jointly sufficient for a solution to the socialplanners problem in sequential formulation. Stokey et al., p. 98-99 prove thistheorem. Note that this theorem does not apply for the case in which theutility function is logarithmic; however, the proof that Stokey et al. give can be

6Often one can find an alternative statement of the TVC.

limt→∞λtkt+1 = 0

where λt is the Lagrange multiplier on the constraint

ct + kt+1 = f(kt)

in the social planner in which consumption is not yet substituted out. From the first ordercondition we have

βtU 0(ct) = λt

βtU 0(f(kt)− kt+1) = λt

Hence the TVC becomes

limt→∞βtU 0(f(kt)− kt+1)kt+1 = 0

This condition is equvalent to the condition given in the main text, as shown by the followingargument (which uses the Euler equation)

0 = limt→∞βtU 0(f(kt)− kt+1)kt+1

= limt→∞βt−1U 0(f(kt−1)− kt)kt

= limt→∞βt−1βU 0(f(kt)− kt+1)f 0(kt)kt

= limt→∞βtU 0(f(kt)− kt+1)f 0(kt)kt

which is the TVC in the main text.

Page 53: Macroeconomics Phd Lectures notes

3.2. OPTIMAL GROWTH: PARETO OPTIMAL ALLOCATIONS 47

extended to the log-case. So although the Euler equations and the TVC maynot be sufficient for every unbounded utility function, for the log-case they are.Also note that we have said nothing about the necessity of the TVC. We

have (loosely) argued that the Euler equations are necessary conditions, but isthe TVC necessary, i.e. does every solution to the sequential planning problemhave to satisfy the TVC? This turns out to be a hard problem, and there isnot a very general result for this. However, for the log-case (with f 0s satisfyingour assumptions), Ekelund and Scheinkman (1985) show that the TVC is infact a necessary condition. Refer to their paper and to the related results byPeleg and Ryder (1972) and Weitzman (1973) for further details. From now onwe assert that the TVC is necessary and sufficient for optimization under theassumptions we made on f, U, but you should remember that these assertionsremain to be proved.For now we take these theoretical results for granted and proceed with our

example of U(c) = ln(c), f(k) = kα. For these particular functional forms, theTVC becomes

limt→∞βtU 0(f(kt)− kt+1)f 0(kt)kt

= limt→∞

αβtkαtkαt − kt+1

= limt→∞

αβt

1− kt+1kαt

= limt→∞

αβt

1− ztWe also repeat the first order difference equation derived from the Euler equa-tions

zt+1 = 1 + αβ − αβ

zt

We can’t solve the Euler equations form zt∞t=0 backwards, but we can solve itforwards, conditional on guessing an initial value for z0.We show that only oneguess for z0 yields a sequence that does not violate the TVC or the nonnegativityconstraint on capital or consumption.

1. z0 < αβ. From Figure 3 we see that in finite time zt < 0, violating thenonnegativity constraint on capital

2. z0 > αβ. Then from Figure 3 we see that limt→∞ zt = 1. (Note that, infact, every z0 > 1 violate the nonnegativity of consumption and hence isnot admissible as a starting value). We will argue that all these pathsviolate the TVC.

3. z0 = αβ. Then zt = αβ for all t > 0. For this path (which obviouslysatisfies the Euler equations) we have that

limt→∞

αβt

1− zt = limt→∞

αβt

1− αβ= 0

Page 54: Macroeconomics Phd Lectures notes

48CHAPTER 3. THE NEOCLASSICALGROWTHMODEL IN DISCRETE TIME

and hence this sequence satisfies the TVC. From the sufficiency of the Eu-ler equation jointly with the TVC we conclude that the sequence zt∞t=0given by zt = αβ is an optimal solution for the sequential social plan-ners problem. Translating into capital sequences yields as optimal policykt+1 = αβkαt , with k0 given. But this is exactly the constant saving ratepolicy that we derived as optimal in the recursive problem.

Now we pick up the unfinished business from point 2. Note that we assertedabove (citing Ekelund and Scheinkman) that for our particular example theTVC is a necessary condition, i.e. any sequence kt+1∞t=0 that does not satisfythe TVC can’t be an optimal solution.

Since all sequences zt∞t=0 in 2. converge to 1, in the TVC both the nomi-nator and the denominator go to zero. Let us linearly approximate zt+1 aroundthe steady state z = 1. This gives

zt+1 = 1 + αβ − αβ

zt= g(zt)

zt+1 ≈ g(1) + (zt − 1)g0(zt)|zt=1= 1 + (zt − 1)

µαβ

z2t

¶|zt=1

= 1 + αβ(zt − 1)(1− zt+1) ≈ αβ(1− zt)

≈ (αβ)t−k+1

(1− zk) for all k

Hence

limt→∞

αβt+1

1− zt+1 ≈ limt→∞

αβt+1

(αβ)t−k+1

(1− zk)

= limt→∞

βk

αt−k(1− zk) =∞

as long as 0 < α < 1. Hence non of the sequences contemplated in 2. can bean optimal solution, and our solution found in 3. is indeed the unique optimalsolution to the infinite-dimensional social planner problem. Therefore in thisspecific case the Euler equation approach, augmented by the TVC works. Butas with the guess-and-verify method this is very unique to specific example athand. Therefore for the general case we can’t rely on pencil and paper, but haveto resort to computational techniques. To make sure that these techniques givethe desired answer, we have to study the general properties of the functionalequation associated with the sequential social planner problem and the relationof its solution to the solution of the sequential problem. We will do this in laterchapters. Before this we will show that, by solving the social planners problemwe have, in effect, solved for a (the) competitive equilibrium in this economy.

Page 55: Macroeconomics Phd Lectures notes

3.3. COMPETITIVE EQUILIBRIUM GROWTH 49

3.3 Competitive Equilibrium Growth

Suppose we have solved the social planners problem for a Pareto efficient al-location c∗t , k∗t+1∞t=0. What we are genuinely interested in are allocations andprices that arise when firms and consumers interact in markets. These mar-kets may be perfectly competitive, in the sense that consumers and firms actas price takers, or main entail strategic interaction between consumers and/orfirms. In this section we will discuss the connection between Pareto optimal al-locations and allocations arising in a competitive equilibrium. There is usuallyno such connection between Pareto optimal allocations and allocations arisingin situations in which agents act strategically. So for the moment we leave thesesituations to the game theorists.For the discussion of Pareto optimal allocations it did not matter who owns

what in the economy, since the planner was allowed to freely redistribute endow-ments across agents. For a competitive equilibrium the question of ownershipis crucial. We make the following assumption on the ownership structure of theeconomy: we assume that consumers own all factors of production (i.e. theyown the capital stock at all times) and rent it out to the firms. We also assumethat households own the firms, i.e. are claimants of the firms profits.Now we have to specify the equilibrium concept and the market structure.

We assume that the final goods market and the factor markets (for labor andcapital services) are perfectly competitive, which means that households as wellas firms take prices are given and beyond their control. We assume that there is asingle market at time zero in which goods fro all future periods are traded. Afterthis market closes, in all future periods the agents in the economy just carry outthe trades they agreed upon in period 0. We assume that all contracts are per-fectly enforceable. This market is often called Arrow-Debreu market structureand the corresponding competitive equilibrium an Arrow-Debreu equilibrium.For each period there are three goods that are traded:

1. The final output good, yt that can be used for consumption ct or invest-ment. Let pt denote the price of the period t final output good, quoted inperiod 0.

2. Labor services nt. Let wt be the price of one unit of labor services deliveredin period t, quoted in period 0, in terms of the period t consumptiongood. Hence wt is the real wage; it tells how many units of the period tconsumption goods one can buy for the receipts for one unit of labor. Thenominal wage is ptwt

3. Capital services kt. Let rt be the rental price of one unit of capital ser-vices delivered in period t, quoted in period 0, in terms of the period tconsumption good. rt is the real rental rate of capital, the nominal rentalrate is ptrt.

Figure 3.4 summarizes the flows of goods and payments in the economy (notethat, since all trade takes place in period 0, no payments are made after period0)

Page 56: Macroeconomics Phd Lectures notes

50CHAPTER 3. THE NEOCLASSICALGROWTHMODEL IN DISCRETE TIME

Firms

y=F(k,n)

Households

Preferences u, β

Endowments eProfits π

Sell output yt

pt

supply labor n ,capital kt t

w , rt t

Figure 3.4:

3.3.1 Definition of Competitive Equilibrium

Now we will define a competitive equilibrium for this economy. Let us first lookat firms. Without loss of generality assume that there is a single, representativefirm that behaves competitively (note: when making this assumption for firms,this is a completely innocuous assumption as long as the technology featuresconstant returns to scale. We will come back to this point). The representativefirm’s problem is , given a sequence of prices pt, wt, rt∞t=0

π = maxyt,kt,nt∞t=0

∞Xt=0

pt(yt − rtkt − wtnt) (3.6)

s.t. yt = F (kt, nt) for all t ≥ 0yt, kt, nt ≥ 0

Page 57: Macroeconomics Phd Lectures notes

3.3. COMPETITIVE EQUILIBRIUM GROWTH 51

Hence firms chose an infinite sequence of inputs kt, nt to maximize total profitsπ. Since in each period all inputs are rented (the firm does not make the capitalaccumulation decision), there is nothing dynamic about the firm’s problem andit will separate into an infinite number of static maximization problems. Morelater.Households face a fully dynamic problem in this economy. They own the

capital stock and hence have to decide how much labor and capital services tosupply, how much to consume and how much capital to accumulate. Takingprices pt, wt, rt∞t=0 as given the representative consumer solves

maxct,it,xt+1,kt,nt∞t=0

∞Xt=0

βtU (ct) (3.7)

s.t.∞Xt=0

pt(ct + it) ≤∞Xt=0

pt(rtkt + wtnt) + π

xt+1 = (1− δ)xt + it all t ≥ 00 ≤ nt ≤ 1, 0 ≤ kt ≤ xt all t ≥ 0

ct, xt+1 ≥ 0 all t ≥ 0x0 given

A few remarks are in order. First, there is only one, time zero budget con-straint, the so-called Arrow-Debreu budget constraint, as markets are only openin period 0. Secondly we carefully distinguish between the capital stock xt andcapital services that households supply to the firm. Capital services are tradedand hence have a price attached to them, the capital stock xt remains in thepossession of the household, is never traded and hence does not have a priceattached to it.7 We have implicitly assumed two things about technology: a)the capital stock depreciates no matter whether it is rented out to the firm ornot and b) there is a technology for households that transforms one unit of thecapital stock at time t into one unit of capital services at time t. The constraintkt ≤ xt then states that households cannot provide more capital services thanthe capital stock at their disposal produces. Also note that we only require thecapital stock to be nonnegative, but not investment. In this sense the capitalstock is “putty-putty”. We are now ready to define a competitive equilibriumfor this economy.

Definition 12 A Competitive Equilibrium (Arrow-Debreu Equilibrium) con-sists of prices pt, wt, rt∞t=0 and allocations for the firm kdt , ndt , yt∞t=0 andthe household ct, it, xt+1, kst , nst∞t=0 such that1. Given prices pt, wt, rt∞t=0, the allocation of the representative firm kdt , ndt , yt∞t=0solves (3.6)

7This is not quite correct: we do not require investment it to be positive. To the extentthat it < −ct is chosen by households, households in fact could transform part of the capitalstock back into final output goods and sell it back to the firm. In equilibrium this will neverhappen of course, since it would require negative production of firms (or free disposal, whichwe ruled out).

Page 58: Macroeconomics Phd Lectures notes

52CHAPTER 3. THE NEOCLASSICALGROWTHMODEL IN DISCRETE TIME

2. Given prices pt, wt, rt∞t=0, the allocation of the representative householdct, it, xt+1, kst , nst∞t=0 solves (3.7)

3. Markets clear

yt = ct + it (Goods Market)

ndt = nst (Labor Market)

kdt = kst (Capital Services Market)

3.3.2 Characterization of the Competitive Equilibrium andthe Welfare Theorems

Let us start with a partial characterization of the competitive equilibrium. Firstof all we simplify notation and denote by kt = k

dt = k

st the equilibrium demand

and supply of capital services. Similarly nt = ndt = nst . It is straightforwardto show that in any equilibrium pt > 0 for all t, since the utility function isstrictly increasing in consumption (and therefore consumption demand wouldbe infinite at a zero price). But then, since the production function exhibitspositive marginal products, rt, wt > 0 in any competitive equilibrium becauseotherwise factor demands would become unbounded.Now let us analyze the problem of the representative firm. As stated earlier,

the firms does not face a dynamic decision problem as the variables chosen atperiod t, (yt, kt, nt) do not affect the constraints nor returns (profits) at laterperiods. The static profit maximization problem for the representative firm isgiven by

maxkt,nt≥0

pt (F (kt, nt)− rtkt − wtnt)

Since the firm take prices as given, the usual “factor price equals marginalproduct” conditions arise

rt = Fk(kt, nt)

wt = Fn(kt, nt)

Note that this implies that the profits the firms earns in period t are equal to

πt = pt (F (kt, nt)− Fk(kt, nt)kt − Fn(kt, nt)nt) = 0This follows from the assumption that the function F exhibits constant returnsto scale (is homogeneous of degree 1)

F (λk,λn) = λF (k, n) for all λ > 0

and from Euler’s theorem8 which implies that

F (kt, nt) = Fk(kt, nt)kt + Fn(kt, nt)nt

8Euler’s theorem states that for any function that is homogeneous of degree k and differ-

Page 59: Macroeconomics Phd Lectures notes

3.3. COMPETITIVE EQUILIBRIUM GROWTH 53

Therefore the total profits of the representative firm are equal to zero in equi-librium. This argument in fact shows that with CRTS the number of firms isindeterminate in equilibrium; it could be one firm, two firms each operating athalf the scale of the one firm or 10 million firms. So this really justifies that theassumption of a single representative firm is without any loss of generality (aslong as we assume that this firm acts as a price taker). It also follows that therepresentative household, as owner of the firm, does not receive any profits inequilibrium.Let’s turn to that infamous representative household. Given that output

and factor prices have to be positive in equilibrium it is clear that the utilitymaximizing choices of the household entail

nt = 1, kt = xt

it = kt+1 − (1− δ)kt

From the equilibrium condition in the goods market we also obtain

F (kt, 1) = ct + kt+1 − (1− δ)kt

f(kt) = ct + kt+1

Since utility is strictly increasing in consumption, we also can conclude that theArrow-Debreu budget constraint holds with equality in equilibrium. Using thefirst results we can rewrite the household problem as

maxct,kt=1∞t=0

∞Xt=0

βtU (ct)

s.t.∞Xt=0

pt(ct + kt+1 − (1− δ)kt) =∞Xt=0

pt(rtkt + wt)

ct, kt+1 ≥ 0 all t ≥ 0k0 given

entiable at x ∈ RL we have

kf(x) =LXi=1

xi∂f(x)

∂xi

Proof. Since f is homogeneous of degree k we have for all λ > 0

f(λx) = λkf(x)

Differentiating both sides with respect to λ yields

LXi=1

xi∂f(λx)

∂xi= kλk−1f(x)

Setting λ = 1 yields

LXi=1

xi∂f(x)

∂xi= kf(x)

Page 60: Macroeconomics Phd Lectures notes

54CHAPTER 3. THE NEOCLASSICALGROWTHMODEL IN DISCRETE TIME

Again the first order conditions are necessary for a solution to the householdoptimization problem. Attaching µ to the Arrow-Debreu budget constraint andignoring the nonnegativity constraints on consumption and capital stock we getas first order conditions9 with respect to ct, ct+1 and kt+1

βtU 0(ct) = µpt

βt+1U 0(ct+1) = µpt+1

µpt = µ(1− δ + rt+1)pt+1

Combining yields the Euler equation

βU 0(ct+1)U 0(ct)

=pt+1pt

=1

1 + rt+1 − δ

(1− δ + rt+1)βU0(ct+1)

U 0(ct)= 1

Note that the net real interest rate in this economy is given by rt+1 − δ; whena household saves one unit of consumption for tomorrow, she can rent it outtomorrow of a rental rate rt+1, but a fraction δ of the one unit depreciates, sothe net return on her saving is rt+1 − δ. In these note we sometimes let rt+1denote the net real interest rate, sometimes the real rental rate of capital; thecontext will always make clear which of the two concepts rt+1 stands for.Now we use the marginal pricing condition and the fact that we defined

f(kt) = F (kt, 1) + (1− δ)kt

rt = Fk(kt, 1) = f0(kt)− (1− δ)

and the market clearing condition from the goods market

ct = f(kt)− kt+1in the Euler equation to obtain

f 0(kt+1)βU 0(f(kt+1)− kt+2)U 0(f(kt)− kt+1) = 1

which is exactly the same Euler equation as in the social planners problem.But as with the social planners problem the households’ maximization problemis an infinite-dimensional optimization problem and the Euler equations are ingeneral not sufficient for an optimum.

9That the nonnegativity constraints on consumption do not bind follows directly from theInada conditions. The nonnegativity constraints on capital could potentially bind if we lookat the household problem in isolation. However, since from the production function kt = 0implies F (kt, 1) = 0 and hence ct = 0 (which can never happen in equilibrium) we take theshortcut and ignore the corners with respect to capital holdings. But you should be awareof the fact that we did something here that was not very koscher, we implicitly imposed anequilibrium condition before carrying out the maximization problem of the household. Thisis OK here, but may lead to a lot of havoc when used in other circumstances.

Page 61: Macroeconomics Phd Lectures notes

3.3. COMPETITIVE EQUILIBRIUM GROWTH 55

Now we assert that the Euler equation, together with the transversality con-dition, is a necessary and sufficient condition for optimization. This conjectureis, to the best of my knowledge, not yet proved (or disproved) for the assump-tions that we made on U, f. As with the social planners problem we assert thatfor the assumptions we made on U, f the Euler conditions with the TVC arejointly sufficient and they are both necessary.10

The TVC for the household problem state that the value of the capital stocksaved for tomorrow must converge to zero as time goes to infinity

limt→∞ ptkt+1 = 0

But using the first order condition yields

limt→∞ ptkt+1 =

1

µlimt→∞βtU 0(ct)kt+1

=1

µlimt→∞βt−1U 0(ct−1)kt

=1

µlimt→∞βt−1βU 0(ct)(1− δ + rt)kt

=1

µlimt→∞βtU 0(f(kt)− kt+1)f 0(kt)kt

where the Lagrange multiplier µ on the Arrow-Debreu budget constraint is pos-itive since the budget constraint is strictly binding. Note that this is exactly thesame TVC as for the social planners problem. Hence an allocation of capitalkt+1∞t=0 satisfies the necessary and sufficient conditions for being a Pareto op-timal allocations if and only if it satisfies the necessary and sufficient conditionsfor being part of a competitive equilibrium (always subject to the caveat aboutthe necessity of the TVC in both problems).This last statement is our version of the fundamental theorems of welfare

economics for the particular economy that we consider. The first welfare the-orem states that a competitive equilibrium allocation is Pareto efficient (un-der very general assumptions). The second welfare theorem states that anyPareto efficient allocation can be decentralized as a competitive equilibriumwith transfers (under much more restrictive assumptions), i.e. there exist pricesand redistributions of initial endowments such that the prices, together withthe Pareto efficient allocation is a competitive equilibrium for the economy withredistributed endowments.In particular, when dealing with an economy with a representative agent (i.e.

when restricting attention to type-identical allocations), whenever the secondwelfare theorem applies we can solve for Pareto efficient allocations by solvinga social planners problem and be sure that all Pareto efficient allocations are

10Note that Stokey et al. in Chapter 2.3, when they discuss the relation between the plan-ning problem and the competitive equilibrium allocation use the finite horizon case, becausefor this case, under the assumptions made the Euler equations are both necessary and suf-ficient for both the planning problem and the household optimization problem, so we don’thave to worry about the TVC.

Page 62: Macroeconomics Phd Lectures notes

56CHAPTER 3. THE NEOCLASSICALGROWTHMODEL IN DISCRETE TIME

competitive equilibrium allocations (since there is nobody to redistribute en-dowments to/from). If, in addition, the first welfare theorem applies we can besure that we found all competitive equilibrium allocations.Also note an important fact. The first welfare theorem is usually easy to

prove, whereas the second welfare theorem is substantially harder, in particularin infinite-dimensional spaces. Suppose we have proved the first welfare theoremand we have established that there exists a unique Pareto efficient allocation(this in general requires representative agent economies and restrictions to type-identical allocations, but in these environments boils down to showing that thesocial planners problem has a unique solution). Then we have established that,if there is a competitive equilibrium, its allocation has to equal the Paretoefficient allocation. Of course we still need to prove existence of a competitiveequilibrium, but this is not surprising given the intimate link between the secondwelfare theorem and the existence proof.Back to our economy at hand. Once we have determined the equilibrium

sequence of capital stocks kt+1∞t=0 we can construct the rest of the competitiveequilibrium. In particular equilibrium allocations are given by

ct = f(kt)− kt+1yt = f(kt)

it = yt − ctnt = 1

for all t ≥ 0. Finally we can construct factor equilibrium prices as

rt = Fk(kt, 1)

wt = Fn(kt, 1)

Finally, the prices of the final output good can be found as follows. As usualprices are determined only up to a normalization, so let us pick p0 = 1. Fromthe Euler equations for the household in then follows that

pt+1 =βU 0(ct+1)U 0(ct)

pt

pt+1pt

=βU 0(ct+1)U 0(ct)

=1

1 + rt+1 − δ

pt+1 =βt+1U 0(ct+1)U 0(c0)

=tY

τ=0

1

1 + rτ+1 − δ

and we have constructed a complete competitive equilibrium, conditional onhaving found kt+1∞t=0. The welfare theorems tell us that we can solve a socialplanner problem to do so, and the next sections will tell us how we can do soby using recursive methods.

3.3.3 Sequential Markets Equilibrium

[To be completed]

Page 63: Macroeconomics Phd Lectures notes

3.3. COMPETITIVE EQUILIBRIUM GROWTH 57

3.3.4 Recursive Competitive Equilibrium

[To be completed]

Page 64: Macroeconomics Phd Lectures notes

58CHAPTER 3. THE NEOCLASSICALGROWTHMODEL IN DISCRETE TIME

Page 65: Macroeconomics Phd Lectures notes

Chapter 4

Mathematical Preliminaries

We want to study functional equations of the form

v(x) = maxy∈Γ(x)

F (x, y) + βv(y)

where r is the period return function (such as the utility function) and Γ is theconstraint set. Note that for the neoclassical growth model x = k, y = k0 andF (k, k0) = U(f(k)− k0) and Γ(k) = k0 ∈ R :0 ≤ k ≤ f(k)In order to so we define the following operator T

(Tv) (x) = maxy∈Γ(x)

F (x, y) + βv(y)

This operator T takes the function v as input and spits out a new function Tv.In this sense T is like a regular function, but it takes as inputs not scalars z ∈ Ror vectors z ∈ Rn, but functions v from some subset of possible functions. Asolution to the functional equation is then a fixed point of this operator, i.e. afunction v∗ such that

v∗ = Tv∗

We want to find out under what conditions the operator T has a fixed point(existence), under what conditions it is unique and under what conditions wecan start from an arbitrary function v and converge, by applying the operator Trepeatedly, to v∗. More precisely, by defining the sequence of functions vn∞n=0recursively by v0 = v and vn+1 = Tvn we want to ask under what conditionslimn→∞ vn = v∗.In order to make these questions (and the answers to them) precise we have

to define the domain and range of the operator T and we have to define whatwe mean by lim . This requires the discussion of complete metric spaces. In thenext subsection I will first define what a metric space is and then what makesa metric space complete.Then I will state and prove the contraction mapping theorem. This theorem

states that an operator T, defined on a metric space, has a unique fixed point if

59

Page 66: Macroeconomics Phd Lectures notes

60 CHAPTER 4. MATHEMATICAL PRELIMINARIES

this operator T is a contraction (I will obviously first define what a contractionis). Furthermore it assures that from any starting guess v repeated applicationsof the operator T will lead to its unique fixed point.Finally I will prove a theorem, Blackwell’s theorem, that provides sufficient

condition for an operator to be a contraction. We will use this theorem to provethat for the neoclassical growth model the operator T is a contraction and hencethe functional equation of our interest has a unique solution.

4.1 Complete Metric Spaces

Definition 13 A metric space is a set S and a function d : S × S → R suchthat for all x, y, z ∈ S1. d(x, y) ≥ 02. d(x, y) = 0 if and only if x = y

3. d(x, y) = d(y, x)

4. d(x, z) ≤ d(x, y) + d(y, z)The function d is called a metric and is used to measure the distance between

two elements in S. The second property is usually referred to as symmetry,the third as triangle inequality (because of its geometric interpretation in RExamples of metric spaces (S, d) include1

Example 14 S = R with metric d(x, y) = |x− y|

Example 15 S = R with metric d(x, y) =

½1 if x 6= y0 otherwise

Example 16 S = l∞ = x = x∞t=0 |xt ∈ R, all t ≥ 0 and supt |xt| <∞ withmetric d(x, y) = supt |xt − yt|Example 17 Let X ⊆ Rl and S = C(X) be the set of all continuous andbounded functions f : X → R. Define the metric d : C(X) × C(X) → R asd(f, g) = supx∈X |f(x)− g(x)|. Then (S, d) is a metric space

1A function f : X → R is said to be bounded if there exists a constant K > 0 such that|f(x)| < K for all x ∈ X.Let S be any subset of R. A number u ∈ R is said to be an upper bound for the set S if

s ≤ u for all s ∈ S. The supremum of S, sup(S) is the smallest upper bound of S.Every set in R that has an upper bound has a supremum (imposed by the completeness

axiom). For sets that are unbounded above some people say that the supremum does notexist, others write sup(S) =∞. We will follow the second convention.Also note that sup(S) = max(S), whenever the latter exists. What the sup buys us is that

it always exists even when the max does not. A simle example

S =

½− 1n: n ∈N

¾For this example sup(S) = 0 whereas max(S) does not exist.

Page 67: Macroeconomics Phd Lectures notes

4.2. CONVERGENCE OF SEQUENCES 61

A few remarks: the space l∞ (with corresponding norm) will be importantwhen we discuss the welfare theorems as naturally consumption allocations formodels with infinitely lived consumers are infinite sequences. Why we want torequire these sequences to be bounded will become clearer later.The space C(X) with norm d as defined above will be used immediately as

we will define the domain of our operator T to be C(X), i.e. T uses as inputscontinuous and bounded functions.Let us prove that some of the examples are indeed metric spaces. For the

first example the result is trivial.

Claim 18 S = R with metric d(x, y) =

½1 if x 6= y0 otherwise

is a metric space

Proof. We have to show that the function d satisfies all three properties inthe definition. The first three properties are obvious. For the forth property: ifx = z, the result follows immediately. So suppose x 6= z. Then d(x, z) = 1. Butthen either y 6= x or y 6= z (or both), so that d(x, y) + d(y, z) ≥ 1

Claim 19 l∞ together with the sup-metric is a metric space

Proof. Take arbitrary x, y, z ∈ l∞. From the basic triangle inequality on Rwe have that |xt − yt| ≤ |xt|+|yt|. Hence, since supt |xt| <∞ and supt |yt| <∞,we have that supt |xt−yt| <∞. Property 1 is obvious. If x = y (i.e. xt = yt forall t), then |xt − yt| = 0 for all t, hence supt |xt − yt| = 0. Suppose x 6= y. Thenthere exists T such that xT 6= yT , hence |xT − yT | > 0, hence supt |xt − yt| > 0Property 3 is obvious since |xt − yt| = |yt − xt|, all t. Finally for property 4.

we note that for all t

|xt − zt| ≤ |xt − yt|+ |yt − zt|

Since this is true for all t, we can apply the sup to both sides to obtain theresult (note that the sup on both sides is finite).

Claim 20 C(X) together with the sup-norm is a metric space

Proof. Take arbitrary f, g ∈ C(X). f = g means that f(x) = g(x) for allx ∈ X. Since f, g are bounded, supx∈X |f(x)| < ∞ and supx∈X |f(x)| < ∞, sosupx∈X |f(x)− g(x)| <∞. Property 1. through 3. are obvious and for property4. we use the same argument as before, including the fact that f, g ∈ C(X)implies that supx∈X |f(x)− g(x)| <∞.

4.2 Convergence of Sequences

The next definition will make precise the meaning of statements of the formlimn→∞ vn = v∗. For an arbitrary metric space (S, d) we have the followingdefinition.

Page 68: Macroeconomics Phd Lectures notes

62 CHAPTER 4. MATHEMATICAL PRELIMINARIES

Definition 21 A sequence xn∞n=0 with xn ∈ S for all n is said to convergeto x ∈ S, if for every ε > 0 there exists a Nε ∈N such that d(xn, x) < ε for alln ≥ Nε. In this case we write limn→∞ xn = x.

This definition basically says that a sequence xn∞n=0 converges to a pointif we, for every distance ε > 0 we can find an index Nε so that the sequence ofxn is not more than ε away from x after the Nε element of the sequence. Alsonote that, in order to verify that a sequence converges, it is usually necessaryto know the x to which it converges in order to apply the definition directly.

Example 22 Take S = R with d(x, y) = |x − y|. Define xn∞n=0 by xn = 1n .

Then limn→∞ xn = 0. This is straightforward to prove, using the definition.Take any ε > 0. Then d(xn, 0) =

1n . By taking Nε =

2ε we have that for n ≥ Nε,

d(xn, 0) =1n ≤ 1

Nε= ε

2 < ε (if Nε =2ε is not an integer, take the next biggest

integer).

For easy examples of sequences it is no problem to guess the limit. Note thatthe limit of a sequence, if it exists, is always unique (you should prove this foryourself). For not so easy examples this may not work. There is an alternativecriterion of convergence, due to Cauchy.2

Definition 23 A sequence xn∞n=0 with xn ∈ S for all n is said to be a Cauchysequence if for each ε > 0 there exists a Nε ∈ N such that d(xn, xm) < ε for alln,m ≥ Nε

Hence a sequence xn∞n=0 is a Cauchy sequence if for every distance ε > 0we can find an index Nε so that the elements of the sequence do not differ bymore than by ε.

Example 24 Take S = R with d(x, y) = |x − y|. Define xn∞n=0 by xn = 1n .

This sequence is a Cauchy sequence. Again this is straightforward to prove. Fixε > 0 and take any n,m ∈ N. Without loss of generality assume that m > n.Then d(xn, xm) =

1n − 1

m < 1n . Pick Nε =

2ε and we have that for n,m ≥ Nε,

d(xn, 0) <1n ≤ 1

Nε= ε

2 < ε. Hence the sequence is a Cauchy sequence.

So it turns out that the sequence in the last example both converges and is aCauchy sequence. This is not an accident. In fact, one can prove the following

Theorem 25 Suppose that (S, d) is a metric space and that the sequence xn∞n=0converges to x ∈ S. Then the sequence xn∞n=0 is a Cauchy sequence.Proof. Since xn∞n=0 converges to x, there existsM ε

2such that d(xn, x) <

ε2

for all n ≥ M ε2. Therefore if n,m ≥ Nε we have that d(xn, xm) ≤ d(xn, x) +

d(xm, x) <ε2 +

ε2 = ε (by the definition of convergence and the triangle inequal-

ity). But then for any ε > 0, pick Nε =M ε2and it follows that for all n,m ≥ Nε

we have d(xn, xm) < ε

2Augustin-Louis Cauchy (1789-1857) was the founder of modern analysis. He wrote about800 (!) mathematical papers during his scientific life.

Page 69: Macroeconomics Phd Lectures notes

4.2. CONVERGENCE OF SEQUENCES 63

Example 26 Take S = R with d(x, y) =

½1 if x 6= y0 otherwise

. Define xn∞n=0 byxn =

1n . Obviously d(xn, xm) = 1 for all n,m ∈ N. Therefore the sequence is

not a Cauchy sequence. It then follows from the preceding theorem (by takingthe contrapositive) that the sequence cannot converge. This example shows that,whenever discussing a metric space, it is absolutely crucial to specify the metric.

This theorem tells us that every convergent sequence is a Cauchy sequence.The reverse does not always hold, but it is such an important property thatwhen it holds, it is given a particular name.

Definition 27 A metric space (S, d) is complete if every Cauchy sequence xn∞n=0with xn ∈ S for all n converges to some x ∈ S.Note that the definition requires that the limit x has to lie within S.We are

interested in complete metric spaces since the Contraction Mapping Theoremdeals with operators T : S → S, where (S, d) is required to be a complete metricspace. Also note that there are important examples of complete metric spaces,but other examples where a metric space is not complete (and for which theContraction Mapping Theorem does not apply).

Example 28 Let S be the set of all continuous, strictly decreasing functionson [1, 2] and let the metric on S be defined as d(f, g) = supx∈[1,2] |f(x)− g(x)|.I claim that (S, d) is not a complete metric space. This can be proved by anexample of a sequence of functions fn∞n=0 that is a Cauchy sequence, but doesnot converge within S. Define fn : [0, 1] → R by fn(x) =

1nx . Obviously all fn

are continuous and strictly decreasing on [1, 2], hence fn ∈ S for all n. Let usfirst prove that this sequence is a Cauchy sequence. Fix ε > 0 and take Nε =

2ε .

Suppose that m,n ≥ Nε and without loss of generality assume that m > n. Then

d(fn, fm) = supx∈[1,2]

¯1

nx− 1

mx

¯= sup

x∈[1,2]

1

nx− 1

mx

= supx∈[1,2]

m− nmnx

=m− nmn

=1− n

m

n

≤ 1

n≤ 1

Nε=

ε

2< ε

Hence the sequence is a Cauchy sequence. But since for all x ∈ [1, 2], limn→∞ fn(x) =0, the sequence converges to the function f, defined as f(x) = 0, for all x ∈ [1, 2].But obviously, since f is not strictly decreasing, f /∈ S. Hence (S, d) is not acomplete metric space. Note that if we choose S to be the set of all continu-ous and decreasing (or increasing) functions on R, then S, together with thesup-norm, is a complete metric space.

Page 70: Macroeconomics Phd Lectures notes

64 CHAPTER 4. MATHEMATICAL PRELIMINARIES

Example 29 Let S = RL and d(x, y) = L

qPLl=1 |xl − yl|L. (S, d) is a complete

metric space. This is easily proved by proving the following three lemmata (whichis left to the reader).

1. Every Cauchy sequence xn∞n=0 in RL is bounded

2. Every bounded sequence xn∞n=0 in RL has a subsequence xni∞i=0 con-verging to some x ∈ RL (Bolzano-Weierstrass Theorem)

3. For every Cauchy sequence xn∞n=0 in RL, if a subsequence xni∞i=0converges to x ∈ RL, then the entire sequence xn∞n=0 converges to x ∈RL.

Example 30 This last example is very important for the applications we areinterested in. Let X ⊆ RL and C(X) be the set of all bounded continuousfunctions f : X → R with d being the sup-norm. Then (C(X), d) is a completemetric space.

Proof. (This follows SLP, pp. 48) We already proved that (C(X), d) is ametric space. Now we want to prove that this space is complete. Let fn∞n=0be an arbitrary sequence of functions in C(X) which is Cauchy. We need toestablish the existence of a function f ∈ C(X) such that for all ε > 0 thereexists Nε satisfying supx∈X |fn(x)− f(x)| < ε for all n ≥ Nε.We will proceed in three steps: a) find a candidate for f, b) establish that the

sequence fn∞n=0 converges to f in the sup-norm and c) show that f ∈ C(X).1. Since fn∞n=0 is Cauchy, for each ε > 0 there existsMε such that supx∈X |fn(x)−fm(x)| < ε for all n,m ≥Mε.Now fix a particular x ∈ X. Then fn(x)∞n=0is just a sequence of numbers. Now

|fn(x)− fm(x)| ≤ supy∈X

|fn(y)− fm(y)| < ε

Hence the sequence of numbers fn(x)∞n=0 is a Cauchy sequence in R.SinceR is a complete metric space, fn(x)∞n=0 converges to some number,call it f(x). By repeating this argument for all x ∈ X we derive ourcandidate function f ; it is the pointwise limit of the sequence of functionsfn∞n=0.

2. Now we want to show that fn∞n=0 converges to f as constructed above.Hence we want to argue that d(fn, f) goes to zero as n goes to infinity.Fix ε > 0. Since fn∞n=0 is Cauchy, it follows that there exists Nε suchthat d(fn, fm) < ε for all n,m ≥ Nε. Now fix x ∈ X. For any m ≥ n ≥ Nε

we have (remember that the norm is the sup-norm)

|fn(x)− f(x)| ≤ |fn(x)− fm(x)|+ |fm(x)− f(x)|≤ d(fn, fm) + |fm(x)− f(x)|≤ ε

2+ |fm(x)− f(x)|

Page 71: Macroeconomics Phd Lectures notes

4.3. THE CONTRACTION MAPPING THEOREM 65

But since fn∞n=0 converges to f pointwise, we have that |fm(x)−f(x)| <ε2 for all m ≥ Nε(x), where Nε(x) is a number that may (and in generaldoes) depend on x. But then, since x ∈ X was arbitrary, |fn(x)−f(x)| < εfor all n ≥ Nε (the key is that this Nε does not depend on x). Thereforesupx∈X |fn(x) − f(x)| = d(fn, f) ≤ ε and hence the sequence fn∞n=0converges to f.

3. Finally we want to show that f ∈ C(X), i.e. that f is bounded andcontinuous. Since fn∞n=0 lies in C(X), all fn are bounded, i.e. there isa sequence of numbers Kn∞n=0 such that supx∈X |fn(x)| ≤ Kn. But bythe triangle inequality, for arbitrary n

supx∈X

|f(x)| = supx∈X

|f(x)− fn(x) + fn(x)|≤ sup

x∈X|f(x)− fn(x)|+ sup

x∈X|fn(x)|

≤ supx∈X

|f(x)− fn(x)|+Kn

But since fn∞n=0 converges to f, there existsNε such that supx∈X |f(x)−fn(x)| < ε for all n ≥ Nε. Fix an ε and take K = KNε + 2ε. It isobvious that supx∈X |f(x)| ≤ K. Hence f is bounded. Finally we provecontinuity of f. Let the Euclidean metric on RL be denoted by ||x −y|| = L

qPLl=1 |xl − yl|L We need to show that for every ε > 0 and every

x ∈ X there exists a δ(ε, x) > 0 such that if ||x − y|| < δ(ε, x) then|f(x)− f(y)| < ε, for all x, y ∈ X. Fix ε and x. Pick a k large enough sothat d(fk, f) <

ε3 (which is possible as fn∞n=0 converges to f). Choose

δ(ε, x) > 0 such that ||x− y|| < δ(ε, x) implies |fk(x)− fk(y)| < ε3 . Since

all fn ∈ C(X), fk is continuous and hence such a δ(ε, x) > 0 exists. Now

|f(x)− f(y)| ≤ |f(x)− fk(x)|+ |fk(x)− fk(y)|+ |fk(y)− f(y)|≤ d(f, fk) + |fk(x)− fk(y)|+ d(fk, f)≤ ε

3+

ε

3+

ε

3= ε

4.3 The Contraction Mapping Theorem

Now we are ready to state the theorem that will give us the existence anduniqueness of a fixed point of the operator T, i.e. existence and uniqueness ofa function v∗ satisfying v∗ = Tv∗. Let (S, d) be a metric space. Just to clarify,an operator T (or a mapping) is just a function that maps elements of S intosome other space. The operator that we are interested in maps functions intofunctions, but the results in this section apply to any metric space. We startwith a definition of what a contraction mapping is.

Page 72: Macroeconomics Phd Lectures notes

66 CHAPTER 4. MATHEMATICAL PRELIMINARIES

Definition 31 Let (S, d) be a metric space and T : S → S be a function map-ping S into itself. The function T is a contraction mapping if there exists anumber β ∈ (0, 1) satisfying

d(Tx, Ty) ≤ βd(x, y) for all x, y ∈ SThe number β is called the modulus of the contraction mapping

A geometric example of a contraction mapping for S = [0, 1], d(x, y) = |x−y|is contained in SLP, p. 50. Note that a function that is a contraction mappingis automatically a continuous function, as the next lemma shows

Lemma 32 Let (S, d) be a metric space and T : S → S be a function mappingS into itself. If T is a contraction mapping, then T is continuous.

Proof. Remember from the definition of continuity we have to show thatfor all s0 ∈ S and all ε > 0 there exists a δ(ε, s0) such that whenever s ∈S, d(s, s0) < δ(ε, s0), then d(Ts, Ts0) < ε. Fix an arbitrary s0 ∈ S and ε > 0and pick δ(ε, s0) = ε. Then

d(Ts, Ts0) ≤ βd(s, s0) ≤ βδ(ε, s0) = βε < ε

We now can state and prove the contraction mapping theorem. Let byvn = Tnv0 ∈ S denote the element in S that is obtained by applying theoperator T n-times to v0, i.e. the n-th element in the sequence starting with anarbitrary v0 and defined recursively by vn = Tvn−1 = T (Tvn−2) = · · · = Tnv0.Then we have

Theorem 33 Let (S, d) be a complete metric space and suppose that T : S → Sis a contraction mapping with modulus β. Then a) the operator T has exactlyone fixed point v∗ ∈ S and b) for any v0 ∈ S, and any n ∈ N we have

d(Tnv0, v∗) ≤ βnd(v0, v

∗)

A few remarks before the proof. Part a) of the theorem tells us that thereis a v∗ ∈ S satisfying v∗ = Tv∗ and that there is only one such v∗ ∈ S. Partb) asserts that from any starting guess v0, the sequence vn∞n=0 as definedrecursively above converges to v∗ at a geometric rate of β. This last part isimportant for computational purposes as it makes sure that we, by repeatedlyapplying T to any (as crazy as can be) initial guess v0 ∈ S, will eventuallyconverge to the unique fixed point and it gives us a lower bound on the speedof convergence. But now to the proof.Proof. First we prove part a) Start with an arbitrary v0. As our candidate

for a fixed point we take v∗ = limn→∞ vn. We first have to establish that thesequence vn∞n=0 in fact converges to a function v∗.We then have to show thatthis v∗ satisfies v∗ = Tv∗ and we then have to show that there is no other vthat also satisfies v = T v

Page 73: Macroeconomics Phd Lectures notes

4.3. THE CONTRACTION MAPPING THEOREM 67

Since by assumption T is a contraction

d(vn+1, vn) = d(Tvn, Tvn−1) ≤ βd(vn, vn−1)= βd(Tvn−1, Tvn−2) ≤ β2d(vn−1, vn−2)= · · · = βnd(v1, v0)

where we used the way the sequence vn∞n=0 was constructed, i.e. the fact thatvn+1 = Tvn. For any m > n it then follows from the triangle inequality that

d(vm, vn) ≤ d(vm, vm−1) + d(vm−1, vn)≤ d(vm, vm−1) + d(vm−1, vm−2) + · · ·+ d(vn+1, vn)≤ βmd(v1, v0) + βm−1d(v1, v0) + · · ·βnd(v1, v0)= βn

¡βm−n−1 + · · ·+ β + 1

¢d(v1, v0)

≤ βn

1− βd(v1, v0)

By making n large we can make d(vm, vn) as small as we want. Hence thesequence vn∞n=0 is a Cauchy sequence. Since (S, d) is a complete metric space,the sequence converges in S and therefore v∗ = limn→∞ vn is well-defined.Now we establish that v∗ is a fixed point of T, i.e. we need to show that

Tv∗ = v∗. But

Tv∗ = T³limn→∞ vn

´= limn→∞T (vn) = lim

n→∞ vn+1 = v∗

Note that the fact that T (limn→∞ vn) = limn→∞ T (vn) follows from the conti-nuity of T.3

Now we want to prove that the fixed point of T is unique. Suppose thereexists another v ∈ S such that v = T v and v 6= v∗. Then there exists c > 0 suchthat d(v, v∗) = a. But

0 < a = d(v, v∗) = d(T v, Tv∗) ≤ βd(v, v∗) = βa

a contradiction. Here the second equality follows from the fact that we assumedthat both v, v∗ are fixed points of T and the inequality follows from the factthat T is a contraction.We prove part b) by induction. For n = 0 (using the convention that T 0v =

v) the claim automatically holds. Now suppose that

d(T kv0, v∗) ≤ βkd(v0, v

∗)

We want to prove that

d(T k+1v0, v∗) ≤ βk+1d(v0, v

∗)

3Almost by definition. Since T is continuous for every ε > 0 there exists a δ(ε) such thatd(vn−v∗) < δ(ε) implies d(T (vn)−T (v∗)) < ε. Hence the sequence T (vn)∞n=0 converges andlimn→∞ T (vn) is well-defined. We showed that limn→∞ vn = v∗. Hence both limn→∞ T (vn)and limn→∞ vn are well-defined. Then obviously limn→∞ T (vn) = T (v∗) = T (limn→∞ vn).

Page 74: Macroeconomics Phd Lectures notes

68 CHAPTER 4. MATHEMATICAL PRELIMINARIES

But

d(T k+1v0, v∗) = d(T

¡T kv0

¢, T v∗) ≤ βd(T kv0, v

∗) ≤ βk+1d(v0, v∗)

where the first inequality follows from the fact that T is a contraction and thesecond follows from the induction hypothesis.This theorem is extremely useful in order to establish that our functional

equation of interest has a unique fixed point. It is, however, not very opera-tional as long as we don’t know how to determine whether a given operator isa contraction mapping. There is some good news, however. Blackwell, in 1965provided sufficient conditions for an operator to be a contraction mapping. Itturns out that these conditions can be easily checked in a lot of applications.Since they are only sufficient however, failure of these conditions does not im-ply that the operator is not a contraction. In these cases we just have to looksomewhere else. Here is Blackwell’s theorem.

Theorem 34 Let X ⊆ RL and B(X) be the space of bounded functions f :X → R with the d being the sup-norm. Let T : B(X) → B(X) be an operatorsatisfying

1. Monotonicity: If f, g ∈ B(X) are such that f(x) ≤ g(x) for all x ∈ X,then (Tf) (x) ≤ (Tg) (x) for all x ∈ X.

2. Discounting: Let the function f + a, for f ∈ B(X) and a ∈ R+ be definedby (f + a)(x) = f(x) + a (i.e. for all x the number a is added to f(x)).There exists β ∈ (0, 1) such that for all f ∈ B(X), a ≥ 0 and all x ∈ X

[T (f + a)](x) ≤ [Tf ](x) + βa

If these two conditions are satisfied, then the operator T is a contractionwith modulus β.

Proof. In terms of notation, if f, g ∈ B(X) are such that f(x) ≤ g(x) forall x ∈ X, then we write f ≤ g.We want to show that if the operator T satisfiesconditions 1. and 2. then there exists β ∈ (0, 1) such that for all f, g ∈ B(X)we have that d(Tf, Tg) ≤ βd(f, g).Fix x ∈ X. Then f(x) − g(x) ≤ supy∈X |f(y) − g(y)|. But this is true for

all x ∈ X. So using our notation we have that f ≤ g + d(f, g) (which meansthat for any value of x ∈ X, adding the constant d(f, g) to g(x) gives somethingbigger than f(x).But from f ≤ g + d(f, g) it follows by monotonicity that

Tf ≤ T [g + d(f, g)]

≤ Tg + βd(f, g)

where the last inequality comes from discounting. Hence we have

Tf − Tg ≤ βd(f, g)

Page 75: Macroeconomics Phd Lectures notes

4.3. THE CONTRACTION MAPPING THEOREM 69

Switching the roles of f and g around we get

−(Tf − Tg) ≤ βd(g, f) = βd(f, g)

(by symmetry of the metric). Combining yields

(Tf) (x)− (Tg) (x) ≤ βd(f, g) for all x ∈ X(Tg) (x)− (Tf) (x) ≤ βd(f, g) for all x ∈ X

Therefore

supx∈X

|(Tf) (x)− (Tg) (x)| = d(Tf, Tg) ≤ βd(f, g)

and T is a contraction mapping with modulus β.Note that do not require the functions in B(X) to be continuous. It is

straightforward to prove that (B(X), d) is a complete metric space once weproved that (B(X), d) is a complete metric space. Also note that we couldrestrict ourselves to continuous and bounded functions and Blackwell’s theoremobviously applies. Note however that Blackwells theorem requires the metricspace to be a space of functions, so we lose generality as compared to theContraction mapping theorem (which is valid for any complete metric space).But for our purposes it is key that, once Blackwell’s conditions are verified wecan invoke the CMT to argue that our functional equation of interest has aunique solution that can be obtained by repeated iterations on the operator T.We can state an alternative version of Blackwell’s theorem

Theorem 35 Let X ⊆ RL and B(X) be the space of bounded functions f :X → R with the d being the sup-norm. Let T : B(X) → B(X) be an operatorsatisfying

1. Monotonicity: If f, g ∈ B(X) are such that f(x) ≤ g(x) for all x ∈ X,then (Tf) (x) ≥ (Tg) (x) for all x ∈ X.

2. Discounting: Let the function f + a, for f ∈ B(X) and a ∈ R+ be definedby (f + a)(x) = f(x) + a (i.e. for all x the number a is added to f(x)).There exists β ∈ (0, 1) such that for all f ∈ B(X), a ≥ 0 and all x ∈ X

[T (f − a)](x) ≤ [Tf ](x) + βa

If these two conditions are satisfied, then the operator T is a contractionwith modulus β.

The proof is identical to the first theorem and hence omitted.As an application of the mathematical structure we developed let us look

back at the neoclassical growth model. The operator T corresponding to ourfunctional equation was

Tv(k) = max0≤k0≤f(k)

U(f(k)− k0) + βv(k0)

Page 76: Macroeconomics Phd Lectures notes

70 CHAPTER 4. MATHEMATICAL PRELIMINARIES

Define as our metric space (B(0,∞), d) the space of bounded functions on (0,∞)with d being the sup-norm. We want to argue that this operator has a uniquefixed point and we want to apply Blackwell’s theorem and the CMT. So let usverify that all the hypotheses for Blackwell’s theorem are satisfied.

1. First we have to verify that the operator T maps B(0,∞) into itself (thisis very often forgotten). So if we take v to be bounded, since we assumedthat U is bounded, then Tv is bounded. Note that you may be in bigtrouble here if U is not bounded.4

2. How about monotonicity. It is obvious that this is satisfied. Supposev ≤ w. Let by gv(k) denote an optimal policy (need not be unique) corre-sponding to v. Then for all k ∈ (0,∞)

Tv(k) = U(f(k)− gv(k)) + βv(gv(k))

≤ U(f(k)− gv(k)) + βw(gv(k))

≤ max0≤k0≤f(k)

U(f(k)− k0) + βw(k0)= Tw(k)

Even by applying the policy gv(k) (which need not be optimal for thesituation in which the value function is w) gives higher Tw(k) than Tv(k).Choosing the policy for w optimally does only improve the value (Tv) (k).

3. Discounting. This also straightforward

T (v + a)(k) = max0≤k0≤f(k)

U(f(k)− k0) + β(v(k0) + a)= max

0≤k0≤f(k)U(f(k)− k0) + βv(k0)+ βa

= Tv(k) + βa

Hence the neoclassical growth model with bounded utility satisfies the Suffi-cient conditions for a contraction and there is a unique fixed point to the func-tional equation that can be computed from any starting guess v0 be repeatedapplication of the T -operator.One can also prove some theoretical properties of the Howard improvement

algorithm using the Contraction Mapping Theorem and Blackwell’s conditions.Even though we could state the results in much generality, we will confineour discussion to the neoclassical growth model. Remember that the Howardimprovement algorithm iterates on feasible policies [To be completed]

4Somewhat surprisingly, in many applications the problem is that u is not bounded below;unboundedness from above is sometimes easy to deal with.We made the assumption that f ∈ C2 f 0 > 0, f 00 < 0, limk&0 f

0(k) = ∞ and

limk→∞ f 0(k) = 1 − δ. Hence there exists a unique k such that f(k) = k. Hence for all

kt > k we have kt+1 ≤ f(kt) < kt. Therefore we can effectively restrict ourselves to capital

stocks in the set [0,max(k0, k)]. Hence, even if u is not bounded above we have that for all

feasible paths policies u(f(k)− k0) ≤ u(f(max(k0, k)) <∞, and hence by sticking a functionv into the operator that is bounded above, we get a Tv that is bounded above. Lack ofboundedness from below is a much harder problem in general.

Page 77: Macroeconomics Phd Lectures notes

4.4. THE THEOREM OF THE MAXIMUM 71

4.4 The Theorem of the Maximum

An important theorem is the theorem of the maximum. It will help us toestablish that, if we stick a continuous function f into our operator T, theresulting function Tf will also be continuous and the optimal policy functionwill be continuous in an appropriate sense.We are interested in problems of the form

h(x) = maxy∈Γ(x)

f(x, y)

The function h gives the value of the maximization problem, conditional on thestate x. We define

G(x) = y ∈ Γ(x) : f(x, y) = h(x)Hence G is the set of all choices y that attain the maximum of f , given the statex, i.e. G(x) is the set of argmax’es. Note that G(x) need not be single-valued.In the example that we study the function f will consist of the sum of the

current return function r and the continuation value v and the constraint setdescribes the resource constraint. The theorem of the maximum is also widelyused in microeconomics. There, most frequently x consists of prices and income,f is the (static) utility function, the function h is the indirect utility function, Γis the budget set and G is the set of consumption bundles that maximize utilityat x = (p,m).Before stating the theorem we need a few definitions. Let X,Y be arbitrary

sets (in what follows we will be mostly concerned with the situations in whichX and Y are subsets of Euclidean spaces. A correspondence Γ : X ⇒ Y mapseach element x ∈ X into a subset Γ(x) of Y. Hence the image of the point xunder Γ may consist of more than one point (in contrast to a function, in whichthe image of x always consists of a singleton).

Definition 36 A compact-valued correspondence Γ : X ⇒ Y is upper-hemicontinuousat a point x if Γ(x) 6= ∅ and if for all sequences xn in X converging to somex ∈ X and all sequences yn in Y such that yn ∈ Γ(xn) for all n, there exists aconvergent subsequence of yn that converges to some y ∈ Γ(x). A correspon-dence is upper-hemicontinuous if it is upper-hemicontinuous at all x ∈ X.A few remarks: by talking about convergence we have implicitly assumed

that X and Y (together with corresponding metrices) are metric spaces. Also,a correspondence is compact-valued, if for all x ∈ X,Γ(x) is a compact set.Also this definition requires Γ to be compact-valued. With this additional re-quirement the definition of upper hemicontinuity actually corresponds to thedefinition of a correspondence having a closed graph. See, e.g. Mas-Colell et al.p. 949-950 for details.

Definition 37 A correspondence Γ : X ⇒ Y is lower-hemicontinuous at apoint x if Γ(x) 6= ∅ and if for every y ∈ Γ(x) and every sequence xn in X

Page 78: Macroeconomics Phd Lectures notes

72 CHAPTER 4. MATHEMATICAL PRELIMINARIES

converging to x ∈ X there exists N ≥ 1 and a sequence yn in Y converging toy such that yn ∈ Γ(xn) for all n ≥ N. A correspondence is lower-hemicontinuousif it is lower-hemicontinuous at all x ∈ X.

Definition 38 A correspondence Γ : X ⇒ Y is continuous if it is both upper-hemicontinuous and lower-hemicontinuous.

Note that a single-valued correspondence (i.e. a function) that is upper-hemicontinuous is continuous. Now we can state the theorem of the maximum.

Theorem 39 Let X ⊆ RL and Y ⊆ RM , let f : X × Y → R be a contin-uous function, and let Γ : X ⇒ Y be a compact-valued and continuous cor-respondence. Then h : X → R is continuous and G : X → Y is nonempty,compact-valued and upper-hemicontinuous.

The proof is somewhat tedious and omitted here (you probably have doneit in micro anyway).

Page 79: Macroeconomics Phd Lectures notes

Chapter 5

Dynamic Programming

5.1 The Principle of Optimality

In the last section we showed that under certain conditions, the functional equa-tion (FE)

v(x) = supy∈Γ(x)

F (x, y) + βv(y)

has a unique solution which is approached from any initial guess v0 at geometricspeed. What we were really interested in, however, was a problem of sequentialform (SP )

w(x0) = supxt+1∞t=0

∞Xt=0

βtF (xt, xt+1)

s.t. xt+1 ∈ Γ(xt)

x0 ∈ X given

Note that I replaced max with sup since we have not made any assumptionsso far that would guarantee that the maximum in either the functional equationor the sequential problem exists. In this section we want to find out under whatconditions the functions v and w are equal and under what conditions optimalsequential policies xt+1∞t=0 are equivalent to optimal policies y = g(x) fromthe recursive problem, i.e. under what conditions the principle of optimalityholds. It turns out that these conditions are very mild.In this section I will try to state the main results and make clear what they

mean; I will not prove the results. The interested reader is invited to consultStokey and Lucas or Bertsekas. Unfortunately, to make our results preciseadditional notation is needed. Let X be the set of possible values that the statex can take. X may be a subset of a Euclidean space, a set of functions orsomething else; we need not be more specific at this point. The correspondenceΓ : X ⇒ X describes the feasible set of next period’s states y, given that today’s

73

Page 80: Macroeconomics Phd Lectures notes

74 CHAPTER 5. DYNAMIC PROGRAMMING

state is x. The graph of Γ, A is defined as

A = (x, y) ∈ X ×X : y ∈ Γ(x)

The period return function F : A→ R maps the set of all feasible combinationsof today’s and tomorrow’s state into the reals. So the fundamentals of ouranalysis are (X,F,β,Γ). For the neoclassical growth model F and β describepreferences and X,Γ describe the technology.We call any sequence of states xt∞t=0 a plan. For a given initial condition

x0, the set of feasible plans Π(x0) from x0 is defined as

Π(x0) = xt∞t=1 : xt+1 ∈ Γ(xt)

Hence Π(x0) is the set of sequences that, for a given initial condition, satisfy allthe feasibility constraints of the economy. We will denote by x a generic elementof Π(x0). The two assumptions that we need for the principle of optimality arebasically that for any initial condition x0 the social planner (or whoever solvesthe problem) has at least one feasible plan and that the total return (the totalutility, say) from all feasible plans can be evaluated. That’s it. More preciselywe haveAssumption 1: Γ(x) is nonempty for all x ∈ XAssumption 2: For all initial conditions x0 and all feasible plans x ∈ Π(x0)

limn→∞

nXt=0

βtF (xt, xt+1)

exists (although it may be +∞ or −∞).Assumption 1 does not require much discussion: we don’t want to deal

with an optimization problem in which the decision maker (at least for someinitial conditions) can’t do anything. Assumption 2 is more subtle. There arevarious ways to verify that assumption 2 is satisfied, i.e. various sets of sufficientconditions for assumption 2 to hold. Assumption 2 holds if

1. F is bounded and β ∈ (0, 1). Note that boundedness of F is not enough.

Suppose β = 1 and F (xt, xt+1) =

½1 if t even−1 if t odd Obviously F is bounded,

but sincePnt=0 β

tF (xt, xt+1) =

½1 if n even0 if n odd

, the limit in assumption 2

does not exist. If β ∈ (0, 1) thenPnt=0 β

tF (xt, xt+1) =

½1− β

n2 + βn if n even

1− βn2 if n odd

and therefore limn→∞Pnt=0 β

tF (xt, xt+1) exists and equals 1. In generalthe joint assumption that F is bounded and β ∈ (0, 1) implies that thesequence yn =

Pnt=0 β

tF (xt, xt+1) is Cauchy and hence converges. In thiscase lim yn = y is obviously finite.

