Case Study on Decision Making Under Uncertainty And Bounded Rationality

Description
Bounded rationality is the idea that in decision-making, rationality of individuals is limited by the information they have, the cognitive limitations of their minds, and the finite amount of time they have to make a decision.

CASE STUDY ON DECISION MAKING UNDER
UNCERTAINTY AND BOUNDED RATIONALITY

Abstract:-

In an attempt to capture the complexity of the economic system many

economists were led to the formulation of complex nonlinear rational expectations

models that in many cases can not be solved analytically. In such cases, numerical

methods need to be employed. In chapter one I review several numerical methods that

have been used in the economic literature to solve non-linear rational expectations

models. I provide a classification of these methodologies and point out their strengths

and weaknesses. I conclude by discussing several approaches used to measure

accuracy of numerical methods.

In the presence of uncertainty, the multistage stochastic optimization literature

has advanced the idea of decomposing a multiperiod optimization problem into many

subproblems, each corresponding to a scenario. Finding a solution to the original

problem involves aggregating in some form the solutions to each scenario and hence

its name, scenario aggregation. In chapter two, I study the viability of scenario

aggregation methodology for solving rational expectation models. Specifically, I

apply the scenario aggregation method to obtain a solution to a finite horizon life
cycle model of consumption. I discuss the characteristics of the methodology and

compare its solution to the analytical solution of the model.

A growing literature in macroeconomics is tweaking the unbounded

rationality assumption in an attempt to find alternative approaches to modeling the

decision making process, that may explain observed facts better or easier. Following

this line of research, in chapter three, I study the impact of bounded rationality on the

level of precautionary savings in a finite horizon life-cycle model of consumption. I

introduce bounded rationality by assuming that the consumer does not have either the

resources or the sophistication to consider all possible future events and to optimize

accordingly over a long horizon. Consequently, he focuses on choosing a

consumption plan over a short span by considering a limited number of possible

scenarios. While under these assumptions the level of precautionary saving in many

cases is below the level that a rational expectations model would predict, there are

also parameterizations of the model for which the reverse is true.















































ii
Table of Contents



Dedication ..................................................................................................................... ii
Table of Contents......................................................................................................... iii
Chapter I. Review of Methods Used for Solving Non-Linear Rational Expectations
Models........................................................................................................................... 1
I.1. Introduction ........................................................................................................ 1
I.2. Generic Model .................................................................................................... 3
I.3. Using Certainty Equivalence; The Extended Path Method ................................ 6
I.3.1. Example ....................................................................................................... 7
I.3.2. Notes on Certainty Equivalence Methods ................................................. 10
I.4. Local Approximation and Perturbation Methods ............................................. 11
I.4.1. Regular and General Perturbation Methods .............................................. 11
I.4.2. Example ..................................................................................................... 13
I.4.3. Flavors of Perturbation Methods ............................................................... 15
I.4.4. Alternative Local Approximation Methods............................................... 16
I.4.5. Notes on Local Approximation Methods .................................................. 18
I.5. Discrete State-Space Methods .......................................................................... 19
I.5.1. Example. Discrete State-Space Approximation Using Value-Function
Iteration ............................................................................................................... 20
I.5.2. Fredholm Equations and Numerical Quadratures ..................................... 21
I.5.3. Example. Using Quadrature Approximations ........................................... 24
I.5.4. Notes on Discrete State-Space Methods.................................................... 26
I.6. Projection Methods........................................................................................... 27
I.6.1. The Concept of Projection Methods.......................................................... 28
I.6.2. Parameterized Expectations....................................................................... 39
I.6.3. Notes on Projection Methods .................................................................... 42
I.7. Comparing Numerical Methods: Accuracy and Computational Burden.......... 44
I.8. Concluding Remarks ........................................................................................ 47
Chapter II. Using Scenario Aggregation Method to Solve a Finite Horizon Life Cycle
Model of Consumption ............................................................................................... 49
II.1. Introduction ..................................................................................................... 49
II.2. A Simple Life-Cycle Model with Precautionary Saving ................................ 50
II.3. The Concept of Scenarios ............................................................................... 52
II.3.1. The Problem ............................................................................................. 52
II.3.2. Scenarios and the Event Tree ................................................................... 53
II.4. Scenario Aggregation...................................................................................... 57
II.5. The Progressive Hedging Algorithm............................................................... 60
II.5.1. Description of the Progressive Hedging Algorithm................................. 61
II.6. Using Scenario Aggregation to Solve a Finite Horizon Life Cycle Model .... 63
II.6.1. The Algorithm .......................................................................................... 65
II.6.2. Simulation Results.................................................................................... 68
II.6.3. The Role of the Penalty Parameter........................................................... 72
II.6.4. More simulations...................................................................................... 74
II.7. Final Remarks ................................................................................................. 76