2. Define F+(x, y) = max0, F (x, y) and F−(x, y) = max0,−F (x, y).

Page 81: Macroeconomics Phd Lectures notes

5.1. THE PRINCIPLE OF OPTIMALITY 75

Then assumption 2 is satisfied if for all x0 ∈ X, all x ∈ Π(x0), either

limn→∞

nXt=0

βtF+(xt, xt+1) < +∞ or

limn→∞

nXt=0

βtF−(xt, xt+1) < +∞

or both. For example, if β ∈ (0, 1) and F is bounded above, then the firstcondition is satisfied, if β ∈ (0, 1) and F is bounded below then the secondcondition is satisfied.

3. Assumption 2 is satisfied if for every x0 ∈ X and every x ∈ Π(x0) thereare numbers (possibly dependent on x0, x) θ ∈ (0, 1β ) and 0 < c < +∞such that for all t

F (xt, xt+1) ≤ cθt

Hence F need not be bounded in any direction for assumption 2 to besatisfied. As long as the returns from the sequences do not grow too fast(at rate higher than 1

β ) we are still fine .

I would conclude that assumption 2 is rather weak (I can’t think of anyinteresting economic example where assumption1 is violated, but let me knowif you come up with one). A final piece of notation and we are ready to statesome theorems.Define the sequence of functions un : Π(x0)→ R by

un(x) =nXt=0

βtF (xt, xt+1)

For each feasible plan un gives the total discounted return (utility) up untilperiod n. If assumption 2 is satisfied, then the function u : Π(x0)→ R

u(x) = limn→∞

nXt=0

βtF (xt, xt+1)

is also well-defined, since under assumption 2 the limit exists. The range ofu is R, the extended real line, i.e. R = R ∪ −∞,+∞ since we allowed thelimit to be plus or minus infinity. From the definition of u it follows that underassumption 2

w(x0) = supx∈Π(x0)

u(x)

Note that by construction, whenever w exists, it is unique (since the supremumof a set is always unique). Also note that the way I have defined w above onlymakes sense under assumption 1. and 2., otherwise w is not well-defined.We have the following theorem, stating the principle of optimality.

Page 82: Macroeconomics Phd Lectures notes

76 CHAPTER 5. DYNAMIC PROGRAMMING

Theorem 40 Suppose (X,Γ, F,β) satisfy assumptions 1. and 2. Then

1. the function w satisfies the functional equation (FE)

2. if for all x0 ∈ X and all x ∈ Π(x0) a solution v to the functional equation(FE) satisfies

limn→∞βnv(xn) = 0 (5.1)

then v = w

I will skip the proof, but try to provide some intuition. The first resultstates that the supremum function from the sequential problem (which is well-defined under assumption 1. and 2.) solves the functional equation. This result,although nice, is not particularly useful for us. We are interested in solving thesequential problem and in the last section we made progress in solving thefunctional equation (not the other way around).

But result 2. is really key. It states a condition under which a solutionto the functional equation (which we know how to compute) is a solution tothe sequential problem (the solution of which we desire). Note that the func-tional equation (FE) may (or may not) have several solution. We haven’t madeenough assumptions to use the CMT to argue uniqueness. However, only oneof these potential several solutions can satisfy (5.1) since if it does, the theo-rem tells us that it has to equal the supremum function w (which is necessarilyunique). The condition (5.1) is somewhat hard to interpret (and SLP don’teven try), but think about the following. We saw in the first lecture that forinfinite-dimensional optimization problems like the one in (SP ) a transversalitycondition was often necessary and (even more often) sufficient (jointly with theEuler equation). The transversality condition rules out as suboptimal plans thatpostpone too much utility into the distant future. There is no equivalent condi-tion for the recursive formulation (as this formulation is basically a two periodformulation, today vs. everything from tomorrow onwards). Condition (5.1)basically requires that the continuation utility from date n onwards, discountedto period 0, should vanish in the time limit. In other words, this puts an upperlimit on the growth rate of continuation utility, which seems to substitute forthe TVC. It is not clear to me how to make this intuition more rigorous, though.

A simple, but quite famous example, shows that the condition (5.1) hassome bite. Consider the following consumption problem of an infinitely livedhousehold. The household has initial wealth x0 ∈ X = R. He can borrow orlend at a gross interest rate 1 + r = 1

β > 1. So the price of a bond that pays offone unit of consumption is q = β. There are no borrowing constraints, so thesequential budget constraint is

ct + βxt+1 ≤ xtand the nonnegativity constraint on consumption, ct ≥ 0. The household values

Page 83: Macroeconomics Phd Lectures notes

5.1. THE PRINCIPLE OF OPTIMALITY 77

discounted consumption, so that her maximization problem is

w(x0) = sup(ct,xt+1)∞t=0

∞Xt=0

βtct

s.t. 0 ≤ ct ≤ xt − βxt+1

x0 given

Since there are no borrowing constraint, the consumer can assure herself infiniteutility by just borrowing an infinite amount in period 0 and then rolling over thedebt by even borrowing more in the future. Such a strategy is called a Ponzi-scheme -see the hand-out. Hence the supremum function equals w(x0) = +∞for all x0 ∈ X. Now consider the recursive formulation (we denote by x currentperiod wealth xt, by y next period’s wealth and substitute out for consumptionct = xt − βxt+1 (which is OK given monotonicity of preferences)

v(x) = supy≤ x

β

x− βy + βv(y)

Obviously the function w(x) = +∞ satisfies this functional equation (just plugin w on the right side, since for all x it is optimal to let y tend to −∞ and hencev(x) = +∞. This should be the case from the first part of the previous theorem.But the function v(x) = x satisfies the functional equation, too. Using it on theright hand side gives, for an arbitrary x ∈ X

supy≤ x

β

x− βy + βy = supy≤ x

β

x = x = v(x)

Note, however that the second part of the preceding theorem does not applyfor v since the sequence xn defined by xn = x0

βn is a feasible plan from x0 > 0and

limn→∞βnv(xn) = lim

n→∞βnxn = x0 > 0

Note however that the second part of the theorem gives only a sufficient con-dition for a solution v to the functional equation being equal to the supremumfunction from (SP ), but not a necessary condition. Also w itself does not satisfythe condition, but is evidently equal to the supremum function. So wheneverwe can use the CMT (or something equivalent) we have to be aware of the factthat there may be several solutions to the functional equation, but at most onethe several is the function that we look for.Now we want to establish a similar equivalence between the sequential prob-

lem and the recursive problem with respect to the optimal policies/plans. Thefirst observation. Solving the functional equation gives us optimal policiesy = g(x) (note that g need not be a function, but could be a correspondence).Such an optimal policy induces a feasible plan xt+1∞t=0 in the following fash-ion: x0 = x0 is an initial condition, x1 ∈ g(x0) and recursively xt+1 = g(xt).The basic question is how a plan constructed from a solution to the functionalequation relates to a plan that solves the sequential problem. We have thefollowing theorem.

Page 84: Macroeconomics Phd Lectures notes

78 CHAPTER 5. DYNAMIC PROGRAMMING

Theorem 41 Suppose (X,Γ, F,β) satisfy assumptions 1. and 2.

1. Let x ∈ Π(x0) be a feasible plan that attains the supremum in the sequentialproblem. Then for all t ≥ 0

w(xt) = F (xt, xt+1) + βw(xt+1)

2. Let x ∈ Π(x0) be a feasible plan satisfying, for all t ≥ 0

w(xt) = F (xt, xt+1) + βw(xt+1)

and additionally1

limt→∞ supβ

tw(xt) ≤ 0 (5.2)

Then x attains the supremum in (SP ) for the initial condition x0.

What does this result say? The first part says that any optimal plan in thesequence problem, together with the supremum function w as value functionsatisfies the functional equation for all t. Loosely it says that any optimal planfrom the sequential problem is an optimal policy for the recursive problem (oncethe value function is the right one).Again the second part is more important. It says that, for the “right”

fixed point of the functional equation w the corresponding policy g generatesa plan x that solves the sequential problem if it satisfies the additional limitcondition. Again we can give this condition a loose interpretation as standingin for a transversality condition. Note that for any plan xt generated from apolicy g associated with a value function v that satisfies (5.1) condition (5.2) isautomatically satisfied. From (5.1) we have

limt→∞βtv(xt) = 0

for any feasible xt ∈ Π(x0), all x0. Also from Theorem 32 v = w. So for anyplan xt generated from a policy g associated with v = w we have

w(xt) = F (xt, xt+1) + βw(xt+1)

and since limt→∞ βtv(xt) exists and equals to 0 (since v satisfies (5.1)), we have

lim supt→∞

βtv(xt) = 0

and hence (5.2) is satisfied. But Theorem 33.2 is obviously not redundant asthere may be situations in which Theorem 32.2 does not apply but 33.2 does.

1The limit superior of a bounded sequence xn is the infimum of the set V of real numbersv such that only a finite number of elements of the sequence strictly exceed v. Hence it is thelargest cluster point of the sequence xn.

Page 85: Macroeconomics Phd Lectures notes

5.1. THE PRINCIPLE OF OPTIMALITY 79

Let us look at the following example, a simple modification of the saving problemfrom before. Now however we impose a borrowing constraint of zero.

w(x0) = maxxt+1∞t=0

∞Xt=0

βt(xt − βxt+1)

s.t. 0 ≤ xt+1 ≤ xtβ

x0 given

Writing out the objective function yields

w0(x0) = (x0 − βx1) + (x1 − βx2) + . . .

= x0

Now consider the associated functional equation

v(x) = max0≤x0≤ x

β

x− βx0 + v(x0)

Obviously one solution of this functional equation is v(x) = x and by Theorem32.1 is rightly follows that w satisfies the functional equation. However, forv condition (5.1) fails, as the feasible plan defined by xt =

x0βtshows. Hence

Theorem 32.2 does not apply and we can’t conclude that v = w (although wehave verified it directly, there may be other examples for which this is not sostraightforward). Still we can apply Theorem 33.2 to conclude that certain plansare optimal plans. Let xt be defined by x0 = x0, xt = 0 all t > 0. Then

lim supt→∞

βtw(xt) = 0

and we can conclude by Theorem 33.2 that this plan is optimal for the sequentialproblem. There are tons of other plans for which we can apply the same logic toshop that they are optimal, too (which shows that we obviously can’t make anyclaim about uniqueness). To show that condition (5.2) has some bite considerthe plan defined by xt =

x0βt. Obviously this is a feasible plan satisfying

w(xt) = F (xt, xt+1) + βw(xt+1)

but since for all x0 > 0

lim supt→∞

βtw(xt) = x0 > 0

Theorem 33.2 does not apply and we can’t conclude that xt is optimal (as infact this plan is not optimal).So basically we have a prescription what to do once we solved our functional

equation: pick the right fixed point (if there are more than one, check the limitcondition to find the right one, if possible) and then construct a plan from the

Page 86: Macroeconomics Phd Lectures notes

80 CHAPTER 5. DYNAMIC PROGRAMMING

policy corresponding to this fixed point. Check the limit condition to make surethat the plan so constructed is indeed optimal for the sequential problem. Done.Note, however, that so far we don’t know anything about the number (unless

the CMT applies) and the shape of fixed point to the functional equation. Thisis not quite surprising given that we have put almost no structure onto oureconomy. By making further assumptions one obtains sharper characterizationsof the fixed point(s) of the functional equation and thus, in the light of thepreceding theorems, about the solution of the sequential problem.

5.2 Dynamic Programming with Bounded Re-turns

Again we look at a functional equation of the form

v(x) = maxy∈Γ(x)

F (x, y) + βv(y)

We will now assume that F : X ×X is bounded and β ∈ (0, 1). We will makethe following two assumptions throughout this sectionAssumption 3: X is a convex subset of RL and the correspondence Γ :

X ⇒ X is nonempty, compact-valued and continuous.Assumption 4: The function F : A→ R is continuous and bounded, and

β ∈ (0, 1)We immediately get that assumptions 1. and 2. are satisfied and hence

the theorems of the previous section apply. Define the policy correspondenceconnected to any solution to the functional equation as

G(x) = y ∈ Γ(x) : v(x) = F (x, y) = βv(y)and the operator T on C(X)

(Tv) (x) = maxy∈Γ(x)

F (x, y) + βv(y)

Here C(X) is the space of bounded continuous functions on X and we use thesup-metric as metric. Then we have the following

Theorem 42 Under Assumptions 3. and 4. the operator T maps C(X) intoitself. T has a unique fixed point v and for all v0 ∈ C(X)

d(Tnv0, v) ≤ βnd(v0, v)

The policy correspondence G belonging to v is compact-valued and upper-hemicontinuous

Now we add further assumptions on the structure of the return function F,with the result that we can characterize the unique fixed point of T better.Assumption 5: For fixed y, F (., y) is strictly increasing in each of its L

components.

Page 87: Macroeconomics Phd Lectures notes

5.2. DYNAMIC PROGRAMMING WITH BOUNDED RETURNS 81

Assumption 6: Γ is monotone in the sense that x ≤ x0 implies Γ(x) ⊆Γ(x0).The result we get out of these assumptions is strict monotonicity of the value

function.

Theorem 43 Under Assumptions 3. to 6. the unique fixed point v of T isstrictly increasing.

We have a similar result in spirit if we make assumptions about the curvatureof the return function and the convexity of the constraint set.Assumption 7: F is strictly concave, i.e. for all (x, y), (x0, y0) ∈ A and

θ ∈ (0, 1)F [θ(x, y) + (1− θ)(x0, y0)] ≥ θF (x, y) + (1− θ)F (x0, y0)

and the inequality is strict if x 6= x0Assumption 8: Γ is convex in the sense that for θ ∈ [0, 1] and x, x0 ∈ X,

the fact y ∈ Γ(x), y0 ∈ Γ(x0)θy + (1− θ)y0 ∈ Γ(θx+ (1− θ)x0)

Again we find that the properties assumed about F extend to the value function.

Theorem 44 Under Assumptions 3.-4. and 7.-8. the unique fixed point of vis strictly concave and the optimal policy is a single-valued continuous function,call it g.

Finally we state a result about the differentiability of the value function,the famous envelope theorem (some people call it the Benveniste-Scheinkmantheorem).2

Assumption 9: F is continuously differentiable on the interior of A.

2You may have seen the envelope theorem stated a bit differently by Prof. Sargent. Hesets up the recursive problem as

V (x) = maxuF (x, u) + βV (y)

s.t. y = g(x, u)

Substituiting for y we get

V (x) = maxuF (x, u) + βV (g(x, u))

The difference between his formulation and mine is that inhis formulation a current periodcontrol variable u is chosen, which, jointly with today’s state x determines next period’s statey. In my formulation we substituted out the control u and chose next period’s state y. Thisyields a different statement of the enveope theorem. Let us briefly derive Prof. Sargent’sstatement.The first order condition (always assuming interiority) is

∂F (x, u)

∂u+ βV 0(g(x, u)

∂g(x, u)

∂u= 0

Let the solution to the FOC be denoted by u = h(x), i.e. h satisfies for every x

∂F (x, h(x))

∂u+ βV 0(g(x, h(x))

∂g(x, h(x))

∂u= 0

Page 88: Macroeconomics Phd Lectures notes

82 CHAPTER 5. DYNAMIC PROGRAMMING

Theorem 45 Under assumptions 3.-4. and 7.-9. if x0 ∈ int(X) and g(x0) ∈int(Γ(x0)), then the unique fixed point of T, v is continuously differentiable atx0 with

∂v(x0)

∂xi=

∂F (x0, g(x0))

∂xi

This theorem gives us an easy way to derive Euler equations from the re-cursive formulation of the neoclassical growth model. Remember the functionalequation

v(k) = max0≤k0≤f(k)

U(f(k)− k0) + βv(k0)

Taking first order conditions with respect to k0 (and ignoring corner solutions)we get

U 0(f(k)− k0) = βv0(k0)

Denote by k0 = g(k) the optimal policy. The problem is that we don’t know v0.But now we can use Benveniste-Scheinkman to obtain

v0(k) = U 0(f(k)− g(k))f 0(k)

Using this in the first order condition we obtain

U 0(f(k)− g(k)) = βv0(k) = βU 0(f(k0)− g(k0))f 0(k0)= βf 0(g(k))U 0(f(g(k))− g(g(k))

Denoting k = kt, g(k) = kt+1 and g(g(k)) = kt+2 we obtain our usual Eulerequation

U 0(f(kt)− kt+1) = βf(kt+1)U0(f(kt+1)− kt+2)

Now we differentiate the value function to obtain

V 0(x) =∂F (x, h(x))

∂x+

∂F (x, h(x))

∂uh0(x)

+βV 0(g(x, h(x)))·∂g(x, h(x))

∂x+

∂g(x, h(x))

∂uh0(x)

¸=

∂F (x, h(x))

∂x+ βV 0(g(x, h(x)))

∂g(x, h(x))

∂x

+h0(x)·∂F (x, h(x))

∂u+ βV 0(g(x, h(x)))

∂g(x, h(x))

∂u

¸Using the first order condtions yields the envelope theorem for Prof. Sargent’s setup of theproblem.

V 0(x) =∂F (x, h(x))

∂x+ βV 0(g(x, h(x)))

∂g(x, h(x))

∂x

Page 89: Macroeconomics Phd Lectures notes

Chapter 6

Models with Uncertainty

In this section we will introduce a basic model with uncertainty, in order toestablish some notation and extend our discussion of efficient economy to thisimportant case. Then, as a fist application, we will look at the stochastic neo-classical growth model, which forms the basis for a particular theory of businesscycles, the so called “Real Business Cycle” (RBC) theory. In this section we willbe a bit loose with our treatment of uncertainty, in that we will not explicitlydiscuss probability spaces that form the formal basis of our representation ofuncertainty.

6.1 Basic Representation of Uncertainty

The basic novelty of models with uncertainty is the formal representation of thisuncertainty and the ensuing description of the information structure that agentshave. We start with the notion of an event st ∈ S. The set S = η1, , . . . , ηNof possible events that can happen in period t is assumed to be finite and thesame for all periods t. If there is no room for confusion we use the notationst = 1 instead of st = η1 and so forth. For example S may consist of all weatherconditions than can happen in the economy, with st = 1 indicating sunshine inperiod t, st = 2 indicating cloudy skies, st = 3 indicating rain and so forth.

1 Asanother example, consider the economy from Section 2, but now with randomendowments. In each period one of the two agents has endowment 0 and theother has endowment 2, but who has what is random, with st = 1 indicatingthat agent 1 has high endowment and st = 2 indicating that agent 2 has highendowment at period t. The set of possible events is given by S = 1, 2An event history st = (s0, s1, . . . st) is a vector of length t+ 1 summarizing

the realizations of all events up to period t. Formally, with St = S×S× . . .×Sdenoting the t+ 1-fold product of S, event history st ∈ St lies in the set of all

1Technically speaking st is a random variable with respect to some underlying probabilityspace (Ω,A, P ), where Ω is some set of basis events with generic element ω, A is a sigmaalgebra on Ω and P is a probability measure.

83

Page 90: Macroeconomics Phd Lectures notes

84 CHAPTER 6. MODELS WITH UNCERTAINTY

possible event histories of length t.

By π(st) let denote the probability of a particular event history. We assumethat π(st) > 0 for all st ∈ St, for all t. For our example economy, if s2 = (1, 1, 2)then π(s2) is the probability that agent 1 has high endowment in period t = 0and t = 1 and agent 2 has high endowment in period 2. Figure 5 summarizes theconcepts introduced so far, for the case in which S = 1, 2 is the set of possibleevents that can happen in every period. Note that the sets St of possible eventsof length t become fairly big very rapidly, which poses computational problemswhen dealing with models with uncertainty.

t=0 t=1 t=2 t=3

0s =1

0s =2

1s =(1,1)

1s =(1,2)

1s =(2,1)

1s =(2,2)

2s =(1,1,1)

2s =(1,1,2)

2s =(1,2,1)

2s =(1,2,2)

2s =(2,1,1)

2s =(2,1,2)

2s =(2,2,1)

2s =(2,2,2)

3s =(1,1,2,1)

3s =(1,1,2,2)

0π(s =2)

0π(s =1)

1π(s =(2,2))

2π(s =(2,2,2))

3π(s =(1,12,2))

Figure 6.1:

All commodities of our economy, instead of being indexed by time t as before,now also have to be indexed by event histories st. In particular, an allocationfor the example economy of Section 2, but now with uncertainty, is given by

(c1, c2) = c1t (st), c2t (st)∞t=0,st∈St

Page 91: Macroeconomics Phd Lectures notes

6.2. DEFINITIONS OF EQUILIBRIUM 85

with the interpretation that cit(st) is consumption of agent i in period t if event

history st has occurred. Note that consumption in period t of agents are allowedto (and in general will) vary with the history of events that have occurred inthe past .Now we are ready to specify to remaining elements of the economy. With

respect to endowments, these also take the general form

(e1, e2) = e1t (st), e2t (st)∞t=0,st∈Stand for the particular example

e1t (st) =

½2 if st = 10 if st = 2

e2t (st) =

½0 if st = 12 if st = 2

i.e. for the particular example endowments in period t only depend on therealization of the event st, not on the entire history. Nothing, however, wouldprevent us from specifying more general endowment patterns.Now we specify preferences. We assume that households maximize expected

lifetime utility where expectations E0 is the expectation operator at period0, prior to any realization of uncertainty (in particular the uncertainty withrespect to s0). Given our notation just established, assuming that preferencesadmit a von-Neumann Morgenstern utility function representation2 we representhouseholds’ preferences by

u(ci) =∞Xt=0

Xst∈St

βtπ(st)U(cit(st))

This completes our description of the simple example economy.

6.2 Definitions of Equilibrium

Again there are two possible market structures that we can work with. TheArrow-Debreu market structure turns out to be easier than the sequential mar-kets market structure, so we will start with it. Again there is an equivalencetheorem between these two economies, once we allow the asset market structurefor the sequential markets market structure to be rich enough.

6.2.1 Arrow-Debreu Market Structure

As usual with Arrow-Debreu, trade takes place at period 0, before any uncer-tainty has been realized (in particular, before s0 has been realized). As withallocations, Arrow-Debreu prices have to be indexed by event histories in addi-tion to time, so let pt(s

t) denote the price of one unit of consumption, quoted

2Felix Kubler will discuss in great length what is required for this.

Page 92: Macroeconomics Phd Lectures notes

86 CHAPTER 6. MODELS WITH UNCERTAINTY

at period 0, delivered at period t if (and only if) event history st has been re-alized. Given this notation, the definition of an AD-equilibrium is identical tothe case without uncertainty, with the exeption that, since goods and prices arenot only indexed by time, but also by histories, we have to sum over both timeand histories in the individual households’ budget constraint.

Definition 46 A (competitive) Arrow-Debreu equilibrium are prices pt(st)∞t=0,st∈Stand allocations (cit(st)∞t=0,st∈St)i=1,2 such that

1. Given pt(st)∞t=0,st∈St , for i = 1, 2, cit(st)∞t=0,st∈St solves

maxcit(st)∞t=0,st∈St

∞Xt=0

Xst∈St

βtπ(st)U(cit(st))(6.1)

s.t.∞Xt=0

Xst∈St

pt(st)cit(s

t) ≤∞Xt=0

Xst∈St

pt(st)eit(s

t) (6.2)

cit(st) ≥ 0 for all t (6.3)

2.

c1t (st) + c2t (s

t) = e1t (st) + e2t (s

t) for all t, all st ∈ St (6.4)

Note that there is again only one budget constraint, and that market clearinghas to hold date, by date, event history by event history. Also note that,when computing equilibria, one can normalize the price of only one commodityto 1, and consumption at the same date, but for different event histories aredifferent commodities. That means that if we normalize p0(s0 = 1) = 1 wecan’t also normalize p0(s0 = 2) = 1. Finally, there are no probabilities in thebudget constraint. Equilibrium prices will reflect the probabilities of differentevent histories, but there is no room for these probabilities in the Arrow-Debreubudget constraint directly.[Characterization of equilibrium allocations: write down individuals prob-

lem, first order condition for both agents, take ratio and argue that this impliesconsumption to be constant over time for both agents; from this show thatprices are proportional to βtπ(st)]The definition of Pareto efficiency is identical to that of the certainty case;

the first welfare theorem goes through without any changes (in particular, theproof is identical, apart from changes in notation). We state both for complete-ness

Definition 47 An allocation (c1t (st), c2t (st))∞t=0,st∈St is feasible if1.

cit(st) ≥ 0 for all t, all st ∈ St, for i = 1, 2

Page 93: Macroeconomics Phd Lectures notes

6.2. DEFINITIONS OF EQUILIBRIUM 87

2.

c1t (st) + c2t (s

t) = e1t (st) + e2t (s

t) for all t, all st ∈ St

Definition 48 An allocation (c1t (st), c2t (st))∞t=0,st∈St is Pareto efficient if itis feasible and if there is no other feasible allocation (c1t (st), c2t (st))∞t=0,st∈Stsuch that

u(ci) ≥ u(ci) for both i = 1, 2

u(ci) > u(ci) for at least one i = 1, 2

Proposition 49 Let (cit(st)∞t=0,st∈St)i=1,2 be a competitive equilibrium allo-

cation. Then (cit(st)∞t=0,st∈St)i=1,2 is Pareto efficient.

[Characterization of Pareto efficient allocations: to be added: 1. write downSocial Planners problem, FOC’s, show that any efficient allocation has to featureconstant consumption for both agents (since aggregate endowment is constant)]

6.2.2 Sequential Markets Market Structure

Now let trade take place sequentially in each period (more precisely, in eachperiod, event-history pair). With certainty, we allowed trade in consumptionand in one-period IOU’s. For the equivalence between Arrow-Debreu and se-quential markets with uncertainty, this is not enough. We introduce one periodcontingent IOU’s, financial contracts bought in period t, that pay out one unitof the consumption good in t + 1 only for a particular realization of st+1 = jtomorrow.3 So let qt(s

t, st+1 = j) denote the price at period t of a contract thatpays out one unit of consumption in period t+1 if (and only if) tomorrow’s eventis st+1 = j. These contracts are often called Arrow securities, contingent claimsor one-period insurance contracts. Let ait+1(s

t, st+1) denote the quantities ofthese Arrow securities bought (or sold) at period t by agent i.The period t, event history st budget constraint of agent i is given by

cit(st) +

Xst+1∈S

qt(st, st+1)a

it+1(s

t, st+1) ≤ eit(st) + ait(st)

Note that agents purchase Arrow securities ait+1(st, st+1)st+1∈S for all contin-gencies st+1 ∈ S that can happen tomorrow, but that, once st+1 is realized,only the ait+1(s

t+1) corresponding to the particular realization of st+1 becomeshis asset position with which he starts the current period. We assume thatai0(s0) = 0 for all s0 ∈ S.We then have the following

3A full set of one-period Arrow securities is sufficient to make markets “sequentially com-plete”, in the sense that any (nonnegative) consumption allocation is attainable with an appro-priate sequence of Arrow security holdings at+1(st, st+1) satisfying all sequential marketsbudget constraints.

Page 94: Macroeconomics Phd Lectures notes

88 CHAPTER 6. MODELS WITH UNCERTAINTY

Definition 50 A SM equilibrium is allocations ³cit(s

t),©ait+1(s

t, st+1)ªst+1∈S

´i=1,2

∞t=0,st∈St ,and prices for Arrow securities qt(st, st+1)∞t=0,st∈St,st+1∈S such that

1. For i = 1, 2, given qt(st, st+1)∞t=0,st∈St,st+1∈S , for all i, cit(st),©ait+1(s

t, st+1)ªst+1∈S∞t=0,st∈St

solves

maxcit(st),ait+1(st,st+1)st+1∈S∞t=0,st∈St

u(ci)

s.t

cit(st) +

Xst+1∈S

qt(st, st+1)a

it+1(s

t, st+1) ≤ eit(st) + ait(s

t)

cit(st) ≥ 0 for all t, st ∈ St

ait+1(st, st+1) ≥ −Ai for all t, st ∈ St

2. For all t ≥ 02Xi=1

cit(st) =

2Xi=1

eit(st) for all t, st ∈ St

2Xi=1

ait+1(st, st+1) = 0 for all t, st ∈ St and all st+1 ∈ S

Note that we have a market clearing condition in the asset market for eachArrow security being traded for period t+ 1. Define

qt(st) =

Xst+1∈S

qt(st, st+1)

The price qt(st) can be interpreted as the price, in period t, event history st,

for buying one unit of consumption delivered for sure in period t + 1 (we buyone unit of consumption for each contingency tomorrow). The risk free interestrate (the counterpart to the interest rate for economies without uncertainty)between periods t and t+ 1 is then given by

1

1 + rt+1(st)= qt(s

t)

6.2.3 Equivalence between Market Structures

[To Be Completed]

6.3 Markov Processes

So far we haven’t specified the exact nature of uncertainty. In particular, inno sense have we assumed that the random variables st and sτ , τ > t are

Page 95: Macroeconomics Phd Lectures notes

6.3. MARKOV PROCESSES 89

independent or dependent in a simple way. Our theory is completely generalalong this dimension; to make it implementable (analytically or numerically),however, one has to assume a particular structure of the uncertainty.In particular, it simplifies matters a lot if one assumes that the st’s follow a

discrete time (time is discrete), discrete state (the number of values st can takeis finite) time homogeneous Markov chain. Let by

π(j|i) = prob(st+1 = j|st = i)denote the conditional probability that the state in t + 1 equals j ∈ S if thestate in period t equals st = i ∈ S. Time homogeneity means that π is notindexed by time. Given that st+1 ∈ S and st ∈ S and S is a finite set, π(.|.) isan N ×N -matrix of the form

π =

π11 π12 · · · ... · · · π1N

π21...

......

......

πi1 · · · · · · πij · · · πiN...

......

πN1 · · · · · · ... · · · πNN

with generic element πij = π(j|i) =prob(st+1 = j|st = i). Hence the i-th rowgives the probabilities of going from state i today to all the possible states to-morrow, and the j-th column gives the probability of landing in state j tomorrowconditional of being in an arbitrary state i today. Since πij ≥ 0 and

Pj πij = 1

for all i (for all states today, one has to go somewhere for tomorrow), the matrixπ is a so-called stochastic matrix.Suppose the probability distribution over states today is given by the N -

dimensional column vector Pt = (p1t , . . . , p

Nt )

T and uncertainty is described bya Markov chain of the from above. Note that

Pi pit = 1. Then the probability

of being in state j tomorrow is given by

pjt+1 =Xi

πijpit

i.e. by the sum of the conditional probabilities of going to state j from state i,weighted by the probabilities of starting out in state i. More compactly we canwrite

Pt+1 = πTPt

A stationary distribution Π of the Markov chain π satisfies

Π = πTΠ

i.e. if you start today with a distribution over states Π then tomorrow youend up with the same distribution over states Π. From the theory of stochastic

Page 96: Macroeconomics Phd Lectures notes

90 CHAPTER 6. MODELS WITH UNCERTAINTY

matrices we know that every π has at least one such stationary distribution.It is the eigenvector (normalized to length 1) associated with the eigenvalueλ = 1 of πT . Note that every stochastic matrix has (at least) one eigenvectorequal to 1. If there is only one such eigenvalue, then there is a unique stationarydistribution, if there are multiple eigenvalues of length 1, then there a multiplestationary distributions (in fact a continuum of them).Note that the Markov assumption restricts the conditional probability dis-

tribution of st+1 to depend only on the realization of st, but not on realizationsof st−1, st−2 and so forth. This obviously is a severe restriction on the possiblerandomness that we allow, but it also means that the nature of uncertainty forperiod t + 1 is completely described by the realization of st, which is crucialwhen formulating these economies recursively. We have to start the Markovprocess out at period 0, so let by Π(s0) denote the probability that the statein period 0 is s0. Given our Markov assumption the probability of a particularevent history can be written as

π(st+1) = π(st+1|st) ∗ π(st|st−1) . . . ∗ π(s1|s0) ∗Π(s0)

[Some examples: π =

µ0.7 0.30.2 0.8

¶and π =

µ1 00 1

¶, show what can

happen, see notes.]

6.4 Stochastic Neoclassical Growth Model

In this section we will briefly consider a stochastic extension to the deterministicneoclassical growth model. You will have fun with this model in the thirdproblem set. The stochastic neoclassical growth model is the workhorse for halfof modern business cycle theory; everybody doing real business cycle theory usesit. I therefore think that it is useful to expose you to this model, even thoughyou may decide not to do RBC-theory in your own research.The economy is populated by a large number of identical households. For

convenience we normalize the number of households to 1. In each period threegoods are traded, labor services nt, capital services kt and the final output goodyt, which can be used for consumption ct or investment it.

1. Technology:

yt = eztF (kt, nt)

where zt is a technology shock. F is assumed to have the usual properties,i.e. has constant returns to scale, positive but declining marginal productsand satisfies the INADA conditions. We assume that the technology shockhas unconditional mean 0 and follows a N -state Markov chain. Let Z =z1, z2, . . . zN be the state space of the Markov chain, i.e. the set of valuesthat zt can take on. Let π = (πij) denote the Markov transition matrixand Π the stationary distribution of the chain (ignore the fact that in

Page 97: Macroeconomics Phd Lectures notes

6.4. STOCHASTIC NEOCLASSICAL GROWTH MODEL 91

some of our applications Π will not be unique). Let π(z0|z) = prob(zt+1 =z0|zt = z). In most of the applications we will take N = 2. The evolutionof the capital stock is given by

kt+1 = (1− δ)kt + it

and the composition of output is given by

yt = ct + it

Note that the set Z takes the role of S in our general formulation ofuncertainty, zt corresponds to st and so forth.

2. Preferences:

E0

∞Xt=0

βtu(ct) with β ∈ (0, 1)

The period utility function is assumed to have the usual properties.

3. Endowment: each household has an initial endowment of capital, k0 andone unit of time in each period. Endowments are not stochastic.

4. Information: The variable zt, the only source of uncertainty in this model,is publicly observable. We assume that in period 0 z0 has not been realized,but is drawn from the stationary distribution Π. All agents are perfectlyinformed that the technology shock follows the Markov chain π with initialdistribution Π.

A lot of the things that we did for the case without uncertainty go throughalmost unchanged for the stochastic model. The only key difference is thatnow commodities have to be indexed not only by time, but also by histories ofproductivity shocks, since goods delivered at different nodes of the event tree aredifferent commodities, even though they have the same physical characteristics.For a lucid discussion of this point see Chapter 7 of Debreu’s (1959) “Theory ofValue”.For the recursive formulation of the social planners problem, note that the

current state of the economy now not only includes the capital stock k that theplanner brings into the current period, but also the current state of the technol-ogy z. This is due to the fact that current production depends on the currenttechnology shock, but also due to the fact that the probability distribution offuture shocks π(z0|z) depends on the current shock, due to the Markov struc-ture of the stochastic shocks. Also note that even if the social planner choosescapital stock k0 for tomorrow today, lifetime utility from tomorrow onwards isuncertain, due to the uncertainty of z0. These considerations, plus the usualobservation that nt = 1 is optimal, give rise to the following Bellman equation

v(k, z) = max0≤k0≤ezF (k,1)+(1−δ)k

(U(ezF (k, 1) + (1− δ)k − k0) + β

Xz0

π(z0|z)v(k0, z0)).

Page 98: Macroeconomics Phd Lectures notes

92 CHAPTER 6. MODELS WITH UNCERTAINTY

[Discussion of Calibration, see notes and Chapter 1 of Cooley]

Page 99: Macroeconomics Phd Lectures notes

Chapter 7

The Two Welfare Theorems

In this section we will present the two fundamental theorems of welfare eco-nomics for economies in which the commodity space is a general (real) vectorspace, which is not necessarily finite dimensional. Since in macroeconomics weoften deal with agents or economies that live forever, usually a finite dimen-sional commodity space is not sufficient for our analysis. The significance of thewelfare theorems, apart from providing a normative justification for studyingcompetitive equilibria is that planning problems characterizing Pareto optimaare usually easier to solve that equilibrium problems, the ultimate goal of ourtheorizing.Our discussion will follow Stokey et al. (1989), which in turn draws heavily

on results developed by Debreu (1954).

7.1 What is an Economy?

We first discuss how what an economy is in Arrow-Debreu language. An econ-omy E = ((Xi, ui)i∈I , (Yj)j∈J) consists of the following elements

1. A list of commodities, represented by the commodity space S.We requireS to be a normed (real) vector space with norm k.k.1

1For completeness we state the following definitions

Definition 51 A real vector space is a set S (whose elements are called vectors) on whichare defined two operations

• Addition + : S × S → S. For any x, y ∈ S, x+ y ∈ S.• Scalar Multiplication · : R×S → S. For any α ∈ R and any x ∈ S, αx ∈ S that satisfythe following algebraic properties: for all x, y ∈ S and all α,β ∈ R(a) x+ y = y + x

(b) (x+ y) + z = x+ (y + z)

(c) α · (x+ y) = α · x+ α · y(d) (α+ β) · x = α · x+ β · x(e) (αβ) · x = α · (β · x)

93

Page 100: Macroeconomics Phd Lectures notes

94 CHAPTER 7. THE TWO WELFARE THEOREMS

2. A finite set of people i ∈ I. Abusing notation I will by I denote both theset of people and the number of people in the economy.

3. Consumption sets Xi ⊆ S for all i ∈ I.We will incorporate the restrictionsthat households endowments place on the xi in the description of theconsumption sets Xi.

4. Preferences representable by utility functions ui : S → R.

5. A finite set of firms j ∈ J. The same remark about notation as aboveapplies.

6. Technology sets Yj ⊆ S for all j ∈ J. Let by

Y =Xj∈J

Yj =

y ∈ S : ∃(yj)j∈J such that y =Xj∈J

yj and yj ∈ Yj for all j ∈ J

denote the aggregate production set.

A private ownership economy E = ((Xi, ui)i∈I , (Yj)j∈J , (θij)i∈I,j∈J) consistsof all the elements of an economy and a specification of ownership of the firmsθij ≥ 0 with

Pi∈I θij = 1 for all j ∈ J. The entity θij is interpreted as the share

of ownership of household ι to firm j, i.e. the fraction of total profits of firm jthat household i is entitled to.With our formalization of the economy we can also make precise what we

mean by an externality. An economy is said to exhibit an externality if householdi’s consumption set Xi or firm j’s production set Yj is affected by the choice ofhousehold k’s consumption bundle xk or firm m’s production plan ym. Unlessotherwise stated we assume that we deal with an economy without externalities.

(f) There exists a null element θ ∈ S such thatx+ θ = x

0 · x = θ

(g) 1 · x = x

Definition 52 A normed vector space is a vector space is a vector space S together with anorm k.k : S → R such that for all x, y ∈ S and α ∈ R(a) kxk ≥ 0, with equality if and only if x = θ

(b) kα · xk = |α| kxk(c) kx+ yk ≤ kxk+ kyk

Note that in the first definition the adjective real refers to the fact that scalar multiplicationis done with respect to a real number. Also note the intimate relation between a norm and ametric defined above. A norm of a vector space S, k.k : S → R induces a metric d : S×S → Rby

d(x, y) = kx− yk

Page 101: Macroeconomics Phd Lectures notes

7.1. WHAT IS AN ECONOMY? 95

Definition 53 An allocation is a tuple [(xi)i∈I , (yj)j∈J ] ∈ SI×J .

In the economy people supply factors of production and demand final outputgoods. We follow Debreu and use the convention that negative componentsof the xi’s denote factor inputs and positive components denote final goods.Similarly negative components of the yj ’s denote factor inputs of firms andpositive components denote final output of firms.

Definition 54 An allocation [(xi)i∈I , (yj)j∈J ] ∈ SI×J is feasible if

1. xi ∈ Xi for all i ∈ I2. yj ∈ Yj for all j ∈ J3. (Resource Balance) X

i∈Ixi =

Xj∈J

yj

Note that we require resource balance to hold with equality, ruling out freedisposal. If we want to allow free disposal we will specify this directly as partof the description of technology.

Definition 55 An allocation [(xi)i∈I , (yj)j∈J ] is Pareto optimal if

1. it is feasible

2. there does not exist another feasible allocation [(x∗i )i∈I , (y∗j )j∈J ] such that

ui(x∗i ) ≥ ui(xi) for all i ∈ I

ui(x∗i ) > ui(xi) for at least one i ∈ I

Note that if I = J = 1 then2 for an allocation [x, y] resource balance requiresx = y, the allocation is feasible if x ∈ X∩Y, and the allocation is Pareto optimalif

x ∈ arg maxz∈X∩Y

u(z)

Also note that the definition of feasibility and Pareto optimality are identical foreconomies E and private ownership economies E. The difference comes in thedefinition of competitive equilibrium and there in particular in the formulationof the resource constraint. The discussion of competitive equilibrium requires

2The assumption that J = 1 is not at all restrictive if we restrict our attention to constantreturns to scale technologies. Then, in any competitive equilibrium profits are zero and thenumber of firms is indeterminate in equilibrium; without loss of generality we then can restrictattention to a single representative firm. If we furthermore restrict attention to identical peopleand type identical allocations, then de facto I = 1. Under which assumptions the restrictionto type identical allocations is justified will be discussed below.

Page 102: Macroeconomics Phd Lectures notes

96 CHAPTER 7. THE TWO WELFARE THEOREMS

a discussion of prices at which allocations are evaluated. Since we deal withpossibly infinite dimensional commodity spaces, prices in general cannot berepresented by a finite dimensional vector. To discuss prices for our generalenvironment we need a more general notion of a price system. This is necessaryin order to state and prove the welfare theorems for infinitely lived economiesthat we are interested in.

7.2 Dual Spaces

A price system attaches to every bundle of the commodity space S a real numberthat indicates how much this bundle costs. If the commodity space is a finite (sayk−) dimensional Euclidean space, then the natural thing to do is to represent aprice system by a k-dimensional vector p = (p1, . . . pk), where pl is the price ofthe l-th component of a commodity vector. The price of an entire point of thecommodity space is then φ(s) =

Pkl=1 slpl. Note that every p ∈ Rk represents

a function that maps S = Rk into R. Obviously, since for a given p and alls, s0 ∈ S and all α,β ∈ R

φ(αs+ βs0) =kXl=1

pl(αsl + βs0l) = αkXl=1

plsl + βkXl=1

pls0l = αφ(s) + βφ(s0)

the mapping associated with p is linear. We will take as a price system for anarbitrary commodity space S a continuous linear functional defined on S. Thenext definition makes the notion of a continuous linear functional precise.

Definition 56 A linear functional φ on a normed vector space S (with asso-ciated norm kkS) is a function φ : S → R that maps S into the reals andsatisfies

φ(αs+ βs0) = αφ(s) + βφ(s0) for all s, s0 ∈ S, all α,β ∈ RThe functional φ is continuous if ksn − skS → 0 implies |φ(sn)− φ(s)|→ 0 forall sn∞n=0 ∈ S, s ∈ S. The functional φ is bounded if there exists a constantM ∈ R such that |φ(s)| ≤M kskS for all s ∈ S. For a bounded linear functionalφ we define its norm by

kφkd = supkskS≤1

|φ(s)|

Fortunately it is rather easy to verify whether a linear functional is contin-uous and bounded. Stokey et al. state and prove a theorem that states that alinear functional is continuous if it is continuous at a particular point s ∈ S andthat it is bounded if (and only if) it is continuous. Hence a linear functional isbounded and continuous if it is continuous at a single point.For any normed vector space S the space

S∗ = φ : φ is a continuous linear functional on S

Page 103: Macroeconomics Phd Lectures notes

7.2. DUAL SPACES 97

is called the (algebraic) dual (or conjugate) space of S.With addition and scalarmultiplication defined in the standard way S∗ is a vector space, and with thenorm kkd defined above S∗ is a normed vector space as well. Note (you shouldprove this3) that even if S is not a complete space, S∗ is a complete space andhence a Banach space (a complete normed vector space). Let us consider severalexamples that will be of interest for our economic applications.

Example 57 For each p ∈ [1,∞) define the space lp by

lp = x = xt∞t=0 : xt ∈ R, for all t; kxkp =Ã ∞Xt=0

|xt|p! 1

p

<∞

with corresponding norm kxkp . For p =∞, the space l∞ is defined correspond-ingly, with norm kxk∞ = supt |xt|. For any p ∈ [1,∞) define the conjugateindex q by

1

p+1

q= 1

For p = 1 we define q =∞.We have the important result that for any p ∈ [1,∞)the dual of lp is lq. This result can be proved by using the following theorem(which in turn is proved by Luenberger (1969), p. 107.)

Theorem 58 Every continuous linear functional φ on lp, p ∈ [1,∞), is repre-sentable uniquely in the form

φ(x) =∞Xt=0

xtyt (7.1)

where y = yt ∈ lq. Furthermore, every element of lq defines an element of thedual of lp, l

∗p in this way, and we have

kφkd = kykq =((P∞t=0 |yt|q)

1q if 1 < p <∞

supt |yt| if p = 1

Let’s first understand what the theorem gives us. Take any space lp (notethat the theorem does NOT make any statements about l∞). Then the theoremstates that its dual is lq. The first part of the theorem states that lq ⊆ l∗p. Takeany element φ ∈ l∗p. Then there exists y ∈ lq such that φ is representable by y. Inthis sense φ ∈ lq. The second part states that any y ∈ lq defines a functional φ onlp by (7.1). Given its definition, φ is obviously continuous and hence bounded.Finally the theorem assures that the norm of the functional φ associated withy is indeed the norm associated with lq. Hence l

∗p ⊆ lq.

3After you are done with this, check Kolmogorov and Fomin (1970), p. 187 (Theorem 1)for their proof.

Page 104: Macroeconomics Phd Lectures notes

98 CHAPTER 7. THE TWO WELFARE THEOREMS

As a result of the theorem, whenever we deal with lp, p ∈ [1,∞) as commod-ity space we can restrict attention to price systems that can be represented by avector p = (p0, p1, . . . pt, . . . ) and hence have a straightforward economic inter-pretation: pt is the price of the good at period t and the cost of a consumptionbundle x is just the sum of the cost of all its components.For reasons that will become clearer later the most interesting commodity

space for infinitely lived economies, however, is l∞. And for this commodityspace the previous theorem does not make any statements. It would suggestthat the dual of l∞ is l1, but this is not quite correct, as the next result shows.

Proposition 59 The dual of l∞ contains l1. There are φ ∈ l∗∞ that are notrepresentable by an element y ∈ l1Proof. For the first part for any y ∈ l1 define φ : l∞ → R by

φ(x) =∞Xt=0

xtyt

We need to show that φ is linear and continuous. Linearity is obvious. Forcontinuity we need to show that for any sequence xn ∈ l∞ and x ∈ l∞,kxn − xk = supt |xnt − xt| → 0 implies |φ(xn) − φ(x)| → 0. Since y ∈ l1 thereexists M such that

P∞t=0 |yt| < M. Since supt |xnt − xt|→ 0, for all δ > 0 there

exists N(δ) such that fro all n > N(δ) we have supt |xnt − xt| < δ. But then forany ε > 0, taking δ(ε) = ε

2M and N(ε) = N(δ(ε)), for all n > N(ε)

|φ(xn)− φ(x)| =

¯¯∞Xt=0

xnt yt −∞Xt=0

xtyt

¯¯

≤∞Xt=0

|yt(xnt − xt)|

≤∞Xt=0

|yt| · |xnt − xt|

≤ Mδ(e) =ε

2< ε

The second part we prove via a counter example after we have proved thesecond welfare theorem.The second part of the proposition is somewhat discouraging in that it asserts

that, when dealing with l∞ as commodity space we may require a price systemthat does not have a natural economic interpretation. It is true that there isa subspace of l∞ for which l1 is its dual. Define the space c0 (with associatedsup-norm) as

c0 = x ∈ l∞ : limt→∞xt = 0

We can prove that l1 is the dual of c0. Since c0 ⊆ l∞ and l1 ⊆ l∗∞, obviouslyl1 ⊆ c∗0. It remains to show that any φ ∈ c∗0 can be represented by a y ∈ l1. [TOBE COMPLETED]

Page 105: Macroeconomics Phd Lectures notes

7.3. DEFINITION OF COMPETITIVE EQUILIBRIUM 99

7.3 Definition of Competitive Equilibrium

Corresponding to our two notions of an economy and a private ownership econ-omy we have two definitions of competitive equilibrium that differ in their spec-ification of the individual budget constraints.

Definition 60 A competitive equilibrium is an allocation [(x0i )i∈I , (y0j )j∈J ] and

a continuous linear functional φ : S → R such that

1. for all i ∈ I, x0i solves maxui(x) subject to x ∈ Xi and φ(x) ≤ φ(x0i )

2. for all j ∈ J, y0j solves maxφ(y) subject to y ∈ Yj3.P

i∈I x0i =

Pj∈J y

0j

In this definition we have obviously ignored ownership of firms. If, however,all Yj are convex cones, the technologies exhibit constant returns to scale, profitsare zero in equilibrium and this definition of equilibrium is equivalent to thedefinition of equilibrium for a private ownership economy (under appropriateassumptions on preferences such as local nonsatiation). Note that condition 1.is equivalent to requiring that for all i ∈ I, x ∈ Xi and φ(x) ≤ φ(x0i ) impliesui(x) ≤ ui(x0i ) which states that all bundles that are cheaper than x0i must notyield higher utility. Again note that we made no reference to the value of anindividuals’ endowment or firm ownership.

Definition 61 A competitive equilibrium for a private ownership economy isan allocation [(x0i )i∈I , (y

0j )j∈J ] and a continuous linear functional φ : S → R

such that

1. for all i ∈ I, x0i solvesmaxui(x) subject to x ∈ Xi and φ(x) ≤Pj∈J θijφ(y

0j )

2. for all j ∈ J, y0j solves maxφ(y) subject to y ∈ Yj3.P

i∈I x0i =

Pj∈J y

0j

We can interpretPj∈J θijφ(y

0j ) as the value of the ownership that household

i holds to all the firms of the economy.

7.4 The Neoclassical Growth Model in Arrow-Debreu Language

Let us look at the neoclassical growth model presented in Section 2. We willadopt the notation so that it fits into our general discussion. Remember thatin the economy the representative household owned the capital stock and therepresentative firm, supplied capital and labor services and bought final out-put from the firm. A helpful exercise would be to repeat this exercise underthe assumption that the firm owns the capital stock. The household had unit

Page 106: Macroeconomics Phd Lectures notes

100 CHAPTER 7. THE TWO WELFARE THEOREMS

endowment of time and initial endowment of k0 of the capital stock. To makeour exercise more interesting we assume that the household values consumptionand leisure according to instantaneous utility function U(c, l), where c is con-sumption and l is leisure. The technology is described by y = F (k, n) where Fexhibits constant returns to scale. For further details refer to Section 2. Let usrepresent this economy in Arrow-Debreu language.

• I = J = 1, θij = 1• Commodity Space S: since three goods are traded in each period (finaloutput, labor and capital services), time is discrete and extends to infinity,a natural choice is S = l3∞ = l∞× l∞× l∞. That is, S consists of all three-dimensional infinite sequences that are bounded in the sup-norm, or

S = s = (s1, s2, s3) = (s1t , s2t , s3t )∞t=0 : sit ∈ R, suptmaxi

¯sit¯<∞

Obviously S, together with the sup-norm, is a (real) normed vector space.We use the convention that the first component of s denotes the outputgood (and hence is required to be positive), whereas the second and thirdcomponents denote labor and capital services, respectively. Again follow-ing the convention these inputs are required to be negative.

• Consumption Set X :

X = x1t , x2t , x3t ∈ S : x30 ≥ −k0,−1 ≤ x2t ≤ 0, x3t ≤ 0, x1t ≥ 0, x1t − (1− δ)x3t + x3t+1 ≥ 0 for all t

We do not distinguish between capital and capital services here; this canbe done by adding extra notation and is an optional homework. Theconstraints indicate that the household cannot provide more capital inthe first period than the initial endowment, can’t provide more than oneunit of labor in each period, holds nonnegative capital stock and is requiredto have nonnegative consumption. Evidently X ⊆ S.

• Utility function u : X → R is defined by

u(x) =∞Xt=0

βtU(x1t − (1− δ)x3t + x3t+1, 1 + x

2t )

Again remember the convention than labor and capital (as inputs) arenegative.

• Aggregate Production Set Y :Y = y1t , y2t , y3t ∈ S : y1t ≥ 0, y2t ≤ 0, y3t ≤ 0, y1t = F (−y3t ,−y2t ) for all t

Note that the aggregate production set reflects the technological con-straints in the economy. It does not contain any constraints that haveto do with limited supply of factors, in particular −1 ≤ y2t is not imposed.

Page 107: Macroeconomics Phd Lectures notes

7.5. A PURE EXCHANGE ECONOMY IN ARROW-DEBREU LANGUAGE101

• An allocation is [x, y] with x, y ∈ S. A feasible allocation is an allocationsuch that x ∈ X, y ∈ Y and x = y. An allocation is Pareto optimal isit is feasible and if there is no other feasible allocation [x∗, y∗] such thatu(x∗) > u(x).

• A price system φ is a continuous linear functional φ : S → R. If φhas inner product representation, we represent it by p = (p1, p2, p3) =(p1t , p2t , p3t )∞t=0.

• A competitive equilibrium for this private ownership economy is an allo-cation [x∗, y∗] and a continuous linear functional such that

1. y∗ maximizes φ(y) subject to y ∈ Y2. x∗ maximizes u(x) subject to x ∈ X and φ(x) ≤ φ(y∗)

3. x∗ = y∗

Note that with constant returns to scale φ(y∗) = 0. With inner productrepresentation of the price system the budget constraint hence becomes

φ(x) = p · x =∞Xt=0

3Xi=1

pitxit ≤ 0

Remembering our sign convention for inputs and mapping p1t = pt, p2t = ptwt,

p3t = ptrt we obtain the same budget constraint as in Section 2.

7.5 A Pure Exchange Economy in Arrow-DebreuLanguage

Suppose there are I individuals that live forever. There is one nonstorableconsumption good in each period. Individuals order consumption allocationsaccording to

ui(ci) =∞Xt=0

βtiU(cit)

They have deterministic endowment streams ei = eit∞t=0. Trade takes place atperiod 0. The standard definition of a competitive (Arrow-Debreu) equilibriumwould go like this:

Definition 62 A competitive equilibrium are prices pt∞t=0 and allocations(cit∞t=0)i∈I such that1. Given pt∞t=0, for all i ∈ I, cit∞t=0 solves maxci≥0 ui(ci) subject to

∞Xt=0

pt(cit − eit) ≤ 0

Page 108: Macroeconomics Phd Lectures notes

102 CHAPTER 7. THE TWO WELFARE THEOREMS

2. Xi∈Icit =

Xi∈Ieit for all t

We briefly want to demonstrate that we can easily write this economy in ourformal language. What goes on is that the household sells his endowment of theconsumption good to the market and buys consumption goods from the market.So even though there is a single good in each period we find it useful to havetwo commodities in each period. We also introduce an artificial technologythat transforms one unit of the endowment in period t into one unit of theconsumption good at period t. There is a single representative firm that operatesthis technology and each consumer owns share θi of the firm, with

Pi∈I θi = 1.

We then have the following representation of this economy

• S = l2∞. We use the convention that the first good is the consumptiongood to be consumed, the second good is the endowment to be sold asinput by consumers. Again we use the convention that final output ispositive, inputs are negative.

• Xi = x ∈ S : x1t ≥ 0,−eit ≤ x2t ≤ 0• ui : Xi → R defined by

ui(x) =∞Xt=0

βtiU(x1t )

• Aggregate production set

Y = y ∈ S : y1t ≥ 0, y2t ≤ 0, y1t = −y2t

• Allocations, feasible allocations and Pareto efficient allocations are definedas before.

• A price system φ is a continuous linear functional φ : S → R. If φ has innerproduct representation, we represent it by p = (p1, p2) = (p1t , p2t )∞t=0.

• A competitive equilibrium [(xi∗)i∈I , y,φ] for this private ownership econ-omy defined as before.

• Note that with constant returns to scale in equilibrium we have φ(y∗) = 0.With inner product representation of the price system in equilibrium alsop1t = p

2t = pt. The budget constraint hence becomes

φ(x) = p · x =∞Xt=0

2Xi=1

pitxit ≤ 0

Page 109: Macroeconomics Phd Lectures notes

7.6. THE FIRST WELFARE THEOREM 103

Obviously (as long as pt > 0 for all t) the consumer will choose xi2t =−eit, i.e. sell all his endowment. The budget constraint then takes thefamiliar form

∞Xt=0

pt(cit − eit) ≤ 0

The purpose of this exercise was to demonstrate that, although in the re-maining part of the course we will describe the economy and define an equilib-rium in the first way, whenever we desire to prove the welfare theorems we canrepresent any pure exchange economy easily in our formal language and use themachinery developed in this section (if applicable).

7.6 The First Welfare Theorem

The first welfare theorem states that every competitive equilibrium allocationis Pareto optimal. The only assumption that is required is that people’s prefer-ences be locally nonsatiated. The proof of the theorem is unchanged from theone you should be familiar with from micro last quarter

Theorem 63 Suppose that for all i, all x ∈ Xi there exists a sequence xn∞n=0in Xi converging to x with u(xn) > u(x) for all n (local nonsatiation). If anallocation [(x0i )i∈I , (y

0j )j∈J ] and a continuous linear functional φ constitute a

competitive equilibrium, then the allocation [(x0i )i∈I , (y0j )j∈J ] is Pareto optimal.

Proof. The proof is by contradiction. Suppose [(x0i )i∈I , (y0j )j∈J ], φ is a

competitive equilibrium.Step 1: We show that for all i, all x ∈ Xi, u(x) ≥ u(x0i ) implies φ(x) ≥ φ(x0i ).

Suppose not, i.e. suppose there exists i and x ∈ Xi with u(x) ≥ u(x0i ) andφ(x) < φ(x0i ). Let xn in Xi be a sequence converging to x with u(xn) > u(x)for all n. Such a sequence exists by our local nonsatiation assumption. Bycontinuity of φ there exists an n such that u(xn) > u(x) ≥ u(x0i ) and φ(xn) <φ(x0i ), violating the fact that x

0i is part of a competitive equilibrium.

Step 2: For all i, all x ∈ Xi, u(x) > u(x0i ) implies φ(x) > φ(x0i ). This followsdirectly from the fact that x0i is part of a competitive equilibrium.Step 3: Now suppose [(x0i )i∈I , (y

0j )j∈J ] is not Pareto optimal. Then there

exists another feasible allocation [(x∗i )i∈I , (y∗j )j∈J ] such that u(x

∗i ) ≥ u(x0i ) for

all i and with strict inequality for some i. Since [(x0i )i∈I , (y0j )j∈J ] is a competitive

equilibrium allocation, by step 1 and 2 we have

φ(x∗i ) ≥ φ(x0i )

for all i, with strict inequality for some i. Summing up over all individuals yieldsXi∈I

φ(x∗i ) >Xi∈I

φ(x0i ) <∞

Page 110: Macroeconomics Phd Lectures notes

104 CHAPTER 7. THE TWO WELFARE THEOREMS

The last inequality comes from the fact that the set of people I is finite andthat for all i, φ(x0i ) is finite (otherwise the consumer maximization problem hasno solution). By linearity of φ we have

φ

ÃXi∈Ix∗i

!=Xi∈I

φ(x∗i ) >Xi∈I

φ(x0i ) = φ

ÃXi∈Ix0i

!

Since both allocations are feasible we have thatXi∈Ix0i =

Xj∈J

y0jXi∈Ix∗i =

Xj∈J

y∗j

and hence

φ

Xj∈J

y∗j

> φ

Xj∈J

y0j

Again by linearity of φ X

j∈Jφ(y∗j ) >

Xj∈J

φ(y0j )

and hence for at least one j ∈ J, φ(y∗j ) > φ(y0j ). But y∗j ∈ Yj and we ob-

tain a contradiction to the hypothesis that [(x0i )i∈I , (y0j )j∈J ] is a competitive

equilibrium allocation.Several remarks are in order. It is crucial for the proof that the set of in-

dividuals is finite, as will be seen in our discussion of overlapping generationseconomies. Also our equilibrium definition seems odd as it makes no referenceto endowments or ownership in the budget constraint. For the preceding the-orem, however, this is not a shortcoming. Since we start with a competitiveequilibrium we know the value of each individual’s consumption allocation. Bylocal nonsatiation each consumer exhausts her budget and hence we implicitlyknow each individual’s income (the value of endowments and firm ownership, ifspecified in a private ownership economy).

7.7 The Second Welfare Theorem

The second welfare theorem provides a converse to the first welfare theorem.Under suitable assumptions it states that for any Pareto-optimal allocationthere exists a price system such that the allocation together with the pricesystem form a competitive equilibrium. It may at first be surprising that thesecond welfare theorem requires much more stringent assumptions than thefirst welfare theorem. Remember, however, that in the first welfare theorem

Page 111: Macroeconomics Phd Lectures notes

7.7. THE SECOND WELFARE THEOREM 105

Aggregate Production Set Y

Set of jointly preferred consumption allocations A

Separating Hyperplane:

Price System Φ

[x,y]

Figure 7.1:

we start with a competitive equilibrium whereas in the proof of the secondwelfare we have to carry out an existence proof. Comparing the assumptionsof the second welfare theorem with those of existence theorems makes clear theintimate relation between them.As in micro we will use a separating hyperplane theorem to establish the

existence of a price system that decentralizes a given allocation [x, y]. Theprice system is nothing else than a hyperplane that separates the aggregateproduction set from the set of consumption allocations that are jointly preferredby all consumers. Figure 6 illustrates this general principle.In lieu of Figure 6 itis not surprising that several convexity assumptions have to be made to provethe second welfare theorem. We will come back to this when we discuss eachspecific assumption. First we state the separating hyperplane that we will usefor our proof. Obviously we can’t use the standard theorems commonly usedin micro4 since our commodity space in a general real vector space (possibly

4See MasColell et al., p. 948. This theorem is usually attributed to Minkowski.

Page 112: Macroeconomics Phd Lectures notes

106 CHAPTER 7. THE TWO WELFARE THEOREMS

infinite dimensional).We will apply the geometric form of the Hahn-Banach theorem. For this we

need the following definition

Definition 64 Let S be a normed real vector space with norm kkS . Define byb(x, ε) = s ∈ S : kx− skS < ε

the open ball of radius ε around x. The interior of a set A ⊆ S, A is defined tobe

A = x ∈ A : ∃ε > 0 with b(x, ε) ⊆ AHence the interior of a set A consists of all the points in A for which we can

find a open ball (no matter how small) around the point that lies entirely in A.We then have the following

Theorem 65 (Geometric Form of the Hahn-Banach Theorem): Let A, Y ⊂ Sbe convex sets and assume that

either Y has an interior point and A ∩ Y = ∅or S is finite dimensional and A ∩ Y = ∅

Then there exists a continuous linear functional φ, not identically zero on S,and a constant c such that

φ(y) ≤ c ≤ φ(x) for all x ∈ A and all y ∈ YFor the proof of the Hahn-Banach theorem in its several forms see Luenberger

(1969), p. 111 and p. 133. For the case that S is finite dimensional this theoremis rather intuitive in light of Figure 6. But since we are interested in commodityspaces with infinite dimensions (typically S = lp, for p ∈ [1,∞]), we usuallyhave to prove that the aggregate production set Y has an interior point inorder to apply the Hahn-Banach theorem. We will two things now: a) prove byexample that the requirement of an interior point is an assumption that cannotbe dispensed with if S is not finite dimensional b) show that this assumptionde facto rules out using S = lp, for p ∈ [1,∞), as commodity space when onewants to apply the second welfare theorem.For the first part consider the following

Example 66 Consider as commodity space

S = xt∞t=0 : xt ∈ R for all t, kxkS =∞Xt=0

βt|xt| <∞

for some β ∈ (0, 1). Let A = θ andY = x ∈ S : |xt| ≤ 1 for all t

Page 113: Macroeconomics Phd Lectures notes

7.7. THE SECOND WELFARE THEOREM 107

Obviously A,B ⊂ S are convex sets. In some sense θ = (0, 0, . . . , 0, . . . ) liesin the middle of Y, but it does not lie in the interior of Y. Suppose it did, thenthere exists ε > 0 such that for all x ∈ S such that

kx− θkS =∞Xt=0

βt|xt| < ε

we have x ∈ Y. But for any ε > 0, define t(ε) = ln( ε2 )

ln(β)+1. Then x = (0, 0, . . . , xt(ε) =

2, 0, . . . ) /∈ Y satisfiesP∞

t=0 βt|xt| = 2βt(ε) < ε. Since this is true for all ε > 0,

this shows that θ is not in the interior of Y, or A∩Y= ∅. A very similar argu-ment shows that no s ∈ S is in the interior of Y, i.e. Y= ∅. Hence the onlyhypothesis for the Hahn-Banach theorem that fails is that Y has an interiorpoint. We now show that the conclusion of the theorem fails. Suppose, to thecontrary, that there exists a continuous linear functional φ on S with φ(s) 6= 0for some s ∈ S and

φ(y) ≤ c ≤ φ(θ) for all y ∈ Y

Obviously φ(θ) = φ(0 · s) = 0 by linearity of φ. Hence it follows that for ally ∈ Y, φ(y) ≤ 0. Now suppose there exists y ∈ Y such that φ(y) < 0. But since−y ∈ Y, by linearity φ(−y) = −φ(y) > 0 a contradiction. Hence φ(y) = 0 forall y ∈ Y. From this it follows that φ(s) = 0 for all s ∈ S (why?), contradictingthe conclusion of the theorem.

As we will see in the proof of the second welfare theorem, to apply theHahn-Banach theorem we have to assure that the aggregate production set hasnonempty interior. The aggregate production set in many application will be(a subset) of the positive orthant of the commodity space. The problem withtaking lp, p ∈ [1,∞) as the commodity space is that, as the next propositionshows, the positive orthant

l+p = x ∈ lp : xt ≥ 0 for all t

has empty interior. The good thing about l∞ is that is has a nonempty interior.This justifies why we usually use it (or its k-fold product space) as commodityspace.

Proposition 67 The positive orthant of lp, p ∈ [0,∞) has an empty interior.The positive orthant of l∞ has nonempty interior.

Proof. For the first part suppose there exists x ∈ l+p and ε > 0 such thatb(x, ε) ⊆ l+p . Since x ∈ lp, xt → 0, i.e. xt <

ε2 for all t ≥ T (ε). Take any τ > T (ε)

and define z as

zt =

½xt if t 6= τ

xt − ε2 if t = τ

Page 114: Macroeconomics Phd Lectures notes

108 CHAPTER 7. THE TWO WELFARE THEOREMS

Evidently zτ < 0 and hence z /∈ l+p . But since

kx− zkp =Ã ∞Xt=0

|xt − zt|p! 1

p

= |xτ − zτ | = ε

2< ε

we have z ∈ b(x, ε), a contradiction. Hence the interior of l+p is empty, theHahn-Banach theorem doesn’t apply and we can’t use it to prove the secondwelfare theorem.For the second part it suffices to construct an interior point of l+∞. Take

x = (1, 1, . . . , 1, . . . ) and ε = 12 . We want to show that b(x, ε) ⊆ l+∞. Take any

z ∈ b(x, ε). Clearly zt ≥ 12 ≥ 0. Furthermore

supt|zt| ≤ 11

2<∞

Hence z ∈ l+∞.Now let us proceed with the statement and the proof of the second welfare

theorem. We need the following assumptions

1. For each i ∈ I, Xi is convex.2. For each i ∈ I, if x, x0 ∈ Xi and ui(x) > ui(x0), then for all λ ∈ (0, 1)

ui(λx+ (1− λ)x0) > ui(x0)

3. For each i ∈ I, ui is continuous.4. The aggregate production set Y is convex

5. Either Y has an interior point or S is finite-dimensional.

Note that the second assumption is sometimes referred to as strict quasi-concavity5 of the utility functions. It implies that the upper contour sets

Aix = z ∈ Xi : ui(z) ≥ ui(x)are convex, for all i, all x ∈ Xi.Without the convexity assumption 1. assumption2 would not be well-defined as without convexXi, λx+(1−λ)x0 /∈ Xi is possible,in which case ui(λx+(1−λ)x0) is not well-defined. I mention this since otherwise1. is not needed for the following theorem. Also note that it is assumption 5that has no counterpart to the theorem in finite dimensions. It only is requiredto use the appropriate separating hyperplane theorem in the proof. With theseassumptions we can state the second welfare theorem

Theorem 68 Let [(x0i ), (y0j )] be a Pareto optimal allocation and assume that

for some h ∈ I there is a xh ∈ Xh with uh(xh) > uh(x0h). Then there exists acontinuous linear functional φ : S → R, not identically zero on S, such that

5To me it seems that quasi-concavity is enough for the theorem to hold as quasi-concavityis equivalent to convex upper contour sets which all one needs in the proof.

Page 115: Macroeconomics Phd Lectures notes

7.7. THE SECOND WELFARE THEOREM 109

1. for all j ∈ J, y0j ∈ argmaxy∈Yj φ(y)2. for all i ∈ I and all x ∈ Xi, ui(x) ≥ ui(x0i ) implies φ(x) ≥ φ(x0i )

Several comments are in order. The theorem states that (under the assump-tions of the theorem) any Pareto optimal allocation can be supported by a pricesystem as a quasi-equilibrium. By definition of Pareto optimality the allocationis feasible and hence satisfies resource balance. The theorem also guaranteesprofit maximization of firms. For consumers, however, it only guarantees thatx0i minimizes the cost of attaining utility ui(x

0i ), but not utility maximization

among the bundles that cost no more than φ(x0i ), as would be required by acompetitive equilibrium. You also may be used to a version of this theoremthat shows that a Pareto optimal allocation can be made into an equilibriumwith transfers. Since here we haven’t defined ownership and in the equilibriumdefinition make no reference to the value of endowments or firm ownership (i.e.do NOT require the budget constraint to hold), we can abstract from trans-fers, too. The proof of the theorem is similar to the one for finite dimensionalcommodity spaces.Proof. Let [(x0i ), (y

0j )] be a Pareto optimal allocation and A

ix0ibe the upper

contour sets (as defined above) with respect to x0i , for all i ∈ I. Also let Aix0i tobe the interior of Ai

x0i, i.e.

Aix0i= z ∈ Xi : ui(z) > ui(x0i )

By assumption 2. the Aix0iare convex and hence Ai

x0iis convex. Furthermore

x0i ∈ Aix0i , so the Aix0iare nonempty. By one of the hypotheses of the theorem

there is some h ∈ I there is a xh ∈ Xh with uh(xh) > uh(x0h). For that h, Ahx0his nonempty. Define

A = Ahx0h+Xi6=h

Aix0i

A is the set of all aggregate consumption bundles that can be split in such a wayas to give every agent at least as much utility and agent h strictly more utilitythan the Pareto optimal allocation [(x0i ), (y

0j )]. As A is the sum of nonempty

convex sets, so is A. Obviously A ⊂ S. By assumption Y is convex. Since[(x0i ), (y

0j )] is a Pareto optimal allocation A ∩ Y = ∅. Otherwise there is an

aggregate consumption bundle x∗ ∈ A∩Y that can be produced (as x∗ ∈ Y ) andPareto dominates x0 (as x∗ ∈ A), contradicting Pareto optimality of [(x0i ), (y0j )].With assumption 5. we have all the assumptions we need to apply the Hahn-Banach theorem. Hence there exists a continuous linear functional φ on S, notidentically zero, and a number c such that

φ(y) ≤ c ≤ φ(x) for all x ∈ A, all y ∈ YIt remains to be shown that [(x0i ), (y

0j )] together with φ satisfy conclusions 1

and 2, i.e. constitute a quasi-equilibrium.

Page 116: Macroeconomics Phd Lectures notes

110 CHAPTER 7. THE TWO WELFARE THEOREMS

First note that the closure of A is A =Pi∈I A

ix0isince by continuity of uh

(assumption 3.) the closure of Ahx0his Ah

x0h. Therefore, since φ is continuous,

c ≤ φ(x) for all x ∈ A =Pi∈I Aix0i.

Second, note that, since [(x0i ), (y0j )] is Pareto optimal, it is feasible and hence

y0 ∈ Y

x0 =Xi∈Ix0i =

Xj∈J

y0j = y0

Obviously x0 ∈ A. Therefore φ(x0) = φ(y0) ≤ c ≤ φ(x0) which implies φ(x0) =φ(y0) = c.To show conclusion 1 fix j ∈ J and suppose there exists yj ∈ Yj such that

φ(yj) > φ(y0j ). For k 6= j define yk = y0k. Obviously y =Pj yj ∈ Y and

φ(y) > φ(y0) = c, a contradiction to the fact that φ(y) ≤ c for all y ∈ Y.Therefore y0j maximizes φ(z) subject to z ∈ Yj , for all j ∈ J.To show conclusion 2 fix i ∈ I and suppose there exists xi ∈ Xi with ui(xi) ≥

ui(x0i ) and φ(xi) < φ(x0i ). For l 6= i define xl = x0l . Obviously x =

Pi xi ∈ A

and φ(x) < φ(x0) = c, a contradiction to the fact that φ(x) ≥ c for all x ∈ A.Therefore x0i minimizes φ(z) subject to ui(z) ≥ ui(x0i ), z ∈ Xi.We now want to provide a condition that assures that the quasi-equilibrium

in the previous theorem is in fact a competitive equilibrium, i.e. is not only costminimizing for the households, but also utility maximizing. This is done in thefollowing

Remark 69 Let the hypotheses of the second welfare theorem be satisfied andlet φ be a continuous linear functional that together with [(x0i ), (y

0j )] satisfies the

conclusions of the second welfare theorem. Also suppose that for all i ∈ I thereexists x0i ∈ Xi such that

φ(x0i) < φ(x0i )

Then [(x0i ), (y0j ),φ] constitutes a competitive equilibrium

Note that, in order to verify the additional condition -the existence of acheaper point in the consumption set for each i ∈ I- we need a candidate pricesystem φ that already passed the test of the second welfare theorem. It isnot, as the assumptions for the second welfare theorem, an assumptions on thefundamentals of the economy alone.Proof. We need to prove that for all i ∈ I, all x ∈ Xi, φ(x) ≤ φ(x0i ) implies

ui(x) ≤ ui(x0i ). Pick an arbitrary i ∈ I, x ∈ Xi satisfying φ(x) ≤ φ(x0i ). Define

xλ = λx0i + (1− λ)x for all λ ∈ (0, 1)

Since by assumption φ(x0i) < φ(x0i ) and φ(x) ≤ φ(x0i ) we have by linearity of φ

φ(xλ) = λφ(x0i) + (1− λ)φ(x) < φ(x0i ) for all λ ∈ (0, 1)

Page 117: Macroeconomics Phd Lectures notes

7.7. THE SECOND WELFARE THEOREM 111

Since xi0 by assumption is part of a quasi-equilibrium and (by convexity of Xiwe have xλ ∈ Xi), ui(xλ) ≥ ui(x0i ) implies φ(xλ) ≥ φ(x0i ), or by contrapositionφ(xλ) < φ(x0i ) implies ui(xλ) < ui(x

0i ) for all λ ∈ (0, 1). But then by continuity

of ui we have ui(x) = limλ→0 ui(xλ) ≤ ui(x0i ) as desired.As shown by an example in Stokey et al. the assumption on the existence

of a cheaper point cannot be dispensed with when wanting to make sure thata quasi-equilibrium is in fact a competitive equilibrium. In Figure 7 we drawthe Edgeworth box of a pure exchange economy. Consumer B’s consumptionset is the entire positive orthant, whereas consumer A’s consumption set is theare above the line marked by −p, as indicated by the broken lines. Both con-sumption sets are convex, the upper contour sets are convex and close as forstandard utility functions satisfying assumptions 2. and 3. Point E clearly rep-resents a Pareto optimal allocation (since at E consumer B’s utility is globallymaximized subject to the allocation being feasible). Furthermore E representsa quasi-equilibrium, since at prices p both consumers minimize costs subjectto attaining at least as much utility as with allocation E. However, at pricesp (obviously the only candidate for supporting E as competitive equilibriumsince tangent to consumer B’s indifference curve through E) agent A obtainshigher utility at allocation E0 with the same cost as with E, hence [E, p] is nota competitive equilibrium. The remark fails because at candidate prices p thereis no consumption allocation for A that is feasible (in XA) and cheaper. Thisdemonstrates that the cheaper-point assumption cannot be dispensed with inthe remark. This concludes the discussion of the second welfare theorem.The last thing we want to do in this section is to demonstrate that our choice

of l∞ as commodity space is not without problems either. We argued earlierthat lp, p ∈ [1,∞) is not an attractive alternative. Now we use the secondwelfare theorem to show that for certain economies the price system needed(whose existence is guaranteed by the theorem) need not lie in l1, i.e. doesnot have a representation as a vector p = (p0, p1, . . . , pt, . . . ). This is bad inthe sense that then the price system we get from the theorem does not have anatural economic interpretation. After presenting such a pathological examplewe will briefly discuss possible remedies.

Example 70 Let S = l∞. There is a single consumer and a single firm. Theaggregate production set is given by

Y = y ∈ S : 0 ≤ yt ≤ 1 + 1t, for all t

The consumption set is given by

X = x ∈ S : xt ≥ 0 for all tThe utility function u : X → R is

u(x) = inftxt

[TO BE COMPLETED]

Page 118: Macroeconomics Phd Lectures notes

112 CHAPTER 7. THE TWO WELFARE THEOREMS

-p

Indifference Curves of B

Indifference Curves of A

0A

0B

E

E’

Figure 7.2:

Page 119: Macroeconomics Phd Lectures notes

7.8. TYPE IDENTICAL ALLOCATIONS 113

7.8 Type Identical Allocations

[TO BE COMPLETED]

Page 120: Macroeconomics Phd Lectures notes

114 CHAPTER 7. THE TWO WELFARE THEOREMS

Page 121: Macroeconomics Phd Lectures notes

Chapter 8

The OverlappingGenerations Model

In this section we will discuss the second major workhorse model of modernmacroeconomics, the Overlapping Generations (OLG) model, due to Allais(1947), Samuelson (1958) and Diamond (1965). The structure of this sectionwill be as follows: we will first present a basic pure exchange version of the OLGmodel, show how to analyze it and contrast its properties with those of a pureexchange economy with infinitely lived agents. The basic differences are that inthe OLG model

• competitive equilibria may be Pareto suboptimal

• (outside) money may have positive value

• there may exist a continuum of equilibria

We will demonstrate these properties in detail via examples. We will thendiscuss the Ricardian Equivalence hypothesis (the notion that, given a stream ofgovernment spending the financing method of the government -taxes or budgetdeficits- does not influence macroeconomic aggregates) for both the infinitelylived agent model as well as the OLG model. Finally we will introduce pro-duction into the OLG model to discuss the notion of dynamic inefficiency. Thefirst part of this section will be based on Kehoe (1989), Geanakoplos (1989), thesecond section on Barro (1974) and the third section on Diamond (1965). Othergood sources of information include Blanchard and Fischer (1989), chapter 3,Sargent and Ljungquist, chapter 8 and Azariadis, chapter 11 and 12.

115

Page 122: Macroeconomics Phd Lectures notes

116 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

8.1 A Simple Pure Exchange Overlapping Gen-erations Model

Let’s start by repeating the infinitely lived agent model to which we will comparethe OLG model. Suppose there are I individuals that live forever. There is onenonstorable consumption good in each period. Individuals order consumptionallocations according to

ui(ci) =∞Xt=1

βt−1i U(cit)

Note that agents start their lives at t = 1 to make this economy comparableto the OLG economies studied below. Agents have deterministic endowmentstreams ei = eit∞t=0. Trade takes place at period 0. The standard definition ofan Arrow-Debreu equilibrium goes like this:

Definition 71 A competitive equilibrium are prices pt∞t=0 and allocations(cit∞t=0)i∈I such that1. Given pt∞t=0, for all i ∈ I, cit∞t=0 solves maxci≥0 ui(ci) subject to

∞Xt=0

pt(cit − eit) ≤ 0

2. Xi∈Icit =

Xi∈Ieit for all t

What are the main shortcomings of this model that have lead to the devel-opment of the OLG model? The first criticism is that individuals apparently donot live forever, so that a model with finitely lived agents is needed. We will seelater that we can give the infinitely lived agent model an interpretation in whichindividuals lived only for a finite number of periods, but, by having an altruisticbequest motive, act so as to maximize the utility of the entire dynasty, whichin effect makes the planning horizon of the agent infinite. So infinite lives initself are not as unsatisfactory as it may seem. But if people live forever, theydon’t undergo a life cycle with low-income youth, high income middle ages andretirement where labor income drops to zero. In the infinitely lived agent modelevery period is like the next (which makes it so useful since this stationarityrenders dynamic programming techniques easily applicable). So in order to an-alyze issues like social security, the effect of taxes on retirement decisions, thedistributive effects of taxes vs. government deficits, the effects of life-cycle sav-ing on capital accumulation one needs a model in which agents experience a lifecycle and in which people of different ages live at the same time in the economy.This is why the OLG model is a very useful tool for applied policy analysis.Because of its interesting (some say, pathological) theoretical properties, it isalso an area of intense study among economic theorists.

Page 123: Macroeconomics Phd Lectures notes

8.1. A SIMPLE PURE EXCHANGEOVERLAPPINGGENERATIONSMODEL117

8.1.1 Basic Setup of the Model

Let us describe the model formally now. Time is discrete, t = 1, 2, 3, . . . andthe economy (but not its people) lives forever. In each period there is a sin-gle, nonstorable consumption good. In each time period a new generation (ofmeasure 1) is born, which we index by its date of birth. People live for twoperiods and then die. By (ett, e

tt+1) we denote generation t’s endowment of the

consumption good in the first and second period of their live and by (ctt, ctt+1)

we denote the consumption allocation of generation t. Hence in time t there aretwo generations alive, one old generation t − 1 that has endowment et−1t andconsumption ct−1t and one young generation t that has endowment ett and con-sumption ctt. In addition, in period 1 there is an initial old generation 0 that hasendowment e01 and consumes c

01. In some of our applications we will endow the

initial generation with an amount of outside money1 m. We will NOT assumem ≥ 0. If m ≥ 0, then m can be interpreted straightforwardly as fiat money,if m < 0 one should envision the initial old people having borrowed from someinstitution (which is, however, outside the model) and m is the amount to berepaid.

In the next Table 1 we demonstrate the demographic structure of the econ-omy. Note that there are both an infinite number of periods as well as well asan infinite number of agents in this economy. This “double infinity” has beencited to be the major source of the theoretical peculiarities of the OLG model(prominently by Karl Shell).

Table 1

TimeG 1 2 . . . t t+ 1e 0 (c01, e

01)

n 1 (c11, e11) (c12, e

12)

e...

. . .

r t− 1 (ct−1t , et−1t )a t (ctt, e

tt) (ctt+1, e

tt+1)

t. t+ 1 (ct+1t+1, et+1t+1)

Preferences of individuals are assumed to be representable by an additivelyseparable utility function of the form

ut(c) = U(ctt) + βU(ctt+1)

and the preferences of the initial old generation is representable by

u0(c) = U(c01)

1Money that is, on net, an asset of the private economy, is “outside money”. This includesfiat currency issued by the government. In contrast, inside money (such as bank deposits) isboth an asset as well as a liability of the private sector (in the case of deposits an asset of thedeposit holder, a liability to the bank).

Page 124: Macroeconomics Phd Lectures notes

118 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

We shall assume that U is strictly increasing, strictly concave and twice contin-uously differentiable. This completes the description of the economy. Note thatwe can easily represent this economy in our formal Arrow-Debreu language fromChapter 7 since it is a standard pure exchange economy with infinite numberof agents and the peculiar preference and endowment structure ets = 0 for alls 6= t, t+1 and ut(c) only depending on ctt, ctt+1. You should complete the formalrepresentation as a useful homework exercise.The following definitions are straightforward

Definition 72 An allocation is a sequence c01, ctt, ctt+1∞t=1. An allocation isfeasible if ctt−1, c

tt ≥ 0 for all t ≥ 1 and

ct−1t + ctt = et−1t + ett for all t ≥ 1

An allocation c01, (ctt, ctt+1)∞t=1 is Pareto optimal if it is feasible and if there isno other feasible allocation c10, (ctt, ctt+1)∞t=1 such that

ut(ctt, c

tt+1) ≥ ut(c

tt, c

tt+1) for all t ≥ 1

u0(c01) ≥ u0(c

01)

with strict inequality for at least one t ≥ 0.We now define an equilibrium for this economy in two different ways, depend-

ing on the market structure. Let pt be the price of one unit of the consumptiongood at period t. In the presence of money (i.e. m 6= 0) we will take moneyto be the numeraire. This is important since we can only normalize the priceof one commoditiy to 1, so with money no further normalizations are admissi-ble. Of course, without money we are free to normalize the price of one othercommodity. Keep this in mind for later. We now have the following

Definition 73 Givenm, an Arrow-Debreu equilibrium is an allocation c01, (ctt, ctt+1)∞t=1and prices pt∞t=1 such that1. Given pt∞t=1, for each t ≥ 1, (ctt, ctt+1) solves

max(ctt,c

tt+1)≥0

ut(ctt, c

tt+1) (8.1)

s.t. ptctt + pt+1c

tt+1 ≤ pte

tt + pt+1e

tt+1 (8.2)

2. Given p1, c01 solves

maxc01

u0(c01)

s.t. p1c01 ≤ p1e

01 +m (8.3)

3. For all t ≥ 1 (Resource Balance or goods market clearing)ct−1t + ctt = e

t−1t + ett for all t ≥ 1

Page 125: Macroeconomics Phd Lectures notes

8.1. A SIMPLE PURE EXCHANGEOVERLAPPINGGENERATIONSMODEL119

As usual within the Arrow-Debreu framework, trading takes place in a hy-pothetical centralized market place at period 0 (even though the generations arenot born yet).2 There is an alternative definition of equilibrium that assumessequential trading. Let rt+1 be the interest rate from period t to period t + 1and stt be the savings of generation t from period t to period t + 1. We willlook at a slightly different form of assets in this section. Previously we dealtwith one-period IOU’s that had price qt in period t and paid out one unit ofthe consumption good in t + 1 (so-called zero bonds). Now we consider assetsthat cost one unit of consumption in period t and deliver 1 + rt+1 units tomor-row. Equilibria with these two different assets are obviously equivalent to eachother, but the latter specification is easier to interpret if the asset at hand isfiat money.We define a Sequential Markets (SM) equilibrium as follows:

Definition 74 Givenm, a sequential markets equilibrium is an allocation c01, (ctt, ctt+1, stt)∞t=1and interest rates rt∞t=1 such that1. Given rt∞t=1 for each t ≥ 1, (ctt, ctt+1, stt) solves

max(ctt,c

tt+1)≥0,stt

ut(ctt, c

tt+1)

s.t. ctt + stt ≤ ett (8.4)

ctt+1 ≤ ett+1 + (1 + rt+1)stt (8.5)

2. Given r1, c01 solves

maxc01

u0(c01)

s.t. c01 ≤ e01 + (1 + r1)m

3. For all t ≥ 1 (Resource Balance or goods market clearing)ct−1t + ctt = e

t−1t + ett for all t ≥ 1 (8.6)

In this interpretation trade takes place sequentially in spot markets for con-sumption goods that open in each period. In addition there is an asset marketthrough which individuals do their saving. Remember that when we wrote downthe sequential formulation of equilibrium for an infinitely lived consumer modelwe had to add a shortsale constraint on borrowing (i.e. st ≥ −A) in order toprevent Ponzi schemes, the continuous rolling over of higher and higher debt.This is not necessary in the OLG model as people live for a finite (two) numberof periods (and we, as usual, assume perfect enforceability of contracts)

2When naming this definition after Arrow-Debreu I make reference to the market structurethat is envisioned under this definition of equilibrium. Others, including Geanakoplos, referto a particular model when talking about Arrow-Debreu, the standard general equilibriummodel encountered in micro with finite number of simultaneously living agents. I hope thisdoes not cause any confusion.

Page 126: Macroeconomics Phd Lectures notes

120 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

Given that the period utility function U is strictly increasing, the budgetconstraints (8.4) and (8.5) hold with equality. Take budget constraint (8.5) forgeneration t and (8.4) for generation t+ 1 and sum them up to obtain

ctt+1 + ct+1t+1 + s

t+1t+1 = e

tt+1 + e

t+1t+1 + (1 + rt+1)s

tt

Now use equation (8.6) to obtain

st+1t+1 = (1 + rt+1)stt

Doing the same manipulations for generation 0 and 1 gives

s11 = (1 + r1)m

and hence, using repeated substitution one obtains

stt = Πtτ=1(1 + rτ )m (8.7)

This is the market clearing condition for the asset market: the amount of saving(in terms of the period t consumption good) has to equal the value of the outsidesupply of assets, Πtτ=1(1+ rτ )m. Strictly speaking one should include condition(8.7) in the definition of equilibrium. By Walras’ law however, either the assetmarket or the good market equilibrium condition is redundant.There is an obvious sense in which equilibria for the Arrow-Debreu economy

(with trading at period 0) are equivalent to equilibria for the sequential marketseconomy. For rt+1 > −1 combine (8.4) and (8.5) into

ctt +ctt+1

1 + rt+1= ett +

ett+11 + rt+1

Divide (8.2) by pt > 0 to obtain

ctt +pt+1ptctt+1 = e

tt +

pt+1ptett+1

Furthermore divide (8.3) by p1 > 0 to obtain

c01 ≤ e01 +m

p1

We then can straightforwardly prove the following proposition

Proposition 75 Let allocation c01, (ctt, ctt+1)∞t=1 and prices pt∞t=1 constitutean Arrow-Debreu equilibrium with pt > 0 for all t ≥ 1. Then there exists a cor-responding sequential market equilibrium with allocations c01, (ctt, ctt+1, stt)∞t=1and interest rates rt∞t=1with

ct−1t = ct−1t for all t ≥ 1ctt = ctt for all t ≥ 1

Page 127: Macroeconomics Phd Lectures notes

8.1. A SIMPLE PURE EXCHANGEOVERLAPPINGGENERATIONSMODEL121

Furthermore, let allocation c01, (ctt, ctt+1, stt)∞t=1 and interest rates rt∞t=1 con-stitute a sequential market equilibrium with rt > −1 for all t ≥ 0. Then there ex-ists a corresponding Arrow-Debreu equilibrium with allocations c01, (ctt, ctt+1)∞t=1and prices pt∞t=1 such that

ct−1t = ct−1t for all t ≥ 1ctt = ctt for all t ≥ 1

Proof. The proof is similar to the infinite horizon counterpart. Givenequilibrium Arrow-Debreu prices pt∞t=1 define interest rates as

1 + rt+1 =ptpt+1

1 + r1 =1

p1

and savings

stt = ett − ctt

It is straightforward to verify that the allocations and prices so constructedconstitute a sequential markets equilibrium.Given equilibrium sequential markets interest rates rt∞t=1 define Arrow-

Debreu prices by

p1 =1

1 + r1

pt+1 =pt

1 + rt+1

Again it is straightforward to verify that the prices and allocations so con-structed form an Arrow-Debreu equilibrium.Note that the requirement on interest rates is weaker for the OLG version

of this proposition than for the infinite horizon counterpart. This is due tothe particular specification of the no-Ponzi condition used. A less stringentcondition still ruling out Ponzi schemes would lead to a weaker condtion in theproposition for the infinite horizon economy also.Also note that with this equivalence we have that

Πtτ=1(1 + rτ )m =m

pt

so that the asset market clearing condition for the sequential markets economycan be written as

ptstt = m

i.e. the demand for assets (saving) equals the outside supply of assets, m. Notethat the demanders of the assets are the currently young whereas the suppliers

Page 128: Macroeconomics Phd Lectures notes

122 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

are the currently old people. From the equivalence we can also see that thereturn on the asset (to be interpreted as money) equals

1 + rt+1 =ptpt+1

=1

1 + πt+1(1 + rt+1)(1 + πt+1) = 1

rt+1 ≈ −πt+1where πt+1 is the inflation rate from period t to t+1. As it should be, the realreturn on money equals the negative of the inflation rate.

8.1.2 Analysis of the Model Using Offer Curves

Unless otherwise noted in this subsection we will focus on Arrow-Debreu equilib-ria. Gale (1973) developed a nice way of analyzing the equilibria of a two-periodOLG economy graphically, using offer curves. First let us assume that the econ-omy is stationary in that ett = w1 and e

tt+1 = w2, i.e. the endowments are time

invariant. For given pt, pt+1 > 0 let by ctt(pt, pt+1) and ctt+1(pt, pt+1) denote

the solution to maximizing (8.1) subject to (8.2) for all t ≥ 1. Given our as-sumptions this solution is unique. Let the excess demand functions y and z bedefined by

y(pt, pt+1) = ctt(pt, pt+1)− ett= ctt(pt, pt+1)− w1

z(pt, pt+1) = ctt+1(pt, pt+1)− w2These two functions summarize, for given prices, all implications that consumeroptimization has for equilibrium allocations. Note that from the Arrow-Debreubudget constraint it is obvious that y and z only depend on the ratio pt+1

pt, but

not on pt and pt+1 separately (this is nothing else than saying that the excessdemand functions are homogeneous of degree zero in prices, as they should be).Varying pt+1

ptbetween 0 and ∞ (not inclusive) one obtains a locus of optimal

excess demands in (y, z) space, the so called offer curve. Let us denote thiscurve as

(y, f(y)) (8.8)

where it is understood that f can be a correspondence, i.e. multi-valued. A pointon the offer curve is an optimal excess demand function for some pt+1

pt∈ (0,∞).

Also note that since ctt(pt, pt+1) ≥ 0 and ctt+1(pt, pt+1) ≥ 0 the offer curveobviously satisfies y(pt, pt+1) ≥ −w1 and z(pt, pt+1) ≥ −w2. Furthermore, sincethe optimal choices obviously satisfy the budget constraint, i.e.

pty(pt, pt+1) + pt+1z(pt, pt+1) = 0

z(pt, pt+1)

y(pt, pt+1)= − pt

pt+1(8.9)

Page 129: Macroeconomics Phd Lectures notes

8.1. A SIMPLE PURE EXCHANGEOVERLAPPINGGENERATIONSMODEL123

Equation (8.9) is an equation in the two unknowns (pt, pt+1) for a given t ≥ 1.Obviously (y, z) = (0, 0) is on the offer curve, as for appropriate prices (whichwe will determine later) no trade is the optimal trading strategy. Equation (8.9)is very useful in that for a given point on the offer curve (y(pt, pt+1), z(pt, pt+1))in y-z space with y(pt, pt+1) 6= 0 we can immediately read off the price ratio atwhich these are the optimal demands. Draw a straight line through the point(y, z) and the origin; the slope of that line equals − pt

pt+1. One should also note

that if y(pt, pt+1) is negative, then z(pt, pt+1) is positive and vice versa. Let’slook at an example

Example 76 Let w1 = ε, w2 = 1 − ε, with ε > 0. Also let U(c) = ln(c) andβ = 1. Then the first order conditions imply

ptctt = pt+1c

tt+1 (8.10)

and the optimal consumption choices are

ctt(pt, pt+1) =1

2

µε+

pt+1pt(1− ε)

¶(8.11)

ctt+1(pt, pt+1) =1

2

µptpt+1

ε+ (1− ε)

¶(8.12)

the excess demands are given by

y(pt, pt+1) =1

2

µpt+1pt(1− ε)− ε

¶(8.13)

z(pt, pt+1) =1

2

µptpt+1

ε− (1− ε)

¶(8.14)

Note that as pt+1pt∈ (0,∞) varies, y varies between − ε

2 and ∞ and z varies

between − (1−ε)2 and ∞. Solving z as a function of y by eliminating pt+1pt

yields

z =ε(1− ε)

4y + 2ε− 1− ε

2for y ∈ (−ε

2,∞) (8.15)

This is the offer curve (y, z) = (y, f(y)). We draw the offer curve in Figure 8

The discussion of the offer curve takes care of the first part of the equilibriumdefinition, namely optimality. It is straightforward to express goods marketclearing in terms of excess demand functions as

y(pt, pt+1) + z(pt−1, pt) = 0 (8.16)

Also note that for the initial old generation the excess demand function is givenby

z0(p1,m) =m

p1

Page 130: Macroeconomics Phd Lectures notes

124 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

z(p ,p )t t+1

y(p ,p )t t+1

Offer Curve z(y)

-w2

-w1

Figure 8.1:

Page 131: Macroeconomics Phd Lectures notes

8.1. A SIMPLE PURE EXCHANGEOVERLAPPINGGENERATIONSMODEL125

so that the goods market equilibrium condition for the first period reads as

y(p1, p2) + z0(p1,m) = 0 (8.17)

Graphically in (y, z)-space equations (8.16) and (8.17) are straight lines throughthe origin with slope −1. All points on this line are resource feasible. Wetherefore have the following procedure to find equilibria for this economy for agiven initial endowment of money m of the initial old generation, using the offercurve (8.8) and the resource feasibility constraints (8.16) and (8.17).

1. Pick an initial price p1 (note that this is NOT a normalization as in theinfinitely lived agent model since the value of p1 determines the real valueof money m

p1the initial old generation is endowed with; we have already

normailzed the price of money). Hence we know z0(p1,m). From (8.17)this determines y(p1, p2).

2. From the offer curve (8.8) we determine z(p1, p2) ∈ f(y(p1, p2)). Note thatif f is a correspondence then there are multiple choices for z.

3. Once we know z(p1, p2), from (8.16) we can find y(p2, p3) and so forth. Inthis way we determine the entire equilibrium consumption allocation

c01 = z0(p1,m) + w2

ctt = y(pt, pt+1) + w1

ctt+1 = z(pt, pt+1) + w2

4. Equilibrium prices can then be found, given p1 from equation (8.9). Anyinitial p1 that induces, in such a way, sequences c

01, (ctt, ctt+1), pt∞t=1 such

that the consumption sequence satisfies ct−1t , ctt ≥ 0 is an equilibriumfor given money stock. This already indicates the possibility of a lot ofequilibria for this model, a fact that we will demonstrate below.

This algorithm can be demonstrated graphically using the offer curve dia-gram. We add the line representing goods market clearing, equation (8.16). Inthe (y, z)-plane this is a straight line through the origin with slope −1. This lineintersects the offer curve at least once, namely at the origin. Unless we havethe degenerate situation that the offer curve has slope −1 at the origin, there is(at least) one other intersection of the offer curve with the goods clearing line.These intersection will have special significance as they will represent stationaryequilibria. As we will see, there is a load of other equilibria as well. We will firstdescribe the graphical procedure in general and then look at some examples.See Figure 9.Given any m (for concreteness let m > 0) pick p1 > 0. This determines

z0 =mp1> 0. Find this quantity on the z-axis, representing the excess demand

of the initial old generation. From this point on the z-axis go horizontally tothe goods market line, from there down to the y-axis. The point on the y-axisrepresents the excess demand function of generation 1 when young. From this

Page 132: Macroeconomics Phd Lectures notes

126 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

z(p ,p ), z(m,p )t t+1 1

y(p ,p )t t+1

Offer Curve z(y)

z0

z1

z2

z3

y1

y2 y

3

Slope=-p /p1 2

Resource constraint y+z=0

Slope=-1

Figure 8.2:

Page 133: Macroeconomics Phd Lectures notes

8.1. A SIMPLE PURE EXCHANGEOVERLAPPINGGENERATIONSMODEL127

point y1 = y(p1, p2) go vertically to the offer curve, then horizontally to thez-axis. The resulting point z1 = z(p1, p2) is the excess demand of generation 1when old. Then back horizontally to the goods market clearing condition anddown yields y2 = y(p2, p3), the excess demand for the second generation and soon. This way the entire equilibrium consumption allocation can be constructed.Equilibrium prices are easily found from equilibrium allocations with (8.9), givenp1. In such a way we construct an entire equilibrium graphically.Let’s now look at some example.

Example 77 Reconsider the example with isoelastic utility above. We foundthe offer curve to be

z =ε(1− ε)

4y + 2ε− 1− ε

2for y ∈ (−ε

2,∞)

The goods market equilibrium condition is

y + z = 0

Now let’s construct an equilibrium for the case m = 0, for zero supply of outsidemoney. Following the procedure outlined above we first find the excess demandfunction for the initial old generation z0(m,p1) = 0 for all p1 > 0. Then fromgoods market y(p1, p2) = −z0(m, p1) = 0. From the offer curve

z(p1, p2) =ε(1− ε)

4y(p1, p2) + 2ε− 1− ε

2

=ε(1− ε)

2ε− 1− ε

2= 0

and continuing we find z(pt, pt+1) = y(pt, pt+1) = 0 for all t ≥ 1. This impliesthat the equilibrium allocation is ct−1t = 1 − ε, ctt = ε. In this equilibrium everyconsumer eats his endowment in each period and no trade between generationstakes place. We call this equilibrium the autarkic equilibrium. Obviously wecan’t determine equilibrium prices from equation (8.9). However, the first orderconditions imply that

pt+1pt

=cttctt+1

1− ε

For m = 0 we can, without loss of generality, normalize the price of the firstperiod consumption good p1 = 1. Note again that only for m = 0 this normaliza-tion is innocuous, since it does not change the real value of the stock of outsidemoney that the initial old generation is endowed with. With this normalizationthe sequence pt∞t=1 defined as

pt =

µε

1− ε

¶t−1

Page 134: Macroeconomics Phd Lectures notes

128 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

together with the autarkic allocation form an (Arrow-Debreu)-equilibrium. Ob-viously any other price sequence pt with pt = αpt for any α > 1, is alsoan equilibrium price sequence supporting the autarkic allocation as equilibrium.This is not, however, what we mean by the possibility of a continuum of equilib-ria in OLG-model, but rather the usual feature of standard competitive equilibriathat the equilibrium prices are only determined up to one normalization. In fact,for this example with m = 0, the autarkic equilibrium is the unique equilibriumfor this economy.3 This is easily seen. Since the initial old generation has nomoney, only its endowments 1−ε, there is no way for them to consume more thantheir endowments. Obviously they can always assure to consume at least theirendowments by not trading, and that is what they do for any p1 > 0 (obviouslyp1 ≤ 0 is not possible in equilibrium). But then from the resource constraintit follows that the first young generation must consume their endowments whenyoung. Since they haven’t saved anything, the best they can do when old is toconsume their endowment again. But then the next young generation is forcedto consume their endowments and so forth. Trade breaks down completely. Forthis allocation to be an equilibrium prices must be such that at these prices allgenerations actually find it optimal not to trade, which yields the prices below.4

Note that in the picture the second intersection of the offer curve with theresource constraint (the first is at the origin) occurs in the forth orthant. Thisneed not be the case. If the slope of the offer curve at the origin is less than one,we obtain the picture above, if the slope is bigger than one, then the secondintersection occurs in the second orthant. Let us distinguish between thesetwo cases more carefully. In general, the price ratio supporting the autarkicequilibrium satisfies

ptpt+1

=U 0(ett)

βU 0(ett+1)=U 0(w1)βU 0(w2)

and this ratio represents the slope of the offer curve at the origin. With thisin mind define the autarkic interest rate (remember our equivalence result from

3The fact that the autarkic is the only equilibrium is specific to pure exchange OLG-modelswith agents living for only two periods. Therefore Samuelson (1958) considered three-periodlived agents for most of his analysis.

4If you look at Sargent and Ljungquist (1999), Chapter 8, you will see that they claim toconstruct several equilibria for exactly this example. Note, however, that their equilibriumdefinition has as feasibility constraint

ct−1t + ctt ≤ et−1t + ett

and all the equilibria apart from the autarkic one constructed above have the feature that fort = 1

c01 + c11 < e

01 + e

11

which violate feasibility in the way we have defined it. Personally I find the free disposalassumption not satisfactory; it makes, however, their life easier in some of the examples tofollow, whereas in my discussion I need more handwaving. You’ll see.

Page 135: Macroeconomics Phd Lectures notes

8.1. A SIMPLE PURE EXCHANGEOVERLAPPINGGENERATIONSMODEL129

above) as

1 + r =U 0(w1)βu0(w2)

Gale (1973) has invented the following terminology: when r < 0 he calls thisthe Samueson case, whereas when r ≥ 0 he calls this the classical case.5 As itwill turn out and will be demonstrated below autarkic equilibria are not Paretooptimal in the Samuelson case whereas they are in the classical case.

8.1.3 Inefficient Equilibria

The preceding example can also serve to demonstrate our first major feature ofOLG economies that sets it apart from the standard infinitely lived consumermodel with finite number of agents: competitive equilibria may be not be Paretooptimal. For economies like the one defined at the beginning of the section thetwo welfare theorems were proved and hence equilibria are Pareto optimal. Nowlet’s see that the equilibrium constructed above for the OLG model may not be.Note that in the economy above the aggregate endowment equals to 1 in

each period. Also note that then the value of the aggregate endowment at theequilibrium prices, given by

P∞t=1 pt. Obviously, if ε < 0.5, then this sum con-

verges and the value of the aggregate endowment is finite, whereas if ε ≥ 0.5,then the value of the aggregate endowment is infinite. Whether the value of theaggregate endowment is infinite has profound implications for the welfare prop-erties of the competitive equilibrium. In particular, using a similar argumentas in the standard proof of the first welfare theorem you can show (and willdo so in the homework) that if

P∞t=1 pt <∞, then the competitive equilibrium

allocation for this economy (and in general for any pure exchange OLG econ-omy) is Pareto-efficient. If, however, the value of the aggregate endowment isinfinite (at the equilibrium prices), then the competitive equilibrium MAY notbe Pareto optimal. In our current example it turns out that if ε > 0.5, thenthe autarkic equilibrium is not Pareto efficient, whereas if ε = 0.5 it is. Sinceinterest rates are defined as

rt+1 =ptpt+1

− 1

ε<0.5 implies rt+1 =1−εε − 1 = 1

ε − 2. Hence ε < 0.5 implies rt+1 > 0 (theclassical case) and ε ≥ 0.5 implies rt+1 < 0. (the Samuelson case). Inefficiency

5More generally, the Samuelson case is defined by the condition that savings of the younggeneration be positive at an interest rate equal to the population growth rate n. So far wehave assumed n = 0, so the Samuelson case requires saving to be positive at zero interestrate. We stated the condition as r < 0. But if the interest rate at which the young don’t save(the autarkic allocation) is smaller than zero, then at the higher interest rate of zero they willsave a positive amount, so that we can define the Samuelson case as in the text, provided thatsavings are strictly increasing in the interest rate. This in turn requires the assumption thatfirst and second period consumption are strict gross substitutes, so that the offer curve is notbackward-bending. In the homework you will encounter an example in which this assumptionis not satisfied.

Page 136: Macroeconomics Phd Lectures notes

130 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

is therefore associated with low (negative interest rates). In fact, Balasko andShell (1980) show that the autarkic equilibrium is Pareto optimal if and only if

∞Xt=1

tYτ=1

(1 + rτ+1) = +∞

where rt+1 is the sequence of autarkic equilibrium interest rates.6 Obviouslythe above equation is satisfied if and only if ε ≤ 0.5.Let us briefly demonstrate the first claim (a more careful discussion is left

for the homework). To show that for ε > 0.5 the autarkic allocation (which isthe unique equilibrium allocation) is not Pareto optimal it is sufficient to findanother feasible allocation that Pareto-dominates it. Let’s do this graphically inFigure 10. The autarkic allocation is represented by the origin (excess demandfunctions equal zero). Consider an alternative allocation represented by theintersection of the offer curve and the resource constraint. We want to arguethat this point Pareto dominates the autarkic allocation. First consider anarbitrary generation t ≥ 1. Note that the indifference curve through any pointmust lie (locally) to the inside of the offer curve. From (8.9) we saw that theprice ratio pt

pt+1at which a point on the offer curve is the optimal choice is a

line through the origin and through the point of question. This line representsnothing else but the budget line at the price ratio pt

pt+1. Since the point on the

offer curve is utility maximizing choices given the prices the indifference curvethrough the point must lie tangent above the line through the point and theorigin. Any other point on this line (including the origin) must be weakly worse

6Rather than a formal proof (which is quite involved), let’s develop some intuition for whylow interest rates are associated with inefficiency. Take the autarkic allocation and try toconstruct a Pareto improvement. In particular, give additional δ0 > 0 units of consumptionto the initial old generation. This obviously improves this generation’s life. From resourcefeasibilty this requires taking away δ0 from generation 1 in their first period of life. To makethem not worse of they have to recieve δ1 in additional consumption in their second period oflife, with δ1 satisfying

δ0U0(e11) = δ1βU

0(e12)

or

δ1 = δ0βU 0(e12)U 0(e11)

= δ0(1 + r2) > 0

and in general

δt = δ0

tYτ=1

(1 + rτ+1)

are the required transfers in the second period of generation t’s life to compensate for thereduction of first period consumption. Obviously such a scheme does not work if the economyends at fine time T since the last generation (that lives only through youth) is worse off. Butas our economy extends forever, such an intergenerational transfer scheme is feasible providedthat the δt don’t grow too fast, i.e. if interest rates are sufficiently small. But if such a transferscheme is feasible, then we found a Pareto improvement over the original autarkic allocation,and hence the autarkic equilibrium allocation is not Pareto efficient.

Page 137: Macroeconomics Phd Lectures notes

8.1. A SIMPLE PURE EXCHANGEOVERLAPPINGGENERATIONSMODEL131

z(p ,p ), z(m,p )t t+1 1

y(p ,p )t t+1

Offer Curve z(y)

Resource constraint y+z=0

Slope=-1

Autarkic Allocation

Pareto-dominating allocation

Indifference Curve through dominating allocation

Indifference Curve through autarkic allocation

z = z0 t

y =y1 t

Figure 8.3:

than this point at given prices ptpt+1

. If we take pt = pt+1 this demonstrates that

the alternative point (which is both on the offer curve as well as the resourceconstraint, the line with slope -1) is at least as good as the autarkic allocationfor all generations t ≥ 1.What about the initial old generation? In the autarkicallocation it has c01 = 1 − ε, or z0 = 0. In the new allocation it has z0 > 0as shown in the figure, so the initial old generation is strictly better off in thisnew allocation. Hence the alternative allocation Pareto-dominates the autarkicequilibrium allocation, which shows that this allocation is not Pareto-optimal.In the homework you are asked to make this argument rigorous by actuallycomputing the alternative allocation and then arguing that it Pareto-dominatesthe autarkic equilibrium.

What in our graphical argument hinges on the assumption that ε > 0.5.Remember that for ε ≤ 0.5 we have said that the autarkic allocation is actuallyPareto optimal. It turns out that for ε < 0.5, the intersection of the resource

Page 138: Macroeconomics Phd Lectures notes

132 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

constraint and the offer curve lies in the fourth orthant instead of in the secondas in Figure 10. It is still the case that every generation t ≥ 1 at least weaklyprefers the alternative to the autarkic allocation. Now, however, this alternativeallocation has z0 < 0, which makes the initial old generation worse off than inthe autarkic allocation, so that the argument does not work. Finally, for ε = 0.5we have the degenerate situation that the slope of the offer curve at the origin is−1, so that the offer curve is tangent to the resource line and there is no secondintersection. Again the argument does not work and we can’t argue that theautarkic allocation is not Pareto optimal. It is an interesting optional exerciseto show that for ε = 0.5 the autarkic allocation is Pareto optimal.

Now we want to demonstrate the second and third feature of OLG modelsthat set it apart from standard Arrow-Debreu economies, namely the possibilityof a continuum of equilibria and the fact that outside money may have positivevalue. We will see that, given the way we have defined our equilibria, thesetwo issues are intimately linked. So now let us suppose that m 6= 0. In ourdiscussion we will assume that m > 0, the situation for m < 0 is symmetric. Wefirst want to argue that for m > 0 the economy has a continuum of equilibria,not of the trivial sort that only prices differ by a constant, but that allocationsdiffer across equilibria. Let us first look at equilibria that are stationary in thefollowing sense:

Definition 78 An equilibrium is stationary if ct−1t = co, ctt = cy and pt+1

pt= a,

where a is a constant.

Given that we made the assumption that each generation has the same en-dowment structure a stationary equilibrium necessarily has to satisfy y(pt, pt+1) =y, z0(m, p1) = z(pt, pt+1) = z for all t ≥ 1. From our offer curve diagram theonly candidates are the autarkic equilibrium (the origin) and any other alloca-tions represented by intersections of the offer curve and the resource line. Wewill discuss the possibility of an autarkic equilibrium with money later. Withrespect to other stationary equilibria, they all have to have prices pt+1pt

= 1, with

p1 such that (mp1,−m

p1) is on the offer curve. For our previous example, for any

m 6= 0 we find the stationary equilibrium by solving for the intersection of offercurve and resource line

y + z = 0

z =ε(1− ε)

4y + 2ε− 1− ε

2

This yields a second order polynomial in y

−y = ε(1− ε)

4y + 2ε− 1− ε

2

whose one solution is y = 0 (the autarkic allocation) and the other solution isy = 1

2 − ε, so that z = −12 + ε. Hence the corresponding consumption allocation

Page 139: Macroeconomics Phd Lectures notes

8.1. A SIMPLE PURE EXCHANGEOVERLAPPINGGENERATIONSMODEL133

has

ct−1t = ctt =1

2for all t ≥ 1

In order for this to be an equilibrium we need

1

2= c01 = (1− ε) +

m

p1

hence p1 =m

ε−0.5 > 0. Therefore a stationary equilibrium (apart from autarky)only exists for m > 0 and ε > 0.5 or m < 0 and ε < 0.5. Also note thatthe choice of p1 is not a matter of normalization: any multiple of p1 will notyield a stationary equilibrium. The equilibrium prices supporting the stationaryallocation have pt = p1 for all t ≥ 1. Finally note that this equilibrium, sinceit features pt+1

pt= 1, has an inflation rate of πt+1 = −rt+1 = 0. It is exactly

this equilibrium allocation that we used to prove that, for ε > 0.5, the autarkicequilibrium is not Pareto-efficient.How about the autarkic allocation? Obviously it is stationary as ct−1t = 1−ε

and ctt = ε for all t ≥ 1. But can it be made into an equilibrium if m 6= 0. If welook at the sequential markets equilibrium definition there is no problem: thebudget constraint of the initial old generation reads

c01 = 1− ε+ (1 + r1)m

So we need r1 = −1. For all other generations the same arguments as withoutmoney apply and the interest sequence satisfying r1 = −1, rt+1 = 1−ε

ε − 1for all t ≥ 1, together with the autarkic allocation forms a sequential marketequilibrium. In this equilibrium the stock of outside money, m, is not valued:the initial old don’t get any goods in exchange for it and future generationsare not willing to ever exchange goods for money, which results in the autarkic,no-trade situation. To make autarky an Arrow-Debreu equilibrium is a bit moreproblematic. Again from the budget constraint of the initial old we find

c01 = 1− ε+m

p1

which, for autarky to be an equilibrium requires p1 =∞, i.e. the price level isso high in the first period that the stock of money de facto has no value. Sincefor all other periods we need pt+1

pt= ε

1−ε to support the autarkic allocation, wehave the obscure requirement that we need price levels to be infinite with well-defined finite price ratios. This is unsatisfactory, but there is no way around itunless we a) change the equilibrium definition (see Sargent and Ljungquist) orb) let the economy extend from the infinite past to the infinite future (insteadof starting with an initial old generation, see Geanakoplos) or c) treat moneysomewhat as a residual, as something almost endogenous (see Kehoe) or d) makesome consumption good rather than money the numeraire (with nonmonetaryequilibria corresponding to situations in which money has a price of zero in termsof real consumption goods). For now we will accept autarky as an equilibrium

Page 140: Macroeconomics Phd Lectures notes

134 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

even with money and we will treat it as identical to the autarkic equilibriumwithout money (because indeed in the sequential markets formulation only r1changes and in the Arrow Debreu formulation only p1 changes, although in anunsatisfactory fashion).