iii
Chapter III. Impact of Bounded Rationality on the Magnitude of Precautionary
Saving ......................................................................................................................... 77
III.1. Introduction.................................................................................................... 77
III.2. Empirical Results on Precautionary Saving................................................... 80
III.3. The Model...................................................................................................... 82
III.3.1. Rule 1 ...................................................................................................... 87
III.3.2. Rule 2 ...................................................................................................... 96
III.3.3. Rule 3 .................................................................................................... 104
III.4. Final Remarks .............................................................................................. 110
Appendices................................................................................................................ 112
Appendix A. Technical notes to chapter 2............................................................ 112
Appendix A1. Definitions for Scenarios, Equivalence Classes and Associated
Probabilities ...................................................................................................... 112
Appendix A2. Description of the Scenario Aggregation Theory ..................... 115
Appendix A3. Solution to a Scenario Subproblem........................................... 118
Appendix B. Technical notes to chapter 3 ............................................................ 121
Appendix B1. Analytical Solution for a Scenario with Deterministic Interest
Rate ................................................................................................................... 121
Appendix B2. Details on the Assumptions in Rule 1 ....................................... 124
Appendix B3. Details on the Assumptions in Rule 2 ....................................... 125

































iv
Chapter I. Review of Methods Used for Solving Non-Linear


Rational Expectations Models




I.1. Introduction

Limitations faced by most linear macroeconomic models coupled with the

growing importance of rational expectations have led many economists, in an attempt to

capture the complexity of the economic system, to turn to non-linear rational expectation

models. Since the majority of these models can not be solved analytically, researchers

have to employ numerical methods in order to be able to compute a solution.

Consequently, the use of numerical methods for solving nonlinear rational expectations

models has been growing substantially in recent years.

For the past decade, several strategies have been used to compute the solutions to

nonlinear rational expectations models. The available numerical methods have several

common features as well as differences, and depending on the criteria used, they may be

grouped in various ways. Following is an ad-hoc categorization
1
that will be used

throughout this chapter.

The first group of methods I consider has as a common feature the fact that the

assumption of certainty equivalence is used at some point in the computation of the

solution.





1





This classification draws on Binder et al. (2000), Burnside (1999.), Marcet et al. (1999),
McGrattan (1999), Novales et al. (1999), Uhlig (1999) and Judd (1992, 1998).



1
The second group of methods has as a common denominator the use of a discrete

state space, or the discretization of an otherwise continuous space of the state variables
2
.

The methods falling into this category are often referred to as discrete state-space

methods. They work well for models with a low number of state variables.

The next set of methods is generically known as the class of perturbation

methods. Since perturbation methods make heavy use of local approximations, in this

presentation, I group them along with some other techniques that use local

approximations under the heading of local approximations and perturbation methods.

The fourth group, labeled here as projection methods consists of a collection of

methodologies that approximate the true value of the conditional expectations of

nonlinear functions with some finite parameterization and then evaluate the initially

undetermined parameters. Several methods included in this group have recently become

very popular in solving nonlinear rational expectations models containing a relatively

small number of state variables
3
.

The layout of the chapter contains the presentation of a generic non-linear rational

expectations model followed by a description of the methods mentioned above.

Throughout the chapter, special cases of the model described in section 2 are used to

show how one can apply the methods discussed here.






2






Examples include Baxter et al. (1990), Christiano (1990a, 1990b), Coleman (1990),
Tauchen (1990) and Taylor and Uhlig (1990), Tauchen and Hussey (1991), Deaton and
Laroque (1992), and Rust (1996)

3

This approach is used, for example, by Binder et al. (2000), Christiano and Fisher
(2000), Judd (1992) and Miranda and Rui (1997).