8.1.4 Positive Valuation of Outside Money

In our construction of the nonautarkic stationary equilibrium we have alreadydemonstrated our second main result of OLG models: outside money may havepositive value. In that equilibrium the initial old had endowment 1 − ε butconsumed c01 =

12 . If ε >

12 , then the stock of outside money, m, is valued in

equilibrium in that the old guys can exchange m pieces of intrinsically worthlesspaper for m

p1> 0 units of period 1 consumption goods.7 The currently young

generation accepts to transfer some of their endowment to the old people forpieces of paper because they expect (correctly so, in equilibrium) to exchangethese pieces of paper against consumption goods when they are old, and henceto achieve an intertemporal allocation of consumption goods that dominatesthe autarkic allocation. Without the outside asset, again, this economy can donothing else but remain in the possibly dismal state of autarky (imagine ε = 1and log-utility). This is why the social contrivance of money is so useful in thiseconomy. As we will see later, other institutions (for example a pay-as-you-gosocial security system) may achieve the same as money.Before we demonstrate that, apart from stationary equilibria (two in the

example, usually at least only a finite number) there may be a continuum ofother, nonstationary equilibria we take a little digression to show for the generalinfinitely lived agent endowment economies set out at the beginning of thissection money cannot have positive value in equilibrium.

Proposition 79 In pure exchange economies with a finite number of infinitelylived agents there cannot be an equilibrium in which outside money is valued.

Proof. Suppose, to the contrary, that there is an equilibrium (cit)i∈I∞t=1, pt∞t=1for initial endowments of outside money (mi)i∈I such that

Pi∈I m

i 6= 0. Giventhe assumption of local nonsatiation each consumer in equilibrium satisfies theArrow-Debreu budget constraint with equality

∞Xt=1

ptcit =

Xt=1

pteit +m

i <∞

Summing over all individuals i ∈ I yields∞Xt=1

ptXi∈I

¡cit − eit

¢=Xi∈Imi

7In finance lingo money in this equilibrium is a “bubble”. The fundamental value of anassets is the value of its dividends, evaluated at the equilibrium Arrow-Debreu prices. Anasset is (or has) a bubble if its price does not equal its fundamental value. Obviuosly, sincemoney doesn’t pay dividends, its fundamental value is zero and the fact that it is valuedpositively in equilibrium makes it a bubble.

Page 141: Macroeconomics Phd Lectures notes

8.1. A SIMPLE PURE EXCHANGEOVERLAPPINGGENERATIONSMODEL135

But resource feasibility requiresPi∈I¡cit − eit

¢= 0 for all t ≥ 1 and henceX

i∈Imi = 0

a contradiction. This shows that there cannot exist an equilibrium in this typeof economy in which outside money is valued in equilibrium. Note that thisresult applies to a much wider class of standard Arrow-Debreu economies thanjust the pure exchange economies considered in this section.

Hence we have established the second major difference between the standardArrow-Debreu general equilibrium model and the OLG model.

Continuum of Equilibria

We will now go ahead and demonstrate the third major difference, the possibilityof a whole continuum of equilibria in OLG models. We will restrict ourselves tothe specific example. Again suppose m > 0 and ε > 0.5.8 For any p1 such thatmp1< ε − 1

2 > 0 we can construct an equilibrium using our geometric methodbefore. From the picture it is clear that all these equilibria have the featurethat the equilibrium allocations over time converge to the autarkic allocation,with z0 > z1 > z2 > . . . zt > 0 and limt→∞ zt = 0 and 0 > yt > . . . y2 > y1 withlimt→∞ yt = 0. We also see from the figure that, since the offer curve lies belowthe -450-line for the part we are concerned with that p1p2 < 1 and

ptpt+1

< pt−1pt

<

. . . < p1p2< 1, implying that prices are increasing with limt→∞ pt = ∞. Hence

all the nonstationary equilibria feature inflation, although the inflation rate isbounded above by π∞ = −r∞ = 1− 1−ε

ε = 2− 1ε > 0. The real value of money,

however, declines to zero in the limit.9 Note that, although all nonstationaryequilibria so constructed in the limit converge to the same allocation (autarky),they differ in the sense that at any finite t, the consumption allocations and priceratios (and levels) differ across equilibria. Hence there is an entire continuum ofequilibria, indexed by p1 ∈ ( m

ε−0.5 ,∞). These equilibria are arbitrarily close toeach other. This is again in stark contrast to standard Arrow-Debreu economieswhere, generically, the set of equilibria is finite and all equilibria are locallyunique.10 For details consult Debreu (1970) and the references therein.

Note that, if we are in the Samuelson case r < 0, then (and only then)

8You should verify that if ε ≤ 0.5, then r ≥ 0 and the only equilibrium with m > 0 isthe autarkic equilibrium in which money has no value. All other possible equilibrium pathseventually violate nonnegativity of consumption.

9But only in the limit. It is crucial that the real value of money is not zero at finite t,since with perfect foresight as in this model generation t would anticipate the fact that moneywould lose all its value, would not accept it from generation t− 1 and all monetary equilibriawould unravel, with only the autarkic euqilibrium surviving.10Generically in this context means, for almost all endowments, i.e. the set of possible values

for the endowments for which this statement does not hold is of measure zero. Local uniquenesmeans that in for every equilibrium price vector there exists ε such that any ε-neighborhoodof the price vector does not contain another equilibrium price vector (apart from the trivialones involving a different normalization).

Page 142: Macroeconomics Phd Lectures notes

136 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

all these equilibria are Pareto-ranked.11 Let the equilibria be indexed by p1.One can show, by similar arguments that demonstrated that the autarkic equi-librium is not Pareto optimal, that these equilibria are Pareto-ranked: letp1, p1 ∈ ( m

ε−0.5 ,∞) with p1 > p1, then the equilibrium corresponding to p1Pareto-dominates the equilibrium indexed by p1. By the same token, the onlyPareto optimal equilibrium allocation is the nonautarkic stationary monetaryequilibrium.

8.1.5 Productive Outside Assets

We have seen that with a positive supply of an outside asset with no intrinsicvalue, m > 0, then in the Samuelson case (for which the slope of the offer curveis smaller than one at the autarkic allocation) we have a continuum of equilibria.Now suppose that, instead of being endowed with intrinsically useless pieces ofpaper the initial old are endowed with a Lucas tree that yields dividends d > 0in terms of the consumption good in each period. In a lot of ways this economyseems a lot like the previous one with money. So it should have the same numberand types of equilibria!? The definition of equilibrium (we will focus on Arrow-Debreu equilibria) remains the same, apart from the resource constraint whichnow reads

ct−1t + ctt = et−1t + ett + d

and the budget constraint of the initial old generation which now reads

p1c01 ≤ p1e01 + d

∞Xt=1

pt

Let’s analyze this economy using our standard techniques. The offer curveremains completely unchanged, but the resource line shifts to the right, nowgoes through the points (y, z) = (d, 0) and (y, z) = (0, d). Let’s look at Figure11.It appears that, as in the case with moneym > 0 there are two stationary and

a continuum of nonstationary equilibria. The point (y1, z0) on the offer curveindeed represents a stationary equilibrium. Note that the constant equilibriumprice ratio satisfies pt

pt+1= α > 1 (just draw a ray through the origin and the

point and compare with the slope of the resource constraint which is −1). Hencewe have, after normalization of p1 = 1, pt =

¡1α

¢t−1and therefore the value of

the Lucas tree in the first period equals

d∞Xt=1

µ1

α

¶t−1<∞

How about the other intersection of the resource line with the offer curve,(y01, z00)? Note that in this hypothetical stationary equilibrium

ptpt+1

= γ < 1, so

11Again we require the assumption that consumption in the first and the second period arestrict gross substitutes, ruling out backward-bending offer curves.

Page 143: Macroeconomics Phd Lectures notes

8.1. A SIMPLE PURE EXCHANGEOVERLAPPINGGENERATIONSMODEL137

z(p ,p ), z(m,p )t t+1 1

y(p ,p )t t+1

Offer Curve z(y)

z0

z’’0

z’’1

z’0

y1

y’1

y’’1

Slope=-p /p1 2

Resource constraint y+z=d

Slope=-1

y’’2

Figure 8.4:

Page 144: Macroeconomics Phd Lectures notes

138 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

that pt =³1γ

´t−1p1. Hence the period 0 value of the Lucas tree is infinite and

the consumption of the initial old exceed the resources available in the economyin period 1. This obviously cannot be an equilibrium. Similarly all equilibriumpaths starting at some point z000 converge to this stationary point, so for allhypothetical nonstationary equilibria we have pt

pt+1< 1 for t large enough and

again the value of the Lucas tree remains unbounded, and these paths cannotbe equilibrium paths either. We conclude that in this economy there exists aunique equilibrium, which, by the way, is Pareto optimal.

This example demonstrates that it is not the existence of a long-lived outsideasset that is responsible for the existence of a continuum of equilibria. What isthe difference? In all monetary equilibria apart from the stationary nonautarkicequilibrium (which exists for the Lucas tree economy, too) the price level goesto infinity, as in the hypothetical Lucas tree equilibria that turned out not tobe equilibria. What is crucial is that money is intrinsically useless and doesnot generate real stuff so that it is possible in equilibrium that prices explode,but the real value of the dividends remains bounded. Also note that we wereto introduce a Lucas tree with negative dividends (the initial old generation isan eternal slave, say, of the government and has to come up with d in everyperiod to be used for government consumption), then the existence of the wholecontinuum of equilibria is restored.

8.1.6 Endogenous Cycles

Not only is there a possibility of a continuum of equilibria in the basic OLG-model, but these equilibria need not take the monotonic form described above.Instead, equilibria with cycles are possible. In Figure 12 we have drawn an offercurve that is backward bending. In the homework you will see an example ofpreferences that yields such a backward bending offer curve, for a rather normalutility function.

Let m > 0 and let p1 be such that z0 =mp1. Using our geometric approach

we find y1 = y(p1, p2) from the resource line, z1 = z(p1, p2) from the offercurve (ignore for the moment the fact that there are several z1 will do; thismerely indicates that the multiplicity of equilibria is of even higher order thanpreviously demonstrated). From the resource line we find y2 = y(p2, p3) andfrom the offer curve z2 = z(p2, p3) = z0. After period t = 2 the economy repeatsthe cycle from the first two periods. The equilibrium allocation is of the form

ct−1t =

½col = z0 − w2 for t oddcoh = z1 − w2 for t even

ctt =

½cyl = y1 − w1 for t oddcyh = y2 − w1 for t even

Page 145: Macroeconomics Phd Lectures notes

8.1. A SIMPLE PURE EXCHANGEOVERLAPPINGGENERATIONSMODEL139

z(p ,p ), z(m,p )t t+1 1

y(p ,p )t t+1

Offer Curve z(y)

z1

z0

y2

y1

Resource constraint y+z=0

Slope=-1

-p /p2 3

-p /p1 2

Figure 8.5:

Page 146: Macroeconomics Phd Lectures notes

140 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

with col < coh, cyl < cyh. Prices satisfy

ptpt+1

=

½αh for t oddαl for t even

πt+1 = −rt+1 =½

πl < 0 for t oddπh > 0 for t even

Consumption of generations fluctuates in a two period cycle, with odd genera-tions eating little when young and a lot when old and even generations havingthe reverse pattern. Equilibrium returns on money (inflation rates) fluctuate,too, with returns from odd to even periods being high (low inflation) and returnsbeing low (high inflation) from even to odd periods. Note that these cycles arepurely endogenous in the sense that the environment is completely stationary:nothing distinguishes odd and even periods in terms of endowments, prefer-ences of people alive or the number of people. It is not surprising that someeconomists have taken this feature of OLG models to be the basis of a theoryof endogenous business cycles (see, for example, Grandmont (1985)). Also notethat it is not particularly difficult to construct cycles of length bigger than 2periods.

8.1.7 Social Security and Population Growth

The pure exchange OLG model renders itself nicely to a discussion of a pay-as-you-go social security system. It also prepares us for the more complicateddiscussion of the same issue once we have introduced capital accumulation.Consider the simple model without money (i.e. m = 0). Also now assume thatthe population is growing at constant rate n, so that for each old person in agiven period there are (1 + n) young people around. Definitions of equilibriaremain unchanged, apart from resource feasibility that now reads

ct−1t + (1 + n)ctt = et−1t + (1 + n)ett

or, in terms of excess demands

z(pt−1, pt) + (1 + n)y(pt, pt+1) = 0

This economy can be analyzed in exactly the same way as before with noticingthat in our offer curve diagram the slope of the resource line is not −1 anymore,but −(1+n). We know from above that, without any government intervention,the unique equilibrium in this case is the autarkic equilibrium. We now wantto analyze under what conditions the introduction of a pay-as-you-go socialsecurity system in period 1 (or any other date) is welfare-improving. We againassume stationary endowments ett = w1 and e

tt+1 = w2 for all t. The social

security system is modeled as follows: the young pay social security taxes ofτ ∈ [0, w1) and receive social security benefits b when old. We assume that thesocial security system balances its budget in each period, so that benefits aregiven by

b = τ(1 + n)

Page 147: Macroeconomics Phd Lectures notes

8.1. A SIMPLE PURE EXCHANGEOVERLAPPINGGENERATIONSMODEL141

Obviously the new unique competitive equilibrium is again autarkic with en-dowments (w1 − τ , w2 + τ(1 + n)) and equilibrium interest rates satisfy

1 + rt+1 = 1 + r =U 0(w1 − τ)

βU 0(w2 + τ(1 + n))

Obviously for any τ > 0, the initial old generation receives a windfall transferof τ(1 + n) > 0 and hence unambiguously benefits from the introduction. Forall other generations, define the equilibrium lifetime utility, as a function of thesocial security system, as

V (τ) = U(w1 − τ) + βU(w2 + τ(1 + n))

The introduction of a small social security system is welfare improving if andonly if V 0(τ), evaluated at τ = 0, is positive. But

V 0(τ) = −U 0(w1 − τ) + βU 0(w2 + τ(1 + n))(1 + n)

V 0(0) = −U 0(w1) + βU 0(w2)(1 + n)

Hence V 0(0) > 0 if and only if

n >U 0(w1)βU 0(w2)

− 1 = r

where r is the autarkic interest rate. Hence the introduction of a (marginal)pay-as-you-go social security system is welfare improving if and only if the pop-ulation growth rate exceeds the equilibrium (autarkic) interest rate, or, to useour previous terminology, if we are in the Samuelson case where autarky is nota Pareto optimal allocation. Note that social security has the same function asmoney in our economy: it is a social institution that transfers resources betweengenerations (backward in time) that do not trade among each other in equilib-rium. In enhancing intergenerational exchange not provided by the market itmay generate allocations that are Pareto superior to the autarkic allocation, inthe case in which individuals private marginal rate of substitution 1+ r (at theautarkic allocation) falls short of the social intertemporal rate of transformation1 + n.

If n > r we can solve for optimal sizes of the social security system analyti-cally in special cases. Remember that for the case with positive money supplym > 0 but no social security system the unique Pareto optimal allocation wasthe nonautarkic stationary allocation. Using similar arguments we can showthat the sizes of the social security system for which the resulting equilibriumallocation is Pareto optimal is such that the resulting autarkic equilibrium in-terest rate is at least equal to the population growth rate, or

1 + n ≤ U 0(w1 − τ)

βU 0(w2 + τ(1 + n))

Page 148: Macroeconomics Phd Lectures notes

142 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

For the case in which the period utility function is of logarithmic form this yields

1 + n ≤ w2 + τ(1 + n)

β(w1 − τ)

τ ≥ β

1 + βw1 − w2

(1 + β)(1 + n)= τ∗(w1, w2, n,β)

Note that τ∗ is the unique size of the social security system that maximizes thelifetime utility of the representative generation. For any smaller size we couldmarginally increase the size and make the representative generation better offand increase the windfall transfers to the initial old. Note, however, that anyτ > τ∗ satisfying τ ≤ w1 generates a Pareto optimal allocation, too: the repre-sentative generation would be better off with a smaller system, but the initial oldgeneration would be worse off. This again demonstrates the weak requirementsthat Pareto optimality puts on an allocation. Also note that the “optimal” sizeof social security is an increasing function of first period income w1, the popu-lation growth rate n and the time discount factor β, and a decreasing functionof the second period income w2.

So far we have assumed that the government sustains the social securitysystem by forcing people to participate.12 Now we briefly describe how sucha system may come about if policy is determined endogenously. We make thefollowing assumptions. The initial old people can decide upon the size of thesocial security system τ0 = τ∗∗ ≥ 0. In each period t ≥ 1 there is a majorityvote as to whether the current system is to be kept or abolished. If the majorityof the population in period t favors the abolishment of the system, then τ t = 0and no payroll taxes or social security benefits are paid. If the vote is in favorof the system, then the young pay taxes τ∗∗ and the old receive (1 + n)τ∗∗.We assume that n > 0, so the current young generation determines currentpolicy. Since current voting behavior depends on expectations about votingbehavior of future generations we have to specify how expectations about thevoting behavior of future generations is determined. We assume the followingexpectations mechanism (see Cooley and Soares (1999) for a more detailed dis-cussion of justifications as well as shortcomings for this specification of formingexpectations):

τet+1 =

½τ∗∗ if τ t = τ∗∗

0 otherwise(8.18)

that is, if young individuals at period t voted down the original social securitysystem then they expect that a newly proposed social security system will bevoted down tomorrow. Expectations are rational if τ et = τ t for all t. Let τ =τ t∞t=0 be an arbitrary sequence of policies that is feasible (i.e. satisfies τ t ∈[0, w1))

12This section is not based on any reference, but rather my own thoughts. Please be awareof this and read with caution.

Page 149: Macroeconomics Phd Lectures notes

8.1. A SIMPLE PURE EXCHANGEOVERLAPPINGGENERATIONSMODEL143

Definition 80 A rational expectations politico-economic equilibrium, given ourexpectations mechanism is an allocation rule c01(τ), (ctt(τ), ctt+1(τ)), price rulept(τ) and policies τ t such that13

1. for all t ≥ 1, for all feasible τ , and given pt(τ),

(ctt, ctt+1) ∈ arg max

(ctt,ctt+1)≥0

V (τ t, τ t+1) = U(ctt) + βU(ctt+1)

s.t. ptctt + pt+1c

tt+1 ≤ pt (w1 − τ t) + pt+1 (w2 + (1 + n)τ t+1)

2. for all feasible τ , and given pt(τ),

c01 ∈ argmaxc01≥0

V (τ0, τ1) = U(c01)

s.t. p1c01 ≤ p1(w2 + (1 + n)τ1)

3.

ct−1t + (1 + n)ctt = w2 + (1 + n)w1

4. For all t ≥ 1

τ t ∈ arg maxθ∈0,τ∗∗

V (θ, τ et+1)

where τet+1 is determined according to (8.18)

5.

τ0 ∈ arg maxθ∈[0,w1)

V (θ, τ1)

6. For all t ≥ 1

τet = τ t

Conditions 1-3 are the standard economic equilibrium conditions for anyarbitrary sequence of social security taxes. Condition 4 says that all agents ofgeneration t ≥ 1 vote rationally and sincerely, given the expectations mechanismspecified. Condition 5 says that the initial old generation implements the bestpossible social security system (for themselves). Note the constraint that theinitial generation faces in its maximization: if it picks θ too high, the first regulargeneration (see condition 4) may find it in its interest to vote the system down.Finally the last condition requires rational expectations with respect to theformation of policy expectations.

13The dependence of allocations and prices on τ is implicit from now on.

Page 150: Macroeconomics Phd Lectures notes

144 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

Political equilibria are in general very hard to solve unless one makes theeconomic equilibrium problem easy, assumes simple voting rules and simplifiesas much as possible the expectations formation process. I tried to do all of theabove for our discussion. So let find an (the!) political economic equilibrium.First notice that for any policy the equilibrium allocation will be autarky sincethere is no outside asset. Hence we have as equilibrium allocations and pricesfor a given policy τ

ct−1t = w2 + (1 + n)τ t

ctt = w1 − τ t

p1 = 1

ptpt+1

=U 0(w1 − τ t)

βU 0(w2 + (1 + n)τ t)

Therefore the only equilibrium element to determine are the optimal policies.Given our expectations mechanism for any choice of τ0 = τ∗∗, when wouldgeneration t vote the system τ∗∗ down when young? If it does, given the ex-pectation mechanism, it would not receive benefits when old (a newly installedsystem would be voted down right away, according to the generations’ expecta-tion). Hence

V (0, τet+1) = V (0, 0) = U(w1) + βU(w2)

Voting to keep the system in place yields

V (τ∗∗, τet+1) = V (τ∗∗, τ∗∗) = U(w1 − τ∗∗) + βU(w2 + (1 + n)τ

∗∗)

and a vote in favor requires

V (τ∗∗, τ∗∗) ≥ V (0, 0) (8.19)

But this is true for all generations, including the first regular generation. Giventhe assumption that we are in the Samuelson case with n > r there exists aτ∗∗ > 0 such that the above inequality holds. Hence the initial old generationcan introduce a positive social security system with τ0 = τ∗∗ > 0 that is notvoted down by the next generation (and hence by no generation) and createspositive transfers for itself. Obviously, then, the optimal choice is to maximizeτ0 = τ∗∗ subject to (8.19), and the equilibrium sequence of policies satisfiesτ t = τ∗∗ where τ∗∗ > 0 satisfies

U(w1 − τ∗∗) + βU(w2 + (1 + n)τ∗∗) = U(w1) + βU(w2)

Page 151: Macroeconomics Phd Lectures notes

8.2. THE RICARDIAN EQUIVALENCE HYPOTHESIS 145

8.2 The Ricardian Equivalence Hypothesis

How should the government finance a given stream of government expenditures,say, for a war? There are two principal ways to levy revenues for a govern-ment, namely to tax current generations or to issue government debt in theform of government bonds the interest and principal of which has to be paidlater.14 The question then arise what the macroeconomic consequences of usingthese different instruments are, and which instrument is to be preferred from anormative point of view. The Ricardian Equivalence Hypothesis claims that itmakes no difference, that a switch from one instrument to the other does notchange real allocations and prices in the economy. Therefore this hypothesis, isalso called Modigliani-Miller theorem of public finance.15 It’s origin dates backto the classical economist David Ricardo (1772-1823). He wrote about howto finance a war with annual expenditures of £20 millions and asked whetherit makes a difference to finance the £20 millions via current taxes or to issuegovernment bonds with infinite maturity (so-called consols) and finance the an-nual interest payments of £1 million in all future years by future taxes (at anassumed interest rate of 5%). His conclusion was (in “Funding System”) that

in the point of the economy, there is no real difference in eitherof the modes; for twenty millions in one payment [or] one million perannum for ever ... are precisely of the same value

Here Ricardo formulates and explains the equivalence hypothesis, but im-mediately makes clear that he is sceptical about its empirical validity

...but the people who pay the taxes never so estimate them, andtherefore do not manage their affairs accordingly. We are too apt tothink, that the war is burdensome only in proportion to what we areat the moment called to pay for it in taxes, without reflecting on theprobable duration of such taxes. It would be difficult to convincea man possessed of £20, 000, or any other sum, that a perpetualpayment of £50 per annum was equally burdensome with a singletax of £1, 000.

Ricardo doubts that agents are as rational as they should, according to “inthe point of the economy”, or that they rationally believe not to live foreverand hence do not have to bear part of the burden of the debt. Since Ricardodidn’t believe in the empirical validity of the theorem, he has a strong opinionabout which financing instrument ought to be used to finance the war

war-taxes, then, are more economical; for when they are paid, aneffort is made to save to the amount of the whole expenditure of the

14I will restrict myself to a discussion of real economic models, in which fiat money is absent.Hence the government cannot levy revenue via seignorage.15When we discuss a theoretical model, Ricardian equivalence will take the form of a theorem

that either holds or does not hold, depending on the assumptions we make. When discussingwhether Ricardian equivalence holds empirically, I will call it a hypothesis.

Page 152: Macroeconomics Phd Lectures notes

146 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

war; in the other case, an effort is only made to save to the amountof the interest of such expenditure.

Ricardo thought of government debt as one of the prime tortures of mankind.Not surprisingly he strongly advocates the use of current taxes. We will, afterhaving discussed the Ricardian equivalence hypothesis, briefly look at the long-run effects of government debt on economic growth, in order to evaluate whetherthe phobia of Ricardo (and almost all other classical economists) about govern-ment debt is in fact justified from a theoretical point of view. Now let’s turn toa model-based discussion of Ricardian equivalence.

8.2.1 Infinite Lifetime Horizon and Borrowing Constraints

The Ricardian Equivalence hypothesis is, in fact, a theorem that holds in a fairlywide class of models. It is most easily demonstrated within the Arrow-Debreumarket structure of infinite horizon models. Consider the simple infinite horizonpure exchange model discussed at the beginning of the section. Now introducea government that has to finance a given exogenous stream of government ex-penditures (in real terms) denoted by Gt∞t=1. These government expendituresdo not yield any utility to the agents (this assumption is not at all restrictivefor the results to come). Let pt denote the Arrow-Debreu price at date 0 ofone unit of the consumption good delivered at period t. The government hasinitial outstanding real debt16 of B1 that is held by the public. Let b

i1 denote

the initial endowment of government bonds of agent i. Obviously we have therestriction X

i∈Ibi1 = B1

Note that bi1 is agent i’s entitlement to period 1 consumption that the govern-ment owes to the agent. In order to finance the government expenditures thegovernment levies lump-sum taxes: let τ it denote the taxes that agent i paysin period t, denoted in terms of the period t consumption good. We define anArrow-Debreu equilibrium with government as follows

Definition 81 Given a sequence of government spending Gt∞t=1 and initialgovernment debt B1 and (b

i1)i∈I an Arrow-Debreu equilibrium are allocations

¡cit¢i∈I∞t=1, prices pt∞t=1 and taxes ¡τ it¢i∈I∞t=1 such that1. Given prices pt∞t=1 and taxes

¡τ it¢i∈I∞t=1 for all i ∈ I, cit∞t=1 solves

maxct∞t=1

∞Xt=1

βt−1U(cit) (8.20)

s.t.∞Xt=1

pt(ct + τ it) ≤∞Xt=1

pteit + p1b

i1

16I.e. the government owes real consumption goods to its citizens.

Page 153: Macroeconomics Phd Lectures notes

8.2. THE RICARDIAN EQUIVALENCE HYPOTHESIS 147

2. Given prices pt∞t=1 the tax policy satisfies∞Xt=1

ptGt + p1B1 =∞Xt=1

Xi∈Iptτ

it

3. For all t ≥ 1 Xi∈Icit +Gt =

Xi∈Ieit

In an Arrow-Debreu definition of equilibrium the government, as the agent,faces a single intertemporal budget constraint which states that the total valueof tax receipts is sufficient to finance the value of all government purchases plusthe initial government debt. From the definition it is clear that, with respect togovernment tax policies, the only thing that matters is the total value of taxesP∞t=1 ptτ

it that the individual has to pay, but not the timing of taxes. It is then

straightforward to prove the Ricardian Equivalence theorem for this economy.

Theorem 82 Take as given a sequence of government spending Gt∞t=1 andinitial government debt B1, (b

i1)i∈I . Suppose that allocations

¡cit¢i∈I∞t=1, prices

pt∞t=1 and taxes ¡τ it¢i∈I∞t=1 form an Arrow-Debreu equilibrium. Let

¡τ it¢i∈I∞t=1

be an arbitrary alternative tax system satisfying

∞Xt=1

ptτit =

∞Xt=1

ptτit for all i ∈ I

Then ¡cit¢i∈I∞t=1, pt∞t=1 and ¡τ it¢i∈I∞t=1 form an Arrow-Debreu equilib-rium.

There are two important elements of this theorem to mention. First, thesequence of government expenditures is taken as fixed and exogenously given.Second, the condition in the theorem rules out redistribution among individuals.It also requires that the new tax system has the same cost to each individual atthe old equilibrium prices (but not necessarily at alternative prices).Proof. This is obvious. The budget constraint of individuals does not

change, hence the optimal consumption choice at the old equilibrium pricesdoes not change. Obviously resource feasibility is satisfied. The governmentbudget constraint is satisfied due to the assumption made in the theorem.A shortcoming of the Arrow-Debreu equilibrium definition and the preced-

ing theorem is that it does not make explicit the substitution between currenttaxes and government deficits that may occur for two equivalent tax systems¡τ it¢i∈I∞t=1 and ¡τ it¢i∈I∞t=1. Therefore we will now reformulate this economysequentially. This will also allow us to see that one of the main assumptions ofthe theorem, the absence of borrowing constraints is crucial for the validity ofthe theorem.

Page 154: Macroeconomics Phd Lectures notes

148 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

As usual with sequential markets we now assume that markets for the con-sumption good and one-period loans open every period. We restrict ourselvesto government bonds and loans with one year maturity, which, in this environ-ment is without loss of generality (note that there is no uncertainty) and willnot distinguish between borrowing and lending between two agents an agentan the government. Let rt+1 denote the interest rate on one period loans fromperiod t to period t + 1. Given the tax system and initial bond holdings eachagent i now faces a sequence of budget constraints of the form

cit +bit+1

1 + rt+1≤ eit − τ it + b

it (8.21)

with bi1 given. In order to rule out Ponzi schemes we have to impose a no Ponzischeme condition of the form bit ≥ −ait(r, ei, τ) on the consumer, which, ingeneral may depend on the sequence of interest rates as well as the endowmentstream of the individual and the tax system. We will be more specific about theexact from of the constraint later. In fact, we will see that the exact specificationof the borrowing constraint is crucial for the validity of Ricardian equivalence.The government faces a similar sequence of budget constraints of the form

Gt +Bt =Xi∈I

τ it +Bt+11 + rt+1

(8.22)

with B1 given. We also impose a condition on the government that rules outgovernment policies that run a Ponzi scheme, or Bt ≥ −At(r,G, τ). The defini-tion of a sequential markets equilibrium is standard

Definition 83 Given a sequence of government spending Gt∞t=1 and initialgovernment debt B1, (b

i1)i∈I a Sequential Markets equilibrium is allocations

³cit, b

it+1

´i∈I∞t=1,

interest rates rt+1∞t=1 and government policies ¡τ it¢i∈I , Bt+1∞t=1 such that

1. Given interest rates rt+1∞t=1 and taxes ¡τ it¢i∈I∞t=1 for all i ∈ I, cit, bit+1∞t=1

maximizes (8.20) subject to (8.21) and bit+1 ≥ −ait(r, ei, τ) for all t ≥ 1.2. Given interest rates rt+1∞t=1, the government policy satisfies (8.22) andBt+1 ≥ −At(r, G) for all t ≥ 1

3. For all t ≥ 1 Xi∈Icit +Gt =

Xi∈IeitX

i∈Ibit+1 = Bt+1

We will particularly concerned with two forms of borrowing constraints. Thefirst is the so called natural borrowing or debt limit: it is that amount that, at

Page 155: Macroeconomics Phd Lectures notes

8.2. THE RICARDIAN EQUIVALENCE HYPOTHESIS 149

given sequence of interest rates, the consumer can maximally repay, by settingconsumption to zero in each period. It is given by

anit(r, e, τ) =∞Xτ=1

eit+τ − τ it+τQt+τ−1j=t+1 (1 + rj+1)

where we defineQtj=t+1(1 + rj+1) = 1. Similarly we set the borrowing limit of

the government at its natural limit

Ant(r, τ) =∞Xτ=1

Pi∈I τ

it+τQt+τ−1

j=t+1 (1 + rj+1)

The other form is to prevent borrowing altogether, setting a0it(r, e) = 0 for alli, t. Note that since there is positive supply of government bonds, such restrictiondoes not rule out saving of individuals in equilibrium. We can make full useof the Ricardian equivalence theorem for Arrow-Debreu economies one we haveproved the following equivalence result

Proposition 84 Fix a sequence of government spending Gt∞t=1 and initialgovernment debt B1, (b

i1)i∈I . Let allocations

¡cit¢i∈I∞t=1, prices pt∞t=1 and

taxes ¡τ it¢i∈I∞t=1 form an Arrow-Debreu equilibrium. Then there exists a cor-

responding sequential markets equilibrium with the natural debt limits ³cit, b

it+1

´i∈I∞t=1,

rt∞t=1, ¡τ it¢i∈I , Bt+1∞t=1 such that

cit = citτ it = τ it for all i, all t

Reversely, let allocations ³cit, b

it+1

´i∈I∞t=1, interest rates rt∞t=1 and govern-

ment policies ¡τ it¢i∈I , Bt+1∞t=1 form a sequential markets equilibrium with nat-ural debt limits. Suppose that it satisfies

rt+1 > −1, for all t ≥ 1∞Xt=1

eit − τ itQt−1j=1(1 + rj+1)

< ∞ for all i ∈ I∞Xτ=1

Pi∈I τ

it+τQt+τ

j=t+1(1 + rj+1)< ∞

Then there exists a corresponding Arrow-Debreu equilibrium ¡cit¢i∈I∞t=1, pt∞t=1,¡τ it¢i∈I∞t=1 such that

cit = citτ it = τ it for all i, all t

Page 156: Macroeconomics Phd Lectures notes

150 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

Proof. The key to the proof is to show the equivalence of the budget setsfor the Arrow-Debreu and the sequential markets structure. Normalize p1 = 1and relate equilibrium prices and interest rates by

1 + rt+1 =ptpt+1

(8.23)

Now look at the sequence of budget constraints and assume that they hold withequality (which they do in equilibrium, due to the nonsatiation assumption)

ci1 +bi2

1 + r2= ei1 − τ i1 + b

i1 (8.24)

ci2 +bi3

1 + r3= ei2 − τ i2 + b

i2 (8.25)

...

cit +bit+1

1 + rt+1= eit − τ it + b

it (8.26)

Substituting for bi2 from (8.25) in (8.24) one gets

ci1 + τ i1 − ei1 +ci2 + τ i2 − ei21 + r2

+bi3

(1 + r2)(1 + r3)= bi1

and in general

TXt=1

ct − etQt−1j=1(1 + rj+1)

+biT+1QT

j=1(1 + rj+1)= bi1

Taking limits on both sides gives, using (8.23)

∞Xt=1

pt(cit + τ it − eit) + lim

T→∞biT+1QT

j=1(1 + rj+1)= bi1

Hence we obtain the Arrow-Debreu budget constraint if and only if

limT→∞

biT+1QTj=1(1 + rj+1)

= limT→∞

pT+1biT+1 ≥ 0

But from the natural debt constraint

pT+1biT+1 ≥ −pT+1

∞Xτ=1

eit+τ − τ it+τQt+τ−1j=t+1 (1 + rj+1)

= −∞X

τ=T+1

pt(eiτ − τ iτ )

= −∞Xτ=1

pt(eiτ − τ iτ ) +

TXτ=1

pt(eiτ − τ iτ )

Page 157: Macroeconomics Phd Lectures notes

8.2. THE RICARDIAN EQUIVALENCE HYPOTHESIS 151

Taking limits with respect to both sides and using that by assumptionP∞t=1

eit−τ itQ t−1j=1(1+rj+1)

=P∞t=1 pt(e

iτ − τ iτ ) <∞ we have

limT→∞

pT+1biT+1 ≥ 0

So at equilibrium prices, with natural debt limits and the restrictions posedin the proposition a consumption allocation satisfies the Arrow-Debreu budgetconstraint (at equilibrium prices) if and only if it satisfies the sequence of budgetconstraints in sequential markets. A similar argument can be carried out forthe budget constraint(s) of the government. The remainder of the proof isthen straightforward and left to the reader. Note that, given an Arrow-Debreuequilibrium consumption allocation, the corresponding bond holdings for thesequential markets formulation are

bit+1 =∞Xτ=1

cit+τ + τ it+τ − eit+τQt+τ−1j=t+1 (1 + rj+1)

As a straightforward corollary of the last two results we obtain the Ricardianequivalence theorem for sequential markets with natural debt limits (under theweak requirements of the last proposition).17 Let us look at a few examples

Example 85 (Financing a war) Let the economy be populated by I = 1000identical people, with U(c) = ln(c), β = 0.5

eit = 1

and G1 = 500 (the war), Gt = 0 for all t > 1. Let b1 = B1 = 0. Consider twotax policies. The first is a balanced budget requirement, i.e. τ1 = 0.5, τ t = 0 forall t > 1. The second is a tax policy that tries to smooth out the cost of the war,i.e. sets τ t = τ = 1

3 for all t ≥ 1. Let us look at the equilibrium for the first taxpolicy. Obviously the equilibrium consumption allocation (we restrict ourselvesto type-identical allocations) has

cit =

½0.5 for t = 11 for t ≥ 1

and the Arrow-Debreu equilibrium price sequence satisfies (after normalizationof p1 = 1) p2 = 0.25 and pt = 0.25∗0.5t−2 for all t > 2. The level of governmentdebt and the bond holdings of individuals in the sequential markets economysatisfy

Bt = bt = 0 for all t

17An equivalence result with even less restrictive assumptions can be proved under thespecification of a bounded shortsale constraint

inftbit <∞

instead of the natural debt limit. See Huang and Werner (1998) for details.

Page 158: Macroeconomics Phd Lectures notes

152 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

Interest rates are easily computed as r2 = 3, rt = 1 for t > 2. The budget con-straint of the government and the agents are obviously satisfied. Now considerthe second tax policy. Given resource constraint the previous equilibrium alloca-tion and price sequences are the only candidate for an equilibrium under the newpolicy. Let’s check whether they satisfy the budget constraints of government andindividuals. For the government

∞Xt=1

ptGt + p1B1 =∞Xt=1

Xi∈Iptτ

it

500 =1

3

∞Xt=1

1000pt

=1000

3(1 + 0.25 +

∞Xt=3

0.25 ∗ 0.5t−2)

= 500

and for the individual

∞Xt=1

pt(ct + τ it) ≤∞Xt=1

pteit + p1b

i1

5

6+4

3

∞Xt=2

pt ≤∞Xt=1

pt

1

3

∞Xt=2

pt =1

6≤ 16

Finally, for this tax policy the sequence of government debt and private bondholdings are

Bt =2000

3, b2 =

2

3for all t ≥ 2

i.e. the government runs a deficit to finance the war and, in later periods,uses taxes to pay interest on the accumulated debt. It never, in fact, retires thedebt. As proved in the theorem both tax policies are equivalent as the equilibriumallocation and prices remain the same after a switch from tax to deficit financeof the war.

The Ricardian equivalence theorem rests on several important assumptions.The first is that there are perfect capital markets. If consumers face bindingborrowing constraints (e.g. for the specification requiring bit+1 ≥ 0), or if,with uncertainty, not a full set of contingent claims is available, then Ricardianequivalence may fail. Secondly one has to require that all taxes are lump-sum.Non-lump sum taxes may distort relative prices (e.g. labor income taxes distortthe relative price of leisure) and hence a change in the timing of taxes may

Page 159: Macroeconomics Phd Lectures notes

8.2. THE RICARDIAN EQUIVALENCE HYPOTHESIS 153

have real effects. All taxes on endowments, whatever form they take, are lump-sum, not, however consumption taxes. Finally a change from one to anothertax system is assumed to not redistribute wealth among agents. This was amaintained assumption of the theorem, which required that the total tax billthat each agent faces was left unchanged by a change in the tax system. In aworld with finitely lived overlapping generations this would mean that a changein the tax system is not supposed to redistribute the tax burden among differentgenerations.Now let’s briefly look at the effect of borrowing constraints. Suppose we

restrict agents from borrowing, i.e. impose bit+1 ≥ 0, for all i, all t. For thegovernment we still impose the old restriction on debt, Bt ≥ −Ant(r, τ). Wecan still prove a limited Ricardian result

Proposition 86 Let Gt∞t=1 and B1, (bi1)i∈I be given and let allocations ³cit, b

it+1

´i∈I∞t=1,

interest rates rt+1∞t=1 and government policies ¡τ it¢i∈I , Bt+1∞t=1 be a Se-

quential Markets equilibrium with no-borrowing constraints for which bit+1 > 0

for all i, t. Let ¡τ it¢i∈I , Bt+1∞t=1 be an alternative government policy such thatbit+1 =

∞Xτ=t+1

ciτ + τ iτ − eiτQτj=t+2(1 + rj+1)

≥ 0 (8.27)

Gt + Bt =Xi∈I

τ it +Bt+11 + rt+1

for all t (8.28)

Bt+1 ≥ −Ant(r, τ) (8.29)∞Xτ=1

τ iτQτ−1j=1 (1 + rj)

=∞Xτ=1

τ iτQτ−1j=1 (1 + rj+1)

(8.30)

Then ³cit, b

it+1

´i∈I∞t=1, rt+1∞t=1 and

¡τ it¢i∈I , Bt+1∞t=1 is also a sequential

markets equilibrium with no-borrowing constraint.

The conditions that we need for this theorem are that the change in the taxsystem is not redistributive (condition (8.30)), that the new government policiessatisfy the government budget constraint and debt limit (conditions (8.28) and(8.29)) and that the new bond holdings of each individual that are required tosatisfy the budget constraints of the individual at old consumption allocationsdo not violate the no-borrowing constraint (condition (8.27)).Proof. This proposition to straightforward to prove so we will sketch it

here only. Budget constraints of the government and resource feasibility areobviously satisfied under the new policy. How about consumer optimization?Given the equilibrium prices and under the imposed conditions both policiesinduce the same budget set of individuals. Now suppose there is an i andallocation cit 6= cit that dominates cit. Since cit was affordable with theold policy, it must be the case that the associated bond holdings under theold policy, bit+1 violated one of the no-borrowing constraints. But then, by

Page 160: Macroeconomics Phd Lectures notes

154 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

continuity of the price functional and the utility function there is an allocationcit with associated bond holdings bit+1 that is affordable under the old policyand satisfies the no-borrowing constraint (take a convex combination of the

cit, bit+1 and the cit, bit+1, with sufficient weight on the cit, bit+1 so as tosatisfy the no-borrowing constraints). Note that for this to work it is crucial that

the no-borrowing constraints are not binding under the old policy for cit, bit+1.You should fill in the mathematical detailsLet us analyze an example in which, because of the borrowing constraints,

Ricardian equivalence fails.

Example 87 Consider an economy with 2 agents, U i = ln(c), βi = 0.5, bi1 =

B1 = 0. Also Gt = 0 for all t and endowments are

e1t =

½2 if t odd1 if t even

e2t =

½1 if t odd2 if t even

As first tax system consider

τ1t =

½0.5 if t odd−0.5 if t even

e2t =

½ −0.5 if t odd0.5 if t even

Obviously this tax system balances the budget. The equilibrium allocation withno-borrowing constraints evidently is the autarkic (after-tax) allocation cit =1.5, for all i, t. From the first order conditions we obtain, taking account thenonnegativity constraint on bit+1 (here λt ≥ 0 is the Lagrange multiplier onthe budget constraint in period t and µt+1 is the Lagrange multiplier on thenonnegativity constraint for bit+1)

βt−1U 0(cit) = λt

βtU(cit+1) = λt+1

λt1 + rt+1

= λt+1 + µt+1

Combining yields

U 0(cit)βU 0(cit+1)

=λtλt+1

= 1 + rt+1 +(1 + rt)µt+1

λt+1

Hence

U 0(cit)βU 0(cit+1)

≥ 1 + rt+1

= 1 + rt+1 if bit+1 > 0

Page 161: Macroeconomics Phd Lectures notes

8.2. THE RICARDIAN EQUIVALENCE HYPOTHESIS 155

The equilibrium interest rates are given as rt+1 ≤ 1, i.e. are indeterminate.Both agents are allowed to save, and at rt+1 > 1 they would do so (which ofcourse can’t happen in equilibrium as there is zero net supply of assets). Forany rt+1 ≤ 1 the agents would like to borrow, but are prevented from doing so bythe no-borrowing constraint, so any of these interest rates is fine as equilibriuminterest rates. For concreteness let’s take rt+1 = 1 for all t.18 Then the totalbill of taxes for the first consumer is 1

3 and −13 for the second agent. Now letsconsider a second tax system that has τ11 =

13 , τ

21 = −13 and τ it = 0 for all

i, t ≥ 2. Obviously now the equilibrium allocation changes to c1t =53 , c

21 =

43 and

cit = eit for all i, t ≥ 2. Obviously the new tax system satisfies the governmentbudget constraint and does not redistribute among agents. However, equilibriumallocations change. Furthermore, equilibrium interest rate change to r2 =

32.5

and rt = 0 for all t ≥ 3. Ricardian equivalence fails.19

8.2.2 Finite Horizon and Operative Bequest Motives

It should be clear from the above discussion that one only obtains a very limitedRicardian equivalence theorem for OLG economies. Any change in the timingof taxes that redistributes among generations is in general not neutral in theRicardian sense. If we insist on representative agents within one generation andpurely selfish, two-period lived individuals, then in fact any change in the timingof taxes can’t be neutral unless it is targeted towards a particular generation,i.e. the tax change is such that it decreases taxes for the currently young onlyand increases them for the old next period. Hence, with sufficient generalitywe can say that Ricardian equivalence does not hold for OLG economies withpurely selfish individuals.

Rather than to demonstrate this obvious point with another example we nowbriefly review Barro’s (1974) argument that under certain conditions finitelylived agents will behave as if they had infinite lifetime. As a consequence,Ricardian equivalence is re-established. Barro’s (1974) article “Are Govern-ment Bonds Net Wealth?” asks exactly the Ricardian question, namely doesan increase in government debt, financed by future taxes to pay the interest onthe debt increase the net wealth of the private sector? If yes, then current con-sumption would increase, aggregate saving (private plus public) would decrease,leading to an increase in interest rate and less capital accumulation. Dependingon the perspective, countercyclical fiscal policy20 is effective against the businesscycle (the Keynesian perspective) or harmful for long term growth (the classicalperspective). If, however, the value of government bonds if completely offset by

18These are the interest rates that would arise under natural debt limits, too.19In general it is very hard to solve for equilibria with no-borrowing constraints analytically,

even in partial equilibrium with fixed exogenous interest rates, even more so in gneral equi-librium. So if the above example seems cooked up, it is, since it is about the only example Iknow how to solve without going to the computer. We will see this more explicitly once wetalk about Deaton’s (1991) EC piece.20By fiscal policy in this section we mean the financing decision of the government for a

given exogenous path of government expenditures.

Page 162: Macroeconomics Phd Lectures notes

156 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

the value of future higher taxes for each individual, then government bonds arenot net wealth of the private sector, and changes in fiscal policy are neutral.Barro identified two main sources for why future taxes are not exactly off-

setting current tax cuts (increasing government deficits): a) finite lives of agentsthat lead to intergenerational redistribution caused by a change in the timingof taxes b) imperfect private capital markets. Barro’s paper focuses on the firstsource of nonneutrality.Barro’s key result is the following: in OLG-models finiteness of lives does not

invalidate Ricardian equivalence as long as current generations are connectedto future generations by a chain of operational intergenerational, altruisticallymotivated transfers. These may be transfers from old to young via bequestsor from young to old via social security programs. Let us look at his formalmodel.21

Consider the standard pure exchange OLG model with two-period livedagents. There is no population growth, so that each member of the old genera-tion (whose size we normalize to 1) has exactly one child. Agents have endow-ment ett = w when young and no endowment when old. There is a governmentthat, for simplicity, has 0 government expenditures but initial outstanding gov-ernment debt B. This debt is denominated in terms of the period 1 (or anyother period) consumption good. The initial old generation is endowed withthese B units of government bonds. We assume that these government bondsare zero coupon bonds with maturity of one period. Further we assume thatthe government keeps its outstanding government debt constant and we assumea constant one-period real interest rate r on these bonds.22 In order to financethe interest payments on government debt the government taxes the currentlyyoung people. The government budget constraint gives

B

1 + r+ τ = B

The right hand side is the old debt that the government has to retire in thecurrent period. On the left hand side we have the revenue from issuing newdebt, B

1+r (remember that we assume zero coupon bonds, so11+r is the price of

one government bond today that pays 1 unit of the consumption good tomorrow)and the tax revenue. With the assumption of constant government debt we find

τ =rB

1 + r

and we assume rB1+r < w.

Now let’s turn to the budget constraints of the individuals. Let by att denotethe savings of currently young people for the second period of their lives and byatt+1 denote the savings of the currently old people for the next generation, i.e.

21I will present a simplified, pure exchange version of his model to more clearly isolate hismain point.22This assumption is justified since the resulting equilibrium allocation (there is no money!)

is the autarkic allocation and hence the interest rate always equals the autarkic interest rate.

Page 163: Macroeconomics Phd Lectures notes

8.2. THE RICARDIAN EQUIVALENCE HYPOTHESIS 157

the old people’s bequests. We require bequests to be nonnegative, i.e. att+1 ≥ 0.In our previous OLG models obviously att+1 = 0 was the only optimal choicesince individuals were completely selfish. We will see below how to induce pos-itive bequests when discussing individuals’ preferences. The budget constraintsof a representative generation are then given by

ctt +att1 + r

= w − τ

ctt+1 +att+11 + r

= att + at−1t

The budget constraint of the young are standard; one may just remember that

assets here are zero coupon bonds: spendingatt1+r on bonds in the current period

yields att units of consumption goods tomorrow. We do not require att to be

positive. When old the individuals have two sources of funds: their own savingsfrom the previous period and the bequests at−1t from the previous generation.They use it to buy own consumption and bequests for the next generation.

The total expenditure for bequests of a currently old individual isatt+11+r and it

delivers funds to her child next period (that has then become old) of att+1. Wecan consolidate the two budget constraints to obtain

ctt +ctt+11 + r

+att+1

(1 + r)2 = w +

at−1t

1 + r− τ

Since the total lifetime resources available to generation t are given by et =

w +at−1t

1+r − τ , the lifetime utility that this generation can attain is determinedby e. The budget constraint of the initial old generation is given by

c01 +a011 + r

= B

With the formulation of preferences comes the crucial twist of Barro. Heassumes that individuals are altruistic and care about the well-being of theirdescendant.23 Altruistic here means that the parents genuinely care about theutility of their children and leave bequests for that reason; it is not that theparents leave bequests in order to induce actions of the children that yieldutility to the parents.24 Preferences of generation t are represented by

ut(ctt, c

tt+1, a

tt+1) = U(c

tt) + βU(ctt+1) + αVt+1(et+1)

where Vt+1(et+1) is the maximal utility generation t+1 can attain with lifetime

resources et+1 = w+att+11+r −τ , which are evidently a function of bequests att+1from

23Note that we only assume that the agent cares only about her immediate descendant, but(possibly) not at all about grandchildren.24This strategic bequest motive does not necessarily help to reestablish Ricardian equiva-

lence, as Bernheim, Shleifer and Summers (1985) show.

Page 164: Macroeconomics Phd Lectures notes

158 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

generation t.25 We make no assumption about the size of α as compared to β,but assume α ∈ (0, 1). The initial old generation has preferences represented by

u0(c01, a

01) = βU(c01) + αV1(e1)

The equilibrium conditions for the goods and the asset market are, respec-tively

ct−1t + ctt = w for all t ≥ 1at−1t + att = B for all t ≥ 1

Now let us look at the optimization problem of the initial old generation

V0(B) = maxc01,a

01≥0

©βU(c01) + αV1(e1)

ªs.t. c01 +

a011 + r

= B

e1 = w +a011 + r

− τ

Note that the two constraints can be consolidated to

c01 + e1 = w +B − τ (8.31)

This yields optimal decision rules c01(B) and a01(B) (or e1(B)). Now assume

that the bequest motive is operative, i.e. a01(B) > 0 and consider the Ricar-dian experiment of government: increase initial government debt marginally by∆B and repay this additional debt by levying higher taxes on the first younggeneration. Clearly, in the OLG model without bequest motives such a changein fiscal policy is not neutral: it increases resources available to the initial oldand reduces resources available to the first regular generation. This will changeconsumption of both generations and interest rate. What happens in the Barroeconomy? In order to repay the ∆B, from the government budget constrainttaxes for the young have to increase by

∆τ = ∆B

since by assumption government debt from the second period onwards remainsunchanged. How does this affect the optimal consumption and bequest choiceof the initial old generation? It is clear from (8.31) that the optimal choices forc01 and e1 do not change as long as the bequest motive was operative before.

26

25To formulate the problem recurively we need separability of the utility function withrespect to time and utility of children. The argument goes through without this, but thenit can’t be clarified using recursive methods. See Barro’s original paper for a more generaldiscussion. Also note that he, in all likelihood, was not aware of the full power of recursivetechniques in 1974. Lucas (1972) seminal paper was probably the first to make full use ofrecursive techiques in (macro) economics.26If the bequest motive was not operative, i.e. if the constraint a01 ≥ 0 was binding, then

by increasing B may result in an increase in c01 and a decrease in e1.

Page 165: Macroeconomics Phd Lectures notes

8.2. THE RICARDIAN EQUIVALENCE HYPOTHESIS 159

The initial old generation receives additional transfers of bonds of magnitude∆B from the government and increases its bequests a01 by (1 + r)∆B so thatlifetime resources available to their descendants (and hence their allocation) isleft unchanged. Altruistically motivated bequest motives just undo the changein fiscal policy. Ricardian equivalence is restored.This last result was just an example. Now let’s show that Ricardian equiva-

lence holds in general with operational altruistic bequests. In doing so we will defacto establish between Barro’s OLG economy and an economy with infinitelylived consumers and borrowing constraints. Again consider the problem of theinitial old generation (and remember that, for a given tax rate and wage thereis a one-to-one mapping between et+1 and a

tt+1

V0(B) = maxc01, a

01 ≥ 0

c01 +a011+r = B

©βU(c01) + αV1(a

01)ª

= maxc01, a

01 ≥ 0

c01 +a011+r = B

βU(c01) + α max

c11, c12, a

12 ≥ 0, a11

c11 +a111+r = w − τ

c12 +a121+r = a

11 + a

01

©U(c11) + βU(c12) + αV2(a

12)ª

But this maximization problem can be rewritten as

maxc01,a

01,c

11,c

12,a

12≥0,a11

©βU(c01) + αU(c11) + αβU(c12) + α2V2(a

12)ª

s.t. c01 +a011 + r

= B

c11 +a111 + r

= w − τ

c12 +a121 + r

= a11 + a01

or, repeating this procedure infinitely many times (which is a valid procedureonly for α < 1), we obtain as implied maximization problem of the initial oldgeneration

max(ct−1t ,ctt,a

t−1t )∞t=1≥0

(βU(c01) +

∞Xt=1

αt¡U(ctt

¢+ βU(ctt+1))

)

s.t. c01 +a011 + r

= B

ctt +ctt+11 + r

+att+1

(1 + r)2= w − τ +

at−1t

1 + r

Page 166: Macroeconomics Phd Lectures notes

160 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

i.e. the problem is equivalent to that of an infinitely lived consumer that faces ano-borrowing constraint. This infinitely lived consumer is peculiar in the sensethat her periods are subdivided into two subperiods, she eats twice a period,ctt in the first subperiod and c

tt+1 in the second subperiod, and the relative

price of the consumption goods in the two subperiods is given by (1+ r). Apartfrom these reinterpretations this is a standard infinitely lived consumer withno-borrowing constraints imposed on her. Consequently one obtains a Ricar-dian equivalence proposition similar to proposition 86, where the requirementof “operative bequest motives” is the equivalent to condition (8.27). More gen-erally, this argument shows that an OLG economy with two period-lived agentsand operative bequest motives is formally equivalent to an infinitely lived agentmodel.

Example 88 Suppose we carry out the Ricardian experiment and increase ini-tial government debt by ∆B. Suppose the debt is never retired, but the requiredinterest payments are financed by permanently higher taxes. The tax increasethat is needed is (see above)

∆τ =r∆B

1 + r

Suppose that for the initial debt level (ct−1t , ctt, at−1t )∞t=1 together with r is an

equilibrium such that at−1t > 0 for all t. It is then straightforward to verifythat (ct−1t , ctt, a

t−1t )∞t=1 together with r is an equilibrium for the new debt level,

where

at−1t = at−1t + (1 + r)∆B for all t

i.e. in each period savings increase by the increased level of debt, plus the pro-vision for the higher required tax payments. Obviously one can construct muchmore complicated tax experiments that are neutral in the Ricardian sense, pro-vided that for the original tax system the non-borrowing constraints never bind(i.e. that bequest motives are always operative). Also note that Barro discussedhis result in the context of a production economy, an issue to which we turnnext.

8.3 Overlapping Generations Models with Pro-duction

So far we have ignored production in our discussion of OLG-models. It maybe the case that some of the pathodologies of the OLG-model appear only inpure exchange versions of the model. Since actual economies feature capitalaccumulation and production, these pathodologies then are nothing to worryabout. However, we will find out that, for example, the possibility of inefficientcompetitive equilibria extends to OLG models with production. The issues ofwhether money may have positive value and whether there exists a continuumof equilibria are not easy for production economies and will not be discussed inthese notes.

Page 167: Macroeconomics Phd Lectures notes

8.3. OVERLAPPING GENERATIONS MODELS WITH PRODUCTION161

8.3.1 Basic Setup of the Model

As much as possible I will synchronize the discussion here with the discrete timeneoclassical growth model in Chapter 2 and the pure exchange OLG model inprevious subsections. The economy consists of individuals and firms. Individu-als live for two periods By N t

t denote the number of young people in period t,by N t−1

t denote the number of old people at period t. Normalize the size of theinitial old generation to 1, i.e. N0

0 = 1.We assume that people do not die early,so N t

t = Ntt+1. Furthermore assume that the population grows at constant rate

n, so that N tt = (1 + n)tN0

0 = (1 + n)t. The total population at period t istherefore given by N t−1

t +N tt = (1 + n)

t(1 + 11+n).

The representative member of generation t has preferences over consumptionstreams given by

u(ctt, ctt+1) = U(c

tt) + βU(ctt+1)

where U is strictly increasing, strictly concave, twice continuously differentiableand satisfies the Inada conditions. All individuals are assumed to be purelyselfish and have no bequest motives whatsoever. The initial old generation haspreferences

u(c01) = U(c01)

Each individual of generation t ≥ 1 has as endowments one unit of time to workwhen young and no endowment when old. Hence the labor force in period t isof size N t

t with maximal labor supply of 1 ∗N tt . Each member of the initial old

generation is endowed with capital stock (1 + n)k1 > 0.Firms has access to a constant returns to scale technology that produces

output Yt using labor input Lt and capital input Kt rented from householdsi.e. Yt = F (Kt, Lt). Since firms face constant returns to scale, profits are zeroin equilibrium and we do not have to specify ownership of firms. Also withoutloss of generality we can assume that there is a single, representative firm, that,as usual, behaves competitively in that it takes as given the rental prices offactor inputs (rt, wt) and the price for its output. Defining the capital-laborratio kt =

Kt

Ltwe have by constant returns to scale

yt =YtLt=F (Kt, Lt)

Lt= F

µKt

Lt, 1

¶= f(kt)

We assume that f is twice continuously differentiable, strictly concave and sat-isfies the Inada conditions.

8.3.2 Competitive Equilibrium

The timing of events for a given generation t is as follows

1. At the beginning of period t production takes place with labor of genera-tion t and capital saved by the now old generation t− 1 from the previousperiod. The young generation earns a wage wt

Page 168: Macroeconomics Phd Lectures notes

162 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

2. At the end of period t the young generation decides how much of the wageincome to consume, ctt, and how much to save for tomorrow, s

tt. The saving

occurs in form of physical capital, which is the only asset in this economy

3. At the beginning of period t + 1 production takes place with labor ofgeneration t + 1 and the saved capital of the now old generation t. Thereturn on savings equals rt+1 − δ, where again rt+1 is the rental rate ofcapital and δ is the rate of depreciation, so that rt+1−δ is the real interestrate from period t to t+ 1.

4. At the end of period t+ 1generation t consumes its savings plus interestrate, i.e. ctt+1 = (1 + rt+1 − δ)stt and then dies.

We now can define a sequential markets equilibrium for this economy

Definition 89 Given k1, a sequential markets equilibrium is allocations forhouseholds c01, (ctt, ctt+1, stt)∞t=1, allocations for the firm (Kt, Lt)∞t=1 and prices(rt, wt)∞t=1 such that1. For all t ≥ 1, given (wt, rt+1), (ctt, ctt+1, stt) solves

maxctt,c

tt+1≥0,stt

U(ctt) + βU(ctt+1)

s.t. ctt + stt ≤ wt

ctt+1 ≤ (1 + rt+1 − δ)stt

2. Given k1 and r1, c01 solves

maxc01≥0

U(c01)

s.t. c01 ≤ (1 + r1 − δ)k1

3. For all t ≥ 1, given (rt, wt), (Kt, Lt) solves

maxKt,Lt≥0

F (Kt, Lt)− rtKt − wtLt

4. For all t ≥ 1(a) (Goods Market)

N tt ctt +N

t−1t ct−1t + Kt+1 − (1− δ)Kt = F (Kt, Lt)

(b) (Asset Market)

N tt stt = Kt+1

(c) (Labor Market)

N tt = Lt

Page 169: Macroeconomics Phd Lectures notes

8.3. OVERLAPPING GENERATIONS MODELS WITH PRODUCTION163

The first two points in the equilibrium definition are completely standard,apart from the change in the timing convention for the interest rate. For firmmaximization we used the fact that, given that the firm is renting inputs ineach period, the firms intertemporal maximization problem separates into asequence of static profit maximization problems. The goods market equilibriumcondition is standard: total consumption plus gross investment equals output.The labor market equilibrium condition is obvious. The asset or capital marketequilibrium condition requires a bit more thought: it states that total savingof the currently young generation makes up the capital stock for tomorrow,since physical capital is the only asset in this economy. Alternatively think ofit as equating the total supply of capital in form the saving done by the nowyoung, tomorrow old generation and the total demand for capital by firms nextperiod.27 It will be useful to single out particular equilibria and attach a certainname to them.

Definition 90 A steady state (or stationary equilibrium) is (k, s, c1, c2, r, w)such that the sequences c01, (ctt, ctt+1, stt)∞t=1, (Kt, Lt)∞t=1 and (rt, wt)∞t=1,defined by

ctt = c1

ct−1t = c2

stt = s

rt = r

wt = w

Kt = k ∗N tt

Lt = N tt

are an equilibrium, for given initial condition k1 = k.

In other words, a steady state is an equilibrium for which the allocation(per capita) is constant over time, given that the initial condition for the initialcapital stock is exactly right. Alternatively it is allocations and prices thatsatisfy all the equilibrium conditions apart from possibly obeying the initialcondition.We can use the goods and asset market equilibrium to derive an equation

that equates saving to investment. By definition gross investment equals Kt+1−(1− δ)Kt, whereas savings equals that part of income that is not consumed, or

Kt+1 − (1− δ)Kt = F (Kt, Lt)−¡N tt ctt +N

t−1t ct−1t

¢But what is total saving equal to? The currently young save N t

t stt, the currently

old dissave st−1t−1Nt−1t−1 = (1 − δ)Kt (they sell whatever capital stock they have

27To define an Arrow-Debreu equilibrium is quite standard here. Let pt the price of theconsumption good at period t, rtpt the nominal rental price of capital and wtpt the nominalwage. Then the household and the firms problems are in the neoclassical growth model, inthe household problem taking into account that agents only live for two periods.

Page 170: Macroeconomics Phd Lectures notes

164 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

left).28 Hence setting investment equal to saving yields

Kt+1 − (1− δ)Kt = Ntt stt − (1− δ)Kt

or our asset market equilibrium condition

N tt stt = Kt+1

Now let us start to characterize the equilibrium It will turn out that we candescribe the equilibrium completely by a first order difference equation in thecapital-labor ratio kt. Unfortunately it will have a rather nasty form in general,so that we can characterize analytic properties of the competitive equilibriumonly very partially. Also note that, as we will see later, the welfare theoremsdo not apply so that there is no social planner problem that will make our liveseasier, as was the case in the infinitely lived consumer model (which I dubbedthe discrete-time neoclassical growth model in Section 3).From now on we will omit the hats above the variables indicating equilibrium

elements as it is understood that the following analysis applies to equilibriumsequences. From the optimization condition for capital for the firm we obtain

rt = FK(Kt, Lt) = FK

µKtLt, 1

¶= f 0(kt)

because partial derivatives of functions that are homogeneous of degree 1 arehomogeneous of degree zero. Since we have zero profits in equilibrium we findthat

wtLt = F (Kt, Lt)− rtKtand dividing by Lt we obtain

wt = f(kt)− f 0(kt)kti.e. factor prices are completely determined by the capital-labor ratio. Investi-gating the households problem we see that its solution is completely character-ized by a saving function (note that given our assumptions on preferences theoptimal choice for savings exists and is unique)

stt = s (wt, rt+1)

= s (f(kt)− f 0(kt)kt, f 0(kt+1))so optimal savings are a function of this and next period’s capital stock. Ob-viously, once we know stt we know ctt and c

tt+1 from the household’s budget

28By definition the saving of the old is their total income minus their total consumption.Their income consists of returns on their assets and hence their total saving ish

(rtst−1t−1 − ct−1t

iNt−1t−1

= −(1− δ)st−1t−1Nt−1t−1 = −(1− δ)Kt

Page 171: Macroeconomics Phd Lectures notes

8.3. OVERLAPPING GENERATIONS MODELS WITH PRODUCTION165

constraint. From Walras law one of the market clearing conditions is redun-dant. Equilibrium in the labor market is straightforward as

Lt = Ntt = (1 + n)

t

So let’s drop the goods market equilibrium condition.29 Then the only con-dition left to exploit is the asset market equilibrium condition

sttNtt = Kt+1

stt =Kt+1

N tt

=N t+1t+1

N tt

Kt+1

N t+1t+1

= (1 + n)Kt+1

Lt+1= (1 + n)kt+1

Substituting in the savings function yields our first order difference equation

kt+1 =s (f(kt)− f 0(kt)kt, f 0(kt+1))

1 + n(8.32)

where the exact form of the saving function obviously depends on the functionalform of the utility function U. As starting value for the capital-labor ratio we

have K1

L1= (1+n)k1

N11

= k1. So in principle we could put equation (8.32) on a

computer and solve for the entire sequence of kt+1∞t=1 and hence for the entireequilibrium. Note, however, that equation (8.32) gives kt+1 only as an implicitfunction of kt as kt+1 appears on the right hand side of the equation as well. Solet us make an attempt to obtain analytical properties of this equation. Before,let’s solve an example.

Example 91 Let U(c) = ln(c), n = 0,β = 1 and f(k) = kα, with α ∈ (0, 1).The choice of log-utility is particularly convenient as the income and substitutioneffects of an interest change cancel each other out; saving is independent of rt+1.As we will see later it is crucial whether the income or substitution effect for aninterest change dominates in the saving decision, i.e. whether

srt+1(wt, rt+1) Q 0

But let’s proceed. The saving function for the example is given by

s(wt, rt+1) =1

2wt

=1

2(kαt − αkαt )

=1− α

2kαt

29In the homework you are asked to do the analysis with dropping the asset market insteadof the goods market equilibrium condition. Keep the present analysis in mind when doingthis question.

Page 172: Macroeconomics Phd Lectures notes

166 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

so that the difference equation characterizing the dynamic equilibrium is givenby

kt+1 =1− α

2kαt

There are two steady states for this differential equation, k0 = 0 and k∗ =¡1−α2

¢ 11−α . The first obviously is not an equilibrium as interest rates are infinite

and no solution to the consumer problem exists. From now on we will ignore thissteady state, not only for the example, but in general. Hence there is a uniquesteady state equilibrium associated with k∗. From any initial condition k1 > 0,there is a unique dynamic equilibrium kt+1∞t=1 converging to k∗ described bythe first order difference equation above.

Unfortunately things are not always that easy. Let us return to the generalfirst order difference equation (8.32) and discuss properties of the saving func-tion. Let, us for simplicity, assume that the saving function s is differentiablein both arguments (wt, rt+1).

30 Since the saving function satisfies the first ordercondition

U 0(wt − s(wt, rt+1)) = βU 0((1 + rt+1 − δ)s(wt, rt+1)) ∗ (1 + rt+1 − δ)

we use the Implicit Function Theorem (which is applicable in this case) to obtain

swt(wt, rt+1) =U 00(wt − s(wt, rt+1))

U 00(wt − s(wt, rt+1)) + βU 00((1 + rt+1 − δ)s(wt, rt+1))(1 + rt+1 − δ)2∈ (0, 1)

srt+1(wt, rt+1) =−βU 0((1 + rt+1 − δ)s(., .))− βU 00((1 + rt+1 − δ)s(., .))(1 + rt+1 − δ)s(., .)

U 00(wt − s(., .)) + βU 00((1 + rt+1 − δ)s(., .))(1 + rt+1 − δ)2R 0

Given our assumptions optimal saving increases in first period income wt, butit may increase or decrease in the interest rate. You may verify from the aboveformula that indeed for the log-case srt+1(wt, rt+1) = 0. A lot of theoretical workfocused on the case in which the saving function increases with the interest rate,which is equivalent to saying that the substitution effect dominates the incomeeffect (and equivalent to assuming that consumption in the two periods are strictgross substitutes).

Equation (8.32) traces out a graph in (kt, kt+1) space whose shape we want tocharacterize. Differentiating both sides of (8.32) with respect to kt we obtain

31

dkt+1dkt

=−swt(wt, rt+1)f 00(kt)kt + srt+1(wt, rt+1)f 0(kt+1)dkt+1dkt

1 + n

30One has to invoke the implicit function theorem (and check its conditions) on the firstorder condition to insure differentiability of the savings function. See Mas-Colell et al. p.940-942 for details.31Again we appeal to the Implicit function theorem that guarantees that kt+1 is a differen-

tiable function of kt with derivative given below.

Page 173: Macroeconomics Phd Lectures notes

8.3. OVERLAPPING GENERATIONS MODELS WITH PRODUCTION167

or rewriting

dkt+1dkt

=−swt(wt, rt+1)f 00(kt)kt

1 + n− srt+1(wt, rt+1)f 00(kt+1)Given our assumptions on f the nominator of the above expression is strictlypositive for all kt > 0. If we assume that srt+1 ≥ 0, then the (kt, kt+1)-locus isupward sloping. If we allow srt+1 < 0, then it may be downward sloping.

Case A

Case B

Case C

45-degree line

kt+1

k* k* k** kB C B t

Figure 8.6:

Figure 13 shows possible shapes of the (kt, kt+1)-locus under the assump-tion that srt+1 ≥ 0. We see that even this assumption does not place a lotof restrictions on the dynamic behavior of our economy. Without further as-sumptions it may be the case that, as in case A there is no steady state withpositive capital-labor ratio. Starting from any initial capital-per worker levelthe economy converges to a situation with no production over time. It may bethat, as in case C, there is a unique positive steady state k∗C and this steady

Page 174: Macroeconomics Phd Lectures notes

168 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

state is globally stable (for state space excluding 0). Or it is possible that thereare multiple steady states which alternate in being locally stable (as k∗B) andunstable (as k∗∗B ) as in case B. Just about any dynamic behavior is possible andin order to deduce further qualitative properties we must either specify specialfunctional forms or make assumptions about endogenous variables, somethingthat one should avoid, if possible.We will proceed however, doing exactly this. For now let’s assume that there

exists a unique positive steady state. Under what conditions is this steady statelocally stable? As suggested by Figure 13 stability requires that the savinglocus intersects the 450-line from above, provided the locus is upward sloping.A necessary and sufficient condition for local stability at the assumed uniquesteady state k∗ is that¯ −swt(w(k∗), r(k∗))f 00(k∗)k∗

1 + n− srt+1(w(k∗), r(k∗))f 00(k∗)¯< 1

If srt+1 < 0 it may be possible that the slope of the saving locus is negative.Under the condition above the steady state is still locally stable, but it exhibitsoscillatory dynamics. If we require that the unique steady state is locally stableand that the dynamic equilibrium is characterized by monotonic adjustment tothe unique steady state we need as necessary and sufficient condition

0 <−swt(w(k∗), r(k∗))f 00(k∗)k∗

1 + n− srt+1(w(k∗), r(k∗))f 00(k∗)< 1

The procedure to make sufficient assumptions that guarantee the existence ofa well-behaved dynamic equilibrium and then use exactly these assumption todeduce qualitative comparative statics results (how does the steady state changeas we change δ, n or the like) is called Samuelson’s correspondence principle,as often exactly the assumptions that guarantee monotonic local stability aresufficient to draw qualitative comparative statics conclusions. Diamond (1965)uses Samuelson’s correspondence principle extensively and we will do so, too,assuming from now on that above inequalities hold.

8.3.3 Optimality of Allocations

Before turning to Diamond’s (1965) analysis of the effect of public debt let usdiscuss the dynamic optimality properties of competitive equilibria. Considerfirst steady state equilibria. Let c∗1, c

∗2 be the steady state consumption levels

when young and old, respectively, and k∗ be the steady state capital labor ratio.Consider the goods market clearing (or resource constraint)

N tt ctt +N

t−1t ct−1t + Kt+1 − (1− δ)Kt = F (Kt, Lt)

Divide by N tt = Lt to obtain

ctt +ct−1t

1 + n+ (1 + n)kt+1 − (1− δ)kt = f(kt) (8.33)

Page 175: Macroeconomics Phd Lectures notes

8.3. OVERLAPPING GENERATIONS MODELS WITH PRODUCTION169

and use the steady state allocations to obtain

c∗1 +c∗21 + n

+ (1 + n)k∗ − (1− δ)k∗ = f(k∗)

Define c∗ = c∗1 +c∗21+n to be total (per worker) consumption in the steady state.

We have that

c∗ = f(k∗)− (n+ δ)k∗

Now suppose that the steady state equilibrium satisfies

f 0(k∗)− δ < n (8.34)

something that may or may not hold, depending on functional forms and pa-rameter values. We claim that this steady state is not Pareto optimal. Theintuition is as follows. Suppose that (8.34) holds. Then it is possible to de-crease the capital stock per worker marginally, and the effect on per capitaconsumption is

dc∗

dk∗= f 0(k∗)− (n+ δ) < 0

so that a marginal decrease of the capital stock leads to higher available over-all consumption. The capital stock is inefficiently high; it is so high that itsmarginal productivity f 0(k∗) is outweighed by the cost of replacing depreciatedcapital, δk∗ and provide newborns with the steady state level of capital perworker, nk∗. In this situation we can again pull the Gamov trick to construct aPareto superior allocation. Suppose the economy is in the steady state at somearbitrary date t and suppose that the steady state satisfies (8.34). Now considerthe alternative allocation: at date t reduce the capital stock per worker to besaved to the next period, kt+1, by a marginal ∆k

∗ < 0 to k∗∗ = k∗ +∆k∗ andkeep it at k∗∗ forever. From (8.33) we obtain

ct = f(kt) + (1− δ)kt − (1 + n)kt+1The effect on per capita consumption from period t onwards is

∆ct = −(1 + n)∆k∗ > 0∆ct+τ = f 0(k∗)∆k∗ + [1− δ − (1 + n)]∆k∗

= [f 0(k∗)− (δ + n)]∆k∗ > 0

In this way we can increase total per capita consumption in every period. Nowwe just divide the additional consumption between the two generations alive ina given period in such a way that make both generations better off, which isstraightforward to do, given that we have extra consumption goods to distributein every period. Note again that for the Gamov trick to work it is crucial to havean infinite hotel, i.e. that time extends to the infinite future. If there is a last

Page 176: Macroeconomics Phd Lectures notes

170 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

generation, it surely will dislike losing some of its final period capital (whichwe assume is eatable as we are in a one sector economy where the good is aconsumption as well as investment good). A construction of a Pareto superiorallocation wouldn’t be possible. The previous discussion can be summarized inthe following proposition

Proposition 92 Suppose a competitive equilibrium converges to a steady statesatisfying (8.34). Then the equilibrium allocation is not Pareto efficient, or, asoften called, the equilibrium is dynamically inefficient.

When comparing this result to the pure exchange model we see the directparallel: an allocation is inefficient if the interest rate (in the steady state) issmaller than the population growth rate, i.e. if we are in the Samuelson case.In fact, we repeat a much stronger result by Balasko and Shell that we quotedearlier, but that also applies to production economies. A feasible allocation isan allocation c01, ctt, ctt+1, kt+1∞t=1 that satisfies all negativity constraints andthe resource constraint (8.33). Obviously from the allocation we can reconstructstt and Kt. Let rt = f

0(kt) denote the marginal products of capital per worker.Maintain all assumptions made on U and f and let nt be the population growthrate from period t− 1 to t. We have the following resultTheorem 93 Cass (1972)32, Balasko and Shell (1980). A feasible allocation isPareto optimal if and only if

∞Xt=1

tYτ=1

(1 + rτ+1 − δ)

(1 + nτ+1)= +∞

As an obvious corollary, alluded to before we have that a steady state equi-librium is Pareto optimal (or dynamically efficient) if and only if f 0(k∗)−δ ≥ n.That dynamic inefficiency is not purely an academic matter is demonstrated

by the following example

Example 94 Consider the previous example with log utility, but now with pop-ulation growth n and time discounting β. It is straightforward to compute thesteady state unique steady state as

k∗ =·

β(1− α)

(1 + β)(1 + n)

¸ 11−α

so that

r∗ =α(1 + β)(1 + n)

β(1− α)

and the economy is dynamically inefficient if and only if

α(1 + β)(1 + n)

β(1− α)− δ < n

32The first reference of this theorem is in fact Cass (1972), Theorem 3.

Page 177: Macroeconomics Phd Lectures notes

8.3. OVERLAPPING GENERATIONS MODELS WITH PRODUCTION171

Let’s pick some reasonable numbers. We have a 2-period OLG model, so let usinterpret one period as 30 years. α corresponds to the capital share of income,so α = .3 is a commonly used value in macroeconomics. The current yearlypopulation growth rate in the US is about 1%, so lets pick (1 + n) = (1 +0.01)30. Suppose that capital depreciates at around 6% per year, so choose (1−δ) = 0.9430. This yields n = 0.35 and δ = 0.843. Then for a yearly subjectivediscount factor βy ≥ 0.998, the economy is dynamically inefficient. Dynamicinefficiency therefore is definitely more than just a theoretical curiousum. If theeconomy features technological progress of rate g, then the condition for dynamicinefficiency becomes (approximately) f 0(k∗) < n+ δ + g. If we assume a yearlyrate of technological progress of 2%, then with the same parameter values forβy ≥ 0.971 we obtain dynamic inefficiency. Note that there is a more immediateway to check for dynamic inefficiency in an actual economy: since in the modelf 0(k∗)− δ is the real interest rate and g+n is the growth rate of real GDP, onemay just check whether the real interest rate is smaller than the growth rate inlong-run averages.

If the competitive equilibrium of the economy features dynamic inefficiencyits citizens save more than is socially optimal. Hence government programsthat reduce national saving are called for. We already have discussed sucha government program, namely an unfunded, or pay-as-you-go social securitysystem. Let’s briefly see how such a program can reduce the capital stock ofan economy and hence leads to a Pareto-superior allocation, provided that theinitial allocation without the system was dynamically inefficient.Suppose the government introduces a social security system that taxes people

the amount τ when young and pays benefits of b = (1 + n)τ when old. Forsimplicity we assume balanced budget for the social security system as well aslump-sum taxation. The budget constraints of the representative individualchange to

ctt + stt = wt − τ

ctt+1 = (1 + rt+1 − δ)stt + (1 + n)τ

We will repeat our previous analysis and first check how individual savings reactto a change in the size of the social security system. The first order conditionfor consumer maximization is

U 0(wt − τ − stt) = βU 0((1 + rt+1 − δ)stt + (1 + n)τ) ∗ (1 + rt+1 − δ)

which implicitly defines the optimal saving function stt = s(wt, rt+1, τ). Againinvoking the implicit function theorem we find that

−U 00(wt − τ − s(., ., .))µ1− ds

¶= βU 00((1 + rt+1 − δ)s(., ., .) + (1 + n)τ) ∗ (1 + rt+1 − δ) ∗

µ(1 + rt+1 − δ)

ds

dτ+ 1 + n

Page 178: Macroeconomics Phd Lectures notes

172 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

or

ds

dτ= sτ =

U 00() + (1 + n)βU 00(.)(1 + rt+1 − δ)

U 00(.) + βU 00(.)(1 + rt+1 − δ)2< 0

Therefore the bigger the pay-as-you-go social security system, the smaller is theprivate savings of individuals, holding factor prices constant. This however, isonly the partial equilibrium effect of social security. Now let’s use the assetmarket equilibrium condition

kt+1 =s(wt, rt+1, τ)

1 + n

=s (f(kt)− f 0(kt)kt, f 0(kt+1, τ)

1 + n

Now let us investigate how the equilibrium (kt, kt+1)-locus changes as τ changes.For fixed kt, how does kt+1(kt) changes as τ changes. Again using the implicitfunction theorem yields

dkt+1dτ

=srt+1f

00(kt+1)dkt+1dτ + sτ

1 + n

and hence

dkt+1dτ

=sτ

1 + n− srt+1f 00(kt+1)The nominator is negative as shown above; the denominator is positive by ourassumption of monotonic local stability (this is our first application of Samuel-

son’s correspondence principle). Hence dkt+1dτ < 0, the locus (always under the

maintained monotonic stability assumption) tilts downwards, as shown in Figure14.We can conduct the following thought experiment. Suppose the economy

converged to its old steady state k∗ and suddenly, at period T, the governmentunanticipatedly announces the introduction of a (marginal) pay-as-you go sys-tem. The saving locus shifts down, the new steady state capital labor ratiodeclines and the economy, over time, converges to its new steady state. Notethat over time the interest rate increases and the wage rate declines. Is the intro-duction of a marginal pay-as-you-go social security system welfare improving?It depends on whether the old steady state capital-labor ratio was inefficientlyhigh, i.e. it depends on whether f 0(k∗)− δ < n or not. Our conclusions aboutthe desirability of social security remain unchanged from the pure exchangemodel.

8.3.4 The Long-Run Effects of Government Debt

Diamond (1965) discusses the effects of government debt on long run capitalaccumulation. He distinguishes between government debt that is held by for-eigners, so-called external debt, and government debt that is held by domestic

Page 179: Macroeconomics Phd Lectures notes

8.3. OVERLAPPING GENERATIONS MODELS WITH PRODUCTION173

45-degree line

kt+1

k’* k* kt

τ up

Figure 8.7:

citizens, so-called internal debt. Note that the second case is identical to Barro’sanalysis if we abstract from capital accumulation and allow altruistic bequestmotives. In fact, in Diamond’s environment with production, but altruistic andoperative bequests a similar Ricardian equivalence result as before applies. Inthis sense Barro’s neutrality result provides the benchmark for Diamond’s anal-ysis of the internal debt case, and we will see how the absence of operativebequests leads to real consequences of different levels of internal debt.

External Debt

Suppose the government has initial outstanding debt, denoted in real terms, ofB1. Denote by bt =

BtLt= Bt

Nttthe debt-labor ratio. All government bonds have

Page 180: Macroeconomics Phd Lectures notes

174 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

maturity of one period, and the government issues new bonds33 so as to keepthe debt-labor ratio constant at bt = b over time. Bonds that are issued inperiod t−1, Bt, are required to pay the same gross interest as domestic capital,namely 1 + rt − δ, in period t when they become due. The government taxesthe current young generation in order to finance the required interest paymentson the debt. Taxes are lump sum and are denoted by τ . The budget constraintof the government is then

Bt(1 + rt − δ) = Bt+1 +Ntt τ

or, dividing by N tt , we get, under the assumption of a constant debt-labor ratio,

τ = (rt − δ − n)bFor the previous discussion of the model nothing but the budget constraint ofyoung individuals changes, namely to

ctt + stt = wt − τ

= wt − (rt − δ − n)bIn particular the asset market equilibrium condition does not change as theoutstanding debt is held exclusively by foreigners, by assumption. As beforewe obtain a saving function s(wt − (rt − δ− n)b, rt+1) as solution to the house-holds optimization problem, and the asset market equilibrium condition readsas before

kt+1 =s(wt − (rt − δ − n)b, rt+1)

1 + n

Our objective is to determine how a change in the external debt-labor ratiochanges the steady state capital stock and the interest rate. This can be an-swered by examining s(). Again we will apply Samuelson’s correspondence prin-ciple. Assuming monotonic local stability of the unique steady state is equivalentto assuming

dkt+1dkt

=−swt(., .)f 00(kt)(kt + b)1 + n− srt+1(., .)f 00(kt+1)

∈ (0, 1) (8.35)

In order to determine how the saving locus in (kt, kt+1) space shifts we applythe Implicit Function Theorem to the asset equilibrium condition to find

dkt+1db

=−swt(., .) (f 0(kt)− δ − n)1 + n− srt+1(., .)f 00(kt+1)

so the sign of dkt+1db equals the negative of the sign of f 0(kt) − δ − n under

the maintained assumption of monotonic local stability. Suppose we are at

33As Diamond (1965) let us specify these bonds as interest-bearing bonds (in contrast tozero-coupon bonds). A bond bought in period t pays (interst plus principal) 1 + rt+1 − δ inperiod t+ 1.

Page 181: Macroeconomics Phd Lectures notes

8.3. OVERLAPPING GENERATIONS MODELS WITH PRODUCTION175

a steady state k∗ corresponding to external debt to labor ratio b∗. Now thegovernment marginally increases the debt-labor ratio. If the old steady statewas not dynamically inefficient, i.e. f 0(k∗) ≥ δ+ n, then the saving locus shiftsdown and the new steady state capital stock is lower than the old one. Diamondgoes on to show that in this case such an increase in government debt leads toa reduction in the utility level of a generation that lives in the new ratherthan the old steady state. Note however that, because of transition generationsthis does not necessarily mean that marginally increasing external debt leadsto a Pareto-inferior allocation. For the case in which the old equilibrium isdynamically inefficient an increase in government debt shifts the saving locusupward and hence increases the steady state capital stock per worker. AgainDiamond shows that now the effects on steady state utility are indeterminate.

Internal Debt

Now we assume that government debt is held exclusively by own citizens. Thetax payments required to finance the interest payments on the outstanding debttake the same form as before. Let’s assume that the government issues newgovernment debt so as to keep the debt-labor ratio Bt

Ltconstant over time at b.

Hence the required tax payments are given by

τ = (rt − δ − n)b

Again denote the new saving function derived from consumer optimization bys(wt−(rt−δ−n)b, rt+1). Now, however, the equilibrium asset market conditionchanges as the savings of the young not only have to absorb the supply of thephysical capital stock, but also the supply of government bonds newly issued.Hence the equilibrium condition becomes

N tt s(wt − (rt − δ − n)b, rt+1) = Kt+1 +Bt+1

or, dividing by N tt = Lt, we obtain

kt+1 =s(wt − (rt − δ − n)b, rt+1)

1 + n− b

Stability and monotonic convergence to the unique (assumed) steady state re-quire that (8.35) holds. To determine the shift in the saving locus in (kt, kt+1)we again implicitly differentiate to obtain

dkt+1

db=−swt(., .)(rt − δ − n) + srt+1f 00(kt+1)dkt+1db

1 + n− 1

and hence

dkt+1

db=− [swt(., .)(f 0(kt)− δ − n)]1 + n− srt+1(., .)f 00(kt+1)

− (1 + n) < −n < 0

Page 182: Macroeconomics Phd Lectures notes

176 CHAPTER 8. THE OVERLAPPING GENERATIONS MODEL

where the first inequality uses (8.35). The curve unambiguously shifts down,leading to a decline in the steady state capital stock per worker. Diamond,again only comparing steady state utilities, shows that if the initial steady statewas dynamically efficient, then an increase in internal debt leads to a reduc-tion in steady state welfare, whereas if the initial steady state was dynamicallyinefficient, then an increase in internal government debt leads to a increase insteady state welfare. Here the intuition is again clear: if the economy has accu-mulated too much capital, then increasing the supply of alternative assets leadsto a interest-driven “crowding out” of demand for physical capital, which is agood thing given that the economy possesses too much capital. In the efficientcase the reverse logic applies. In comparison with the external debt case weobtain clearer welfare conclusions for the dynamically inefficient case. For ex-ternal debt an increase in debt is not necessarily good even in the dynamicallyinefficient case because it requires higher tax payments, which, in contrast tointernal debt, leave the country and therefore reduce the available resources tobe consumed (or invested). This negative effect balances against the positiveeffect of reducing the inefficiently high capital stock, so that the overall effectsare indeterminate. In comparison to Barro (1974) we see that without operativebequests the level of outstanding government bonds influences real equilibriumallocations: Ricardian equivalence breaks down.

Page 183: Macroeconomics Phd Lectures notes

Chapter 9

Continuous Time GrowthTheory

I do not see how one can look at figures like these without seeingthem as representing possibilities. Is there some action a govern-ment could take that would lead the Indian economy to grow likeIndonesia’s or Egypt’s? If so, what exactly? If not, what is it aboutthe nature of India that makes it so? The consequences for humanwelfare involved in questions like these are simply staggering: Onceone starts to think about them, it is hard to think about anythingelse. [Lucas 1988, p. 5]

So much for motivation. We are doing growth in continuous time since Ithink you should know how to deal with continuous time models as a significantfraction of the economic literature employs continuous time, partly because incertain instances the mathematics becomes easier. In continuous time, variablesare functions of time and one can use calculus to analyze how they change overtime.

9.1 Stylized Growth and Development Facts

Data! Data! Data! I can’t make bricks without clay. [SherlockHolmes]

In this part we will briefly review the main stylized facts characterizingeconomic growth of the now industrialized countries and the main facts charac-terizing the level and change of economic development of not yet industrializedcountries.

177

Page 184: Macroeconomics Phd Lectures notes

178 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

9.1.1 Kaldor’s Growth Facts

The British economist Nicholas Kaldor pointed out the following stylized growthfacts (empirical regularities of the growth process) for the US and for most otherindustrialized countries.

1. Output (real GDP) per worker y = YL and capital per worker k =

KL grow

over time at relatively constant and positive rate. See Figure 9.1.

2. They grow at similar rates, so that the ratio between capital and output,KY is relatively constant over time

3. The real return to capital r (and the real interest rate r − δ) is relativelyconstant over time.

4. The capital and labor shares are roughly constant over time. The capitalshare α is the fraction of GDP that is devoted to interest payments oncapital, α = rK

Y . The labor share 1 − α is the fraction of GDP that isdevoted to the payments to labor inputs; i.e. to wages and salaries andother compensations: 1− α = wL

Y . Here w is the real wage.

These stylized facts motivated the development of the neoclassical growthmodel, the Solow growth model, to be discussed below. The Solow model hasspectacular success in explaining the stylized growth facts by Kaldor.

9.1.2 Development Facts from the Summers-Heston DataSet

In addition to the growth facts we will be concerned with how income (perworker) levels and growth rates vary across countries in different stages of theirdevelopment process. The true test of the Solow model is to what extent it canexplain differences in income levels and growth rates across countries, the socalled development facts. As we will see in our discussion of Mankiw, Romerand Weil (1992) the verdict is mixed.Now we summarize the most important facts from the Summers and Heston’s

panel data set. This data set follows about 100 countries for 30 years andhas data on income (production) levels and growth rates as well as populationand labor force data. In what follows we focus on the variable income perworker. This is due to two considerations: a) our theory (the Solow model)will make predictions about exactly this variable b) although other variablesare also important determinants for the standard of living in a country, incomeper worker (or income per capita) may be the most important variable (forthe economist anyway) and other determinants of well-being tend to be highlypositively correlated with income per worker.Before looking at the data we have to think about an important measurement

issue. Income is measured as GDP, and GDP of a particular country is measuredin the currency of that particular country. In order to compare income betweencountries we have to convert these income measures into a common unit. One

Page 185: Macroeconomics Phd Lectures notes

9.1. STYLIZED GROWTH AND DEVELOPMENT FACTS 179

1965 1970 1975 1980 1985 1990 1995 20008

8.1

8.2

8.3

8.4

8.5

8.6

8.7

8.8

8.9

9Real GDP in the United States 1967-1999

Year

Log

of re

al G

DP

GDP

Trend

Figure 9.1:

option would be exchange rates. These, however, tend to be rather volatile andreactive to events on world financial markets. Economists which study growthand development tend to use PPP-based exchange rates, where PPP stands forPurchasing Power Parity. All income numbers used by Summers and Heston(and used in these notes) are converted to $US via PPP-based exchange rates.Here are the most important facts from the Summers and Heston data set:

1. Enormous variation of per capita income across countries: the poorestcountries have about 5% of per capita GDP of US per capita GDP. Thisfact makes a statement about dispersion in income levels. When we lookat Figure 9.2, we see that out of the 104 countries in the data set, 37in 1990 and 38 in 1960 had per worker incomes of less than 10% of theUS level. The richest countries in 1990, in terms of per worker income,are Luxembourg, the US, Canada and Switzerland with over $30,000, thepoorest countries, without exceptions, are in Africa. Mali, Uganda, Chad,

Page 186: Macroeconomics Phd Lectures notes

180 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

0 0.2 0.4 0.6 0.8 1 1.2 1.40

5

10

15

20

25

30

35

40Distribution of Relative Per Worker Income

Income Per Worker Relative to US

Num

ber o

f Cou

ntrie

s

19601990

Figure 9.2:

Central African Republic, Burundi, Burkina Faso all have income perworker of less than $1000. Not only are most countries extremely poorcompared to the US, but most of the world’s population is poor relativeto the US.

2. Enormous variation in growth rates of per worker income. This fact makesa statement about changes of levels in per capita income. Figure 9.3shows the distribution of average yearly growth rates from 1960 to 1990.The majority of countries grew at average rates of between 1% and 3%(these are growth rates for real GDP per worker). Note that some coun-tries posted average growth rates in excess of 6% (Singapore, Hong Kong,Japan, Taiwan, South Korea) whereas other countries actually shrunk,i.e. had negative growth rates (Venezuela, Nicaragua, Guyana, Zambia,Benin, Ghana, Mauretania, Madagascar, Mozambique, Malawi, Uganda,Mali). We will sometimes call the first group growth miracles, the sec-ond group growth disasters. Note that not only did the disasters’ relative

Page 187: Macroeconomics Phd Lectures notes

9.1. STYLIZED GROWTH AND DEVELOPMENT FACTS 181

-0.03 -0.02 -0.01 0 0.01 0.02 0.03 0.04 0.05 0.060

5

10

15

20

25Distribution of Average Growth Rates (Real GDP) Between 1960 and 1990

Average Growth Rate

Num

ber o

f Cou

ntrie

s

Figure 9.3:

position worsen, but that these countries experienced absolute declines inliving standards. The US, in terms of its growth experience in the last 30years, was in the middle of the pack with a growth rate of real per workerGDP of 1.4% between 1960 and 1990.

3. Growth rates determine economic fate of a country over longer periods oftime. How long does it take for a country to double its per capita GDPif it grows at average rate of g% per year? A good rule of thumb: 70/gyears (this rule of thumb is due to Nobel Price winner Robert E. Lucas(1988)).1 Growth rates are not constant over time for a given country.This can easily be demonstrated for the US. GDP per worker in 1990

1Let yT denote GDP per capita in period T and y0 denote period 0 GDP per capita in aparticular country. Suppose the growth rate of GDP per capita is constant at g, i.e. 100∗g%.Then

yT = y0egT

Page 188: Macroeconomics Phd Lectures notes

182 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

was $36,810. If GDP would always have grown at 1.4%, then for the USGDP per worker would have been about $9,000 in 1900, $2,300 in 1800,$570 in 1700, $140 in 1600, $35 in 1500 and so forth. Economic historians(and common sense) tells us that nobody can survive on $35 per year(estimates are that about $300 are necessary as minimum income levelfor survival). This indicates that the US (or any other country) cannothave experienced sustained positive growth for the last millennium or so.In fact, prior to the era of modern economic growth, which started inEngland in the late 1800th century, per worker income levels have beenalmost constant at subsistence levels. This can be seen from Figure 9.4,which compiles data from various historical sources. The start of modern

GDP per Capita (in 1985 US $): Western Europe and its Offsprings

02000

40006000

800010000

1200014000

16000

050

010

0014

0016

1018

2018

7019

1319

5019

7319

89

Tim e

GDP per Capita

Figure 9.4:

economic growth is sometimes referred to as the Industrial Revolution.It is the single most significant economic event in history and has, likeno other event, changed the economic circumstances in which we live.Hence modern economic growth is a quite recent phenomenon, and sofar has occurred only in Western Europe and its offsprings (US, Canada,

Suppose we want to double GDP per capita in T years. Then

2 =yT

y0= egT

or

ln(2) = gT

T ∗ =ln(2)

g=100 ∗ ln(2)g(in %)

Since 100 ∗ ln(2) ≈ 70, the rule of thumb follows.

Page 189: Macroeconomics Phd Lectures notes

9.2. THE SOLOW MODEL AND ITS EMPIRICAL EVALUATION 183

Australia and New Zealand) as well as recently in East Asia.

4. Countries change their relative position in the (international) income dis-tribution. Growth disasters fall, growth miracles rise, in the relative cross-country income distribution. A classical example of a growth disaster isArgentina. At the turn of the century Argentina had a per-worker incomethat was comparable to that in the US. In 1990 the per-worker incomeof Argentina was only on a level of one third of the US, due to a healthygrowth experience of the US and a disastrous growth performance of Ar-gentina. Countries that dramatically moved up in the relative incomedistribution include Italy, Spain, Hong Kong, Japan, Taiwan and SouthKorea, countries that moved down are New Zealand, Venezuela, Iran,Nicaragua, Peru and Trinidad&Tobago.

In the next section we have two tasks: to construct a model, the Solowmodel, that a) can successfully explain the stylized growth facts b) investigateto which extent the Solow model can explain the development facts.

9.2 The Solow Model and its Empirical Evalua-tion

The basic assumptions of the Solow model are that there is a single good pro-duced in our economy and that there is no international trade, i.e. the economyis closed to international goods and factor flows. Also there is no government.It is also assumed that all factors of production (labor, capital) are fully em-ployed in the production process. We assume that the labor force, L(t) growsat constant rate n > 0, so that, by normalizing L(0) = 1 we have that

L(t) = entL(0) = ent

The model consists of two basic equations, the neoclassical aggregate productionfunction and a capital accumulation equation.

1. Neoclassical aggregate production function

Y (t) = F (K(t), A(t)L(t))

We assume that F has constant returns to scale, is strictly concave andstrictly increasing, twice continuously differentiable, F (0, .) = F (., 0) = 0and satisfies the Inada conditions. Here Y (t) is total output, K(t) is thecapital stock at time t and A(t) is the level of technology at time t. Wenormalize A(0) = 1, so that a worker in period t provides the same laborinput as A(t) workers in period 0. We call A(t)L(t) labor input in laborefficiency units (rather than in raw number of bodies) or effective labor atdate t. We assume that

A(t) = egt

Page 190: Macroeconomics Phd Lectures notes

184 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

i.e. the level of technology increases at continuous rate g > 0.We interpretthis as technological progress: due to the invention of new technologies or“ideas” workers get more productive over time. This exogenous techno-logical progress, which is not explained within the model is the key drivingforce of economic growth in the Solow model. One of the main criticismsof the Solow model is that it does not provide an endogenous explana-tion for why technological progress, the driving force of growth, arises.Romer (1990) and Jones (1995) pick up exactly this point. We modeltechnological progress as making labor more effective in the productionprocess. This form of technological progress is called labor augmentingor Harrod-neutral technological progress.2 In order to analyze the modelwe seek a representation in variables that remain stationary over time,so that we can talk about steady states and dynamics around the steadystate. Obviously, since the number of workers as well as technology growsexponentially, total output and capital (even per capita or per worker)will tend to grow. However, expressing all variables of the model in pereffective labor units there is hope to arrive at a representation of the modelin which the endogenous variables are stationary. Hence we divide bothsides of the production function by the effective labor input A(t)L(t) toobtain (using the constant returns to scale assumption)3

ξ(t) =Y (t)

A(t)L(t)=F (K(t), A(t)L(t))

A(t)L(t)= F

µK(t)

A(t)L(t), 1

¶= f(κ(t))

(9.1)

where ξ(t) = Y (t)A(t)L(t) is output per effective labor input and κ(t) =

K(t)A(t)L(t)

is the capital stock perfect labor input. From the assumptions made onF it follows that f is strictly increasing, strictly concave, twice continu-ously differentiable, f(0) = 0 and satisfies the Inada condition. Equation(9.1) summarizes our assumptions about the production technology of theeconomy.

2. Capital accumulation equation and resource constraint

K(t) = sY (t)− δK(t) (9.2)

K(t) + δK(t) = Y (t)− C(t) (9.3)

The change of the capital stock in period t, K(t) is given by gross invest-ment in period t, sY (t) minus the depreciation of the old capital stock

2Alternative specifications of the production functions are F (AK,L) in which case tech-nological progress is called capital augmenting or Solow neutral technological progress, andAF (K,L) in which case it is called Hicks neutral technological progress. For the way we willdefine a balanced growth path below it is only Harrod-neutral technological progress (at leastfor general production functions) that guarantees the existence of a balanced growth path inthe Solow model.

3In terms of notation I will use uppercase variables for aggregate variables, lowercase forper-worker variables and the corresponding greek letter for variables per effective labor units.Since there is no greek y I use ξ for per capita output

Page 191: Macroeconomics Phd Lectures notes

9.2. THE SOLOW MODEL AND ITS EMPIRICAL EVALUATION 185

δK(t). We assume δ ≥ 0. Since we have a closed economy model grossinvestment is equal to national saving (which is equal to saving of theprivate sector, since there is no government). Here s is the fraction oftotal output (income) in period t that is saved, i.e. not consumed. Theimportant assumption implicit in equation (9.2) is that households save aconstant fraction s of output (income), regardless of the level of income.This is a strong assumption about the behavior of households that is notendogenously derived from within a model of utility-maximizing agents(and the Cass-Koopmans-Ramsey model relaxes exactly this assumption).Remember that the discrete time counterpart of this equation was

Kt+1 −Kt = sYt − δKt

Kt+1 − (1− δ)Kt = Yt − Ct

Now we can divide both sides of equation (9.2) by A(t)L(t) to obtain

K(t)

A(t)L(t)= sξ(t)− δκ(t) (9.4)

Expanding the left hand side of equation (9.4) gives

K(t)

A(t)L(t)=K(t)

K(t)

K(t)

A(t)L(t)=K(t)

K(t)κ(t) (9.5)

But

κ(t)

κ(t)=K(t)

K(t)− L(t)L(t)

− A(t)A(t)

=K(t)

K(t)− n− g

Hence

K(t)

K(t)=

κ(t)

κ(t)+ n+ g (9.6)

Combining equations (9.5) and (9.6) with (9.4) yields

K(t)

A(t)L(t)=

K(t)

K(t)κ(t) =

µκ(t)

κ(t)+ n+ g

¶κ(t) (9.7)

κ(t) + κ(t)(n+ g) = sξ(t)− δκ(t) (9.8)

κ(t) = sξ(t)− (n+ g + δ)κ(t) (9.9)

This is the capital accumulation equation in per-effective worker terms. Combin-ing this equation with the production function gives the fundamental differentialequation of the Solow model

κ(t) = sf(κ(t))− (n+ g + δ)κ(t) (9.10)

Technically speaking this is a first order nonlinear ordinary differential equation,and it completely characterizes the evolution of the economy for any initial

Page 192: Macroeconomics Phd Lectures notes

186 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

condition κ(0) = K(0). Once we have solved the differential equation for thecapital per effective labor path κ(t)t∈[0,∞) the rest of the endogenous variablesare simply given by

k(t) = κ(t)A(t) = egtκ(t)

K(t) = e(n+g)tκ(t)

y(t) = egtf(κ(t))

Y (t) = e(n+g)tf(κ(t))

C(t) = (1− s)e(n+g)tf(κ(t))c(t) = (1− s)egtf(κ(t))

9.2.1 The Model and its Implications

Analyzing the qualitative properties of the model amounts to analyzing the dif-ferential equation (9.10). Unfortunately this differential equation is nonlinear,so there is no general method to explicitly solve for the function κ(t). We can,however, analyze the differential equation graphically. Before doing this, how-ever, let us look at a (I think the only) particular example for which we actuallycan solve the equation analytically

Example 95 Let f(κ) = κα (i.e. F (K,AL) = Kα(AL)1−α). The fundamentaldifferential equation becomes

κ(t) = sκ(t)α − (n+ g + δ)κ(t) (9.11)

with κ(0) > 0 given. A steady state of this equation is given by κ(t) = κ∗ forwhich κ(t) = 0 for all t. There are two steady states, the trivial one at κ = 0(which we will ignore from now on, as it is only reached if κ(0) = 0) and the

unique positive steady state κ∗ =³

sn+g+δ

´ 11−α

. Now let’s solve the differential

equation. This equation is, in fact, a special case of the so-called Bernoulliequation. Let’s do the following substitution of variables. Define v(t) = κ(t)1−α.Then

v(t) = (1− α)κ(t)−α ∗ κ(t) = (1− α)κ(t)

κ(t)α

Dividing both sides of (9.11) by κ(t)α

1−α yields

(1− α)κ(t)

κ(t)α= (1− α)s− (1− α)(n+ g + δ)κ(t)1−α

and now making the substitution of variables

v(t) = (1− α)s− (1− α)(n+ g + δ)v(t)

Page 193: Macroeconomics Phd Lectures notes

9.2. THE SOLOW MODEL AND ITS EMPIRICAL EVALUATION 187

which is a linear ordinary first order (nonhomogeneous) differential equation,which we know how to solve.4 The general solution to the homogeneous equationtakes the form

vg(t) = Ce−(1−α)(n+g+δ)t

where C is an arbitrary constant. A particular solution to the nonhomogeneousequation is

vp(t) =s

n+ g + δ= v∗ = (κ∗)1−α

Hence all solutions to the differential equation take the form

v(t) = vg(t) + vp(t)

= v∗ + Ce−(1−α)(n+g+δ)t

Now we use the initial condition v(0) = κ(0)1−α to determine the constant C

v(0) = v∗ + CC = v(0)− v∗

Hence the solution to the initial value problem is

v(t) = v∗ + (v(0)− v∗) e−(1−α)(n+g+δ)t

and substituting back κ for v we obtain

κ(t)1−α = (κ∗)1−α +³κ(0)1−α − (κ∗)1−α

´e−(1−α)(n+g+δ)t

and hence

κ(t) =

·s

n+ g + δ+

µκ(0)1−α − s

n+ g + δ

¶e−(1−α)(n+g+δ)t

¸ 11−α

Note that limt→∞ κ(t) =h

sn+g+δ

i 11−α

= κ∗ regardless of the value of κ(0) > 0.In other words the unique steady state capital per labor efficiency unit is locally(globally if one restricts attention to strictly positive capital stocks) asymptoti-cally stable

For a general production function one can’t solve the differential equationexplicitly and has to resort to graphical analysis. In Figure 9.5 we plot thetwo functions (n + δ + g)κ(t) and sf(κ(t)) against κ(t). Given the propertiesof f it is clear that both curves intersect twice, once at the origin and once

4An excellent reference for economists is Gandolfo, G. “Economic Dynamics: Methods andModels”. There are thousands of math books on differential equations, e.g. Boyce, W. andDiPrima, R. “Elementary Differential Equations and Boundary Value Problems”

Page 194: Macroeconomics Phd Lectures notes

188 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

at a unique positive κ∗ and (n + δ + g)κ(t) < sf(κ(t)) for all κ(t) < κ∗ and(n+ δ + g)κ(t) > sf(κ(t)) for all κ(t) > κ∗. The steady state solves

sf(κ∗)k∗

= n+ δ + g

Since the change in κ is given by the difference of the two curves, for κ(t) < κ∗

κ increases, for κ(t) > κ∗ it decreases over time and for κ(t) = κ∗ it remainsconstant. Hence, as for the example above, also in the general case there existsa unique positive steady state level of the capital-labor-efficiency ratio that islocally asymptotically stable. Hence in the long run κ settles down at κ∗ for anyinitial condition κ(0) > 0. Once the economy has settled down at κ∗, output,

κ(0) κ* κ(t)

(n+g+δ)κ(t)

sf(κ(t))

.κ(0)

Figure 9.5:

consumption and capital per worker grow at constant rates g and total output,capital and consumption grow at constant rates g+n. A situation in which theendogenous variables of the model grow at constant (not necessarily the same)

Page 195: Macroeconomics Phd Lectures notes

9.2. THE SOLOW MODEL AND ITS EMPIRICAL EVALUATION 189

rates is called a Balanced Growth Path (henceforth BGP). A steady state is abalanced growth path with growth rate of 0.

9.2.2 Empirical Evaluation of the Model

Kaldor’s Growth Facts

Can the Solow model reproduce the stylized growth facts? The prediction ofthe model is that in the long run output per worker and capital per worker bothgrow at positive and constant rate g, the growth rate of technology. Thereforethe capital-labor ratio k is constant, as observed by Kaldor. The other twostylized facts have to do with factor prices. Suppose that output is produced bya single competitive firm that faces a rental rate of capital r(t) and wage ratew(t) for one unit of raw labor (i.e. not labor in efficiency units). The firm rentsboth input at each instant in time and solves

maxK(t),L(t)≥0

F (K(t), A(t)L(t))− r(t)K(t)− w(t)L(t)

Profit maximization requires

r(t) = FK(K(t), A(t)L(t))

w(t) = A(t)FL(K(t), A(t)L(t))

Given that F is homogenous of degree 1, FK and FL are homogeneous of degreezero, i.e.

r(t) = FK

µK(t)

A(t)L(t), 1

¶w(t) = A(t)FL

µK(t)

A(t)L(t), 1

¶In a balanced grow path K(t)

A(t)L(t) = κ(t) = κ∗ is constant, so the real rental rateof capital is constant and hence the real interest rate is constant. The wagerate increases at the rate of technological progress, g. Finally we can computecapital and labor shares. The capital share is given as

α =r(t)K(t)

Y (t)

which is constant in a balanced growth path since the rental rate of capitalis constant and Y (t) and K(t) grow at the same rate g + n. Hence the uniquebalanced growth path of the Solow model, to which the economy converges fromany initial condition, reproduces all four stylized facts reported by Kaldor. Inthis dimension the Solow model is a big success and Solow won the Nobel pricefor it in 1989.

Page 196: Macroeconomics Phd Lectures notes

190 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

The Summers-Heston Development Facts

How can we explain the large difference in per capita income levels across coun-tries? Assume first that all countries have access to the same production tech-nology, face the same population growth rate and have the same saving rate.Then the Solow model predicts that all countries over time converge to the samebalanced growth path represented by κ∗. All countries’ per capita income con-verges to the path y(t) = A(t)κ∗, equal for all countries under the assumptionof the same technology, i.e. same A(t) process. Hence, so the prediction of themodel, eventually per worker income (GDP) is equalized internationally. Thefact that we observe large differences in per worker incomes across countries inthe data must then be due to different initial conditions for the capital stock,so that countries differ with respect to their relative distance to the commonBGP. Poorer countries are just further away from the BGP because they startedwith lesser capital stock, but will eventually catch up. This implies that poorercountries temporarily should grow faster than richer countries, according to themodel. To see this, note that the growth rate of output per worker γy(t) is givenby

γy(t) =y(t)

y(t)= g +

f 0(κ(t))κ(t)f(κ(t))

= g +f 0(κ(t))f(κ(t))

(sf(κ(t))− (n+ δ + g)κ(t))

Since f 0(κ(t))f(κ(t)) is positive and decreasing in κ(t) and (sf(κ(t))− (n+ δ + g)κ(t))

is decreasing in κ(t) for two countries with κ1(t) < κ2(t) < κ∗ we have γ1y(t) >γ2y(t) > 0, i.e. countries that a further away from the balanced growth pathgrow more rapidly. The hypothesis that all countries’ per worker income even-tually converges to the same balanced growth path, or the somewhat weakerhypothesis that initially poorer countries grow faster than initially richer coun-tries is called absolute convergence. If one imposes the assumptions of equalityof technology and savings rates across countries, then the Solow model predictsabsolute convergence. This implication of the model has been tested empiricallyby several authors. The data one needs is a measure of “initially poor vs. rich”and data on growth rates from “initially” until now. As measure of “initiallypoor vs. rich” the income per worker (in $US) of different countries at someinitial year has been used.In Figure 9.6 we use data for a long time horizon for 16 now industrialized

countries. Clearly the level of GDP per capita in 1885 is negatively correlatedwith the growth rate of GDP per capita over the last 100 years across countries.So this figure lends support to the convergence hypothesis. We get the samequalitative picture when we use more recent data for 22 industrialized countries:the level of GDP per worker in 1960 is negatively correlated with the growth ratebetween 1960 and 1990 across this group of countries, as Figure 9.7 shows. Thisresult, however, may be due to the way we selected countries: the very fact thatthese countries are industrialized countries means that they must have caught

Page 197: Macroeconomics Phd Lectures notes

9.2. THE SOLOW MODEL AND ITS EMPIRICAL EVALUATION 191

up with the leading country (otherwise they wouldn’t be called industrializedcountries now). This important point was raised by Bradford deLong (1988)

Growth Rate Versus Initial Per Capita GDP

Per Capita GDP, 1885

Gro

wth

Rat

e of

Per

Cap

ita G

DP

, 188

5-19

94

0 1000 2000 3000 4000 50001

1.5

2

2.5

3JPN

FINNOR

ITL

SWE

CAN

FRA

DNK

AUTGER

BEL

USA

NLD

NZL

GBR

AUS

Figure 9.6:

Let us take deLongs point seriously and look at the correlation betweeninitial income levels and subsequent growth rates for the whole cross-sectionalsample of Summers-Heston. Figure 9.8 doesn’t seem to support the convergencehypothesis: for the whole sample initial levels of GDP per worker are prettymuch uncorrelated with consequent growth rates. In particular, it doesn’t seemto be the case that most of the very poor countries, in particular in Africa, arecatching up with the rest of the world, at least not until 1990 (or until 2002 forthat matter).So does Figure 9.8 constitute the big failure of the Solow model? After all, for

the big sample of countries it didn’t seem to be the case that poor countries growfaster than rich countries. But isn’t that what the Solow model predicts? Notexactly: the Solow model predicts that countries that are further away from their

Page 198: Macroeconomics Phd Lectures notes

192 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

Growth Rate Versus Initial Per Capita GDP

Per Worker GDP, 1960

Gro

wth

Rat

e of

Per

Cap

ita G

DP

, 196

0-19

90

0 0.5 1 1.5 2 2.5x 104

0

1

2

3

4

5

TUR

POR

JPN

GRC

ESP

IRL

AUT

ITL

FIN

FRA

GERBEL

NORGBR

DNK

NLD

SWEAUS

CANCHE

NZL

USA

Figure 9.7:

balanced growth path grow faster than countries that are closer to their balancedgrowth path (always assuming that the rate of technological progress is thesame across countries). This hypothesis is called conditional convergence. The“conditional” means that we have to condition on characteristics of countriesthat may make them have different steady states κ∗(s, n, δ) (they still shouldgrow at the same rate eventually, after having converged to their steady states)to determine which countries should grow faster than others. So the fact thatpoor African countries grow slowly even though they are poor may be, accordingto the conditional convergence hypothesis, due to the fact that they have a lowbalanced growth path and are already close to it, whereas some richer countriesgrow fast since they have a high balanced growth path and are still far fromreaching it.

To test the conditional convergence hypothesis economists basically do the

Page 199: Macroeconomics Phd Lectures notes

9.2. THE SOLOW MODEL AND ITS EMPIRICAL EVALUATION 193

Growth Rate Versus Initial Per Capita GDP

Per Worker GDP, 1960

Gro

wth

Rat

e of

Per

Cap

ita G

DP

, 196

0-19

90

0 0.5 1 1.5 2 2.5x 104

-4

-2

0

2

4

6

LUX

USACANCHE

BELNLD

ITAFRA

AUS

GERNOR

SWE

FIN

GBR

AUT

ESP

NZL

ISLDNK

SGP

IRLISR

HKG

JPN

TTO

OAN

CYPGRC

VEN

MEX

PRT

KOR

SYRJORMYS

DZA

CHLURY

FJI

IRN

BRA

MUSCOL

YUG

CRIZAF

NAM

SYC

ECUTUN

TUR

GAB PANCSKGTM DOM

EGY

PERMAR

THA

PRY

LKASLV

BOL

JAM

IDN

BGDPHL

PAK

COG

HND

NIC

IND CIV

PNG

GUY

CIV

CMR

ZWESEN

CHN

NGA

LSO

ZMBBENGHA

KENGMBMRT

GIN

TGO

MDGMOZ RWA

GNB COM

CAF

MWITCD

UGAMLIBDI BFA

LSO

MLI

BFAMOZ

CAF

Figure 9.8:

following: they compute the steady state output per worker5 that a countryshould possess in a given initial period, say 1960, given n, s, δ measured inthis country’s data. Then they measure the actual GDP per worker in thisperiod and build the difference. This difference indicates how far away thisparticular country is away from its balanced growth path. This variable, thedifference between hypothetical steady state and actual GDP per worker is thenplotted against the growth rate of GDP per worker from the initial period tothe current period. If the hypothesis of conditional convergence were true, thesetwo variables should be negatively correlated across countries: countries thatare further away from their from their balanced growth path should grow faster.Jones’ (1998) Figure 3.8 provides such a plot. In contrast to Figure 9.8 he findsthat, once one conditions on country-specific steady states, poor (relative to

5Which is proportional to the balanced growth path for output per worker (just multiplyit by the constant A(1960)).