2
I.2. Generic Model


I start by presenting a generic model in discrete time that will be used along the

way to exemplify the application of some of the methods discussed in this chapter. I

assume that the problem consists of maximizing the expected present discounted value of

an objective function:

max E
?
¿ |
tt
(u
t
) O
0
? ·

subject to
u
t
?
?t =0


x
t
=
h
(x
t
÷
1
,u
t
, y
t
)

f (x
t
, x
t
÷
1
) > 0
?
?
(1.2.1)




(1.2.2)

(1.2.3)


where u
t
and x
t
denote the values of the control and state variables u and x

respectively, at the beginning of period t . y
t
is a vector of forcing variables, | e (0,1) a

constant discount factor while t represents the objective function. I further assume that

t
(
?
) is twice continuously differentiable, strictly increasing, and strictly concave with

respect to u
t
.
E
(? O
0
) denotes the mathematical expectations operator, conditional on


the information set at the beginning of period 0, O
0
. At any point in time, t , the

information set is given by O
t =
{u
t
,u
t
÷
1
,...; x
t
, x
t
÷
1
,...; y
t
, y
t
÷
1
,...}
4
. Finally, y
t
is assumed


to be generated by a first-order process

y
t
=
q
( y
t
÷
1
, z
t
) , (1.2.4)





4





The elements of the information set point to the fact that variables become known at the
beginning of the period. During the chapter this assumption may change to allow for an
easier setup of the problem.



3
where the elements of z
t
are distributed independently and identically across t and are

drawn from a distribution with a finite number of parameters.

The preceding generic optimization problem covers various examples of models

in economics, including the life-cycle model of consumption under uncertainty with or

without liquidity constraints, stochastic growth model with or without irreversible

investment and certain versions of asset pricing models. The present specification does

not cover models that have more than one control variable. However, some of the

techniques presented in this chapter could be used to solve such models.

If the underlying assumptions are such that the Bellman principle holds, one can

use the Bellman equation method to solve the dynamic programming problem. The

Bellman equation for the problem described by (1.2.1) - (1.2.2) is given by

V
( x
t
, y
t
) = muax
t
(u
t
) + | E ?
V
(
h
( x
t
,u
t
+
1
, y
t
+
1
), y
t
+
1
) | O
t
?
t
{
?
?
}
(1.2.5)

where
V
(
?
) is the value function. An alternative way to solve the model is to use the

Euler equation method. If u can be expressed as a function of x , i.e. u
t
=
g
(x
t
, x
t
÷
1
, y
t
) ,

the Euler equation for period t for the same problem is:

t'
u
?
g
( x
t
, x
t
÷
1
, y
t
)? g
'
x ( x
t
, x
t
÷
1
, y
t
) +
? ?
t

+ | E t'
u
?
g
( x
t
, x
t
+
1
, y
t
+
1
)? g
'
x
t
( x
t
, x
t
+
1
, y
t
+
1
) | O
t
= 0
{
? ?
}
(1.2.6)
So far, it has been assumed that the inequality constraint was not binding. If one

considers the possibility of constraint (1.2.3) being binding, then one must employ either

the Kuhn-Tucker method or the penalty function method. In the case of the former, the

Euler equation for period t becomes:

t'
u
?
g
( x
t
, x
t
÷
1
, y
t
)? g
'
x ( x
t
, x
t
÷
1
, y
t
) + ·
t
f
x
'
( x
t
, x
t
÷
1
) + ·
t
+
1
f
x
'
( x
t
, x
t
+
1
) +
? ?
t t t
(1.2.7)
+ | E t ?
g
( x
t
, x
t
+
1
, y
t
+
1
)? g
{
'
u
?



4
?
'
x
t
( x
t
, x
t
+
1
, y
t
+
1
) | O
t
} = 0
where ·
t
and ·
t
+
1
are Lagrange multipliers. The additional Kuhn-Tucker conditions are

given by:

·
t
> 0,
f
( x
t
, x
t
÷
1
) > 0, ·
t f
( x
t
, x
t
÷
1
) = 0 (1.2.8)
Alternatively, one can use penalty methods to account for the inequality constraint. One

approach is to modify the objective function by introducing a penalty term
5
. Then the

new objective function becomes:

E
e
9 |
t

t
t
(u
t
) + µ min
f
(x
t
, x
t
÷
1
), 0
?
O
0
÷
c

3

©
t =
0
e _
( )
?
?
?
?
where · is the penalty parameter. Consequently, the Bellman equation is given by:


V
( x
t
, y
t
) = muax
t
(u
t
) + · min (
f
( x
t
, x
t
÷
1
),
0
)3 + | E ?
V
(
h
( x
t
,u
t
+
1
, y
t
+
1
), y
t
+
1
) | O
t
?
t
{
?
?
}
(1.2.9)

Let u
*
t =
d
( x
t
, y
t
) denote the solution of the problem. When an analytical solution

for
d
(
?
) can not be computed, numerical techniques need to be used. Three main


approaches have been used in the literature to solve the problem (1.2.1) - (1.2.4) and to

obtain an approximation of the solution. First approach consists of modifying the

specification of the problem (1.2.1) - (1.2.2) so that it becomes easier to solve, as is the

case with the linear quadratic approximation
6
. Second approach is to employ methods

that seek to approximate the value and policy functions by using the Bellman equation
7
.




5

6





This approach is used by McGrattan (1990).

This approach has been used, among others, by Christiano (1990b) and McGrattan
(1990).

7

Examples of this approach are: Christiano (1990a), Rust (1997), Santos and Vigo
(1998), Tauchen (1990).



5
Finally, the third approach focuses on approximating certain terms appearing in the Euler

equation such as decision functions or expectations
8
.

These approaches have shaped the design of numerical algorithms used in solving

dynamic non-linear rational expectation models. In the next few sections, I will present

several of the numerical methods employed by researchers in their attempt to solve

functional equations such as the Euler and Bellman equations (1.2.5) - (1.2.9) presented

above.





I.3. Using Certainty Equivalence; The Extended Path Method

Certainty equivalence has been used especially for its convenience since it may

allow researchers to compute an analytical solution for their models. It has also been used

to compute the steady state of a model as a prerequisite for applying some linearization or

log-linearization around its equilibrium state
9
or to provide a starting point for more

complex algorithms
10
. One methodology that received a lot of attention in the literature is

the extended path method developed by Fair and Taylor (1983). Solving a model such as

(1.2.1) - (1.2.3) usually leads to a functional equation such as a Bellman or an Euler

equation.



8



Examples of this approach are Binder et al. (2000), Christiano and Fisher (2000), Judd
(1992), Marcet (1994), Mc-Grattan (1996).

9

This is the case in the linear quadratic approach where the law of motion is linearized
and the objective function is replaced by a quadratic approximation around the
deterministic steady state.

10

Certainty equivalence has also been used to provide starting values or temporary values
in algorithms used to solve models leading to nonlinear stochastic equations as in early
work by Chow (1973, 1976), Bitros and Kelejian (1976) and Prucha and Nadiri (1984).



6
Let

F x
t
, x
t
÷
1
,u
t
,u
t
÷
1
, y
t
, y
t
÷
1
, E
t
t
'
?
h
( x
t
, x
t
+
1
, y
t
+
1
)? h '
x
t
, E
t
t ?
h
( x
t
, x
t
+
1
, y
t
+
1
)? = 0
(
{
? ?
}
? ?
)

(1.3.1)

denote such a functional equation for period t . As before, x
t
is the state variable, u
t
is

the control variable, y
t
is a vector of forcing variables,
t
(
?
) is the objective function, t'

is the derivative of t with respect to the control variable, and E
t
is the conditional

expectations operator based on information available through period t . F is a function

that may be nonlinear in variables and expectations. For numerous models if the

expectations terms appearing in F were known, (1.3.1) could be easily solved. Since that

is not the case, the approach of the extended path method is to first set current and future

values of the forcing variables to their expected values. This is equivalent to assuming

that all future values of z
t
in equation (1.2.4) are zero. Then equation (1.3.1) becomes:

F x
t
, x
t
÷
1
,u
t
,u
t
÷
1
, y
t
, y
t
÷
1
,t
'
?
h
( x
t
, x
t
+
1
, E
t
y
t
+
1
)? h '
x
t
,t ?
h
( x
t
, x
t
+
1
, E
t
y
t
+
1
)? ,... = 0
(
? ? ? ?
) (1.3.2)
Then, the idea is to expand the horizon and iterate over solution paths. Let us consider an

example to see how this method can be applied.


I.3.1. Example
11


Consider the following problem where the social planner or a representative agent

maximizes an objective function

max E
e
9 |
tt
(u
t
) O
0
÷ •
u
t
c
©t=0
?
?
(1.3.3)

11

The application of the extended path method in this example draws to some extent on
the model presented in Gagnon (1990).