Page 200: Macroeconomics Phd Lectures notes

194 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

their steady) tend to grow faster than rich countries. So again, the Solow modelis quite successful qualitatively.Now we want to go one step further and ask whether the Solow model can

predict the magnitude of cross-country income differences once we allow param-eters that determine the steady state to vary across countries. Such a quanti-tative exercise was carried out in the influential paper by Mankiw, Romer andWeil (1992). The authors “want to take Robert Solow seriously”, i.e. inves-tigate whether the quantitative predictions of his model are in line with thedata. More specifically they ask whether the model can explain the enormouscross-country variation of income per worker. For example in 1985 per workerincome of the US was 31 times as high as in Ethiopia.There is an obvious way in which the Solow model can account for this

number. Suppose we constrain ourselves to balanced growth paths (i.e. ignorethe convergence discussion that relies on the assumptions that countries havenot yet reached their BGP’s). Then, by denoting yUS(t) as per worker incomein the US and yETH(t) as per worker income in Ethiopia in time t we find thatalong BGP’s, with assumed Cobb-Douglas production function

yUS(t)

yETH(t)=

AUS(t)

AETH(t)∗

³sUS

nUS+gUS+δUS

´ α1−α³

sETH

nETH+gETH+δETH

´ α1−α

(9.12)

One easy way to get the income differential is to assume large enough differences

in levels of technology AUS(t)AETH(t)

. One fraction of the literature has gone this route;

the hard part is to justify the large differences in levels of technology whentechnology transfer is relatively easy between a lot of countries.6 The otherfraction, instead of attributing the large income differences to differences in Aattributes the difference to variation in savings (investment) and populationgrowth rates. Mankiw et al. take this view. They assume that there is infact no difference across countries in the production technologies used, so thatAUS(t) = AETH(t) = A(0)egt, gETH = gUS and δETH = δUS . Assumingbalanced growth paths and Cobb-Douglas production we can write

yi(t) = A(0)egtµ

si

ni + δ + g

¶ α1−α

where i indexes a country. Taking natural logs on both sides we get

ln(yi(t)) = ln(A(0)) + gt+α

1− αln(si)− α

1− αln(ni + δ + g)

Given this linear relationship derived from the theoretical model it very tempt-ing to run this as a regression on cross-country data. For this, however, we needa stochastic error term which is nowhere to be detected in the model. Mankiwet al. use the following assumption

ln(A(0)) = a+ εi (9.13)

6See, e.g. Parente and Prescott (1994, 1999).

Page 201: Macroeconomics Phd Lectures notes

9.2. THE SOLOW MODEL AND ITS EMPIRICAL EVALUATION 195

where a is a constant (common across countries) and εi is a country specificrandom shock to the (initial) level of technology that may, according to theauthors, represent not only variations in production technologies used, but alsoclimate, institutions, endowments with natural resources and the like. Usingthis and assuming that the time period for the cross sectional data on which theregression is run is t = 0 (if t = T that only changes the constant7) we obtainthe following linear regression

ln(yi) = a+α

1− αln(si)− α

1− αln(ni + δ + g) + εi

ln(yi) = a+ b1 ln(si) + b2 ln(n

i + δ + g) + εi (9.14)

Note that the variation in εi across countries, according to the underlying model,are attributed to variations in technology. Hence the regression results will tellus how much of the variation in cross-country per-worker income is due to vari-ations in investment and population growth rates, and how much is due torandom differences in the level of technology. This is, if we take (9.13) literally,how the regression results have to be interpreted. If we want to estimate (9.14)by OLS, the identifying assumption is that the εi are uncorrelated with theother variables on the right hand side, in particular the investment and popu-lation growth rate. Given the interpretation the authors offered for εi I inviteyou all to contemplate whether this is a good assumption or not. Note that theregression equation also implies restrictions on the parameters to be estimated:if the specification is correct, then one expects the estimated b1 = −b2. Onemay also impose this constraint a priori on the parameter values and do con-strained OLS. Apparently the results don’t change much from the unrestrictedestimation. Also, given that the production function is Cobb-Douglas, α has theinterpretation as capital share, which is observable in the data and is thought tobe around .25-0.5 for most countries, one would expect b1 ∈ [13 , 1] a priori. Thisis an important test for whether the specification of the regression is correct.With respect to data, yi is taken to be real GDP divided by working age

population in 1985, ni is the average growth rate of the working-age population8

from 1960 − 85 and s is the average share of real investment9 from real GDPbetween 1960− 85. Finally they assume that g + δ = 0.05 for all countries.Table 2 reports their results for the unrestricted OLS-estimated regression

on a sample of 98 countries (see their data appendix for the countries in thesample)

Table 2

a b1 b2 R2

5.48(1.59)

1.42(0.14)

−1.48(0.12)

0.59

7Note that we do not use the time series dimension of the data, only the cross-sectional,i.e. cross-country dimension.

8This implicitly assumes a constant labor force participation rate from 1960− 85.9Private as well as government (gross) investment.

Page 202: Macroeconomics Phd Lectures notes

196 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

The basic results supporting the Solow model are that the bi have the rightsign, are highly statistically significant and are of similar size. Most impor-tantly, a major fraction of the cross-country variation in per-worker incomes,namely about 60% is accounted for by the variations in the explanatory vari-ables, namely investment rates and population growth rates. The rest, given theassumptions about where the stochastic error term comes from, is attributed torandom variations in the level of technology employed in particular countries.That seems like a fairly big success of the Solow model. However, the size

of the estimates bi indicates that the implied required capital shares on aver-age have to lie around 2

3 rather than13 usually observed in the data. This

is both problematic for the success of the model and points to a direction ofimprovement of the model.Let’s first understand where the high coefficients come from. Assume that

nUS = nETH = n (variation in population growth rates is too small to makea significant difference) and rewrite (9.12) as (using the assumption of sametechnology, the differences are assumed to be of stochastic nature)

yUS(t)

yETH(t)=

µsUS

sETH

¶ α1−α

To generate a spread of incomes of 31, for α = 13 one needs a ratio of investment

rates of 961 which is obviously absurdly high. But for α = 23 one only requires

a ratio of 5.5. In the data, the measured ratio is about 3.9 for the US versusEthiopia. This comes pretty close (population growth differentials would almostdo the rest). Obviously this is a back of the envelope calculation involving onlytwo countries, but it demonstrates the core of the problem: there is substantialvariation in investment and population growth rates across countries, but ifthe importance of capital in the production process is as low as the commonlybelieved α = 1

3 , then these variations are nowhere nearly high enough to generatethe large income differentials that we observe in the data. Hence the regressionforces the estimated α up to two thirds to make the variations in si (and ni)matter sufficiently much.So if we can’t change the data to give us a higher capital share and can’t

force the model to deliver the cross-country spread in incomes given reasonablecapital shares, how can we rescue the model? Mankiw, Romer and Weil doa combination of both. Suppose you reinterpret the capital stock as broadlycontaining not only the physical capital stock, but also the stock of humancapital and you interpret part of labor income as return to not just raw physicallabor, but as returns to human capital such as education, then possibly a capitalshare of two thirds is reasonable. In order to do this reinterpretation on the data,one better first augments the model to incorporate human capital as well.So now let the aggregate production function be given by

Y (t) = K(t)αH(t)β (A(t)L(t))1−α−β

where H(t) is the stock of human capital. We assume α + β < 1, since ifα + β = 1, there are constant returns to scale in accumulable factors alone,

Page 203: Macroeconomics Phd Lectures notes

9.2. THE SOLOW MODEL AND ITS EMPIRICAL EVALUATION 197

which prevents the existence of a balanced growth path (the model basicallybecomes an AK-model to be discussed below. We will specify below how tomeasure human capital (or better: investment into human capital) in the data.The capital accumulation equations are now given by

K(t) = skY (t)− δK(t)

H(t) = shY (t)− δH(t)

Expressing all equations in per-effective labor units yields (where η(t) = H(t)A(t)L(t)

ξ(t) = κ(t)aη(t)β

κ(t) = skξ(t)− (n+ δ + g)κ(t)

η(t) = shξ(t)− (n+ δ + g)η(t)

Obviously a unique positive steady state exists which can be computed as before

κ∗ =

Ãs1−βk sβhn+ δ + g

! 11−α−β

η∗ =

µsαk s

1−αh

n+ δ + g

¶ 11−α−β

ξ∗ = (κ∗)α (η∗)β

and the associated balanced growth path has

y(t) = A(0)egtξ∗

= A(0)egt

Ãs1−βk sβhn+ δ + g

! α1−α−β µ

sαks1−αh

n+ δ + g

¶ β1−α−β

Taking logs yields

ln(y(t)) = ln(A(0)) + gt+ b1 ln(sk) + b2 ln(sh) + b3 ln(n+ δ + g)

where b1 =α

1−α−β , b2 =β

1−α−β and b3 = − α+β1−α−β . Making the same assump-

tions about how to bring a stochastic component into the completely determin-istic model yields the regression equation

ln(yi) = a+ b1 ln(sik) + b2 ln(s

ih) + b3 ln(n

i + δ + g) + εi

The main problem in estimating this regression (apart from the validity of theorthogonality assumption of errors and instruments) is to construct reasonabledata for the savings rate of human capital. Ideally we would measure all theresources flowing into investment that increases the stock of human capital,including investment into education, health and so forth. For now let’s limit

Page 204: Macroeconomics Phd Lectures notes

198 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

attention to investment into education. Mankiw et al.’s measure of the invest-ment rate of education is the fraction of the total working age population thatgoes to secondary school, as found in data collected by the UNESCO, i.e.

sh =S

L

where S is the number of people in the labor force that go to school (and forgowages as unskilled workers) and L is the total labor force. Why may this bea good proxy for the investment share of output into education? Investmentexpenditures for education include new buildings of the universities, salariesof teachers, and most significantly, the forgone wages of the students in school.Let’s assume that forgone wages are the only input for human capital investment(if the other inputs are proportional to this measure, the argument goes throughunchanged). Let the people in school forgo wages wL as unskilled workers. Totalforgone earnings are then wLS and the investment share of output into humancapital is wLS

Y . But the wage of an unskilled worker is given (under perfectcompetition) by its marginal product

wL = (1− α− β)K(t)αH(t)βA(t)1−α−βL(t)−α−β

so that

wLS

Y=wLLS

Y L= (1− α− β)

S

L= (1− α− β)sh

so that the measure that the authors employ is proportional to a “theoreticallymore ideal” measure of the human capital savings rate. Noting that ln((1−α−β)sh) = ln(1 − α − β) + ln(sh) one immediately see that the proportionalityfactor will only affect the estimate of the constant, but not the estimates of thebi.The results of estimating the augmented regression by OLS are given in

Table 3

Table 3

a b1 b2 b3 R2

6.89(1.17)

0.69(0.13)

0.66(0.07)

−1.73(0.41)

0.78

The results are quite remarkable. First of all, almost 80% of the variation ofcross-country income differences is explained by differences in savings rates inphysical and human capital This is a huge number for cross-sectional regressions.Second, all parameter estimates are highly significant and have the right sign.In addition we (i.e. Mankiw, Romer and Weil) seem to have found a remedyfor the excessively high implied estimates for α. Now the estimates for bi implyalmost precisely α = β = 1

3 and the one overidentifying restriction on the b0is

can’t be rejected at standard confidence levels (although b3 is a bit high). The

Page 205: Macroeconomics Phd Lectures notes

9.3. THE RAMSEY-CASS-KOOPMANS MODEL 199

final verdict is that with respect to explaining cross-country income differencesan augmented version of the Solow model does remarkably well. This is, asusual subject to the standard quarrels that there may be big problems withdata quality and that their method is not applicable for non-Cobb-Douglastechnology. On a more fundamental level the Solow model has methodologicalproblems and Mankiw et al.’s analysis leaves several questions wide open:

1. The assumption of a constant saving rate is a strong behavioral assumptionthat is not derived from any underlying utility maximization problem ofrational agents. Our next topic, the discussion of the Cass-Koopmans-Ramsey model will remedy exactly this shortcoming

2. The driving force of economic growth, technological progress, is model-exogenous; it is assumed, rather than endogenously derived. We will pickthis up in our discussion of endogenous growth models.

3. The cross-country variation of per-worker income is attributed to varia-tions in investment rates, which are taken to be exogenous. What is thenneeded is a theory of why investment rates differ across countries. I canprovide you with interesting references that deal with this problem, butwe will not talk about this in detail in class.

But now let’s turn to the first of these points, the introduction of endogenousdetermination of household’s saving rates.

9.3 The Ramsey-Cass-Koopmans Model

In this section we discuss the first logical extension of the Solow model. Insteadof assuming that households save at a fixed, exogenously given rate s we willanalyze a model in which agents actually make economic decisions; in particu-lar they make the decision how much of their income to consume in the currentperiod and how much to save for later. This model was first analyzed by theBritish mathematician and economist Frank Ramsey. He died in 1930 at age29 from tuberculosis, not before he wrote two of the most influential economicspapers ever to be written. We will discuss a second pathbreaking idea of hisin our section on optimal fiscal policy. Ramsey’s ideas were taken up indepen-dently by David Cass and Tjelling Koopmans in 1965 and have now become thesecond major workhorse model in modern macroeconomics, besides the OLGmodel discussed previously. In fact, in Section 3 of these notes we discussedthe discrete-time version of this model and named it the neoclassical growthmodel. Now we will in fact incorporate economic growth into the model, whichis somewhat more elegant to do in continuous time, although there is nothingconceptually difficult about introducing growth into the discrete-time version-a useful exercise.

Page 206: Macroeconomics Phd Lectures notes

200 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

9.3.1 Mathematical Preliminaries: Pontryagin’s MaximumPrinciple

Intriligator, Chapter 14

9.3.2 Setup of the Model

Our basic assumptions made in the previous section are carried over. There isa representative, infinitely lived family (dynasty) in our economy that grows atpopulation growth rate n > 0 over time, so that, by normalizing the size of thepopulation at time 0 to 1 we have that L(t) = ent is the size of the family (orpopulation) at date t. We will treat this dynasty as a single economic agent.There is no uncertainty in this economy and all agents are assumed to haveperfect foresight.Production takes place with a constant returns to scale production function

Y (t) = F (K(t), A(t)L(t))

where the level of technology grows at constant rate g > 0, so that, normalizingA(0) = 1 we find that the level of technology at date t is given by A(t) = egt.The aggregate capital stock evolves according to

K(t) = F (K(t), A(t)L(t))− δK(t)− C(t) (9.15)

i.e. the net change in the capital stock is given by that fraction of output thatis not consumed by households, C(t) or by depreciation δK(t). Alternatively,this equation can be written as

K(t) + δK(t) = F (K(t), A(t)L(t))− C(t)

which simply says that aggregate gross investment K(t)+δK(t) equals aggregatesaving F (K(t), A(t)L(t))−C(t) (note that the economy is closed and there is nogovernment). As before this equation can be expressed in labor-intensive form:

define c(t) = C(t)L(t) as consumption per capita (or worker) and ζ(t) = C(t)

A(t)L(t)

as consumption per labor efficiency unit (the Greek symbol is called a “zeta”).Then we can rewrite (9.10) as, using the same manipulations as before

κ(t) = f(κ(t))− ζ(t)− (n+ δ + g)κ(t) (9.16)

Again f is assumed to have all the properties as in the previous section. Weassume that the initial endowment of capital is given by K(0) = κ(0) = κ0 > 0So far we just discussed the technology side of the economy. Now we want

to describe the preferences of the representative family. We assume that thefamily values streams of per-capita consumption c(t)t∈[0,∞) by

u(c) =

Z ∞0

e−ρtU(c(t))dt

Page 207: Macroeconomics Phd Lectures notes

9.3. THE RAMSEY-CASS-KOOPMANS MODEL 201

where ρ > 0 is a time discount factor. Note that this implicitly discounts utilityof agents that are born at later periods. Ramsey found this to be unethicaland hence assumed ρ = 0. Here U(c) is the instantaneous utility or felicityfunction.10 In most of our discussion we will assume that the period utilityfunction is of constant relative risk aversion (CRRA) form, i.e.

U(c) =

½c1−σ1−σ if σ 6= 1ln(c) if σ = 1

Under our assumption of CRRA11 we can rewrite

e−ρtU(c(t)) = e−ρtc(t)1−σ

1− σ

= e−ρt(ζ(t)egt)

1−σ

1− σ

= e−(ρ−g(1−σ))tζ(t)1−σ

1− σ

and we assume ρ > g(1− σ). Define ρ = ρ− g(1− σ). We therefore can rewritethe utility function of the dynasty as

u(ζ) =

Z ∞0

e−ρtζ(t)1−σ

1− σdt (9.17)

=

Z ∞0

e−ρtU(ζ(t))dt (9.18)

where σ = 1 is understood to be the log-case. As before note that, once weknow the variables κ(t) and ζ(t) we can immediately determine per capita con-sumption c(t) = ζ(t)egt and the per capita capital stock k(t) = κ(t)egt andoutput y(t) = egtf(κ(t)). Aggregate consumption, output and capital stock canbe deduced similarly.This completes the description of the environment. We will now, in turn, de-

scribe Pareto optimal and competitive equilibrium allocations and argue (heuris-tically) that they coincide.

10An alternative, so-called Benthamite (after British philosopher Jeremy Bentham) felicityfunction would read as L(t)U(c(t)). Since L(t) = ent we immediately see

e−ρtL(t)U(c(t))

= e−(ρ−n)tU(c(t))

and hence we would have the same problem with adjusted time discount factor, and we wouldneed to make the additional assumption that ρ > n.11Some of the subsequent analysis could be carried out with more general assumptions on

the period utility functions. However for the existence of a balanced growth path one has toassume CRRA, so I don’t see much of a point in higher degree of generality that in some pointof the argument has to be dispensed with anyway.For an extensive discussion of the properties of the CRRA utility function see the appendix

to Chapter 2 and HW1.

Page 208: Macroeconomics Phd Lectures notes

202 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

9.3.3 Social Planners Problem

The first question is how a social planner would allocate consumption and savingover time. Note that in this economy there is a single agent, so the problem ofthe social planner is reduced from the OLG model to only intertemporal (andnot also intergenerational) allocation of consumption. An allocation is a pair offunctions κ(t) : [0,∞)→ R and ζ(t) : [0,∞)→ R.

Definition 96 An allocation (κ, ζ) is feasible if it satisfies κ(0) = κ0, κ(t), ζ(t) ≥0 and (9.16) for all t ∈ [0,∞).

Definition 97 An allocation (κ∗, ζ∗) is Pareto optimal if it is feasible and ifthere is no other feasible allocation (κ, ζ) such that u(ζ) > u(ζ∗).

It is obvious that (κ∗, ζ∗) is Pareto optimal, if and only if it solves the socialplanner problem

max(κ,ζ)≥0

Z ∞0

e−ρtU(ζ(t))dt (9.19)

s.t. κ(t) = f(κ(t))− ζ(t)− (n+ δ + g)κ(t)

κ(0) = κ0

This problem can be solved using Pontryagin’s maximum principle. Thestate variable in this problem is κ(t) and the control variable is ζ(t). Let by λ(t)denote the co-state variable corresponding to κ(t). Forming the present valueHamiltonian and ignoring nonnegativity constraints12 yields

H(t,κ, ζ,λ) = e−ρtU(ζ(t)) + λ(t) [f(κ(t))− ζ(t)− (n+ δ + g)κ(t)]

Sufficient conditions for an optimal solution to the planners problem (9.19) are13

∂H(t,κ, ζ,λ)∂ζ(t)

= 0

λ(t) = −∂H(t,κ, ζ,λ)∂κ(t)

limt→∞λ(t)κ(t) = 0

The last condition is the so-called transversality condition (TVC). This yields

e−ρtU 0(ζ(t)) = λ(t) (9.20)

λ(t) = − (f 0(κ(t))− (n+ δ + g))λ(t) (9.21)

limt→∞λ(t)κ(t) = 0 (9.22)

12Given the functional form assumptions this is unproblematic.13I use present value Hamiltonians. You should do the same derivation using current value

Hamiltonians, as, e.g. in Intriligator, Chapter 16.

Page 209: Macroeconomics Phd Lectures notes

9.3. THE RAMSEY-CASS-KOOPMANS MODEL 203

plus the constraint

κ(t) = f(κ(t))− ζ(t)− (n+ δ + g)κ(t) (9.23)

Now we eliminate the co-state variable from this system. Differentiating (9.20)with respect to time yields

λ(t) = e−ρtU00(ζ(t))ζ(t)− ρe−ρtU 0(ζ(t))

or, using (9.20)

λ(t)

λ(t)=

ζ(t)U00(ζ(t))

U 0(ζ(t))− ρ (9.24)

Combining (9.24) with (9.21) yields

ζ(t)U00(ζ(t))

U 0(ζ(t))= − (f 0(κ(t))− (n+ δ + g + ρ)) (9.25)

or multiplying both sides by ζ(t) yields

ζ(t)ζ(t)U

00(ζ(t))

U 0(ζ(t))= − (f 0(κ(t))− (n+ δ + g + ρ)) ζ(t)

Using our functional form assumption on the utility function U(ζ) = ζ1−σ1−σ we

obtain for the coefficient of relative risk aversion − ζ(t)U00(ζ(t))

U 0(ζ(t)) = σ and hence

ζ(t) =1

σ(f 0(κ(t))− (n+ δ + g + ρ)) ζ(t)

Note that for the isoelastic case (σ = 1) we have that ρ = ρ and hence theequation becomes

ζ(t) = (f 0(κ(t))− (n+ δ + g + ρ)) ζ(t)

The transversality condition can be written as

limt→∞λ(t)κ(t) = lim

t→∞ e−ρtU 0(ζ(t))κ(t) = 0

Hence any allocation (κ, ζ) that satisfies the system of nonlinear ordinary dif-ferential equations

ζ(t) =1

σ(f 0(κ(t))− (n+ δ + g + ρ)) ζ(t) (9.26)

κ(t) = f(κ(t))− ζ(t)− (n+ δ + g)κ(t) (9.27)

with the initial condition κ(0) = κ0 and terminal condition (TVC)

limt→∞ e

−ρtU 0(ζ(t))κ(t) = 0

is a Pareto optimal allocation. We now want to analyze the dynamic system(9.26)− (9.27) in more detail.

Page 210: Macroeconomics Phd Lectures notes

204 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

Steady State Analysis

Before analyzing the full dynamics of the system we look at the steady stateof the optimal allocation. A steady state satisfies ζ(t) = κ(t) = 0. Hence fromequation (9.26) we have14, denoting steady state capital and consumption perefficiency units by ζ∗ and κ∗

f 0(κ∗) = (n+ δ + g + ρ) (9.28)

The unique capital stock κ∗ satisfying this equation is called the modified goldenrule capital stock.The “modified” comes from the following consideration. Suppose there is

no technological progress, then the modified golden rule capital stock κ∗ = k∗

satisfies

f 0(k∗) = (n+ δ + ρ) (9.29)

The golden rule capital stock is that capital stock per worker kg that maxi-mizes per-capita consumption. The steady state capital accumulation condition(without technological progress) is (see (9.27))

c = f(k)− (n+ δ)k

Hence the original golden rule capital stock satisfies15

f 0(kg) = n+ δ

and hence k∗ < kg. The social planner optimally chooses a capital stock perworker k∗ below the one that would maximize consumption per capita. So eventhough the planner could increase every person’s steady state consumption byincreasing the capital stock, taking into account the impatience of individualsthe planner finds it optimal not to do so.Equation (9.28) or (9.29) indicate that the exogenous parameters governing

individual time preference, population and technology growth determine theinterest rate and the marginal product of capital. The production technologythen determines the unique steady state capital stock and the unique steadystate consumption from (9.27) as

ζ∗ = f(κ∗)− (n+ δ + g)κ∗

The Phase Diagram

It is in general impossible to solve the two-dimensional system of differentialequations analytically, even for the simple example for which we obtain an

14There is the trivial steady state κ∗ = ζ∗ = 0. We will ignore this steady state from nowon, as it only is optimal when κ(0) = κ0.15Note that the golden rule capital stock had special significance in OLG economies. In

particular, any steady state equilibrium with capital stock above the golden rule was shownto be dynamically inefficient.

Page 211: Macroeconomics Phd Lectures notes

9.3. THE RAMSEY-CASS-KOOPMANS MODEL 205

analytical solution in the Solow model. A powerful tool when analyzing thedynamics of continuous time economies turn out to be so-called phase diagrams.Again, the dynamic system to be analyzed is

ζ(t) =1

σ(f 0(κ(t))− (n+ δ + g + ρ)) ζ(t)

κ(t) = f(κ(t))− ζ(t)− (n+ δ + g)κ(t)

with initial condition κ(0) = κ0 and terminal transversality condition limt→∞ e−ρtU 0(ζ(t))κ(t) =0. We will analyze the dynamics of this system in (κ, ζ) space. For any givenvalue of the pair (κ, ζ) ≥ 0 the dynamic system above indicates the change ofthe variables κ(t) and ζ(t) over time. Let us start with the first equation.

The locus of values for (κ, ζ) for which ζ(t) = 0 is called an isocline; it is thecollection of all points (κ, ζ) for which ζ(t) = 0. Apart from the trivial steadystate we have ζ(t) = 0 if and only if κ(t) satisfies f 0(κ(t))− (n+ δ+ g+ ρ) = 0,or κ(t) = κ∗. Hence in the (κ, ζ) plane the isocline is a vertical line at κ(t) = κ∗.Whenever κ(t) > κ∗ (and ζ(t) > 0), then ζ(t) < 0, i.e. ζ(t) declines. Weindicate this in Figure 9.9 with vertical arrows downwards at all points (κ, ζ)for which κ < κ∗. Reversely, whenever κ < κ∗ we have that ζ(t) > 0, i.e. ζ(t)increases. We indicate this with vertical arrows upwards at all points (κ, ζ) atwhich κ < κ∗. Similarly we determine the isocline corresponding to the equationκ(t) = f(κ(t))− ζ(t)− (n+ δ + g)κ(t). Setting κ(t) = 0 we obtain all points in(κ, ζ)-plane for which κ(t) = 0, or ζ(t) = f(κ(t))− (n+δ+g)κ(t). Obviously forκ(t) = 0 we have ζ(t) = 0. The curve is strictly concave in κ(t) (as f is strictlyconcave), has its maximum at κg > κ∗ solving f 0(κg) = (n+ δ + g) and againintersects the horizontal axis for κ(t) > κg solving f(κ(t)) = (n + δ + g)κ(t).Hence the isocline corresponding to κ(t) = 0 is hump-shaped with peak at κg.

For all (κ, ζ) combinations above the isocline we have ζ(t) > f(κ(t))− (n+δ + g)κ(t), hence κ(t) < 0 and hence κ(t) is decreasing. This is indicated byhorizontal arrows pointing to the left in Figure 9.9. Correspondingly, for all(κ, ζ) combinations below the isocline we have ζ(t) < f(κ(t))− (n+ δ + g)κ(t)and hence κ(t) > 0; i.e. κ(t) is increasing, which is indicated by arrows pointingto the right.

Note that we have one initial condition for the dynamic system, κ(0) = κ0.The arrows indicate the direction of the dynamics, starting from κ(0). However,one initial condition is generally not enough to pin down the behavior of thedynamic system over time, i.e. there may be several time paths of (κ(t), ζ(t))that are an optimal solution to the social planners problem. The question is,basically, how the social planner should choose ζ(0). Once this choice is madethe dynamic system as described by the phase diagram uniquely determines theoptimal path of capital and consumption. Possible such paths are traced out inFigure 9.10.

We now want to argue two things: a) for a given κ(0) > 0 any choice ζ(0) ofthe planner leading to a path not converging to the steady state (κ∗, ζ∗) cannotbe an optimal solution and b) there is a unique stable path leading to the steadystate. The second property is called-saddle-path stability of the steady state and

Page 212: Macroeconomics Phd Lectures notes

206 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

ζ (t)

.ζ (t)=0

ζ *

κ* κ(t)

.κ(t)=0

Figure 9.9:

the unique stable path is often called a saddle path (or a one-dimensional stablemanifold).

Let us start with the first point. There are three possibilities for any pathstarting with arbitrary κ(0) > 0; they either go to the unique steady state, theylead to the point E (as trajectories starting from points A or C), or they go topoints with κ = 0 such as trajectories starting at B or D. Obviously trajectorieslike A and C that don’t converge to E violate the nonnegativity of consumptionζ(t) = 0 in finite amount of time. But a trajectory converging asymptoticallyto E violates the transversality condition

limt→∞ e

−ρtU 0(ζ(t))κ(t) = 0

As the trajectory converges to E, κ(t) converges to a κ > κg > κ∗ > 0 and from

Page 213: Macroeconomics Phd Lectures notes

9.3. THE RAMSEY-CASS-KOOPMANS MODEL 207

ζ (t)

.ζ (t)=0

ζ *

κ(0) κ* κ(t)

.κ(t)=0

ζ (0)

A

B

C

D

Saddle path

Saddle path

E

Figure 9.10:

(9.25) we have, since dU 0(ζ(t))dt = ζ(t)U

00(ζ(t))

dU 0(ζ(t))dt

U 0(ζ(t))= −f 0(κ(t)) + (n+ δ + g + ρ) > ρ > 0

i.e. the growth rate of marginal utility of consumption is bigger than ρ asthe trajectory approaches A. Given that κ approaches κ it is clear that thetransversality condition is violated for all those trajectories.Now consider trajectories like B or D. If, in finite amount of time, the

trajectory hits the ζ-axis, then κ(t) = ζ(t) = 0 from that time onwards, which,given the Inada conditions imposed on the utility function can’t be optimal. Itmay, however, be possible that these trajectories asymptotically go to (κ, ζ) =(0,∞). That this can’t happen can be shown as follows. From (9.27) we have

κ(t) = f(κ(t))− ζ(t)− (n+ δ + g)κ(t)

Page 214: Macroeconomics Phd Lectures notes

208 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

which is negative for all κ(t) < κ∗. Differentiating both sides with respect totime yields

dκ(t)

dt=d2κ(t)

dt2= (f 0(κ(t))− (n+ δ + g)) κ(t)− ζ(t) < 0

since along a possible asymptotic path ζ(t) > 0. So not only does κ(t) decline,but it declines at increasing pace. Asymptotic convergence to the ζ-axis, how-ever, would require κ(t) to decline at a decreasing pace. Hence all paths like Bor D have to reach κ(t) = 0 at finite time and therefore can’t be optimal. Thesearguments show that only trajectories that lead to the unique positive steadystate (κ∗, ζ∗) can be optimal solutions to the planner problemIn order to prove the second claim that there is a unique such path for

each possible initial condition κ(0) we have to analyze the dynamics around thesteady state.

Dynamics around the Steady State

We can’t solve the system of differential equations explicitly even for simpleexamples. But from the theory of linear approximations we know that in aneighborhood of the steady state the dynamic behavior of the nonlinear systemis characterized by the behavior of the linearized system around the steady state.Remember that the first order Taylor expansion of a function f : Rn → Raround a point x∗ ∈ Rn is given by

f(x) = f(x∗) +∇f(x∗) · (x− x∗)where ∇f(x∗) ∈ Rn is the gradient (vector of partial derivatives) of f at x∗. Inour case we have x∗ = (κ∗, ζ∗), and two functions f, g defined as

ζ(t) = f(κ(t), ζ(t)) =1

σ(f 0(κ(t))− (n+ δ + g + ρ)) ζ(t)

κ(t) = g(κ(t), ζ(t)) = f(κ(t))− ζ(t)− (n+ δ + g)κ(t)

Obviously we have f(κ∗, ζ∗) = g(κ∗, ζ∗) = 0 since (κ∗, ζ∗) is a steady state.Hence the linear approximation around the steady state takes the formµ

ζ(t)κ(t)

¶≈

µ1σ (f

0(κ(t))− (n+ δ + g + ρ)) 1σf

00(κ(t))ζ(t)−1 f 0(κ(t))− (n+ δ + g)

¶¯(ζ(t),κ(t))=(ζ∗,κ∗)

·µ

ζ(t)− ζκ(t)− κ

=

µ0 1

σf00(κ∗)ζ∗

−1 ρ

¶·µ

ζ(t)− ζ∗

κ(t)− κ∗

¶(9

This two-dimensional linear difference equation can now be solved analyti-cally. It is easiest to obtain the qualitative properties of this system by reducingit two a single second order differential equation. Differentiate the second equa-tion with respect to time to obtain

κ(t) = −ζ(t) + ρκ(t)

Page 215: Macroeconomics Phd Lectures notes

9.3. THE RAMSEY-CASS-KOOPMANS MODEL 209

Defining β = − 1σf

00(κ∗)ζ∗ > 0 and substituting in from (9.30) for ζ(t) yields

κ(t) = β (κ(t)− κ∗) + ρκ(t)

κ(t)− ρκ(t)− βκ(t) = −βκ∗ (9.31)

We know how to solve this second order differential equation; we just have tofind the general solution to the homogeneous equation and a particular solutionto the nonhomogeneous equation, i.e.

κ(t) = κg(t) + κp(t)

It is straightforward to verify that a particular solution to the nonhomogeneousequation is given by κp(t) = κ∗. With respect to the general solution to thehomogeneous equation we know that its general form is given by

κg(t) = C1eλ1t + C2e

λ2t

where C1, C2 are two constants and λ1,λ2 are the two roots of the characteristicequation

λ2 − ρλ− β = 0

λ1,2 =ρ

2±sβ +

ρ2

4

We see that the two roots are real, distinct and one is bigger than zero and oneis less than zero. Let λ1 be the smaller and λ2 be the bigger root. The fact thatone of the roots is bigger, one is smaller than one implies that locally aroundthe steady state the dynamic system is saddle-path stable, i.e. there is a uniquestable manifold (path) leading to the steady state. For any value other thanC2 = 0 we will have limt→∞ κ(t) =∞ (or −∞) which violates feasibility. Hencewe have that

κ(t) = κ∗ + C1eλ1t

(remember that λ1 < 0). Finally C1 is determined by the initial conditionκ(0) = κ0 since

κ(0) = κ∗ + C1C1 = κ(0)− κ∗

and hence the solution for κ is

κ(t) = κ∗ + (κ(0)− κ∗) eλ1t

and the corresponding solution for ζ can be found from

κ(t) = −ζ(t) + ζ∗ + ρ (κ(t)− κ∗)ζ(t) = ζ∗ + ρ (κ(t)− κ∗)− κ(t)

Page 216: Macroeconomics Phd Lectures notes

210 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

by simply using the solution for κ(t). Hence for any given κ(0) there is aunique optimal path (κ(t), ζ(t)) which converges to the steady state (κ∗, ζ∗).Note that the speed of convergence to the steady state is determined by |λ1| =¯ρ2 −

q− 1

σf00(κ∗)ζ∗ + ρ2

4

¯which is increasing in − 1

σ and decreasing in ρ. The

higher the intertemporal elasticity of substitution, the more are individuals will-ing to forgo early consumption for later consumption an the more rapid doescapital accumulation towards the steady state occur. The higher the effectivetime discount rate ρ, the more impatient are households and the stronger theyprefer current over future consumption, inducing a lower rate of capital accu-mulation.So far what have we showed? That only paths converging to the unique

steady state can be optimal solutions and that locally, around the steady statethis path is unique, and therefore was referred to as saddle path. This also meansthat any potentially optimal path must hit the saddle path in finite time.Hence there is a unique solution to the social planners problem that is graph-

ically given as follows. The initial condition κ0 determines the starting pointof the optimal path κ(0). The planner then optimally chooses ζ(0) such as tojump on the saddle path. From then on the optimal sequences (κ(t), ζ(t))t∈[0,∞)are just given by the segment of the saddle path from κ(0) to the steady state.Convergence to the steady state is asymptotic, monotonic (the path does notjump over the steady state) and exponential. This indicates that eventually,once the steady state is reached, per capita variables grow at constant rates gand aggregate variables grow at constant rates g + n:

c(t) = egtζ∗

k(t) = egtκ∗

y(t) = egtf(κ∗)C(t) = e(n+g)tζ∗

K(t) = e(n+g)tκ∗

Y (t) = e(n+g)tf(κ∗)

Hence the long-run behavior of this model is identical to that of the Solow model;it predicts that the economy converges to a balanced growth path at which allper capita variables grow at rate g and all aggregate variables grow at rate g+n.In this sense we can understand the Cass-Koopmans-Ramsey model as a microfoundation of the Solow model, with predictions that are quite similar.

9.3.4 Decentralization

In this subsection we want to demonstrate that the solution to the social plan-ners problem does correspond to the (unique) competitive equilibrium allocationand we want to find prices supporting the Pareto optimal allocation as a com-petitive equilibrium.In the decentralized economy there is a single representative firm that rents

capital and labor services to produce output. As usual, whenever the firm

Page 217: Macroeconomics Phd Lectures notes

9.3. THE RAMSEY-CASS-KOOPMANS MODEL 211

does not own the capital stock its intertemporal profit maximization problem isequivalent to a continuum of static maximization problems

maxK(t),L(t)≥0

F (K(t), A(t) + (t))− r(t)K(t)− w(t)L(t) (9.32)

taking w(t) and r(t), the real wage rate and rental rate of capital, respectively,as given.The representative household (dynasty) maximizes the family’s utility by

choosing per capita consumption and per capita asset holding at each instantin time. Remember that preferences were given as

u(c) =

Z ∞0

e−ρtU(c(t))dt (9.33)

The only asset in this economy is physical capital16 on which the return isr(t)− δ. As before we could introduce notation for the real interest rate i(t) =r(t)−δ but we will take a shortcut and use r(t)−δ in the period household budgetconstraint. This budget constraint (in per capita terms, with the consumptiongood being the numeraire) is given by

c(t) + a(t) + na(t) = w(t) + (r(t)− δ) a(t) (9.34)

where a(t) = A(t)L(t) are per capita asset holdings, with a(0) = κ0 given. Note that

the term na(t) enters because of population growth: in order to, say, keep theper-capita assets constant, the household has to spend na(t) units to account forits growing size.17 As with discrete time we have to impose a condition on thehousehold that rules out Ponzi schemes. At the same time we do not preventthe household from temporarily borrowing (for the households a is perceived asan arbitrary asset, not necessarily physical capital). A standard condition thatis widely used is to require that the household debt holdings in the limit do nogrow at a faster rate than the interest rate, or alternatively put, that the timezero value of household debt has to be nonnegative in the limit.

limt→∞ a(t)e

− R t0(r(τ)−δ−n)dτ ≥ 0 (9.35)

Note that with a path of interest rates r(t) − δ, the value of one unit of theconsumption good at time t in units of the period consumption good is givenby e−

R t0(r(τ)−δ)dτ . We immediately have the following definition of equilibrium

16Introducing a second asset, say government bonds, is straightforward and you should doit as an exercise.17The household’s budget constraint in aggregate (not per capita) terms is

C(t) + A(t) = L(t)w(t) + (r(t)− δ)A(t)

Dividing by L(t) yields

c(t) +A(t)

L(t)= w(t) + (r(t)− δ) a(t)

and expandingA(t)L(t)

gives the result in the main text.

Page 218: Macroeconomics Phd Lectures notes

212 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

Definition 98 A sequential markets equilibrium are allocations for the house-hold (c(t), a(t))t∈[0,∞), allocations for the firm (K(t), L(t))t∈[0,∞) and prices(r(t), w(t))t∈[0,∞) such that

1. Given prices (r(t), w(t))t∈[0,∞) and κ0, the allocation (c(t), a(t))t∈[0,∞)maximizes (9.33) subject to (9.34), for all t, and (9.35) and c(t) ≥ 0.

2. Given prices (r(t), w(t))t∈[0,∞), the allocation (K(t), L(t))t∈[0,∞) solves(9.32)

3.

L(t) = ent

L(t)a(t) = K(t)

L(t)c(t) + K(t) + δK(t) = F (K(t), L(t))

This definition is completely standard; the three market clearing conditionsare for the labor market, the capital market and the goods market, respec-tively. Note that we can, as for the discrete time case, define an Arrow-Debreuequilibrium and show equivalence between Arrow-Debreu equilibria and sequen-tial market equilibria under the imposition of the no Ponzi condition (9.35). Aheuristic argument will do here. Rewrite (9.34) as

c(t) = w(t) + (r(t)− δ) a(t)− a(t)− na(t)

then multiply both sides by e−R t0(r(τ)−n−δ)dτ and integrate from t = 0 to t = T

to getZ T

0

c(t)e−R t0(r(τ)−n−δ)dτdt =

Z T

0

w(t)e−R t0(r(τ)−n−δ)dτdt (9.36)

−Z T

0

[a(t)− (r(t)− n− δ) a(t)] e−R t0(r(τ)−n−δ)dτdt

But if we define

F (t) = a(t)e−R t0(r(τ)−n−δ)dτ

then

F 0(t) = a(t)e−R t0(r(τ)−n−δ)dτdt− a(t)e−

R t0(r(τ)−n−δ)dτ [r(t)− δ − n]

= [a(t)− (r(t)− n− δ) a(t)] e−R t0(r(τ)−n−δ)dτ

so that (9.36) becomesZ T

0

c(t)e−R t0(r(τ)−n−δ)dτdt =

Z T

0

w(t)e−R t0(r(τ)−n−δ)dτdt+ F (0)− F (T )

=

Z T

0

w(t)e−R t0(r(τ)−n−δ)dτdt+ a(0)− a(T )e−

R T0(r(τ)−n−δ)dτ

Page 219: Macroeconomics Phd Lectures notes

9.3. THE RAMSEY-CASS-KOOPMANS MODEL 213

Now taking limits with respect to T and using (9.35) yieldsZ ∞0

c(t)e−R t0(r(τ)−n−δ)dτdt =

Z ∞0

w(t)e−R t0(r(τ)−n−δ)dτdt+ a(0)

or defining Arrow-Debreu prices as p(t) = e−R t0(r(τ)−δ)dτ we haveZ ∞

0

p(t)C(t)dt =

Z ∞0

p(t)L(t)w(t)dt+ a(0)L(0)

where C(t) = L(t)c(t) and we used the fact that L(0) = 1. But this is a stan-dard Arrow-Debreu budget constraint. Hence by imposing the correct no Ponzicondition we have shown that the collection of sequential budget constraints isequivalent to the Arrow Debreu budget constraint with appropriate prices

p(t) = e−R t0(r(τ)−δ)dτ

The rest of the proof that the set of Arrow-Debreu equilibrium allocations equalsthe set of sequential markets equilibrium allocations is obvious.18

We now want to characterize the equilibrium; in particular we want to showthat the resulting dynamic system is identical to that arising for the socialplanner problem, suggesting that the welfare theorems hold for this economy.From the firm’s problem we obtain

r(t) = FK(K(t), A(t)L(t)) = FK

µK(t)

A(t)L(t), 1

¶(9.37)

= f 0(κ(t))

and by zero profits in equilibrium

w(t)L(t) = F (K(t), A(t)L(t))− r(t)K(t) (9.38)

ω(t) =w(t)

A(t)= f(κ(t))− f 0(κ(t))κ(t)

w(t) = A(t) (f(κ(t))− f 0(κ(t))κ(t))From the goods market equilibrium condition we find as before (by dividing byA(t)L(t))

L(t)c(t) + K(t) + δK(t) = F (K(t), L(t))

κ(t) = f(κ(t))− (n+ δ + n)κ(t)− ζ(t) (9.39)

Now we analyze the household’s decision problem. First we rewrite the utilityfunction and the household’s budget constraint in intensive form. Making the

18Note that no equilbrium can exist for prices satisfying

limt→∞ p(t)L(t) = lim

t→∞ e− R t0 (r(τ)−δ−n)dτ > 0

because otherwise labor income of the family is unbounded.

Page 220: Macroeconomics Phd Lectures notes

214 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

assumption that the period utility is of CRRA form we again obtain (9.17).With respect to the individual budget constraint we obtain (again by dividingby A(t))

c(t) + a(t) + na(t) = w(t) + (r(t)− δ) a(t)

α(t) = ω(t) + (r(t)− (δ + n+ g))α(t)− ζ(t)

where α(t) = a(t)A(t) . The individual state variable is the per-capita asset holdings

in intensive form α(t) and the individual control variable is ζ(t). Forming theHamiltonian yields

H(t,α, ζ,λ) = e−ρtU(ζ(t)) + λ(t) [ω(t) + (r(t)− (δ + n+ g))α(t)− ζ(t)]

The first order condition yields

e−ρtU 0(ζ(t)) = λ(t) (9.40)

and the time derivative of the Lagrange multiplier is given by

λ(t) = − [r(t)− (δ + n+ g)]λ(t) (9.41)

The transversality condition is given by

limt→∞λ(t)α(t) = 0

Now we proceed as in the social planners case. We first differentiate (9.40) withrespect to time to obtain

e−ρtU 00(ζ(t))ζ(t)− ρe−ρtU 0(ζ(t)) = λ(t)

and use this and (9.40) to substitute out for the costate variable in (9.41) toobtain

λ(t)

λ(t)= − [r(t)− (δ + n+ g)]

= −ρ+ U00(ζ(t))ζ(t)U 0(ζ(t))

= −ρ− σζ(t)

ζ(t)

or

ζ(t) =1

σ[r(t)− (δ + n+ g + ρ)] ζ(t)

Note that this condition has an intuitive interpretation: if the interest rate ishigher than the effective subjective time discount factor, the individual values

Page 221: Macroeconomics Phd Lectures notes

9.4. ENDOGENOUS GROWTH MODELS 215

consumption tomorrow relatively higher than the market and hence ζ(t) > 0,i.e. consumption is increasing over time.Finally we use the profit maximization conditions of the firm to substitute

r(t) = f 0(κ(t)) to obtain

ζ(t) =1

σ[f 0(κ(t))− (δ + n+ g + ρ)] ζ(t)

Combining this with the resource constraint (9.39) gives us the same dynamicsystem as for the social planners problem, with the same initial condition κ(0) =κ0. And given that the capital market clearing condition reads L(t)a(t) = K(t)or α(t) = κ(t) the transversality condition is identical to that of the socialplanners problem. Obviously the competitive equilibrium allocation coincideswith the (unique) Pareto optimal allocation; in particular it also possesses thesaddle path property. Competitive equilibrium prices are simply given by

r(t) = f 0(κ(t))w(t) = A(t) (f(κ(t))− f 0(κ(t))κ(t))

Note in particular that real wages are growing at the rate of technologicalprogress along the balanced growth path. This argument shows that in con-trast to the OLG economies considered before here the welfare theorems apply.In fact, this section should be quite familiar to you; it is nothing else but arepetition of Chapter 3 in continuous time, executed to make you familiar withcontinuous time optimization techniques. In terms of economics, the currentmodel provides a micro foundation of the basic Solow model. It removes theproblem of a constant, exogenous saving rate. However the engine of growthis, as in the Solow model, exogenously given technological progress. The nextstep in our analysis is to develop models that do not assume economic growth,but rather derive it as an equilibrium phenomenon. These models are thereforecalled endogenous growth models (as opposed to exogenous growth models).

9.4 Endogenous Growth Models

The second main problem of the Solow model, which is shared with the Cass-Koopmans model of growth is that growth is exogenous: without exogenoustechnological progress there is no sustained growth in per capita income andconsumption. In this sense growth in these models is more assumed ratherthan derived endogenously as an equilibrium phenomenon. The key assumptiondriving the result, that, absent technological progress the economy will convergeto a no-growth steady state is the assumption of diminishing marginal productto the production factor that is accumulated, namely capital. As economiesgrow they accumulate more and more capital, which, with decreasing marginalproducts, yields lower and lower returns. Absent technological progress thisforce drives the economy to the steady state. Hence the key to derive sustainedgrowth without assuming it being created by exogenous technological progress

Page 222: Macroeconomics Phd Lectures notes

216 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

is to pose production technologies in which marginal products to accumulablefactors are not driven down as the economy accumulates these factors.We will start our discussion of these models with a stylized version of the

so called AK-model, then turn to models with externalities as in Romer (1986)and Lucas (1988) and finally look at Romer’s (1990) model of endogenous tech-nological progress.

9.4.1 The Basic AK-Model

Even though the basic AK-model may seem unrealistic it is a good first step toanalyze the basic properties of most one-sector competitive endogenous growthmodels. The basic structure of the economy is very similar to the Cass-Koopmansmodel. Assume that there is no technological progress. The representativehousehold again grows in size at population growth rate n > 0 and its prefer-ences are given by

U(c) =

Z ∞0

e−ρtc(t)1−σ

1− σdt

Its budget constraint is again given by

c(t) + a(t) + na(t) = w(t) + (r(t)− δ) a(t)

with initial condition a(0) = k0.We impose the same condition to rule out Ponzischemes as before

limt→∞ a(t)e

− R t0(r(τ)−δ−n)dτ ≥ 0

The main difference to the previous model comes from the specification of tech-nology. We assume that output is produced by a constant returns to scaletechnology only using capital

Y (t) = AK(t)

The aggregate resource constraint is, as before, given by

K(t) + δK(t) + C(t) = Y (t)

This completes the description of the model. The definition of equilibrium iscompletely standard and hence omitted. Also note that this economy does notfeature externalities, tax distortions or the like that would invalidate the welfaretheorems. So we could, in principle, solve a social planners problem to obtainequilibrium allocations and then find supporting prices. Given that for thiseconomy the competitive equilibrium itself is straightforward to characterize wewill take a shot at it directly.Let’s first consider the household problem. Forming the Hamiltonian and

carrying out the same manipulations as for the Cass-Koopmans model yields as

Page 223: Macroeconomics Phd Lectures notes

9.4. ENDOGENOUS GROWTH MODELS 217

Euler equation (note that there is no technological progress here)

c(t) =1

σ[r(t)− (n+ δ + ρ)] c(t)

γc(t) =c(t)

c(t)=1

σ[r(t)− (n+ δ + ρ)]

The transversality condition is given as

limt→∞λ(t)a(t) = lim

t→∞ e−ρtc(t)−σa(t) (9.42)

The representative firm’s problem is as before

maxK(t),L(t)≥0

AK(t)− r(t)K(t)− w(t)L(t)

and yields as marginal cost pricing conditions

r(t) = A

w(t) = 0

Hence the marginal product of capital and therefore the real interest rate areconstant across time, independent of the level of capital accumulated in theeconomy. Plugging into the consumption Euler equation yields

γc(t) =c(t)

c(t)=1

σ[A− (n+ δ + ρ)]

i.e. the consumption growth rate is constant (always, not only along a balancedgrowth path) and equal to A− (n+ δ + ρ). Integrating both sides with respectto time, say, until time t yields

c(t) = c(0)e1σ [A−(n+δ+ρ)]t (9.43)

where c(0) is an endogenous variable that yet needs to be determined. We nowmake the following assumptions on parameters

[A− (n+ δ + ρ)] > 0 (9.44)

1− σ

σ

·A− (n+ δ)− ρ

1− σ

¸= φ < 0 (9.45)

The first assumption, requiring that the interest rate exceeds the populationgrowth rate plus the time discount rate, will guarantee positive growth of percapita consumption. It basically requires that the production technology is pro-ductive enough to generate sustained growth. The second assumption assuresthat utility from a consumption stream satisfying (9.43) remains bounded sinceZ ∞

0

e−ρtc(t)1−σ

1− σdt =

Z ∞0

e−ρtc(0)1−σe

1−σσ [A−(n+δ+ρ)]t

1− σdt

=c(0)1−σ

1− σ

Z ∞0

e[1−σσ [A−(n+δ)− ρ

1−σ ]]tdt

< ∞ if and only if1− σ

σ

·A− (n+ δ)− ρ

1− σ

¸< 0

Page 224: Macroeconomics Phd Lectures notes

218 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

From the aggregate resource constraint we have

K(t) + δK(t) + C(t) = AK(t)

c(t) + k(t) = Ak(t)− (n+ δ)k(t) (9.46)

Dividing both sides by k(t) yields

γk(t) =k(t)

k(t)= A− (n+ δ)− c(t)

k(t)

In a balanced growth path γk(t) is constant over time, and hence k(t) is pro-portional to c(t), which implies that along a balanced growth path

γk(t) = γc(t) = A− (n+ δ + ρ)

i.e. not only do consumption and capital grow at constant rates (this is bydefinition of a balanced growth path), but they grow at the same rate A− (n+δ + ρ). We already saw that consumption always grows at a constant rate inthis model. We will now argue that capital does, too, right away from t = 0. Inother words, we will show that transition to the (unique) balanced growth pathis immediate.Plugging in for c(t) in equation (9.46) yields

k(t) = −c(0)e 1σ [A−(n+δ+ρ)]t +Ak(t)− (n+ δ)k(t)

which is a first order nonhomogeneous differential equation. The general solutionto the homogeneous equation is

kg(t) = C1e(A−n−δ)t

A particular solution to the nonhomogeneous equation is (verify this by plugginginto the differential equation)

kp(t) =−c(0)e 1σ [A−(n+δ+ρ)]t

φ

Hence the general solution to the differential equation is given by

k(t) = C1e(A−n−δ)t − c(0)

φe1σ [A−(n+δ+ρ)]t

where φ = 1−σσ

hA− (n+ δ)− ρ

1−σi< 0. Now we use that in equilibrium a(t) =

k(t). From the transversality condition we have that, using (9.43)

limt→∞ e

−ρtc(t)−σk(t) = limt→∞ e

−ρtc(0)−σe−[A−(n+δ+ρ)]t·C1e

(A−n−δ)t − c(0)φe1σ [A−(n+δ+ρ)]t

¸= c(0)−σ

·C1 lim

t→∞ e[−ρ−A+n+δ+ρ+A−n−δ]t − c(0)

φlimt→∞ e

[−ρ−A+n+δ+ρ+ 1σ [A−(n+δ+ρ)

= c(0)−σ·C1 − c(0)

φlimt→∞ e

1−σσ [A−(n+δ)− ρ

1−σ ]¸= 0 if and only if C1 = 0

Page 225: Macroeconomics Phd Lectures notes

9.4. ENDOGENOUS GROWTH MODELS 219

because of the assumed inequality in (9.45). Hence

k(t) = −c(0)φe1σ [A−(n+δ+ρ)]t = −c(t)

φ

i.e. the capital stock is proportional to consumption. Since we already foundthat consumption always grows at a constant rate γc = A− (n+ δ+ ρ), so doesk(t). The initial condition k(0) = k0 determines the level of capital, consump-tion c(0) = −φk(0) and output y(0) = Ak(0) that the economy starts from;subsequently all variables grow at constant rate γc = γk = γy. Note that in thismodel the transition to a balanced growth path from any initial condition k(0)is immediate.In this simple model we can explicitly compute the saving rate for any point

in time. It is given by

s(t) =Y (t)− C(t)

Y (t)=Ak(t)− c(t)Ak(t)

= 1 +φ

A= s ∈ (0, 1)

i.e. the saving rate is constant over time (as in the original Solow model andin contrast to the Cass-Koopmans model where the saving rate is only constantalong a balanced growth path).In the Solow and Cass-Koopmans model the growth rate of the economy

was given by γc = γk = γy = g, the growth rate of technological progress. Inparticular, savings rates, population growth rates, depreciation and the subjec-tive time discount rate affect per capita income levels, but not growth rates.In contrast, in the basic AK-model the growth rate of the economy is affectedpositively by the parameter governing the productivity of capital, A and neg-atively by parameters reducing the willingness to save, namely the effectivedepreciation rate δ + n and the degree of impatience ρ. Any policy affectingthese parameters in the Solow or Cass-Koopmans model have only level, but nogrowth rate effects, but have growth rate effects in the AK-model. Hence theformer models are sometimes referred to as “income level models” whereas theothers are referred to as “growth rate models”.With respect to their empirical predictions, the AK-model does not predict

convergence. Suppose all countries share the same characteristics in terms oftechnology and preferences, and only differ in terms of their initial capital stock.The Solow and Cass-Koopmans model then predict absolute convergence in in-come levels and higher growth rates in poorer countries, whereas the AK-modelpredicts no convergence whatsoever. In fact, since all countries share the samegrowth rate and all economies are on the balanced growth path immediately,initial differences in per capita capital and hence per capita income and con-sumption persist forever and completely. The absence of decreasing marginalproducts of capital prevents richer countries to slow down in their growth pro-cess as compared to poor countries. If countries differ with respect to theircharacteristics, the Solow and Cass-Koopmans model predict conditional con-vergence. The AK-model predicts that different countries grow at differentrates. Hence it may be possible that the gap between rich and poor countries

Page 226: Macroeconomics Phd Lectures notes

220 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

widen or that poor countries take over rich countries. Hence one importanttest of these two competing theories of growth is an empirical exercise to deter-mine whether we in fact see absolute and/or conditional convergence. Note thatwe discuss the predictions of the basic AK-model with respect to convergenceat length here because the following, more sophisticated models will share thequalitative features of the simple model.

9.4.2 Models with Externalities

The main assumption generating sustained growth in the last chapter was thepresence of constant returns to scale with respect to production factors thatare, in contrast to raw labor, accumulable. Otherwise eventually decreasingmarginal products set in and bring the growth process to a halt. One obviousunsatisfactory element of the previous model was that labor was not neededfor production and that therefore the capital share equals one. Even if oneinterprets capital broadly as including physical capital, this assumption may berather unrealistic. We, i.e. the growth theorist faces the following dilemma:on the one hand we want constant returns to scale to accumulable factors, onthe other hand we want labor to claim a share of income, on the third handwe can’t deal with increasing returns to scale on the firm level as this destroysexistence of competitive equilibrium. (At least) two ways out of this problemhave been proposed: a) there may be increasing returns to scale on the firmlevel, but the firm does not perceive it this way because part of its inputs comefrom positive externalities beyond the control of the firm b) a departure fromperfect competition towards monopolistic competition. We will discuss the maincontributions in both of these proposed resolutions.

Romer (1986)

We consider a simplified version of Romer’s (1986) model. This model is verysimilar in spirit and qualitative results to the one in the previous section. How-ever, the production technology is modified in the following form. Firms areindexed by i ∈ [0, 1], i.e. there is a continuum of firms of measure 1 that behavecompetitively. Each firm produces output according to the production function

yi(t) = F (ki(t), li(t)K(t))

where ki(t) and li(t) are labor and capital input of firm i, respectively, andK(t) =

Rki(t)di is the average capital stock in the economy at time t. We

assume that firm i, when choosing capital input ki(t), does not take into accountthe effect of ki(t) on K(t).