7
subject to

x
t
=
h
(x
t
÷
1
,u
t
, y
t
) (1.3.4)

where y
t
is a Gaussian AR (
1
) process with the law of motion y
t
= µ y
t
÷
1
+ z
t
where z
t

is i.i.d. N 0,o
2
. It is further assumed that u can be expressed as a function of x , i.e.
( )

u
t
=
g
(x
t
, x
t
÷
1
, y
t
) . Then the Euler equation for period t is:

0 = t ' ?
g
( x
t
, x
t
÷
1
, y
t
)? ? g '
x
t
( x
t
, x
t
÷
1
, y
t
)
? ?
+ | E t ' ?
g
( x
t
+
1
, x
t
, y
t
+
1
)? ? g '
x
t
( x
t
+
1
, x
t
, y
t
+
1
) O
t
{
? ?
}
(1.3.5)
If the expectation term were known in equation (1.3.5), it would be easy to find a

solution. The idea of the extended path method is to expand the horizon and then iterate

over solution paths. As in Fair and Taylor (1983), I consider the horizon t,...,t + k +1 and

assume that x
t
÷
1
and y
t
÷
1
are given and that z
t
+
s
= 0 for s = 1,..., k +1 . Following is an

algorithm that would implement the extended path methodology. The first step is to

choose initial values for x
t
+
s
and y
t
+
s
for s = 1,..., k +1 and denote them by ˆ
t
+
s
and x



t
+
s
. Then, for period t , the Euler equation becomes:

0 = t ' ?
g
( x
t
, x
t
÷
1
, y
t
)? ? g '
x
t
( x
t
, x
t
÷
1
, y
t
)
? ?
(1.3.6)
+ |t ' ?
g

t
+
1
, x
t
, ˆ
t
+
1
)? ? g '
x
t
( x
t
+
1
, x
t
, ˆ
t
+
1
)
?x y?
ˆ y
Similarly, for period t + s , the Euler equation is given by:

0 = t ' ?
g
( x
t
+
s
, x
t
+s÷
1
, y
t
+
s
)? ? g '
x
t+
s
( x
t
+
s
, x
t
+s÷
1
, y
t
+
s
)
? ?
(1.3.7)
+ |t ' ?
g

t
+s+
1
, x
t
+
s
, ˆ
t
+s+
1
)? ? g '
x
t+
s

t
+
1
, x
t
+
s
, ˆ
t
+s+
1
)
In addition,
?x
y
?
x y

y
t
+
s
= µ y
t
+s÷
1
+ z
t
+s (1.3.8)

u
t
+
s
=
g
(x
t
+
s
, x
t
+s÷
1
, y
t
+
s
) (1.3.9)




8
Therefore, for period t + s , equations (1.3.7) - (1.3.9) define a system where x
t
+s÷
1
, y
t
+s÷
1
,


ˆ
t
+s+
1
, ˆ
t
+s+
1
are known so one can determine the unknowns x
t
+
s
, y
t
+
s
and u
t
+
s
. Let x
t
j+
s
,
x y

y
t
j+
s
and u
t
j+
s
denote the solutions of the system for s = 0,..., k +1 , where j represents the
iteration for a fixed horizon, in this case t,...,t + k +1. If the solutions x
t
j+s
{}

k +1

k +1
s =0
, y
t
j+s
{}

k +1
s =0
and u
t
j+s
{}

s =0
obtained in iteration j are not satisfactory then proceed with the next

iteration
where

t
j++s
1
}
s
=
1
=
{x
t
j+
s
}
s
=
1
,

t
j++s
1
}
s
=
1
=
{y
t
j+
s
}
s
=
1
. Notice that the horizon remains
k +1 k +1 k +1 k +1
x y

the same for iteration j +1. The iterations will continue until a satisfactory solution is

obtained. At this point, the methodology calls for the extension of the horizon without

modifying the starting period. Fair and Taylor extend the horizon by a number of periods

that is limited to the number of endogenous variables. This is in essence an ad-hoc rule.

In the present example, the horizon is extended by 2 periods, that is, t,...,t + k + 3. The

same steps are followed for the new horizon with the exception of the end criterion,

which should consist of a comparison between the last obtained solution, using the

t,...,t + k + 3 horizon, and the solution provided using the previous horizon, t,...,t + k +1.

The expansion of the horizon continues until a satisfactory solution is obtained. At that

point, the procedure will start over with a new starting period and a new horizon. In our

example the next starting period should be t +1 and the initial horizon t +1,...,t + k + 2 .

One of the less mentioned caveats of this method is that no general convergence

proofs for the algorithm are available. In addition, the method relies on the certainty

equivalence assumption even though the model is nonlinear. Since expectations of

functions are treated as functions of the expectations in future periods in equation (1.3.2),




9
the solution is only approximate unless function F is linear. This assumption is similar

to the one used in the case of linear-quadratic approximation to rational expectations

models that has been proposed, for example, by Kydland and Prescott (1982).

In the spirit of Fair and Taylor, Fuhrer and Bleakley (1996), following an

algorithm from an unpublished paper by Anderson and Moore (1986), sketch a

methodology for finding the solution for nonlinear dynamic rational expectations models.



I.3.2. Notes on Certainty Equivalence Methods

All the methods that use certainty equivalence either as a main step or as a

preliminary step in finding a solution, incur an approximation error due to the assumption

of perfect foresight. The magnitude of this error depends on the degree of nonlinearity of

the model being solved. Fair (2003), while acknowledging its limitations, argues that the

use of certainty equivalence may provide good approximations for many

macroeconometric models.

In the case of the extended path algorithm, the error propagates through each level

of iteration and therefore it forces the use of strong convergence criteria. Due to this fact,

the extended path algorithm tends to be computationally intensive. Other methodologies

that only use certainty equivalence as a preliminary step as in the case of linearization

methods or linear quadratic approaches are not subject to the same computational burden.

In conclusion, while there are cases where certainty equivalence may be used to

obtain good approximations, one needs to be careful when using this methodology since

there are no guarantees when it comes to accuracy.







10
I.4.Local Approximation and Perturbation Methods


Economic modeling problems have used a variety of approximation methods in

the absence of a closed form solution. One of the most used approximation methods,

coming in different flavors, is the local approximation. In particular, the first order

approximation has been extensively used in economic modeling. Formally, a function

a(x) is a first order approximation of b(x) around x
0
if a(x
0
) = b(x
0
) and the

derivatives at x
0
are the same, a '(x
0
) = b '(x
0
) . In certain instances, first order

approximations may not be enough so one would have to compute higher order

approximations. Perturbation methods often use high order local approximation and

therefore rely heavily on two very well own theorems, Taylor's theorem and implicit

function theorem.



I.4.1. Regular and General Perturbation Methods

Perturbation methods are formally addressed by Judd (1998). In this section,

following Judd's framework, I try to highlight the basic idea of regular perturbation

methods. I start by assuming that the Euler equation of the model under consideration is

given by:

F
( u,
c
) = 0 (1.4.1)

where
u
(
c
) is the policy I want to solve for and c is a parameter. Further on, I assume

that a solution to (1.4.1) exists, that F is differentiable,
u
(
c
) is a smooth function and

u
(
0
) can be easily determined or is known. Differentiating equation (1.4.1) leads to:

F
u
(
u
(
c
),
c
)u
'
(
c
) + F
c
(
u
(
c
),
c
) = 0 (1.4.2)




11
Making c = 0 in equation (1.4.2) allows one to compute u
'
(
0
) :

F
c
(
u
(
0
),
0
)
u
'
(
0
) = ÷
F
u
(
u
(
0
),
0
)
(1.4.3)

The necessary condition for the computation of u
'
(
0
) is that F
u
(
u
(
0
),
0
) = 0 . Assuming

that indeed F
u
(
u
(
0
),
0
) = 0 , it means that now u
'
(
0
) is known and one can compute the

first order Taylor expansion, of
u
(
c
) around c = 0 :

F
c
(
u
(
0
),
0
)
u
(
c
) ?
u
(

0
) ÷
F
u
(
u
(
0
),
0
)
c
(1.4.4)

This is a linear approximation of
u
(
c
) around c = 0 . In order to be able to compute

higher order approximations of
u
(
c
) one needs to know at least the value of u
''
(
0
) . That


can be found by differentiating (1.4.2):


u
''
(
0
) = ÷

F
uu
(
u
(
0
),
0
)(u
'
(
0
))2 + 2F
u
c
(
u
(
0
),
0
)u
'
(
0
) + F
cc
(
u
(
0
),
0
)
F
u
(
u
(
0
),
0
)