19 We make the usual assumption on F : constant

19Since we assume that there is a continuum of firms this assumption is completely rigourousas Z 1

0ki(t)di =

Z 1

0ki(t)di

as long as ki(t) = ki(t) for all but countably many agents.

Page 227: Macroeconomics Phd Lectures notes

9.4. ENDOGENOUS GROWTH MODELS 221

returns to scale with respect to the two inputs ki(t) and li(t)K(t), positive butdecreasing marginal products (we will denote by F1 the partial derivative withrespect to the first, by F2 the partial derivative with respect to the secondargument), and Inada conditions.

Note that F exhibits increasing returns to scale with respect to all threefactors of production

F (λki(t),λli(t)λK(t)) = F (λki(t),λ2li(t)K(t)) > λF (ki(t), li(t)K(t)) for all λ > 1

F (λki(t),λ [li(t)K(t)]) = λF (ki(t), li(t)K(t))

but since the firm does not realize its impact on K(t), a competitive equilibriumwill exist in this economy. It will, however, in general not be Pareto optimal.This is due to the externality in the production technology of the firm: a higheraggregate capital stock makes individual firm’s workers more productive, butfirms do not internalize this effect of the capital input decision on the aggregatecapital stock. As we will see, this will lead to less investment and a lower capitalstock than socially optimal.

The household sector is described as before, with standard preferences andinitial capital endowments k(0) > 0. For simplicity we abstract from populationgrowth (you should work out the model with population growth). However weassume that the representative household in the economy has a size of L identicalpeople (we will only look at type identical allocations). We do this in order todiscuss “scale effects”, i.e. the dependence of income levels and growth rates onthe size of the economy.

Since this economy is not quite as standard as before we define a competitiveequilibrium

Definition 99 A competitive equilibrium are allocations (c(t), a(t))t∈[0,∞) forthe representative household, allocations (ki(t), li(t))t∈[0,∞),i∈[0,1] for firms, anaggregate capital stock K(t)t∈[0,∞) and prices (r(t), w(t))t∈[0,∞) such that

1. Given (r(t), w(t))t∈[0,∞) (c(t), a(t))t∈[0,∞) solve

max(c(t),a(t))t∈[0,∞)

Z ∞0

e−ρtc(t)1−σ

1− σdt

s.t. c(t) + a(t) = w(t) + (r(t)− δ) a(t) with a(0) = k(0) given

c(t) ≥ 0

limt→∞ a(t)e

− R t0(r(τ)−δ)dτ ≥ 0

2. Given r(t), w(t) and K(t) for all t and all i, ki(t), li(t) solve

maxki(t),li(t)≥0

F (ki(t), li(t)K(t))− r(t)ki(t)− w(t)li(t)

Page 228: Macroeconomics Phd Lectures notes

222 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

3. For all t

Lc(t) + bK(t) + K(t)δ(t) =

Z 1

0

F (ki(t), li(t)K(t))diZ 1

0

li(t)di = LZ 1

0

ki(t)di = La(t)

4. For all t

Z 1

0

ki(t)di = K(t)

The first element of the equilibrium definition is completely standard. Inthe firm’s maximization problem the important feature is that the equilibriumaverage capital stock is taken as given by individual firms. The market clearingconditions for goods, labor and capital are straightforward. Finally the lastcondition imposes rational expectations: what individual firms perceive to bethe average capital stock in equilibrium is the average capital stock, given thefirms’ behavior, i.e. equilibrium capital demand.

Given that all L households are identical it is straightforward to define aPareto optimal allocation and it is easy to see that it must solve the following

Page 229: Macroeconomics Phd Lectures notes

9.4. ENDOGENOUS GROWTH MODELS 223

social planners problem20

max(c(t),K(t))t∈[0,∞)≥0

Z ∞0

e−ρtc(t)1−σ

1− σdt

s.t. Lc(t) + K(t) + δK(t) = F (K(t),K(t)L) with K(0) = Lk(0) given

Note that the social planner, in contrast to the competitively behaving firms,internalizes the effect of the average (aggregate) capital stock on labor produc-tivity. Let us start with this social planners problem. Forming the Hamiltonianand manipulation the optimality conditions yields as socially optimal growth

20The social planner has the power to dictate how much each firm produces and how muchinputs to allocate to that firm. Since production has no intertemporal links it is obviousthat the planners maximization problem can solved in two steps: first the planner decides onaggregate variables c(t) and K(t) and then she decides how to allocate aggregate inputs Land K(t) between firms. The second stage of this problem is therefore

maxli(t),ki(t)≥0

Z 1

0F

µki(t), li(t)

·Z 1

0kj(t)dj

¸¶di

s.t.

Z 1

0ki(t) = K(t)Z 1

0li(t) = L(t)

i.e. given the aggregate amount of capital chosen the planner decides how to best allocate it.Let µ and λ denote the Lagrange multipliers on the two constraints.First order conditions with respect to li(t) imply that

F2 (ki(t), li(t)K(t))K(t) = λ

or, since F2 is homogeneous of degree zero

F2

µki(t)

li(t),K(t)

¶K(t) = λ

which indicates that the planner allocates inputs so that each firm has the same capital laborratio. Denote this common ratio by

φ =ki(t)

li(t)for all i ∈ [0, 1]

=K(t)

L

But then total output becomesZ 1

0F

µki(t), li(t)

·Z 1

0kj(t)dj

¸¶di =

Z 1

0ki(t)F (1,

K(t)

φ)di

= F (1,K(t)

φ)K(t)

F (K(t),K(t)L)

How much production the planner allocates to each firm hence does not matter; the onlyimportant thing is that she equalizes capital-labor ratios across firms. Once she does, theproduction possibilies for any given choice of K(t) are given by F (K(t),K(t)L).

Page 230: Macroeconomics Phd Lectures notes

224 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

rate for consumption

γSPc (t) =c(t)

c(t)=1

σ[F1(K(t),K(t)L) + F2(K(t),K(t)L)L− (δ + ρ)]

Note that, since F is homogeneous of degree one, the partial derivatives arehomogeneous of degree zero and hence

F1(K(t),K(t)L) + F2(K(t),K(t)L)L = F1(1,K(t)L

K(t)) + F2(1,

K(t)L

K(t))L

= F1(1, L) + F2(1, L)L

and hence the growth rate of consumption

c(t)

c(t)=1

σ[F1(1, L) + F2(1, L)L− (δ + ρ)]

is constant over time. By dividing the aggregate resource constraint by K(t) wefind that

Lc(t)

K(t)+K(t)

K(t)+ δ = F (1, L)

and hence along a balanced growth path γSPK = γSPk = γSPc . As before the tran-sition to the balanced growth path is immediate, which can be shown invokingthe transversality condition as before.Now let’s turn to the competitive equilibrium. From the household problem

we immediately obtain as Euler equation

γCEc (t) =c(t)

c(t)=1

σ[r(t)− (δ + ρ)]

The firm’s profit maximization condition implies

r(t) = F1(ki(t), li(t)K(t))

But since all firms are identical and hence choose the same allocations21 we havethat

ki(t) = k(t) =

Z 1

0

k(t)di = K(t)

li(t) = L

and hence

r(t) = F1(K(t),K(t)L) = F1(1, L)

21This is without loss of generality. As long as firms choose the same capital-labor ratio(which they have to in equilibrium), the scale of operation of any particular firm is irrelevant.

Page 231: Macroeconomics Phd Lectures notes

9.4. ENDOGENOUS GROWTH MODELS 225

Hence the growth rate of per capita consumption in the competitive equilibriumis given by

γCEc (t) =c(t)

c(t)=1

σ[F1(1, L)− (δ + ρ)]

and is constant over time, not only in the steady state. Doing the same ma-nipulation with resource constraint we see that along a balanced growth paththe growth rate of capital has to equal the growth rate of consumption, i.e.γCEK = γCEk = γCEc . Again, in order to obtain sustained endogenous growth wehave to assume that the technology is sufficiently productive, or

F1(1, L)− (δ + ρ) > 0

Using arguments similar to the ones above we can show that in this economytransition to the balanced growth path is immediate, i.e. there are no transitiondynamics.Comparing the growth rates of the competitive equilibrium with the socially

optimal growth rates we see that, since F2(1, L)L > 0 the competitive economygrows inefficiently slow, i.e. γCEc < γSPc . This is due to the fact that competi-tive firms do not internalize the productivity-enhancing effect of higher averagecapital and hence under-employ capital, compared to the social optimum. Putotherwise, the private returns to investment (saving) are too low, giving rise tounderinvestment and slow capital accumulation. Compared to the competitiveequilibrium the planner chooses lower period zero consumption and higher in-vestment, which generates a higher growth rate. Obviously welfare is higher inthe socially optimal allocation than under the competitive equilibrium alloca-tion (since the planner can always choose the competitive equilibrium allocation,but does not find it optimal in general to do so). In fact, under special func-tional form assumptions on F we could derive both competitive and sociallyoptimal allocations directly and compare welfare, showing that the lower initialconsumption level that the social planner dictates is more than offset by thesubsequently higher consumption growth.An obvious next question is what type of policies would be able to remove

the inefficiency of the competitive equilibrium? The answer is obvious once werealize the source of the inefficiency. Firms do not take into account the exter-nality of a higher aggregate capital stock, because at the equilibrium interestrate it is optimal to choose exactly as much capital input as they do in a com-petitive equilibrium. The private return to capital (i.e. the private marginalproduct of capital in equilibrium equals F1(1, L) whereas the social return equalsF1(1, L) + F2(1, L)L. One way for the firms to internalize the social returns intheir private decisions is to pay them a subsidy of F2(1, L)L for each unit ofcapital hired. The firm would then face an effective rental rate of capital of

r(t)− F2(1, L)Lper unit of capital hired and would hire more capital. Since all factor paymentsgo to private households, total capital income from a given firm is given by

Page 232: Macroeconomics Phd Lectures notes

226 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

[r(t) + F2(1, L)L] ki(t), i.e. given by the (now lower) return on capital plus thesubsidy. The higher return on capital will induce the household to consume lessand save more, providing the necessary funds for higher capital accumulation.These subsidies have to be financed, however. In order to reproduce the so-cial optimum as a competitive equilibrium with subsidies it is important not tointroduce further distortions of private decisions. A lump sum tax on the repre-sentative household in each period will do the trick, not however a consumptiontax (at least not in general) or a tax that taxes factor income at different rates.

The empirical predictions of the Romer model with respect to the conver-gence discussion are similar to the predictions of the basic AK-model and hencenot further discussed. An interesting property of the Romer model and a wholeclass of models following this model is the presence of scale effects. Realiz-ing that F1(1, L) = F1(

1L , 1) and F1(1, L) + F2(1, L)L = F (1, L) (by Euler’s

theorem) we find that

∂γCEc∂L

= − 1

σL2F11(1, L) > 0

∂γSPc∂L

=F2(1, L)

σ> 0

i.e. that the growth rate of a country should grow with its size (more precisely,with the size of its labor force). This result is basically due to the fact that thehigher the number of workers, the more workers benefit form the externality ofthe aggregate (average) capital stock. Note that this scale effect would vanish if,instead of the aggregate capital stock K the aggregate capital stock per workerKL would generate the externality. The prediction of the model that countrieswith a bigger labor force are predicted to grow faster has led some people todismiss this type of endogenous models as empirically relevant. Others havetried, with some, but not big success, to find evidence for a scale effect in thedata. The question seems unsettled for now, but I am sceptical whether thisprediction of the model(s) can be identified in the data.

Lucas (1988)

Whereas Romer (1986) stresses the externalities generated by a high economy-wide capital stock, Lucas (1988) focuses on the effect of externalities generatedby human capital. You will write a good thesis because you are around abunch of smart colleagues with high average human capital from which you canlearn. In other respects Lucas’ model is very similar in spirit to Romer (1986),unfortunately much harder to analyze. Hence we will only sketch the mainelements here.

The economy is populated by a continuum of identical, infinitely lived house-holds that are indexed by i ∈ [0, 1]. They value consumption according to stan-dard CRRA utility. There is a single consumption good in each period. Individ-uals are endowed with hi(0) = h0 units of human capital and ki(0) = k0 unitsof physical capital. In each period the households make the following decisions

Page 233: Macroeconomics Phd Lectures notes

9.4. ENDOGENOUS GROWTH MODELS 227

• what fraction of their time to spend in the production of the consumptiongood, 1 − si(t) and what fraction to spend on the accumulation of newhuman capital, si(t). A household that spends 1 − si(t) units of time inthe production of the consumption good and has a level of human capitalof hi(t) supplies (1 − si(t))hi(t) units of effective labor, and hence totallabor income is given by (1− si(t))hi(t)w(t)

• how much of the current labor income to consume and how much to savefor tomorrow

The budget constraint of the household is then given as

ci(t) + ai(t) = (r(t)− δ)ai(t) + (1− si(t))hi(t)w(t)

Human capital is assumed to accumulate according to the accumulation equa-tion

hi(t) = θhi(t)si(t)− δhi(t)

where θ > 0 is a productivity parameter for the human capital productionfunction. Note that this formulation implies that the time cost needed to acquirean extra 1% of human capital is constant, independent of the level of humancapital already acquired. Also note that for human capital to the engine ofsustained endogenous economic growth it is absolutely crucial that there are nodecreasing marginal products of h in the production of human capital; if therewere then eventually the growth in human capital would cease and the growthin the economy would stall.A household then maximizes utility by choosing consumption ci(t), time al-

location si(t) and asset levels ai(t) as well as human capital levels hi(t), subjectto the budget constraint, the human capital accumulation equation, a standardno-Ponzi scheme condition and nonnegativity constraints on consumption aswell as human capital, and the constraint si(t) ∈ [0, 1]. There is a single repre-sentative firm that hires labor L(t) and capital K(t) for rental rates r(t) andw(t) and produces output according to the technology

Y (t) = AK(t)αL(t)1−αH(t)β

where α ∈ (0, 1),β > 0. Note that the firm faces a production externality in thatthe average level of human capital in the economy, H(t) =

R 10hi(t)di enters the

production function positively. The firm acts competitive and treats the average(or aggregate) level of human capital as exogenously given. Hence the firm’sproblem is completely standard. Note, however, that because of the externalityin production (which is beyond the control of the firm and not internalizedby individual households, although higher average human capital means higherwages) this economy again will feature inefficiency of competitive equilibriumallocations; in particular it is to be expected that the competitive equilibriumfeatures underinvestment in human capital.

Page 234: Macroeconomics Phd Lectures notes

228 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

The market clearing conditions for the goods market, labor market andcapital market areZ 1

0

ci(t)di+ K(t) + δK(t) = AK(t)αL(t)1−αH(t)βZ 1

0

(1− si(t)hi(t)) di = L(t)Z 1

0

ai(t)di = K(t)

Rational expectations require that the average level of human capital that isexpected by firms and households coincides with the level that households infact choose, i.e. Z 1

0

hi(t)di = H(t)

The definition of equilibrium is then straightforward as is the definition of aPareto optimal allocation (if, since all agents are ex ante identical, we confineourselves to type-identical allocations, i.e. all individuals have the same welfareweights in the objective function of the social planner). The social plannersproblem that solves for Pareto optimal allocations is given as

max(c(t),s(t),H(t),K(t))t∈[0,∞)≥0

Z ∞0

e−ρtc(t)1−σ

1− σdt

s.t. c(t) + K(t) + δK(t) = AK(t)α((1− s(t)H(t))1−αH(t)β with K(0) = k0 givenH(t) = θH(t)s(t)− δH(t) with H(0) = h0 given

s(t) ∈ [0, 1]

This model is already so complex that we can’t do much more than simplydetermine growth rates of the competitive equilibrium and a Pareto optimum,compare them and discuss potential policies that may remove the inefficiencyof the competitive equilibrium. In this economy a balanced growth path is anallocation (competitive equilibrium or social planners) such that consumption,physical and human capital and output grow at constant rates (which need notequal each other) and the time spent in human capital accumulation is constantover time.Let’s start with the social planner’s problem. In this model we have two

state variables, namely K(t) and H(t), and two control variables, namely s(t)and c(t). Obviously we need two co-state variables and the whole dynamicalsystem becomes more messy. Let λ(t) be the co-state variable for K(t) and µ(t)the co-state variable for H(t). The Hamiltonian is µ

H(c(t), s(t),K(t),H(t),λ(t), µ(t), t)= e−ρt

c(t)1−σ

1− σ+ λ(t)

hAK(t)α ((1− s(t)H(t))1−αH(t)β − δK(t)− c(t)

i+µ(t) [θH(t)s(t)− δH(t)]

Page 235: Macroeconomics Phd Lectures notes

9.4. ENDOGENOUS GROWTH MODELS 229

The first order conditions are

e−ρtc(t)−σ = λ(t) (9.47)

µ(t)θH(t) = λ(t)(1− α)

"AK(t)α ((1− s(t)H(t))1−αH(t)β

(1− s(t))

#(9.48)

The co-state equations are

λ(t) = −λ(t)α"AK(t)α ((1− s(t)H(t))1−αH(t)β

K(t)− δ

#(9.49)

µ(t) = −λ(t)(1− α+ β)

"AK(t)α ((1− s(t)H(t))1−αH(t)β

H(t)

#− µ(t) [θs(t)− δ]

(9.50)

Define Y (t) = AK(t)α ((1− s(t)H(t))1−αH(t)β . Along a balanced growth pathwe have

Y (t)

Y (t)= γY (t) = γY

c(t)

c(t)= γc(t) = γc

K(t)

K(t)= γK(t) = γK

H(t)

H(t)= γH(t) = γH

λ(t)

λ(t)= γλ(t) = γλ

µ(t)

µ(t)= γµ(t) = γµ

s(t) = s

Let’s focus on BGP’s. From the definition of Y (t) we have (by log-differentiating)

γY = aγK + (1− α+ β)γH (9.51)

From the human capital accumulation equation we have

γH = θs− δ (9.52)

From the Euler equation we have

γc =1

σ

·αY (t)

K(t)− (δ + ρ)

¸(9.53)

Page 236: Macroeconomics Phd Lectures notes

230 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

and hence

γY = γK (9.54)

From the resource constraint it then follows that

γc = γY = γK (9.55)

and therefore

γK =1− α+ β

1− αγH (9.56)

From the first order conditions we have

γλ = −ρ− σγc (9.57)

γµ = γλ + γY − γH (9.58)

Divide (9.47) by µ(t) and isolate λ(t)µ(t) to obtain

λ(t)

µ(t)=

θH(t)(1− s(t))(1− α)Y (t)

Do the same with (9.50) to obtain

λ(t)

µ(t)= − ¡γµ + γH

¢ H(t)

(1− α+ β)Y (t)

Equating the last two equations yields

−¡γµ + γH

¢(1− α+ β)

=θ(1− s)(1− α)

Using (9.58) and (9.55) and (9.52) and (9.56) we finally arrive at

γc =1

σ

·(θ − δ)(1− α+ β)

1− α− ρ

¸The other growth rates and the time spent with the accumulation of humancapital can then be easily deduced form the above equations. Be aware of thealgebra.In general, due to the externality the competitive equilibrium will not be

Pareto optimal; in particular, agents may underinvest into human capital. Fromthe firms problem we obtain the standard conditions (from now on we leave outthe i index for households

r(t) = αY (t)

K(t)

w(t) = (1− α)Y (t)

L(t)= (1− α)

Y (t)

(1− s(t))h(t)

Page 237: Macroeconomics Phd Lectures notes

9.4. ENDOGENOUS GROWTH MODELS 231

Form the Lagrangian for the representative household with state variablesa(t), h(t) and control variables s(t), c(t)

H = e−ρtc(t)1−σ

1− σ+ λ(t) [(r(t)− δ) a(t) + (1− s(t))h(t)w(t)− c(t)]

+µ(t) [θh(t)s(t)− δh(t)]

The first order conditions are

e−ρtc(t)−σ = λ(t) (9.59)

λ(t)h(t)w(t) = µ(t)θh(t) (9.60)

and the derivatives of the co-state variables are given by

λ(t) = −λ(t)(r(t)− δ) (9.61)

µ(t) = −λ(t)(1− s(t))w(t)− µ(t)(θs(t)− δ) (9.62)

Imposing balanced growth path conditions gives

γc =1

σ(−γλ − ρ)

γλ = γµ − γw = γµ − γY + γhγc = γY = γK

γh =1− α

1− α+ βγY

Hence

γλ = γµ −µ

β

1− α+ β

¶γc

Using (9.60) and (9.62) we find

γµ = δ − θ

and hence

γc =1

σ(θ −

µβ

1− α+ β

¶γc − (ρ+ ρ))

γCEc =1

σ + β1−α+β

(θ − (ρ+ ρ))

Compare this to the growth rate a social planner would choose

γSPc =1

σ

·(θ − δ)(1− α+ β)

1− α− ρ

¸We note that if β = 0 (no externality), then both growth rates are identical ((asthey should since then the welfare theorems apply). If, however β > 0 and the

Page 238: Macroeconomics Phd Lectures notes

232 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

externality from human capital is present, then if both growth rates are positive,tedious algebra can show that γCEc < γSPc . The competitive economy growsslower than optimal since the private returns to human capital accumulation arelower than the social returns (agents don’t take the externality into account) andhence accumulate to little human capital, lowering the growth rate of humancapital.

9.4.3 Models of Technological Progress Based on Monop-olistic Competition: Variant of Romer (1990)

In this section we will present a model in which technological progress, and henceeconomic growth, is the result of a conscious effort of profit maximizing agentsto invent new ideas and sell them to other producers, in order to recover theircosts for invention.22 We envision a world in which competitive software firmshire factor inputs to produce new software, which is then sold to intermediategoods producers who use it in the production of a new intermediate good, whichin turn is needed for the production of a final good which is sold to consumers.In this sense the Romer model (and its followers, in particular Jones (1995))are sometimes referred to as endogenous growth models, whereas the previousgrowth models are sometimes called only semi-endogenous growth models.

Setup of the Model

Production in the economy is composed of three sectors. There is a final goodsproducing sector in which all firms behave perfectly competitive. These firmshave the following production technology

Y (t) = L(t)1−αÃZ A(t)

0

xi(t)1−µdi

! α1−µ

where Y (t) is output, L(t) is labor input of the final goods sector and xi(t) isthe input of intermediate good i in the production of final goods. 1

µ is elasticity

of substitution between two inputs (i.e. measures the slope of isoquants), withµ = 0 being the special case in which intermediate inputs are perfect substi-tutes. For µ → ∞ we approach the Leontieff technology. Evidently this is aconstant returns to scale technology, and hence, without loss of generality wecan normalize the number of final goods producers to 1.At time t there is a continuum of differentiated intermediate goods indexed

by i ∈ [0, A(t)], where A(t) will evolve endogenously as described below. LetA0 > 0 be the initial level of technology. Technological progress in this modeltakes the form of an increase in the variety of intermediate goods. For 0 < µ < 1this will expand the production possibility frontier (see below). We will assumethis restriction on µ to hold.

22I changed and simplified the model a bit, in order to obtain analytic solutions and makeresults coparable to previous sections. The model is basically a continuous time version of themodel described in Jones and Manuelli (1998), section 6.

Page 239: Macroeconomics Phd Lectures notes

9.4. ENDOGENOUS GROWTH MODELS 233

Each differentiated product is produced by a single, monopolistically com-petitive firm. This firm has bought the patent for producing good i and is theonly firm that is entitled to produce good i. The fact, however, that the in-termediate goods are substitutes in production limits the market power of thisfirm. Each intermediate goods firm has the following constant returns to scaleproduction function to produce the intermediate good

xi(t) = ali(t)

where li(t) is the labor input of intermediate goods producer i at date t anda > 0 is a technology parameter, common across firms, that measures laborproductivity in the intermediate goods sector. We assume that the intermediategoods producers act competitively in the labor marketFinally there is a sector producing new “ideas”, patents to new intermediate

products. The technology for this sector is described by

A(t) = bX(t)

Note that this technology faces constant returns to scale in the production ofnew ideas in that X(t) is the only input in the production of new ideas. Theparameter b measures the productivity of the production of new ideas: if theideas producers buy X(t) units of the final good for their production of newideas, they generate bX(t) new ideas.

Planner’s Problem

Before we go ahead and more fully describe the equilibrium concept for thiseconomy we first want to solve for Pareto-optimal allocations. As usual wespecify consumer preferences as

u(c) =

Z ∞0

e−ρtc(t)1−σ

1− σdt

The social planner then solves23

maxc(t),li(t),xi(t),A(t),L(t),X(t)≥0

Z ∞0

e−ρtc(t)1−σ

1− σdt

s.t. c(t) +X(t) = L(t)1−αÃZ A(t)

0

xi(t)1−µdi

! α1−µ

L(t) +

Z A(t)

0

li(t)di = 1

xi(t) = ali(t) for all i ∈ [0, A(t)]A(t) = bX(t)

23Note that there is no physical capital in this model. Romer (1990) assumes that inter-mediate goods producers produce a durable intermediate good that they then rent out everyperiod. This makes the intermediate goods capital goods, which slightly complicates theanalysis of the model. See the original article for further details.

Page 240: Macroeconomics Phd Lectures notes

234 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

This problem can be simplified substantially. Since µ ∈ (0, 1) it is obviousthat xi(t) = xj(t) = x(t) for all i, j ∈ [0, A(t)] and li(t) = lj(t) = l(t) forall i, j ∈ [0, A(t)].24. Also use the fact that L(t) = 1 − A(t)l(t) to obtain theconstraint set

c(t) +X(t) = L(t)1−αÃZ A(t)

0

xi(t)1−µdi

! α1−µ

= L(t)1−αÃ(al(t))

1−µZ A(t)

0

di

! α1−µ

= L(t)1−αÃA(t)

µa1− L(t)A(t)

¶1−µ! α1−µ

= aαL(t)1−α (1− L(t))αA(t) αµ1−µ

A(t) = bX(t)

Finally we note that the optimal allocation of labor solves the static problem of

maxL(t)∈[0,1]

L(t)1−α (1− L(t))α

with solution L(t) = 1− α. So finally we can write the social planners problemas

u(c) =

Z ∞0

e−ρtc(t)1−σ

1− σdt

s.t. c(t) +A(t)

b= CA(t)η (9.63)

where C = aa(1− α)1−ααα and η = αµ1−µ > 0 and with A(0) = A0 given. Note

that if 0 < µ < 1, this model boils down to the standard Cass-Koopmans model,whereas if η = 1 we obtain the basic AK-model. Finally, if η > 1 the model

24Suppose there are only two intermediate goods and one wants to

maxl1(t),l2(t)≥0

Ã2Xi=1

ali(t)1−µ

! α1−µ

s.t. l1(t) + l2(t) = L

For µ ∈ (0, 1) the isoquant Ã2Xi=1

ali(t)1−µ

! α1−µ

= C > 0

is strictly convex, with slope strictly bigger than one in absolute value. Given the above con-straint, the maximum is interior and the first order conditions imply l1(t) = l2(t) immediately.The same logic applies to the integral, where, strictly speaking, we have to add an “almosteverywhere” (since sets of Lebesgue measure zero leave the integral unchanged). Note thatfor µ ≤ 0 the above argument doesn’t work as we have corner solutions.

Page 241: Macroeconomics Phd Lectures notes

9.4. ENDOGENOUS GROWTH MODELS 235

will exhibit accelerating growth. Forming the Hamiltonian and manipulatingthe first order conditions yields

γc(t) =1

σ

£bηCA(t)η−1 − ρ

¤Hence along a balanced growth path A(t)η−1 has to remain constant over time.From the ideas accumulation equation we find

A(t)

A(t)=bX(t)

A(t)

which implies that along a balanced growth path X and A grow at the samerate. Dividing () by A(t) yields

c(t)

A(t)+A(t)

bA(t)= CA(t)η−1

which implies that c grows at the same rate as A and X.

We see that for η < 1 the economy behaves like the neoclassical growthmodel: from A(0) = A0 the level of technology converges to the steady state A

satisfying

bηC

(A∗)1−η= ρ

X∗ = 0

c∗ = C (A∗)η

Without exogenous technological progress sustained economic growth in percapita income and consumption is infeasible; the economy is saddle path stableas the Cass-Koopmans model.

If η = 1, then the balanced growth path growth rate is

γc(t) =1

σ[bηC − ρ] > 0

provided that the technology producing new ideas, manifested in the parameterb, is productive enough to sustain positive growth. Now the model behaves asthe AK-model, with constant positive growth possible and immediate conver-gence to the balanced growth path. Note that a condition equivalent to (9.45)is needed to ensure convergence of the utility generated by the consumptionstream. Finally, for η > 1 (and A0 > 1) we can show that the growth rate ofconsumption (and income) increases over time. Remember again that η = αµ

1−µ ,which, a priori, does not indicate the size of η. What empirical predictions themodel has therefore crucially depends on the magnitudes of the capital share αand the intratemporal elasticity of substitution between inputs, µ.

Page 242: Macroeconomics Phd Lectures notes

236 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

Decentralization

We have in mind the following market structure. There is a single representativefinal goods producing firm that faces the constant returns to scale productiontechnology as discussed above. The firm sells final output at time t for pricep(t) and hires labor L(t) for a (nominal) wage w(t). It also buys intermedi-ate goods of all varieties for prices pi(t) per unit. The final goods firm actscompetitively in all markets. The final goods producer makes zero profits inequilibrium (remember CRTS). The representative producer of new ideas ineach period buys final goods X(t) as inputs for price p(t) and sells a new ideato a new intermediate goods producer for price κ(t). The idea producer behavescompetitively and makes zero profits in equilibrium (remember CRTS). Thereis free entry in the intermediate goods producing sector. Each new intermediategoods producer has to pay the fixed cost κ(t) for the idea and will earn subse-quent profits π(τ), τ ≥ t since he is a monopolistic competition, by hiring laborli(t) for wage w(t) and selling output xi(t) for price pi(t). Each intermediate pro-ducer takes as given the entire demand schedule of the final producer xdi (

−→p (t)),where −→p = (p,w, (pi)i∈[0,A(t)]. We denote by −→p −1 all prices but the price ofintermediate good i. Free entry drives net profits to zero, i.e. equates κ(t) andthe (appropriately discounted) stream of future profits. Now let’s define a mar-ket equilibrium (note that we can’t call it a competitive equilibrium anymorebecause the intermediate goods producers are monopolistic competitors).

Definition 100 A market equilibrium is prices (p(t), κ(t), pi(t)i∈[0,A(t), w(t))t∈[0,∞),allocations for the household c(t)t∈[0,∞), demands for the final goods producer(L(−→p (t)), xdi (−→p (t))i∈[0,A(t)])t∈[0,∞), allocations for the intermediate goods pro-ducers ((xsi (t), li(t))i∈[0,A(t))t=[0,∞) and allocations for the idea producer (A(t), X(t))t=[0,∞)such that

1. Given κ(0), (p(t), w(t))t∈[0,∞), c(t)t∈[0,∞) solves

maxc(t)≥0

Z ∞0

e−ρtc(t)1−σ

1− σdt

s.t.

Z ∞0

p(t)c(t)dt =

Z ∞0

w(t)dt+ κ(0)A0

2. For each i, t, given−→p −i(t), w(t), and xdi (−→p (t), (xsi (t), li(t), pi(t)) solves

πi(t) = maxxi(t),li(t),pi(t)≥0

pi(t)xdi (−→p (t))− w(t)li(t)

s.t. xi(t) = xdi (−→p (t))

xi(t) = ali(t)

3. For each t, and each −→p ≥ 0, (L(−→p (t)), xdi (−→p (t)) solves

maxL(t),xi(t)≥0

p(t)L(t)1−αÃZ A(t)

0

xi(t)1−µdi

! α1−µ

− w(t)L(t)−Z A(t)

0

pi(t)xi(t)di

Page 243: Macroeconomics Phd Lectures notes

9.4. ENDOGENOUS GROWTH MODELS 237

4. Given (p(t), c(t))t∈[0,∞, (A(t), X(t))t=[0,∞) solves

max

Z ∞0

c(t)A(t)−Z ∞0

p(t)X(t)dt

s.t. A(t) = bX(t) with A(0) = A0 given

5. For all t

L(−→p (t))1−α

ÃZ A(t)

0

xdi (−→p (t))1−µdi

! α1−µ

= X(t) + c(t)

xsi (t) = xdi (−→p (t)) for all i ∈ [0, A(t)]

L(t) +

Z A(t)

0

li(t)di = 1

6. For all t, all i ∈ A(t)

κ(t) =

Z ∞t

πi(τ)dτ

Several remarks are in order. First, note that in this model there is no phys-ical capital. Hence the household only receives income from labor and fromselling initial ideas (of course we could make the idea producers own the ini-tial ideas and transfer the profits from selling them to the household). Thekey equilibrium condition involves the intermediate goods producers. They, byassumption, are monopolistic competitors and hence can set prices, taking asgiven the entire demand schedule of the final goods producer. Since the inter-mediate goods are substitutes in production, the demand for intermediate goodi depends on all intermediate goods prices. Note that the intermediate goodsproducer can only set quantity or price, the other is dictated by the demand ofthe final goods producer. The required labor input follows from the productiontechnology. Since we require the entire demand schedule for the intermediategoods producers we require the final goods producer to solve its maximizationproblem for all conceivable (positive) prices. The profit maximization require-ment for the ideas producer is standard (remember that he behave perfectlycompetitive by assumption). The equilibrium conditions for final goods, inter-mediate goods and labor market are straightforward. The final condition isthe zero profit condition for new entrants into intermediate goods production,stating that the price of the pattern must equal to future profits.It is in general very hard to solve for an equilibrium explicitly in these type

of models. However, parts of the equilibrium can be characterized quite sharply;in particular optimal pricing policies of the intermediate goods producers. Sincethe differentiated product model is widely used, not only in growth, but alsoin monetary economics and particularly in trade, we want to analyze it morecarefully.

Page 244: Macroeconomics Phd Lectures notes

238 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

Let’s start with the final goods producer. First order conditions with respectto L(t) and xi(t) entail

25

w(t) = (1− α)p(t)L(t)−αÃZ A(t)

0

xi(t)1−µdi

! α1−µ

=(1− α)p(t)Y (t)

L(t)(9.64)

pi(t) = αp(t)L(t)1−αÃZ A(t)

0

xi(t)1−µdi

! α1−µ−1

xi(t)−µ (9.65)

or

xi(t)µpi(t) = αp(t)L(t)1−α

ÃZ A(t)

0

xi(t)1−µdi

! α1−µ−1

for all i ∈ [0, A(t)]

=αp(t)Y (t)R A(t)

0xi(t)1−µdi

Hence the demand for input xi(t) is given by

xi(t) =

µp(t)

pi(t)

¶ 1µ

ÃαY (t)R A(t)

0xi(t)1−µdi

! 1µ

(9.66)

=

µp(t)

pi(t)

¶ 1µ

αY (t)µ+α−1αµ L(t)

(1−µ)(1−α)αµ (9.67)

As it should be, demand for intermediate input i is decreasing in its relative

price p(t)pi(t)

. Now we proceed to the profit maximization problem of the typical

intermediate goods firm. Taking as given the demand schedule derived above,the firm solves (using the fact that xi(t) = ali(t)

maxpi(t)

pi(t)xi(t)− w(t)xi(t)a

= xi(t)

µpi(t)− w(t)

a

¶The first order condition reads (note that pi(t) enters xi(t) as shown in (9.67)

xi(t)− 1

µpi(t)xi(t)

µpi(t)− w(t)

a

¶= 0

and hence

1 =1

µ− w(t)

µapi(t)

pi(t) =w(t)

a(1− µ) (9.68)

25Strictly speaking we should worry about corners. However, by assumption µ ∈ (0, 1) willassure that for equilibrium prices corners don’t occur

Page 245: Macroeconomics Phd Lectures notes

9.4. ENDOGENOUS GROWTH MODELS 239

A perfectly competitive firm would have price pi(t) equal marginal costw(t)a .

The pricing rule of the monopolistic competitor is very simple, he charges aconstant markup 1

1−µ > 1 over marginal cost. Note that the markup is thelower the lower µ. For the special case in which the intermediate goods areperfect substitutes in production, µ = 0 and there is no markup over marginalcost. Perfect substitutability of inputs forces the monopolistic competitor tobehave as under perfect competition. On the other hand, the closer µ gets to1 (in which case the inputs are complements), the higher the markup the firmscan charge. Note that this pricing policy is valid not only in a balanced growthpath. indicating that

Another important implication is that all firms charge the same price, andtherefore have the same scale of production. So let x(t) denote this common

output of firms and p(t) = w(t)a(1−µ) the common price of intermediate producers.

Profits of every monopolistic competitor are given by

π(t) = p(t)x(t)− w(t)x(t)a

= µx(t)p(t)

=µαp(t)Y (t)

A(t)(9.69)

We see that in the case of perfect substitutes profits are zero, whereas profitsincrease with declining degree of substitutability between intermediate goods.26

Using the above results in equations (9.64) and (9.66) yields

w(t)L(t) = (1− α)p(t)Y (t) (9.70)

A(t)x(t)p(t) = αp(t)Y (t) (9.71)

We see that for the final goods producer factor payments to labor, w(t)L(t) andto intermediate goods, A(t)x(t)p(t), exhaust the value of production p(t)Y (t)so that profits are zero as they should be for a perfectly competitive firm withconstant returns to scale. From the labor market equilibrium condition we find

L(t) = 1− A(t)x(t)a

(9.72)

and output is given from the production function as

Y (t) = L(t)1−αx(t)αA(t)α

1−µ (9.73)

and is used for consumption and investment into new ideas

Y (t) = c(t) +X(t) (9.74)

26This is not a precise argument. One has to consider the general equilibrium effects ofchanges in µ on p(t), Y (t), A(t) which is, in fact, quite tricky.

Page 246: Macroeconomics Phd Lectures notes

240 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

We assumed that the ideas producer is perfectly competitive. Then it followsimmediately, given the technology

A(t) = bX(t)

A(t) = A(0) +

Z t

0

X(τ)dτ (9.75)

that

κ(t) =p(t)

b(9.76)

The zero profit-free entry condition then reads (using (9.69))

p(t)

κ= µα

Z ∞t

p(τ)Y (τ)

A(τ)dτ (9.77)

Finally, let us look at the household maximization problem. Note that, inthe absence of physical capital or any other long-lived asset household problemdoes not have any state variable. Hence the household problem is a standardmaximization problem, subject to a single budget constraint. Let λ be theLagrange multiplier associated with this constraint. The first order conditionreads

e−ρtc(t)−σ = λp(t)

Differentiating this condition with respect to time yields

−σe−ρtc(t)−σ−1c(t)− ρe−ρtc(t)−σ = λp(t)

and hence

c(t)

c(t)=1

σ

µ− p(t)p(t)− ρ

¶(9.78)

i.e. the growth rate of consumption equals the rate of deflation minus the timediscount rate. In summary, the entire market equilibrium is characterized by the10 equations (9.68) and (9.70) to (9.78) in the 10 variables x(t), c(t),X(t), Y (t),L(t), A(t),κ(t), p(t), w(t), p(t), with initial condition A(0) = A0. Since it is, inprinciple, extremely hard to solve this entire system we restrict ourselves to afew more interesting results.First we want to solve for the fraction of labor devoted to the production of

final goods, L(t). Remember that the social planner allocated a fraction 1−α ofall labor to this sector. From (9.72) we have that L(t) = 1− A(t)x(t)

a . Dividing(9.71) by (9.70) yields

α

1− α=

A(t)x(t)p(t)

w(t)L(t)=

A(t)x(t)

aL(t)(1− µ)A(t)x(t)

a=

α(1− µ)L(t)1− α

Page 247: Macroeconomics Phd Lectures notes

9.4. ENDOGENOUS GROWTH MODELS 241

and hence

L(t) = 1− A(t)x(t)a

= 1− α(1− µ)L(t)1− α

L(t) =1− α

1− αµ> 1− α

Hence in the market equilibrium more workers work in the final goods sectorand less in the intermediate goods sector than socially optimal. The intuitionfor this is simple: since the intermediate goods sector is monopolistically com-petitive, prices are higher than optimal (than social shadow prices) and outputis lower than optimal; differently put, final goods producers substitute awayfrom expensive intermediate goods into labor. Obviously labor input in theintermediate goods sector is lower than in the social optimum and hence

AME(t)xME(t) < ASP (t)xSP (t)

Again these relationships hold always, not just in the balanced growth path.

Now let’s focus on a balanced growth path where all variables grow at con-stant, possibly different rate. Obviously, since L(t) = 1−α

1−αµ we have that gL = 0.From the labor market equilibrium gA = −gx. From constant markup pricing wehave gw = gp. From (9.75) we have gA = gX and from the resource constraint(9.74) we have gA = gX = gc = gY . Then from (9.70) and (9.71) we have that

gw = gY + gP

gp = gY + gP

From the production function we find that

gY = αgx +α

1− µgA

=αµ

1− µgA

Hence a balanced growth path exists if and only if gY = 0 or η = αµ1−µ = 1.

The first case corresponds to the standard Solow or Cass-Koopmans model: ifη < 1 the model behaves as the neoclassical growth model with asymptoticconvergence to the no-growth steady state (unless there is exogenous techno-logical progress). The case η = 1 delivers (as in the social planners problem)a balanced growth path with sustained positive growth, whereas η > 1 yieldsexplosive growth (for the appropriate initial conditions).

Let’s assume η = 1 for the moment. Then gY = gA and hence Y (t)A(t) =

Y (0)A0

=constant. The no entry-zero profit condition in the BGP can be written

Page 248: Macroeconomics Phd Lectures notes

242 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

as, since p(τ) = p(t)egp(τ−t) for all τ ≥ tp(t)

b= µα

Y (0)

A0

Z ∞t

p(t)gp(τ−t)dτ

1 = −bµαY (0)A0gp

gp =p(t)

p(t)= −

µbµα

Y (0)

A0

¶< 0

Finally, from the consumption Euler equation

gc =1

σ

µbµα

Y (0)

A0− ρ

¶But now note that

Y (0) = L(0)1−αx(0)αA(0)α

1−µ

= L(0)1−α (x(0)A(0))αA(0)αµ1−µ

= L(0)1−α (x(0)A(0))αA(0)

under the assumption that η = 1. Hence, using (9.72)

Y (0)

A0= L(0)1−α (x(0)A(0))α

= L(0)1−α (a(1− L(0))α= L(0)aα

=1− α

1− αµaa

Therefore finally

gc = gY = gA =1

σ

µbaaµα

1− α

1− αµ− ρ

¶is the competitive equilibrium growth rate in the balanced growth path underthe assumption that η = 1. Comparing this to the growth rate that the socialplanner would choose

γc(t) =1

σ

£baa(1− α)1−ααα − ρ

¤We see that for µα ≤ 1 the social planner would choose a higher balancedgrowth path growth rate than the market equilibrium BGP growth rate. Themarket power of the intermediate goods producers leads to lower production ofintermediate goods and hence less resources for consumption and new inventions,which drive growth in this model.27

27Note however that there is an effect of market power in the opposite direction. Since in themarket equilibrium the intermediate goods producers make profits due to their (competitive)monopoly position, and the ideas inventors can extract these profits by selling new designs,due to the free entry condition, they have too big an incentive to invent new intermediategoods, relative to the social optimum. For big µ and big α this may, in fact, lead to aninefficiently high growth rate in the market equilibrium.

Page 249: Macroeconomics Phd Lectures notes

9.4. ENDOGENOUS GROWTH MODELS 243

This completes our discussion of endogenous growth theory. The Romer-typemodel discussed last can, appropriately interpreted, nest the standard Solow-Cass-Koopmans type neoclassical growth models as well as the early AK-typegrowth models. In addition it achieves to make the growth rate of the economytruly endogenous: the economy grows because inventors of new ideas consciouslyexpend resources to develop new ideas and sell them to intermediate producersthat use them in the production of a new product.

Page 250: Macroeconomics Phd Lectures notes

244 CHAPTER 9. CONTINUOUS TIME GROWTH THEORY

Page 251: Macroeconomics Phd Lectures notes

Chapter 10

Bewley Models

In this section we will look at a class of models that take a first step at explain-ing the distribution of wealth in actual economies. So far our models abstractedfrom distributional aspects. As standard in macro up until the early 90’s ourmodels had representative agents, that all faced the same preferences, endow-ments and choices, and hence received the same allocations. Obviously, in suchenvironments one cannot talk meaningfully about the income distribution, thewealth distribution or the consumption distribution. One exception was theOLG model, where, at a given point of time we had agents that differed by age,and hence differed in their consumption and savings decisions. However, withonly two (groups of) agents the cross sectional distribution of consumption andwealth looks rather sparse, containing only two points at any time period.

We want to accomplish two things in this section. First, we want to summa-rize the main empirical facts about the current U.S. income and wealth distri-bution. Second we want to build a class of models which are both tractable andwhose equilibria feature a nontrivial distribution of wealth across agents. Thebasic idea is the following. The is a continuum of agents that are ex ante iden-tical and all have a stochastic endowment process that follow a Markov chain.Then endowments are realized in each period, and it so happens that someagents are lucky and get good endowment realizations, others are unlucky andget bad endowment realizations. The aggregate endowment is constant acrosstime. If there was a complete set of Arrow-Debreu contingent claims, then peo-ple would simple insure each other against the endowment shocks and we wouldbe back at the standard representative agent model. We will assume that peo-ple cannot insure against these shocks (for reasons exogenous from the model),in that we close down all insurance markets. The only financial instrumentthat agents, by assumption, can use to hedge against endowment uncertaintyare one period bonds (or IOU’s) that yield a riskless return r. In other words,agents can only self -insure by borrowing and lending at a risk free rate r. Inaddition, we impose tight limits on how much people can borrow (otherwise,it turns out, self-insurance (almost) as good as insuring with Arrow-Debreuclaims). As a result, agents will accumulate wealth, in the form of bonds, to

245

Page 252: Macroeconomics Phd Lectures notes

246 CHAPTER 10. BEWLEY MODELS

hedge against endowment uncertainty. Those agents with a sequence of goodendowment shocks will have a lot of wealth, those with a sequence of bad shockswill have low wealth (or even debt). Hence the model will use as input an ex-ogenously specified stochastic endowment (income) process, and will deliver asoutput an endogenously derived wealth distribution.To analyze these models we will need to keep track of the characteristics of

each agent at a given point of time, which, in most cases, is at least the currentendowment realization and the current wealth position. Since these differ acrossagents, we need an entire distribution (measure) to keep track of the state of theeconomy. Hence the richness of the model with respect to distributional aspectscomes at a cost: we need to deal with entire distributions as state variables,instead of just numbers as the capital stock. Therefore the preparation withrespect to measure theory in recitation

10.1 Some Stylized Facts about the Income andWealth Distribution in the U.S.

In this section we describe the main stylized facts characterizing the U.S. incomeand wealth distribution.1 For data on the income and wealth distribution wehave to look beyond the national income and product accounts (NIPA) data,since NIPA only contains aggregated data. What we need are data on incomeand wealth of a sample of individual families.

10.1.1 Data Sources

For the U.S. there are three main data sets

• the Survey of Consumer Finances (SCF). The SCF is conducted in threeyear intervals; the four available surveys are for the years 1989, 1992,1995 and 1998. It is conducted by the National Opinion Research cen-ter at the University of Chicago and sponsored by the Federal Reservesystem. It contains rich information about U.S. households’ income andwealth. In each survey about 4,000 households are asked detailed ques-tions about their labor earnings, income and wealth. One part of thesample is representative of the U.S. population, to give an accurate de-scription of the entire population. The second part oversamples rich house-holds, to get a more precise idea about the precise composition of thisgroups’ income and wealth composition. As we will see, this group ac-counts for the majority of total household wealth, and hence it is partic-ularly important to have good information about this group. The mainadvantage of the SCF is the level of detail of information about incomeand 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

1This section summarizes the basic findings of Diaz-Gimenez, Quadrini and Rios-Rull(1997).

Page 253: Macroeconomics Phd Lectures notes

10.1. SOME STYLIZED FACTS ABOUTTHE INCOMEANDWEALTHDISTRIBUTION IN THEU.S.247

and wealth accumulation cannot be documented on the household levelwith this data set. For further information and some of the data seehttp://www.federalreserve.gov/pubs/oss/oss2/98/scf98home.html

• the Panel Study of Income Dynamics (PSID). It is conducted by the Sur-vey Research Center of the University of Michigan and mainly sponsoredby the National Science Foundation. The PSID is a panel data set thatstarted with a national sample of 5,000 U.S. households in 1968. Thesame sample individuals are followed over the years, barring attrition dueto death or nonresponse. New households are added to the sample ona consistent basis, making the total sample size of the PSID about 8700households. The income and wealth data are not as detailed as for theSCF, but its panel dimension allows to construct measures of income andwealth dynamics, since the same households are interviewed year afteryear. Also the PSID contains data on consumption expenditures, al-beit only food consumption. In addition, in 1990, a representative na-tional sample of 2,000 Latinos, differentially sampled to provide adequatenumbers of Puerto Rican, Mexican-American, and Cuban-Americans, wasadded to the PSID database. This provides a host of information for stud-ies on discrimination. For further information and the complete data setsee http://www.isr.umich.edu/src/psid/index.html

• the Consumer Expenditure Survey (CEX) or (CES). The CEX is con-ducted by the U.S. Bureau of the Census and sponsored by the Bureauof Labor statistics. The first year the survey was carried out was 1980.The CEX is a so-called rotating panel: each household in the sample isinterviewed for four consecutive quarters and then rotated out of the sur-vey. Hence in each quarter 20% of all households is rotated out of thesample and replaced by new households. In each quarter about 3000 to5000 households are in the sample, and the sample is representative ofthe U.S. population. The main advantage of the CEX is that it containsvery detailed information about consumption expenditures. Informationabout income and wealth is inferior to the SCF and PSID, also the paneldimension is significantly shorter than for the PSID (one household is onlyfollowed for 4 quarters). Given our focus on income and wealth we willnot use the CEX here, but anyone writing a paper about consumption willfind the CEX an extremely useful data set. For further information andthe complete data set see http://www.stats.bls.gov/csxhome.htm.

10.1.2 Main Stylized Facts

We will look at facts for three variables, earnings, income and wealth. Let’s firstdefine how we measure these variables in the data.

Definition 101 We define the following variables as

Page 254: Macroeconomics Phd Lectures notes

248 CHAPTER 10. BEWLEY MODELS

1. Earnings: Wages, Salaries of all kinds, plus a fraction 0.864 of businessincome (such as income from professional practices, business and farmsources)

2. Income: All kinds of household revenues before taxes, including: wagesand salaries, a fraction of business income (as above), interest income,dividends, gains or losses from the sale of stocks, bonds, and real es-tate, rent, trust income and royalties from any other investment or busi-ness, unemployment and worker compensation, child support and alimony,aid to dependent children, aid to families with dependent children, foodstamps and other forms of welfare and assistance, income form social se-curity and other pensions, annuities, compensation for disabilities andretirement programs, income from all other sources including settlements,prizes, scholarships and grants, inheritances, gifts and so forth.

3. Wealth: Net worth of households, defined as the value of all real and fi-nancial assets of all kinds net of all the kinds of debts. Assets consideredare: residences and other real estate, farms and other businesses, checkingaccounts, certificates of deposit, and other bank accounts, IRA/Keogh ac-counts, money market accounts, mutual funds, bonds and stocks, cash andcall money at the stock brokerage, all annuities, trusts and managed in-vestment accounts, vehicles, the cash value of term life insurance policies,money owed by friends, relatives and businesses, pension plans accumu-lated in accounts.

So, roughly, earnings correspond to labor income before taxes, income corre-sponds to household income before taxes and wealth corresponds to marketableassets. Now turn to some stylized facts about the distribution of these variablesacross U.S. households

Measures of Concentration

In this section we use data from the SCF. We measure the dispersion of theearnings, income and wealth distribution in a cross section of households byseveral measures. Let the sample of size n, assumed to be representative of thepopulation, be given by x1, x2, . . . xn, where x is the variable of interest (i.e.earnings, income or wealth). Define by

x =1

n

nXi=1

xi

std(x) =

vuut 1

n

nXi=1

(xi − x)2

Page 255: Macroeconomics Phd Lectures notes

10.1. SOME STYLIZED FACTS ABOUTTHE INCOMEANDWEALTHDISTRIBUTION IN THEU.S.249

the mean and the standard deviation. A commonly reported measure of disper-sion is the coefficient of variation cv(x)

cv(x) =std(x)

x

A second commonly used measure is the Gini coefficient and the associatedLorenz curve. To derive the Lorenz curve we do the following. We first orderx1, x2, . . . xn by size in ascending order, yielding y1, y2, . . . yn. The Lorenzcurve then plots i

n , i ∈ 1, 2 . . . n against zi =P i

j=1 yjPnj=1 yj

. In other words, it plots

the percentile of households against the fraction of total wealth (if x measureswealth) that this percentile of households holds. For example, if n = 100 theni = 5 corresponds to the 5 percentile of households. Note that, since the yiare ordered ascendingly, zi ≤ 1, zi+1 ≥ zi and that zn = 1. The closer theLorenz curve is to the 45 degree line, the more equal is x distributed. The Ginicoefficient is two times the area between the Lorenz curve and the 45 degreeline. If x ≥ 0, then the Gini coefficient falls between zero and 1, with higher Ginicoefficients indicating bigger concentration of earnings, income or wealth. Asextremes, for complete equality of x the Gini coefficient is zero and for completeconcentration (the richest person has all earnings, income, wealth) the Ginicoefficient is 1.2.Figure 26 and Table 4 summarize the main stylized facts with respect to the

concentration of earnings, income and wealth from the 1992 SCF

Table 4

Variable Mean Gini cv Top 1%Bottom 40% Loc. of Mean Mean

Median

Earnings $ 33, 074 0.63 4.19 211 65% 1.65Income $ 45, 924 0.57 3.86 84 71% 1.72Wealth $ 184, 308 0.78 6.09 875 80% 3.61

We observe the following stylized facts

• There is substantial variability in earnings, income and wealth across U.S.households. The standard deviation of earnings, for example, is about $140, 000. The top 1% of earners on average earn 21, 100% more than thebottom 40%, the corresponding number for income is still 8, 400%.

• Wealth is by far the most concentrated of the three variables, followedby earnings and income. That income is most equally distributed acrosshouseholds makes sense as income includes payments from governmentinsurance programs. The distribution would be even less dispersed if wewould look at income after taxes, due to the progressivity of the tax code.

2Strictly speaking it approaches 1, as n→∞ with complete concentration.

Page 256: Macroeconomics Phd Lectures notes

250 CHAPTER 10. BEWLEY MODELS

0 10 20 30 40 50 60 70 80 90 100-20

0

20

40

60

80

100Lorenz Curves for Earnings, Income and Wealth for the US in 1992

% of Households

% o

f Ear

ning

s, In

com

e, W

ealth

Hel

d

Earnings

Income

Wealth

Figure 10.1:

Page 257: Macroeconomics Phd Lectures notes

10.1. SOME STYLIZED FACTS ABOUTTHE INCOMEANDWEALTHDISTRIBUTION IN THEU.S.251

Since wealth is accumulated past income minus consumption, it also makesintuitive sense that it is most most concentrated. For example, the top1% households of the wealth distribution hold about 30% of total wealth.

• The distribution of all three variables is skewed. If the distributions weresymmetric, the median would equal the mean and the mean would belocated at the 50-percentile of the distribution. For all three variables themean is substantially higher than the median, which indicates skewnessto the right. In accordance with the last stylized fact, the distribution ofwealth is most skewed, followed by the distribution of earnings and thedistribution of income.

It is also instructive to look at the correlation between earnings, income andwealth. In Table 5 we compute the pairwise correlation coefficients betweenearnings, income and wealth. Remember that the correlation coefficient betweentwo variables x, y is given by

ρ(x, y) =cov(x, y)

std(x) ∗ std(y)

=1n

Pni=1(xi − x)(yi − y)q

1n

Pni=1(xi − x)2 ∗

q1n

Pni=1(yi − y)2

Table 5

Variables ρ(x, y)

Earnings and Income 0.928Earnings and Wealth 0.230Income and Wealth 0.321

We see that earnings and income as almost perfectly correlated, which isnatural since the largest fraction of household income consists of earnings. Thealmost perfect correlation indicates that transfer payments and capital income,at least on average, do not constitute a major component of household income.In fact, on average 72% of total income is accounted for by earnings in thesample. On the other hand, wealth is only weakly positively correlated withincome and earnings. Wealth is the consequence of past income, and only tothe extent that current and past income and earnings are positively correlatedshould wealth and earnings (income) by naturally positively correlated.3

Measures of Mobility

Not only is there a lot of variability in earnings, income and wealth acrosshouseholds, but also a lot of dynamics within the corresponding distribution.

3Wealth is measured as stock at the end of the period, so current income (earnings) con-tibute to wealth accumulation during the period).

Page 258: Macroeconomics Phd Lectures notes

252 CHAPTER 10. BEWLEY MODELS

Some poor households get rich, some rich households get poor over time. InTable 6 we report mobility matrices for earnings, income and wealth. Thetables are read as follows: a particular row indicates the probability of movingfrom a particular quintile in 1984 to a particular quintile in 1989. Note thatfor these matrices we used data from the PSID since the SCF does not have apanel dimension and hence does not contain information about households attwo different points of time, which is obviously necessary for studies of income,earnings and wealth mobility.

Table 6

1984 1989 Quintile

Measure Quintile 1st 2nd 3rd 4th 5th

Earnings 1st 85.5% 11.6% 1.4% 0.6% 0.5%2nd 16.8% 40.9% 30.0% 7.1% 3.4%3rd 7.1% 12.0% 47.0% 26.2% 7.6%4th 7.5% 6.8% 17.5% 46.5% 21.7%5th 5.8% 4.1% 5.5% 18.3% 66.3%

Income 1st 71.0% 17.9% 7.0% 2.9% 1.3%2nd 19.5% 43.8% 22.9% 10.1% 3.7%3rd 5.1% 25.5% 37.2% 24.9% 7.3%4th 2.5% 10.7% 23.4% 42.5% 20.8%5th 1.9% 2.1% 9.5% 20.3% 66.3%

Wealth 1st 66.7% 23.4% 6.6% 2.9% 0.4%2nd 25.4% 46.6% 20.4% 5.4% 2.3%3rd 5.8% 24.4% 44.9% 20.5% 4.6%4th 1.8% 4.6% 22.4% 49.6% 21.6%5th 0.7% 0.8% 5.7% 21.6% 71.2%

In Table 7 we condition the sample on two factors. The first matrix computestransition probabilities of earnings for people with positive earnings in both 1984and 1989, i.e. filters out households all of which members are either retiredor unemployed in either of the years. This is done to get a clearer look atearnings mobility of those actually working. The second matrix shows transitionprobabilities for households with heads of so-called prime age, age between 35-45.

Table 7

Page 259: Macroeconomics Phd Lectures notes

10.2. THE CLASSIC INCOME FLUCTUATION PROBLEM 253

1984 1989 Quintile

Type of Household Quintile 1st 2nd 3rd 4th 5th

with positive 1st 58.8% 25.1% 9.0% 5.1% 2.0%earnings in 2nd 20.2% 45.6% 21.6% 8.6% 4.0%both 1984 and 1989 3rd 9.7% 20.2% 40.4% 21.9% 7.8%

4th 7.7% 6.1% 20.0% 45.9% 20.4%5th 3.6% 2.9% 9.0% 18.4% 66.1%

with heads 1st 63.3% 27.2% 4.0% 3.3% 2.3%35-45 years old 2nd 23.6% 44.3% 22.3% 7.3% 2.4%

3rd 4.7% 16.7% 47.0% 25.1% 6.6%4th 6.9% 8.1% 20.2% 44.6% 20.1%5th 1.1% 4.0% 6.4% 19.1% 69.3%

We find the following stylized facts

• There is substantial persistence of labor earnings, in particular at thelowest and highest quintile. For the lowest quintile this may be due toretirees and long-term unemployed. Stratifying the sample as in Table 7indicates that this may be part of the explanation that 85.8% of all thehouseholds that were in the lowest earnings quintile in 1984 are in thelowest earnings quintile in 1989. But even looking at Table 7 there seemsto be substantial persistence of earnings at the low and high end, withpersistence being even more accentuated for prime-age households.

• The persistence properties of income are similar to those of earnings, whichis understandable given the high correlation between income and earnings

• Wealth seems to be more persistent than income and earnings.

Now let us start building a model that tries to explain the U.S. wealthdistribution, taking as given the earnings distribution, i.e. treating the earningsdistribution as an input in the model.

10.2 The Classic Income Fluctuation Problem

Bewley models study economies where a large number of agents face the classicincome fluctuation problem: they face a stochastic, exogenously given incomeand interest rate process and decide how to allocate consumption over time,i.e. how much of current income to consume and how much to save. So beforediscussing the full-blown general equilibrium dynamics of the model, let’s reviewthe basic results on the partial equilibrium income fluctuation problem.

Page 260: Macroeconomics Phd Lectures notes

254 CHAPTER 10. BEWLEY MODELS

The problem is to

maxct,at+1Tt=0

E0

TXt=0

βtu(ct) (10.1)

s.t. ct + at+1 = yt + (1 + rt)at

at+1 ≥ −b, ct ≥ 0a0 given

aT+1 = 0 if T finite

Here ytTt=0 and rtTt=0 are stochastic processes, b is a constant borrowingconstraint and T is the life horizon of the agent, where T = ∞ corresponds tothe standard infinitely lived agent model. We will make the assumptions thatu is strictly increasing, strictly concave and satisfies the Inada conditions. Notethat we will have to make further assumptions on the processes yt and rtto assure that the above problem has a solution.Given that this section is thought of as a preparation for the general equilib-

rium of the Bewley model, and given that we will have to constrain ourselves tostationary equilibria, we will from now on assume that rtTt=0 is deterministicand constant sequence, i.e. rt = r ∈ (−1,∞), for all t.

10.2.1 Deterministic Income

Suppose that the income stream is deterministic, with yt ≥ 0 for all t and yt > 0for some t. Also assume that r > 0 and

TXt=0

yt(1 + r)t

+ (1 + r)a0 <∞

This constraint is obviously satisfied if T is finite. If T is infinite this constraint issatisfied whenever the sequence yt∞t=0 is bounded and r > 0, although weakerrestrictions are sufficient, too.4

Note that under this assumption we can consolidate the budget constraintsto one Arrow-Debreu budget constraint

a0 +TXt=0

yt(1 + r)t+1

=TXt=0

ct(1 + r)t+1

and the implicit asset holdings at period t+ 1 are (by summing up the budgetconstraints from period t+ 1 onwards)

at+1 =TX

τ=t+1

cτ(1 + r)τ−t

−TX

τ=t+1

yτ(1 + r)τ−t

4For example, that yt grows at a rate lower than the interest rate. Note that our assump-tions serve two purposes, to make sure that the income fluctuation problem has a solutionand that it can be derived from the Arrow Debreu budget constraint. One can weaken theassumptions if one is only interested in one of these purposes.

Page 261: Macroeconomics Phd Lectures notes

10.2. THE CLASSIC INCOME FLUCTUATION PROBLEM 255

Natural Debt Limit

Let the borrowing constraint be specified as follows

b = − supt

TXτ=t+1

yτ(1 + r)τ−t

<∞

where the last inequality follows from our assumptions made above.5 The key ofspecifying the borrowing constraint in this form is that the borrowing constraintwill never be binding. Suppose it would at some date T . Then cT+τ = 0 forall τ > 0, since the household has to spend all his income on repaying his debtand servicing the interest payments, which obviously cannot be optimal, giventhe assumed Inada conditions. Hence the optimal consumption allocation iscompletely characterized by the Euler equations

u0(ct) = β(1 + r)u0(ct+1)

and the Arrow-Debreu budget constraint

a0 +TXt=0

yt(1 + r)t+1

=TXt=0

ct(1 + r)t+1

Define discounted lifetime income as Y = a0 +PTt=0

yt(1+r)t+1 , then the optimal

consumption choices take the form

ct = ft(r, Y )

i.e. only depend on the interest rate and discounted lifetime income, and partic-ular do not depend on the timing of income. This is the simplest statement of thepermanent income-life cycle (PILCH) hypothesis by Friedman and Modigliani(and Ando and Brumberg). Obviously, since we discuss a model here the hy-pothesis takes the form of a theorem.

For example, take u(c) = c1−σ1−σ , then the first order condition becomes

c−σt = β(1 + r)c−σt+1

ct+1 = [β(1 + r)]1σ ct

and hence

ct = [β(1 + r)]tσ c0

Provided that a = [β(1+r)]1σ

1+r < 1 (which we will assume from now on)6 we find

5For finite T it would make sense to define time-specific borrowing limits bt+1. This ex-tension is straightforward and hence omitted.

6We also need to assume that

β [b(1 + r)]1σ < 1

to assure that the sum of utilities converges for T =∞. Obviously, for finite T both assump-tions are not necessary for the following analysis.

Page 262: Macroeconomics Phd Lectures notes

256 CHAPTER 10. BEWLEY MODELS

that

c0 =(1 + r)(1− a)1− aT+1 Y

= f0,TY

ct =(1 + r)(1− a)1− aT+1 [β(1 + r)]

tσ Y

= ft,TY

where ft,T is the marginal propensity to consume out of lifetime income inperiod t if the lifetime horizon is T. We observe the following

1. If T > T then ft,T < ft,T . A longer lifetime horizon reduces the marginalpropensity to consume out of lifetime income for a given lifetime income.This is obvious in that consumption over a longer horizon has to be fi-nanced with given resources.

2. If 1 + r < 1β then consumption is decreasing over time. If 1 + r >

1β then

consumption is increasing over time. If 1 + r = 1β then consumption is

constant over time. The more patient the agent is, the higher the growthrate of consumption. Also, the higher the interest rate the higher thegrowth rate of consumption.

3. If 1 + r = 1β then ft,T = fT , i.e. the marginal propensity to consume is

constant for all time periods.

4. If in addition σ = 1 (iso-elastic utility) and T = ∞, then f∞ = r, i.e.the household consumes the annuity value of discounted lifetime income:ct = rY for all t ≥ 0. This is probably the most familiar statement of thePILCH hypothesis: agents should consume permanent income rY in eachperiod.

Potentially Binding Borrowing Limits

Let us make the same assumptions as before, but now assume that the borrowingconstraint is tighter than the natural borrowing limit. For simplicity assumethat the consumer is prevented from borrowing completely, i.e. assume b =0, and assume that yt > 0 for all t ≥ 0. We also assume that the sequenceytTt=0 is constant at yt = y. Now in the optimization problem we have to takethe borrowing constraints into account explicitly. Forming the Lagrangian anddenoting by λt the Lagrange multiplier for the budget constraint at time t andby µt the Lagrange multiplier for the non-negativity constraint for at+1 we have,ignoring non-negativity constraints for consumption

L =TXt=0

βtu(ct) + λt(yt + (1 + r)at − at+1 − ct) + µtat+1

Page 263: Macroeconomics Phd Lectures notes

10.2. THE CLASSIC INCOME FLUCTUATION PROBLEM 257

The first order conditions are

βtu0(ct) = λt

βt+1u0(ct+1) = λt+1

−λt + µt + (1 + r)λt+1 = 0

and the complementary slackness conditions are

at+1µt = 0

or equivalently

at+1 > 0 implies µt = 0

Combining the first order conditions yields

u0(ct) ≥ β(1 + r)u0(ct+1)= if at+1 > 0

Now suppose that β(1 + r) < 1. We will show that in Bewley models in generalequilibrium the endogenous interest rate r indeed satisfies this restriction. Wedistinguish two situations

1. The household is not borrowing-constrained in the current period, i.e.at+1 > 0. Then under the assumption made about the interest rate ct+1 <ct, i.e. consumption is declining.

2. The household is borrowing constrained, i.e. at+1 = 0. He would liketo borrow and have higher consumption today, at the expense of lowerconsumption tomorrow, but can’t transfer income from tomorrow to todaybecause of the imposed constraint.

To deduce further properties of the optimal consumption-asset accumula-tion decision we now make the income fluctuation problem recursive. For thedeterministic problem this may seem more complicated, but it turns out to beuseful and also a good preparation for the stochastic case. The first questionalways is what the correct state variable of the problem is. At a given point oftime the past history of income is completely described by the current wealthlevel at that the agents brings into the period. In addition what matters for hiscurrent consumption choice is his current income yt. So we pose the followingfunctional equation(s)7

vt(at, y) = maxat+1,ct≥0

u(ct) + βvt+1(at+1, y)s.t. ct + at+1 = y + (1 + r)at

7In the case of finite T these are T distinct functional equations.

Page 264: Macroeconomics Phd Lectures notes

258 CHAPTER 10. BEWLEY MODELS

with a0 given. If the agent’s time horizon is finite and equal to T , we takevT+1(aT+1, y) ≡ 0. If T = ∞, then we can skip the dependence of the valuefunctions and the resulting policy functions on time.8

The next steps would be to show the following

1. Show that the principle of optimality applies, i.e. that a solution to thefunctional equation(s) one indeed solves the sequential problem defined in(10.1).

2. Show that there exists a unique (sequence) of solution to the functionalequation.

3. Prove qualitative properties of the unique solution to the functional equa-tion, such as v (or the vt) being strictly increasing in both arguments,strictly concave and differentiable in its first argument.

We will skip this here; most of the arguments are relatively straightforwardand follow from material in Chapter 3 of these notes.9 Instead we will assertthese propositions to be true and look at some results that they buy us. First weobserve that at and y only enter as sum in the dynamic programming problem.Hence we can define a variable xt = (1 + r)at + y, which we call, after Deaton(1991) “cash at hand”, i.e. the total resources of the agent available for con-sumption or capital accumulation at time t. The we can rewrite the functionalequation as

vt(xt) = maxct,at+1≥0

u(ct) + βvt+1(xt+1)s.t. ct + at+1 = xt

xt+1 = (1 + r)at+1 + y

or more compactly

vt(xt) = max0≤at+1≤xt

u(xt − at+1) + βvt+1((1 + r)at+1 + y)

8For the finite lifetime case, we could have assumed deterministically flucuating endow-ments ytTt=0, since we index value and policy function by time. For T =∞ in order to havethe value function independent of time we need stationarity in the underlying environment,i.e. a constant income (in fact, with the introduction of further state variables we can handledeterministic cycles in endowments).

9What is not straightforward is to demonstrate that we have a bounded dynamic program-ming problem, which obviously isn’t a problem for finite T, but may be for T = ∞ since wehave not assumed u to be bounded. One trick that is often used is to put bounds on thestate space for (y, at) and then show that the solution to the functional equation with theadditional bounds does satisfy the original functional equation. Obviously for yt = y we havealready assumed boundedness, but for the endogenous choices at we have to verify that is isinnocuous. It is relatively easy to show that there is an upper bound for a, say a such thatif at > a, then at+1 < at for arbitrary yt. This will bound the value function(s) from above.To prove boundedness from below is substantially more difficult since one has to bound con-sumption ct away from zero even for at = 0. Obviously for this we need the assumption thaty > 0.

Page 265: Macroeconomics Phd Lectures notes

10.2. THE CLASSIC INCOME FLUCTUATION PROBLEM 259

The advantage of this formulation is that we have reduced the problem to onestate variable. As it will turn out, the same trick works when the exogenousincome process is stochastic and i.i.d over time. If, however, the stochasticincome process follows a Markov chain with nonzero autocorrelation we willhave to add back current income as one of the state variables, since currentincome contains information about expected future income.It is straightforward to show that the value function(s) for the reformulated

problem has the same properties as the value function for the original problem.Again we invite the reader to fill in the details. We now want to show someproperties of the optimal policies. We denote by at+1(xt) the optimal assetaccumulation and by ct(xt) the optimal consumption policy for period t inthe finite horizon case (note that, strictly speaking, we also have to index thesepolicies by the lifetime horizon T , but we keep T fixed for now) and by a0(x), c(x)the optimal policies in the infinite horizon case. Note again that the infinitehorizon model is significantly simpler than the finite horizon case. As long asthe results for the finite and infinite horizon problem to be stated below areidentical, it is understood that the results both apply to the finiteFrom the first order condition, ignoring the nonnegativity constraint on con-

sumption we get

u0(ct(xt)) ≥ β(1 + r)v0t+1((1 + r)at+1(xt) + y) (10.2)

= if at+1(xt) > 0

and the envelope condition reads

v0t(xt) = u0(ct(xt)) (10.3)

The first result is straightforward and intuitive

Proposition 102 Consumption is strictly increasing in cash at hand, or c0t(xt) >0. There exists an xt such that at+1(xt) = 0 for all xt ≤ xt and a0t+1(xt) > 0for all xt > xt. It is understood that xt may be +∞. Finally c0t(x) ≤ 1 anda0t+1(xt) < 1.

Proof. For the first part differentiate the envelope condition with respectto xt to obtain

v00t (xt) = u00(ct(xt)) ∗ c0t(xt)

and hence

c0t(xt) =v00t (xt)

u00(ct(xt))> 0

since the value function is strictly concave.10

10Note that we implicitly assumed that the value function is twice differentiable and thepolicy function is differentiable. For general condition under which this is true, see Santos(1991). I strongly encourage students interested in these issues to take Mordecai Kurz’s Econ284.

Page 266: Macroeconomics Phd Lectures notes

260 CHAPTER 10. BEWLEY MODELS

Suppose the borrowing constraint is not binding, then from differentiatingthe first order condition with respect to xt we get

u00(ct(xt)) ∗ c0(xt) = β(1 + r)2v00t+1((1 + r)at+1(xt) + yt+1) ∗ a0t+1(xt)and hence

a0t+1(xt) =u00(ct(xt)) ∗ c0(xt)

β(1 + r)2v00t+1((1 + r)at+1(xt) + yt+1)> 0

Now suppose that at+1(xt) = 0. We want to show that if xt < xt, thenat+1(xt) = 0. Suppose not, i.e. suppose that at+1(xt) > at+1(xt) = 0. From thefirst order condition

u0(ct(xt)) = β(1 + r)v0t+1((1 + r)at+1(xt) + yt+1)u0(ct(xt)) ≥ β(1 + r)v0t+1((1 + r)at+1(xt) + yt+1)

But since v0t+1 is strictly decreasing (as vt+1 is strictly concave) we have

β(1 + r)v0t+1((1 + r)at+1(xt) + yt+1) < β(1 + r)v0t+1((1 + r)at+1(xt) + yt+1)

and on the other hand, since we already showed that ct(xt) is strictly increasing

ct(xt) < ct(xt)

and hence

u0(ct(xt)) > u0(ct(xt))

Combining we find

u0(ct(xt)) > u0(ct(xt)) ≥ β(1 + r)v0t+1((1 + r)at+1(xt) + yt+1)> β(1 + r)v0t+1((1 + r)at+1(xt) + yt+1) = u

0(ct(xt))

a contradiction since u0 is positive.Finally, to show that c0t(xt) ≤ 1 and a0t+1(xt) < 1 we differentiate the identity

in the region xt > xt

ct(xt) + at+1(xt) = xt

with respect to xt to obtain

c0t(xt) + a0t+1(xt) = 1

and since both function are strictly increasing, the desired result follows (Notethat for xt ≤ xt we have a0t+1(xt) = 0 and c0t(xt) = 1).The last result basically state that the more cash at hand an agent has,

coming into the period, the more he consumes and the higher his asset accumu-lation, provided that the borrowing constraint is not binding. It also states that

Page 267: Macroeconomics Phd Lectures notes

10.2. THE CLASSIC INCOME FLUCTUATION PROBLEM 261

there is a cut-off level for cash at hand below which the borrowing constraintis always binding. Obviously for all xt ≤ xt we have ct(xt) = xt, i.e. the agentconsumes all his income (current income plus accumulated assets plus interestrate).For the infinite lifetime case we can say more.

Proposition 103 Let T =∞. If a0(x) > 0 then x0 < x. a0(y) = 0. There existsa x > y such that a0(x) = 0 for all x ≤ x

Proof. If a0(x) > 0 then from envelope and FOC

v0(x) = (1 + r)βv0(x0)< v0(x0)

since (1+ r)β < 1 by our maintained assumption. Since v is strictly concave wehave x > x0.For second part, suppose that a0(y) > 0. Then from the first order condition

and strict concavity of the value function

v0(y) = (1 + r)βv0((1 + r)a0(y) + y)< v0((1 + r)a0(y) + y)< v0(y)

a contradiction. Hence a0(y) = 0 and c(y) = y.The last part we also prove by contradiction. Suppose a0(x) > 0 for all

x > y. Pick arbitrary such x and define the sequence xt∞t=0 recursively by

x0 = x

xt = (1 + r)a0(xt−1) + y ≥ y

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. Butthen a0(xt) > 0 by assumption. Hence

v0(x0) = (1 + r)βv0(x1)= [(1 + r)β]tv0(xt)< [(1 + r)β]tv0(y)= [(1 + r)β]tu0(y)

where the inequality follows from the fact that xt > y and the strict concavity ofv. the last equality follows from the envelope theorem and the fact that a0(y) = 0so that c(y) = y.But since v0(x0) > 0 and u0(y) > 0 and (1 + r)β < 1, we have that there

exists finite t such that v0(x0) > [(1 + r)β]tu0(y), a contradiction.This last result bounds the optimal asset holdings (and hence cash at hand)

from above for T = ∞. Since computational techniques usually rely on the

Page 268: Macroeconomics Phd Lectures notes

262 CHAPTER 10. BEWLEY MODELS

finiteness of the state space we want to make sure that for our theory thestate space can be bounded from above. For the finite lifetime case there is noproblem. The most an agent can save is by consuming 0 in each period andhence

at+1(xt) ≤ xt ≤ (1 + r)t+1a0 +tX

j=0

(1 + r)jy

which is bounded for any finite lifetime horizon T <∞.The last theorem says that cash at hand declines over time or is constant at

y, in the case the borrowing constraint binds. The theorem also shows that theagent eventually becomes credit-constrained: there exists a finite τ such thatthe agent consumes his endowment in all periods following τ . This follows fromthe fact that marginal utility of consumption has to decline at geometric rateβ(1 + r) if the agent is unconstrained and from the fact that once he is credit-constrained, he remains credit constrained forever. This can be seen as follows.First x ≥ y by the credit constraint. Suppose that a0(x) = 0 but a0(x0) > 0.Since x0 = a0(x)+y = y we have that x0 ≤ x. Thus from the previous propositiona0(x0) ≤ a0(x) = 0 and hence the agent remains credit-constrained forever.For the infinite lifetime horizon, under deterministic and constant income

we have a full qualitative characterization of the allocation: If a0 = 0 then theconsumer consumes his income forever from time 0. If a0 > 0, then cash at handand hence consumption is declining over time, and there exists a time τ(a0) suchthat for all t > τ(a0) the consumer consumes his income forever from thereon,and consequently does not save anything.

10.2.2 Stochastic Income and Borrowing Limits

Now we discuss the income fluctuation problem that the typical consumer inour Bewley economy faces. We assume that T = ∞ and (1 + r)β < 1. Theconsumer is assumed to have a stochastic income process yt∞t=0. We assumethat yt ∈ Y = y1, . . . yN, i.e. the income can take only a finite number ofvalues. We will first assume that yt is i.i.d over time, with

Π(yj) = prob(yt = yj)

We will then extend our discussion to the case where the endowment processfollows a Markov chain with transition function π

πij = prob(yt+1 = yj if yt = yi)

In this case we will assume that the transition matrix has a unique stationarymeasure11 associated with it, which we will denote by Π and we will assume

11Remember that a stationary measure Π (distribution) associated with Markov transitionmatrix π is an N × 1 vector that satisfies

Π0π = Π0

Given that π is a stochastic matix it has a (not necessarily unique) stationary distributionassociated with it.

Page 269: Macroeconomics Phd Lectures notes

10.2. THE CLASSIC INCOME FLUCTUATION PROBLEM 263

that the agent at period t = 0 draws the initial income from Π. We continue toassume that the borrowing limit is at b = 0 and that β(1 + r) < 1.For the i.i.d case the dynamic programming problem takes the form (we will

focus on infinite horizon from now on)

v(x) = max0≤a0≤x

u(x− a0) + βXj

Π(yj)v((1 + r)a0 + yj)

with first order condition

u0(x− a0(x)) ≥ β(1 + r)Xj

Π(yj)v0((1 + r)a0(x) + yj)

= if a0(x) > 0

and envelope condition

v0(x) = u0(x− a0(x)) = u0(c(x))Denote by X

j

Π(yj)v0((1 + r)a0(x) + yj) = Ev0(x0)

Note that we need the expectation operator since, even though a0(x) is a de-terministic choice, y0 is stochastic and hence x0 is stochastic. Again taking forgranted that we can show the value function to be strictly increasing, strictlyconcave and twice differentiable we go ahead and characterize the optimal poli-cies. The proof of the following proposition is identical to the deterministiccase.

Proposition 104 Consumption is strictly increasing in cash at hand, i.e. c0(x) ∈(0, 1]. Optimal asset holdings are either constant at the borrowing limit or strictly

increasing in cash at hand, i.e. a0(x) = 0 or da0(x)dx ∈ (0, 1)

It is obvious that a0(x) ≥ 0 and hence x0(x, y0) = (1 + r)a0(x) + y0 ≥ y1 sowe have y1 > 0 as a lower bound on the state space for x. We now show thatthere is a level x > y1 for cash at hand such that for all x ≤ x we have thatc(x) = x and a0(x) = 0

Proposition 105 There exists x ≥ y1 such that for all x ≤ x we have c(x) = xand a0(x) = 0

Proof. Suppose, to the contrary, that a0(x) > 0 for all x ≥ y1. Then, usingthe first order condition and the envelope condition we have for all x ≥ y1

v(x) = β(1 + r)Ev0(x0) ≤ β(1 + r)v0(y1) < v0(y1)

Picking x = y1 yields a contradiction.

Page 270: Macroeconomics Phd Lectures notes

264 CHAPTER 10. BEWLEY MODELS

Hence there is a cutoff level for cash at hand below which the consumerconsumes all cash at hand and above which he consumes less than cash at handand saves a0(x) > 0. So far the results are strikingly similar to the deterministiccase. Unfortunately here it basically ends, and therefore our analytical ability tocharacterize the optimal policies. In particular, the very important propositionshowing that there exists x such that if x ≥ x then x0 < x does not go throughanymore, which is obviously quite problematic for computational considerations.In fact we state, without a proof, a result due to Schechtman and Escudero(1977)

Proposition 106 Suppose the period utility function is of constant absoluterisk aversion form u(c) = −e−c, then for the infinite life income fluctuationproblem, if Π(y = 0) > 0 we have xt → +∞ almost surely, i.e. for almost everysample path y0(ω), y2(ω), . . . of the stochastic income process.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 toSchechtman and Escudero for the proof of the following results. Intuitively whywould cash at hand go off to infinity even if the agents are impatient relative tothe market interest rate, i.e. even if β(1+ r) < 1? If agents are very risk averse,face borrowing constraints and a positive probability of having very low incomefor a long time, they may find it optimal to accumulated unbounded funds overtime to self-insure against the eventuality of this unlikely, but very bad eventto happen. It turns out that if one assumes that the risk aversion of the agentis sufficiently bounded, then one can rule this out.

Proposition 107 Suppose that the marginal utility function has the propertythat there exist finite eu0 such that

limc→∞ (logc u

0(c)) = eu0

Then there exists a x such that x0 = (1 + r)a0(x) + yN ≤ x for all x ≥ x.Proof. See Schechtman and Escudero (1977), Theorems 3.8 and 3.9The number eu0 is called the asymptotic exponent of u

0. Note that if theutility function is of CRRA form with risk aversion parameter σ, then since

logc c−σ = −σ logc c = −σ

we have eu0 = −σ and hence for these utility function the previous propositionapplies. Also note that for CARA utility function

logc e−c = −c logc e = −

c

ln(c)

− limc→∞

c

ln(c)= −∞

Page 271: Macroeconomics Phd Lectures notes

10.2. THE CLASSIC INCOME FLUCTUATION PROBLEM 265

45 degree line

45 degree line

_ ~y x x x1

y1

yN

a’(x)

c(x)

x’=a’(x)+yN

x’=a’(x)+y1

Figure 10.2:

and hence the proposition does not apply.So under the proposition of the previous theorem we have the result that

cash at hand stays in the bounded set X = [y1, x].12 Consumption equals cash

at hand for x ≤ x and is lower than x for x > x, with the rest being spent oncapital accumulation a0(x) > 0. Figure 27 shows the situation.Finally consider the case where income is correlated over time and follows

a Markov chain with transition π. Now the trick of reducing the state to thesingle variable cash at hand does not work anymore. This was only possible sincecurrent income y and past saving (1 + r)a entered additively in the constraintset of the Bellman equation, but neither variable appeared separately. Withserially correlated income, however, current income influences the probabilitydistribution of future income. There are several possibilities of choosing thestate space for the Bellman equation. One can use cash at hand and current

12If x0 = (1 + r)a0 + y0 happens to be bigger than x, then pick x0 = x0.

Page 272: Macroeconomics Phd Lectures notes

266 CHAPTER 10. BEWLEY MODELS

income, (x, y), or asset holdings and current income (a, y). Obviously both waysare equivalent and I opted for the later variant, which leads to the functionalequation

v(a, y) = maxc,a0≥0

u(c) + βXy0∈Y

π(y0|y)v(a0, y0)

s.t. c+ a0 = y + (1 + r)a

What can we say in general about the properties of the optimal policy functionsa0(a, y) and c(a, y). Huggett (1993) proves a proposition similar to the onesabove showing that c(a, y) is strictly increasing in a and that a0(a, y) is constantat the borrowing limit or strictly increasing (which implies a cutoff a(y) asbefore, which now will depend on current income y). What turns out to be verydifficult to prove is the existence of an upper bound of the state space, a suchthat a0(a, y) ≤ a if a ≥ a. Huggett proves this result for the special case thatN = 2, assumptions on the Markov transition function and CRRA utility. Seehis Lemmata 1-3 in the appendix. I am not aware of any more general result forthe non-iid case. With respect to computation in more general cases, we haveto cross our fingers and hope that a0(a, y) eventually (i.e. for finite a) crossesthe 450-line for all y.Until now we basically have described the dynamic properties of the optimal

decision rules of a single agent. The next task is to explicitly describe ourBewley economy, aggregate the decisions of all individuals in the economy andfind the equilibrium interest rate for this economy.

10.3 Aggregation: Distributions as State Vari-ables

10.3.1 Theory

Now let us proceed with the aggregation across individuals. First we describethe economy formally. We consider a pure exchange economy with a continuumof agents of measure 1. Each individual has the same stochastic endowmentprocess yt∞t=0 where yt ∈ Y = y1, y2, . . . yN. The endowment process isMarkov. Let π(y0|y) denote the probability that tomorrow’s endowment takesthe value y0 if today’s endowment takes the value y. We assume a law of largenumbers to hold: not only is π(y0|y) the probability of a particular agent ofa transition form y to y0 but also the deterministic fraction of the populationthat has this particular transition.13 Let Π denote the stationary distributionassociated with π, assumed to be unique. We assume that at period 0 theincome of all agents, y0, is given, and that the distribution of incomes across

13Whether and under what conditions we can assume such a law of large numbers createda heated discussion among theorists. See Judd (1985), Feldman and Gilles (1985) and Uhlig(1996) for further references.

Page 273: Macroeconomics Phd Lectures notes

10.3. AGGREGATION: DISTRIBUTIONS AS STATE VARIABLES 267

the population is given by Π. Given our assumptions, then, the distribution ofincome in all future periods is also given by Π. In particular, the total income(endowment) in the economy is given by

y =Xy

yΠ(y)

Hence, although there is substantial idiosyncratic uncertainty about a particularindividual’s income, the aggregate income in the economy is constant over time,i.e. there is no aggregate uncertainty.Each agent’s preferences over stochastic consumption processes are given by

u(c) = E0

∞Xt=0

βtu(ct)

with β ∈ (0, 1). In period t the agent can purchase one period bonds that paynet real interest rate rt+1 tomorrow. Hence an agent that buys one bond today,at the cost of one unit of today’s consumption good, receives (1+ rt+1) units ofconsumption goods for sure tomorrow. Hence his budget constraint at period treads as

ct + at+1 = yt + (1 + rt)at

We impose an exogenous borrowing constraint on bond holdings: at+1 ≥ −b.The agent starts out with initial conditions (a0, y0). Let Φ0(a0, y0) denote theinitial distribution over (a0, y0) across households. In accordance with our pre-vious assumption the marginal distribution of Φ0 with respect to y0 is assumedto be Π. We assume that there is no government, no physical capital or nosupply or demand of bonds from abroad. Hence the net supply of assets in thiseconomy is zero.At each point of time an agent is characterized by her current asset position

at and her current income yt. These are her individual state variables. Whatdescribes the aggregate state of the economy is the cross-sectional distributionover individual characteristics Φt(at, yt). We are now ready to define an equi-librium. We could define a sequential markets equilibrium and it is a goodexercise to do so, but instead 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 aggregatestate variable.First we need to define an appropriate measurable space on which the mea-

sures Φ are defined. Define the set A = [−b,∞) of possible asset holdings andby Y the set of possible income realizations. Define by P(Y ) the power setof 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 allprobability measures on the measurable space M = (Z,B(Z)). Why all this?Because our measures Φ will be required to elements ofM. Now we are readyto define a recursive competitive equilibrium. At the heart of any RCE is the

Page 274: Macroeconomics Phd Lectures notes

268 CHAPTER 10. BEWLEY MODELS

recursive formulation of the household problem. Note that we have to includeall state variables in the household problem, in particular the aggregate statevariable, since the interest rate r will depend on Φ. Hence the household problemin recursive formulation is

v(a, y;Φ) = maxc≥0,a0≥−b

u(c) + βXy0∈Y

π(y0|y)v(a0, y0;Φ0)

s.t. c+ a0 = y + (1 + r(Φ))

Φ0 = H(Φ)

The function H :M→M is called the aggregate “law of motion”. Now let usproceed to the equilibrium definition.

Definition 108 A recursive competitive equilibrium is a value function v : Z×M → R, policy functions a0 : Z ×M → R and c : Z ×M → R, a pricingfunction r :M→ R and an aggregate law of motion H :M→M such that

1. v, a0, c are measurable with respect to B(Z), v satisfies the household’sBellman equation and a0, c are the associated policy functions, given r().

2. For all Φ ∈M Zc(a, y;Φ)dΦ =

ZydΦZ

a0(a, y;Φ)dΦ = 0

3. The aggregate law of motion H is generated by the exogenous Markovprocess π and the policy function a0 (as described below)

Several remarks are in order. Condition 2. requires that asset and goodsmarkets clear for all possible measures Φ ∈M. Similarly for the requirementsin 1. As usual, one of the two market clearing conditions is redundant by Walras’law. Also note that the zero on the right hand side of the asset market clearingcondition indicates that bonds are in zero net supply in this economy: wheneversomebody borrows, another private household holds the loan.Now let us specify what it means thatH is generated by π and a0. H basically

tells us how a current measure over (a, y) translates into a measure Φ0 tomorrow.So H has to summarize how individuals move within the distribution over assetsand income from one period to the next. But this is exactly what a transitionfunction tells us. So define the transition function Q : Z × B(Z)→ [0, 1] by14

Q((a, y), (A,Y)) =Xy0∈Y

½π(y0|y) if a0(a, y) ∈ A

0 else

14Note that, since a0 is also a function of Φ, Q is implicitly a function of Φ, too.

Page 275: Macroeconomics Phd Lectures notes

10.3. AGGREGATION: DISTRIBUTIONS AS STATE VARIABLES 269

for all (a, y) ∈ Z and all (A,Y) ∈ B(Z). Q((a, y), (A,Y)) is the probability thatan agent with current assets a and current income y ends up with assets a0 inA tomorrow and income y0 in Y tomorrow. Suppose that Y is a singleton, sayY = y1. The probability that tomorrow’s income is y0 = y1, given today’sincome is π(y0|y). The transition of assets is non-stochastic as tomorrows assetsare chosen today according to the function a0(a, y). So either a0(a, y) falls intoA or it does not. Hence the probability of transition from (a, y) to y1 × A isπ(y0|y) if a0(a, y) falls into A and zero if it does not fall into A. If Y containsmore than one element, then one has to sum over the appropriate π(y0|y).How does the function Q help us to determine tomorrow’s measure over

(a, y) from today’s measure? Suppose Q where a Markov transition matrix fora finite state Markov chain and Φt would be the distribution today. Then tofigure out the distribution Φt tomorrow we would just multiply Q by Φt, or

Φt+1 = QTΦt

But a transition function is just a generalization of a Markov transition matrixto uncountable state spaces. For the finite state space we use sums

Φj,t+1 =NXi=1

QTijΦi,t

in our case we use the same idea, but integrals

Φ0(A,Y) = (H(Φ)) (A,Y) =ZQ((a, y), (A,Y))Φ(da× dy)

The fraction of people with income in Y and assets in A is that fraction ofpeople today, as measured by Φ, that transit to (A,Y), as measured by Q.In general there no presumption that tomorrow’s measure Φ0 equals today’s

measure, since we posed an arbitrary initial distribution over types, Φ0. If thesequence of measures Φt generated by Φ0 and H is not constant, then obvi-ously interest rates rt = r(Φt) are not constant, decision rules are not constantover time and the computation of equilibria is difficult in general. Therefore weare frequently interested in stationary RCE’s:

Definition 109 A stationary RCE is a value function v : Z → R, policy func-tions a0 : Z → R and c : Z → R, an interest rate r∗ and a probability measureΦ∗ such that

1. v, a0, c are measurable with respect to B(Z), v satisfies the household’sBellman equation and a0, c are the associated policy functions, given r∗

2. Zc(a, y)dΦ∗ =

ZydΦ∗Z

a0(a, y)dΦ∗ = 0

Page 276: Macroeconomics Phd Lectures notes

270 CHAPTER 10. BEWLEY MODELS

3. For all (A,Y) ∈ B(Z)

Φ∗(A,Y) =ZQ((a, y), (A,Y))dΦ∗ (10.4)

where Q is the transition function induced by π and a0 as described above

Note the big simplification: value functions, policy functions and prices arenot any longer indexed by measures Φ, all conditions have to be satisfied only forthe equilibrium measure Φ∗. The last requirement states that the measure Φ∗

reproduces itself: starting with distribution over incomes and assets Φ∗ todaygenerates the same distribution tomorrow. In this sense a stationary RCE isthe equivalent of a steady state, only that the entity characterizing the steadystate is not longer a number (the aggregate capital stock, say) but a rathercomplicated infinite-dimensional object, namely a measure.What can we do theoretically about such an economy? Ideally one would like

to prove existence and uniqueness of a stationary RCE. This is pretty hard andwe will not go into the details. Instead I will outline of an algorithm to computesuch an equilibrium and indicate where the crucial steps in proving existenceare. In the last homework some (optional) questions guide you through animplementation of this algorithm.Finding a stationary RCE really amounts to finding an interest rate r∗ that

clears the asset market. I propose the following algorithm

1. Fix an r ∈ (−1, 1β −1). For a fixed r we can solve the household’s recursiveproblem (e.g. by value function iteration). This yields a value function vrand decision rules a0r, cr, which obviously depend on the r we picked.

2. The policy function a0r and π induce a Markov transition function Qr.Compute the unique stationary measure Φr associated with this transitionfunction from (10.4). The existence of such unique measure needs proving;here the property of a0r that for sufficiently large a, a0(a, y) ≤ a is crucial.Otherwise assets of individuals wander off to infinity over time and astationary measure over (a, y) does not exist.

3. Compute average net asset demand

Ear =

Za0r(a, y)dΦr

Note that Ear is just a number. If this number happens to equal zero, weare done and have found a stationary RCE. If not we update our guess forr and start from 1. anew.

So the key steps, apart from proving the existence and uniqueness of astationary measure in proving the existence of an RCE is to show that, as afunction of r, Ear is continuous in r, negative for small r and positive for larger. If one also wants to prove uniqueness of a stationary RCE, one in addition

Page 277: Macroeconomics Phd Lectures notes

10.3. AGGREGATION: DISTRIBUTIONS AS STATE VARIABLES 271

has to show that Ear is strictly increasing in r, i.e. that households want tosave more the higher the interest rate. Continuity of Ear is quite technical, butbasically requires to show that a0r is continuous in r; proving strict monotonicityof Ear requires proving monotonicity of a

0r with respect to r. I will spare you the

details, some of which are not so well-understood yet (in particular if income isMarkov rather than i.i.d). That Ear is negative for r = −1. If r = −1, agentscan borrow without repaying anything, and obviously all agents will borrow upto the borrowing limit. Hence Ea−1 = −b < 0On the other hand, as r approaches ρ = 1

β −1 from below, Ear goes to +∞.The result that for r = ρ asset holdings wander off to infinity almost surely wasproved by Chamberlain and Wilson (1984) using the martingale convergencetheorem; this is well beyond this course. Let’s give a heuristic argument for thecase in which income is i.i.d. In this case the first order condition and envelopecondition reads

u0(c(a, y))) ≥Xy0

π(y0)v0(a0(a, y), y0)

= if a0(a, y) > −bv0(a, y) = u0(c(a, y))

Suppose there exists an amax such that a0(amax, y) ≤ a for all y ∈ Y. But then

v0(amax, ymax) ≥Xy0

π(y0)v0(a0(amax, ymax), y0)

>Xy0

π(y0)v0(amax, y0)

>Xy0

π(y0)v0(amax, ymax)

= v0(amax, ymax)

a contradiction. The inequalities follow from strict concavity of the value func-tion in its first argument and the fact that higher income makes the marginalutility form wealth decline. Hence asset holdings wander off to infinity almostsurely and Ear =∞.What goes on is that without uncertainty and β(1+r) = 1the consumer wants to keep a constant profile of marginal utility over time. Withuncertainty, since there is a positive probability of getting a sufficiently long se-quence of bad income, this requires arbitrarily high asset holdings.15 Figure 28summarizes the results.The average asset demand curve, as a function of the interest rate, is up-

ward sloping, is equal to −b for sufficiently low r, asymptotes towards ∞ as r

15This argument was loose in the sense that if a0(a, y) does not cross the 45 degree line,then no stationary asset distribution exists and, strictly speaking, Ear is not well-defined.What one can show, however, is that in the income fluctuation problem with β(1+ r) = 1 foreach agent at+1 →∞ almost surely, meaning that in the limit asset holdings become infinite.If we understand this time limit as the stationary situation, then Ear =∞ for r = ρ.

Page 278: Macroeconomics Phd Lectures notes

272 CHAPTER 10. BEWLEY MODELS

Ear

-b

r

0

r*

ρ

Figure 10.3:

Page 279: Macroeconomics Phd Lectures notes

10.3. AGGREGATION: DISTRIBUTIONS AS STATE VARIABLES 273

approaches ρ = 1β − 1 from below. The Ear curve intersects the zero-line at a

unique r∗, the unique stationary equilibrium interest rate.

This completes our description of the theoretical features of the Bewleymodels. Now we will turn to the quantitative results that applications of thesemodels have delivered.

10.3.2 Numerical Results

In this subsection we report results obtained from numerical simulations of themodel described above. In order to execute these simulations we first have topick the exogenous parameters characterizing the economy. The parameters in-clude the parameters specifying preferences, (σ,β) (we assume constant relativerisk aversion utility function), the exogenous borrowing limit b and the param-eters specifying the income process, i.e. the transition matrix π and the statesthat the income process can take, Y.

We envision the model period as 1 year, so we choose β = 0.97. As coefficientof relative risk aversion we choose σ = 2. As borrowing limit we choose b = 1.We will normalize the income process so that average (aggregate) income in theeconomy is 1. Hence the borrowing constraint permits borrowing up to 100% ofaverage yearly income. For the income process we do the following. We followthe labor literature and assume that log-income follows an AR(1) process

log yt = ρ log yt−1 + σ(1− ρ2)12 εt

where εt is distributed normally with mean zero and variance 1.16 We then

use the procedure by Tauchen and Hussey to discretize this continuous statespace process into a discrete Markov chain.17 For ρ and σε we used numbersestimated by Heaton and Lucas (1996) who found ρ = 0.53 and σε = 0.296.Wepicked the number of states to be N = 5. The resulting income process looks

16For the process spaecified above ρ is the autocrooelation of the process

ρ =cov(log yt, log yt−1)

var(log yt)

and σε is the unconditional standard deviation of the process

σε =pvar(log yt))

The numbers were estimated from PSID data.17The details of this rather standard approach in applied work are not that important here.

Aiyagari’s working paper version of the paper has a very good description of the procedure;see me if you would like a copy.

Page 280: Macroeconomics Phd Lectures notes

274 CHAPTER 10. BEWLEY MODELS

as follows.

π =

0.27 0.55 0.17 0.01 0.000.07 0.45 0.41 0.06 0.000.01 0.22 0.53 0.22 0.010.00 0.06 0.41 0.45 0.070.00 0.01 0.17 0.55 0.27

Y = 0.40, 0.63, 0.94, 1.40, 2.19Π = [0.03, 0.24, 0.45, 0.24, 0.03]

Remember that Π is the stationary distribution associated with π.What are the key endogenous variables of interest. First, the interest rate

in this economy is computed to be r − 0.5%. Second, the model delivers anendogenous distribution over asset holdings. This distribution is shown in Figure29.We see that the richest (in terms of wealth) people in this economy holdabout six times average income, whereas the poorest people are pushed to theborrowing constraint. About 6% of the population appears to be borrowingconstrained. How does this economy compare to the data. First, the averagelevel of wealth in the economy is zero, by construction, since the net supply ofassets is zero. This is obviously unrealistic, and we will come back to this below.How about the dispersion of wealth. The Lorenz curve and the Gini coefficientdo not make much sense here, since too many people hold negative wealth byconstruction. The standard deviation is about 0.93. Since average assets arezero, we can’t compute the coefficient of variation of wealth. However, sinceaverage income is 1, the ratio of the standard deviation of wealth to averageincome is 0.93, whereas in the data it is 33 (where we used earnings instead ofincome). Hence the model underachieves in terms of wealth dispersion. This ismostly due to two reasons, one that has to do with the model and one that has todo with our parameterization. How much dispersion in income did we stick intothe model? The coefficient of variation of the income process that we used in themodel is 0.355 instead of 4.19 in the data (again we used earnings for the data).So we didn’t we use a more dispersed income process, or in other words, why didHeaton and Lucas find the numbers in the data that we used? Remember thatin our model all people are ex ante identical and income differences result inex-post differences of luck. In the data earnings of people differ not only becauseof chance, but because of observable differences. Heaton and Lucas filtered outdifferences in income that have to do with deterministic factors like age, sex,race etc.18 Why do we use their numbers? Because they filter out exactly thosecomponents of income dispersion that our model abstracts from.Even if one would rig the income numbers to be more dispersed, the model

would fail to reproduce the amount of wealth dispersion, largely because it failsto generate the fat upper tail of the wealth distribution. There have been severalsuggestions to cure this failure; for example to introduce potential entrepreneursthat have to accumulate a lot of wealth before financing investment projects, A

18The details of their procedure are more appropriately discussed by the econometriciansof the department than by me, so I punt here.

Page 281: Macroeconomics Phd Lectures notes

10.3. AGGREGATION: DISTRIBUTIONS AS STATE VARIABLES 275

-2 0 2 4 6 8 100

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

0.1Stationary Asset Distribution

Amount of Assets (as Fraction of Average Yearly Income)

Per

cent

of t

he P

opul

atio

n

Figure 10.4:

Page 282: Macroeconomics Phd Lectures notes

276 CHAPTER 10. BEWLEY MODELS

somewhat successful strategy has been to introduce stochastic discount factors;let β follow a Markov chain with persistence. Some days people wake up andare really impatient, other days they are patient. This seems to do the trick.Instead of picking up these extensions we want to study how the model reacts

to changes in parameter values. Most interestingly, what happens if we loosenthe borrowing constraint? Suppose we increase the borrowing limit from 1 yearsaverage income to 2 years average income. Then people can borrow more andsome previously constrained people will do so. On the other hand agents canalways freely save. Hence for a given interest rate the net demand for bonds, ornet saving should go down, the Ear-curve shifts to the left and the equilibriuminterest rate should increase. The new equilibrium interest rate is r = 1.7%.Now about 1.5% of population is borrowing constrained. The richest peoplehold about eight times average income as wealth. The ratio of the standarddeviation of wealth to average income is 1.65 now, increased from 0.93 with aborrowing limit of b = 1. Figure 30 shows the equilibrium asset distribution.

Page 283: Macroeconomics Phd Lectures notes

10.3. AGGREGATION: DISTRIBUTIONS AS STATE VARIABLES 277

-2 0 2 4 6 8 100

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

0.1Stationary Asset Distribution

Amount of Assets (as Fraction of Average Yearly Income)

Per

cent

of t

he P

opul

atio

n

Figure 10.5:

Page 284: Macroeconomics Phd Lectures notes

278 CHAPTER 10. BEWLEY MODELS

Page 285: Macroeconomics Phd Lectures notes

Chapter 11

Fiscal Policy

11.1 Positive Fiscal Policy

11.2 Normative Fiscal Policy

11.2.1 Optimal Policy with Commitment

11.2.2 The Time Consistency Problem and Optimal FiscalPolicy without Commitment

[To Be Written]

279

Page 286: Macroeconomics Phd Lectures notes

280 CHAPTER 11. FISCAL POLICY

Page 287: Macroeconomics Phd Lectures notes

Chapter 12

Political Economy andMacroeconomics

[To Be Written]

281

Page 288: Macroeconomics Phd Lectures notes

282 CHAPTER 12. POLITICAL ECONOMY AND MACROECONOMICS

Page 289: Macroeconomics Phd Lectures notes

Chapter 13

References

1. Introduction

• Ljungqvist, L. and T. Sargent (2000): Recursive Macroeconomic The-ory, MIT Press, Preface.

2. Arrow-Debreu Equilibria, Sequential Markets Equilibria and Pareto Op-timality in Simple Dynamic Economies

• Kehoe, T. (1989): “Intertemporal General Equilibrium Models,” inF. Hahn (ed.) The Economics of Missing Markets, Information andGames, Claredon Press

• Ljungqvist and Sargent, Chapter 7.• Negishi, T. (1960): “Welfare Economics and Existence of an Equi-librium for a Competitive Economy,” Metroeconomica, 12, 92-97.

3. The Neoclassical Growth Model in Discrete Time

• Cooley, T. and E. Prescott (1995): “Economic Growth and BusinessCycles,” in T. Cooley (ed.) Frontiers of Business Cycle Research,Princeton University Press.

• Prescott, E. and R. Mehra (1980): “Recursive Competitive Equi-librium: the Case of Homogeneous Households,” Econometrica, 48,1356-1379.

• Stokey, N. and R. Lucas, with E. Prescott (1989): Recursive Methodsin Economic Dynamics, Harvard University Press, Chapter 2.

4. Mathematical Preliminaries for Dynamic Programming

• Stokey et al., Chapter 3.

5. Discrete Time Dynamic Programming

283

Page 290: Macroeconomics Phd Lectures notes

284 CHAPTER 13. REFERENCES

• Ljungqvist and Sargent, Chapter 2 and 3.• Stokey et al., Chapter 4.

6. Models with Uncertainty

• Stokey et al., Chapter 7.

7. The Welfare Theorems in Infinite Dimensions

• Debreu, G. (1983): “Valuation Equilibrium and Pareto Optimum, inMathematical Economics: Twenty Papers of Gerard Debreu, Cam-bridge University Press.

• Stokey et al., Chapter 15 and 16

8. Overlapping Generations Economies: Theory and Applications

• Barro, R. (1974): “Are Government Bonds Net Wealth?,” Journal ofPolitical Economy, 82, 1095-1117.

• Blanchard and Fischer, Chapter 3.• Conesa, J. and D. Krueger (1999): “Social Security Reform withHeterogeneous Agents,” Review of Economic Dynamics, 2, 757-795.

• Diamond, P. (1965): “National Debt in a Neo-Classical GrowthModel,”American Economic Review, 55, 1126-1150.

• Gale, D. (1973): “Pure Exchange Equilibrium of Dynamic EconomicModels,” Journal of Economic Theory, 6, 12-36.

• Geanakoplos, J (1989): “Overlapping Generations Model of GeneralEquilibrium,” in J. Eatwell, M. Milgrate and P. Newman (eds.) TheNew Palgrave: General Equilibrium

• Kehoe, T. (1989): “Intertemporal General Equilibrium Models,” inF. Hahn (ed.) The Economics of Missing Markets, Information andGames, Claredon Press

• Ljungquist and Sargent, Chapter 8 and 9.• Samuelson (1958): “An Exact Consumption Loan Model of Inter-est, With or Without the Social Contrivance of Money,” Journal ofPolitical Economy, 66, 476-82.

• Wallace, N. (1980): “The Overlapping Generations Model of FiatMoney,” in J.H. Kareken and N. Wallace (eds.) Models of MonetaryEconomies, Federal Reserve Bank of Minneapolis.

9. Growth Models in Continuous Time and their Empirical Evaluation

• Barro, R. (1990): “Government Spending in a Simple Model ofEndogenous Growth,” Journal of Political Economy, 98, S103-S125.

Page 291: Macroeconomics Phd Lectures notes

285

• Barro, R. and Sala-i-Martin, X. (1995): Economic Growth, McGraw-Hill, Chapters 1,2,4,6 and Appendix

• Blanchard and Fischer, Chapter 2• Cass, David (1965): “Optimum Growth in an Aggregative Model ofCapital Accumulation,” Review of Economic Studies, 32, 233-240

• Chari, V.V., Kehoe, P. and McGrattan, E. (1997): “The Povertyof Nations: A Quantitative Investigation,” Federal Reserve Bank ofMinneapolis Staff Report 204.

• Intriligator, M. (1971): Mathematical Optimization and EconomicTheory, Englewood Cliffs, Chapters 14 and 16.

• Jones (1995): “R&D-Based Models of Economic Growth,” Journalof Political Economy, 103, 759-784.

• Lucas, R. (1988): “On the Mechanics of Economic Development,”Journal of Political Economy, Journal of Monetary Economics

• Mankiw, G., Romer, D. and Weil (1992): “A Contribution to theEmpirics of Economic Growth,” Quarterly Journal of Economics,107, 407-437.

• Ramsey, Frank (1928): “A Mathematical Theory of Saving,” Eco-nomic Journal, 38, 543-559.

• Rebelo, S. (1991): “Long-Run Policy Analysis and Long-Run Growth,”Journal of Political Economy, 99, 500-521.

• Romer (1986): “Increasing Returns and Long Run Growth,” Journalof Political Economy, 94, 1002-1037.

• Romer (1990): “Endogenous Technological Change,” Journal of Po-litical Economy, 98, S71-S102.

• Romer, D. (1996): Advanced Macroeconomics, McGraw-Hill, Chap-ter 2 and 3

• Ljungquist and Sargent, Chapter 11.

10. Models with Heterogeneous Agents

• Aiyagari, R. (1994): “Uninsured Risk and Aggregate Saving,” Quar-terly Journal of Economics, 109, 659-684.

• Aiyagari, A. (1995): “Optimal Capital Income Taxation with Incom-plete Markets, Borrowing Constraints, and Constant Discounting,”Journal of Political Economy, 103, 1158-1175.

• Aiyagari R. and McGrattan, E. (1998): “The Optimum Quantity ofDebt,” Journal of Monetary Economics, 42, 447-469

• Carroll, C. (1997): “Buffer-Stock Saving and the Life Cycle/PermanentIncome Hypothesis,” Quarterly Journal of Economics, 112, 1-55.

Page 292: Macroeconomics Phd Lectures notes

286 CHAPTER 13. REFERENCES

• Deaton, A. (1991): “Saving and Liquidity Constraints,” Economet-rica, 59, 1221-1248.

• Diaz-Jimenez, J., V. Quadrini and J.V. Rios-Rull (1997), “Dimen-sions of Inequality: Facts on the U.S. Distributions of Earnings, In-come, and Wealth,” Federal Reserve Bank of Minneapolis QuarterlyReview, Spring.

• Huggett, M. (1993): “The Risk-Free Rate in Heterogeneous-AgentIncomplete-Insurance Economies,” Journal of Economic Dynamicsand Control, 17, 953-969.

• Krusell, P. and Smith, A. (1998): “Income and Wealth Heterogeneityin the Macroeconomy,” Journal of Political Economy, 106, 867-896.

• Rios-Rull, V. (1999): “Computation of Equilibria in Heterogeneous-Agent Models,” in: R. Marimon and A. Scott (eds.) ComputationalMethods for the Study of Dynamic Economies, Oxford UniversityPress, 238-265.

• Sargent and Ljungquist, Chapter 14.• Schechtman, J. (1976): “An Income Fluctuation Problem,” Journalof Economic Theory, 12, 218-241.

• Schechtman, J. and Escudero, V. (1977): “Some Results on “AnIncome Fluctuation Problem”,” Journal of Economic Theory, 16,151-166.

• Stokey et al., Chapters 7-14.

11. Fiscal Policy with and without Commitment

• Aiyagari, R., Christiano, L., and Eichenbaum, M. (1992): “The Out-put, Employment and Interest Rate Effects of Government Consump-tion,” Journal of Monetary Economics, 30, 73-86.

• Barro, R. (1974): “Are Government Bonds Net Wealth?,” Journal ofPolitical Economy, 82, 1095-1117.

• Barro, R. (1979): “On the Determination of the Public Debt,” Jour-nal of Political Economy, 87, 940-971.

• Barro, R. (1981): “Output Effects of Government Purchases,” Jour-nal of Political Economy, 89, 1086-1121.

• Bassetto, M. (1998): “Optimal Taxation with Heterogeneous Agents,”mimeo.

• Baxter, M. and King, R. (1993): “Fiscal Policy in General Equilib-rium,” American Economic Review, 83, 315-334.

• Blanchard and Fischer, chapter 11• Chamley (1986): “Optimal taxation of Capital Income in GeneralEquilibrium with Infinite Lives,” Econometrica, 54, 607-622.

Page 293: Macroeconomics Phd Lectures notes

287

• Chari, V.V., Christiano, L. and Kehoe, P. (1995): “Policy Analysisin Business Cycle Models,” in: T. Cooley (ed.) Frontiers of BusinessCycle Research, Princeton University Press, 357-392.

• Chari, V.V and Kehoe, P. (1990): “Sustainable Plans,” Journal ofPolitical Economy, 98, 783-802.

• Chari, V.V and Kehoe, P. (1993a): “Sustainable Plans and MutualDefault,” Review of Economic Studies, 60, 175-195.

• Chari, V.V and Kehoe, P. (1993b): “Sustainable Plans and Debt,”Journal of Economic Theory, 61, 230-261.

• Chari, V.V. and Kehoe, P. (1999): “Optimal Monetary and FiscalPolicy,” Federal Reserve Bank of Minneapolis Staff Report 251.

• Klein and Rios-Rull (1999): “Time-Consistent Optimal Fiscal Pol-icy,” mimeo.

• Kydland, F. and Prescott, E. (1977): “Rules Rather than Discretion:The Inconsistency of Optimal Plans,” Journal of Political Economy,85, 473-492.

• Ljungquist and Sargent (1999), chapter 12 and 16• Ohanian, L. (1997): The Macroeconomic Effects of War Finance inthe United States: World war II and the Korean War,” AmericanEconomic Review, 87, 23-40.

• Stokey, N. (1989): “Reputation and Time Consistency,” AmericanEconomic Review, 79, 134-139.

• Stokey, N. (1991): “Credible Public Policy,” Journal of EconomicDynamics and Control, 15, 626-656.

12. Political Economy and Macroeconomics

• Alesina, A. and Rodrik, D. (1994): “Distributive Politics and Eco-nomic Growth,” Quarterly Journal of Economics, 109, 465-490.

• Bassetto, M. (1999): “Political Economy of Taxation in an Overlapping-Generations Economy,” mimeo.

• Boldrin, M. and Rustichini, A. (1998): “Political Equilibria withSocial Security,” mimeo.

• Cooley, T. and Soares, J. (1999): “A Positive Theory of Social Se-curity Based on Reputation, Journal of Political Economy, 107, 135-160.

• Imrohoroglu, A., Merlo, A. and Rupert, P. (1997): “On the Politi-cal Economy of Income Redistribution and Crime,” Federal ReserveBank of Minneapolis Staff Report 216.

• Krusell, P., Quadrini, V. and Rios-Rull, V. (1997): “Politico-EconomicEquilibrium and Economic Growth,” Journal of Economic Dynamicsand Control, 21, 243-72.

Page 294: Macroeconomics Phd Lectures notes

288 CHAPTER 13. REFERENCES

• Mulligan, C. and Sala-i-Martin, X. (1998): “Gerontocracy, Retire-ment and Social Security,” NBER Working Paper 7117.

• Persson and Tabellini (1994): “Is Inequality Harmful for Growth?,”American Economic Review, 84, 600-619.