(1.4.5)

The necessary condition for the computation of u
''
(
0
) is, once again, that

F
u
(
u
(
0
),
0
) = 0 . In addition, second order derivatives shall exist. Then the second order

approximation of
u
(
c
) around c = 0 is given by:


u
(
c
) ?
u
(

0
) ÷

F
c
(
u
(
0
),

0
)





1
c
2
F
uu
(
u
(
0
),
0
)(u
'
(
0
)) + 2F
u
c
(
u
(
0
),
0
)u
'
(
0
) + F
cc
(
u
(
0
),
0
) 2
F
u
(
u
(
0
),
0
)
2
F
u
(
u
(
0
),
0
)
In general, higher order approximations of
u
(
c
) can be computed if higher

derivatives of
F
(u,
c
) with respect with u exist and if F
u
(
u
(
0
),
0
) = 0 . The advantage


of regular perturbation methods based on an implicit function formulation is that one






12
directly computes the Taylor expansions in terms of whatever variables one wants to use,

and that expansion is the best possible asymptotically.



I.4.2. Example

Consider the following optimization problem

max E
e
9 |
tt
(u
t
) | O
0
÷ •

subject to
u
t
c
©t=0

x
t
=
h
(x
t
÷
1
,u
t
÷
1
, y
t
)
?
?
(1.4.6)



(1.4.7)

with y
t
= y
t
÷
1
+ c z
t
, where u
t
is the control variable, x
t
is the state variable, c is a scalar


parameter and z
t
is a stochastic variable drawn from a distribution with zero mean and

unit variance. x
t
, u
t
, c and z
t
are all scalars. The Bellman equation is given by:

V (x
t
) = muax
t
(u
t
) + | E ?V (
h
( x
t
,u
t
+
1
,c z
t
+
1
)) | O
t
?
t
{
?
?
}
(1.4.8)
Then the first order condition is:

0 = t
u
(u
t
) + | E ?V
'
(
h
( x
t
,u
t
,c z
t
+
1
)) h
u
( x
t
,u
t
,c z
t
+
1
)?
?
Differentiating the Bellman equation with respect to x
t
, one obtains:
?
(1.4.9)

V
'
(x
t
) = | E ?V
'
(
h
( x
t
,u
t
,c z
t
+
1
)) h
x
( x
t
,u
t
,c z
t
+
1
)?
? ?
(1.4.10)

Let the control law
U
( x,
c
) be the solution of this problem. Then the above equation


becomes:

V
'
(x) = | E
?
V
'
h ( x,
U
( x,
c
),c
z
) h
x
?
?
( )
?

The idea is to first solve for steady state in the deterministic case, which here is

equivalent to c = 0 , and then find a Taylor expansion for
U
( x,
c
) around c = 0 .





13

Assuming that there exists a steady state defined by (x
*
,u
*
) such that x
*
= h x
*
,u
*
, one
( )

can use the following system to obtain steady state solutions:

x
*
= h x
*
, u
*
( )
(1.4.11)

0 = t
u
u
*
+ |V ' h x
*
,u
*
h
u
x
*
,u
*
() (( )) ( ) (1.4.12)

V ' x
*
= |V ' h x
*
,u
*
h
x
x
*
,u
*
() (( )) ( ) (1.4.13)

V x
*
= t u
*
+ | V x
*
() () () (1.4.14)
Further assuming local uniqueness and stability for the steady state, equations (1.4.11)-
(1.4.14) provide the solutions for the four steady state quantities x
*
, u
*
, V x
*
, and ()
V ' x
*
. Given that the time subscript for all variables is the same, I drop it for the ()

moment. Going back to equations (1.4.9) - (1.4.10), in the deterministic case, that is, for

c = 0 , one obtains:

0 = t
u
(
U
(
x
)) + |V ' ?h ( x,
U
(
x
))? h
u
( x,
U
(
x
))
? ?
(1.4.15)

V
'
(x) = |V
'
?
h
( x,
U
(
x
))? h
x
( x,
U
(
x
))
? ?
(1.4.16)
Differentiating (1.4.15) and (1.4.16) with respect to x yields

0 = t
uu
U
'
x + |V
"
(
h
)(h
x
+ h
u
U
'
x ) h
u
+ |V
'
(
h
)(h
ux
+ h
uu
U
'
x ) (1.4.17)

V " = |V
"
(
h
)(h
x
+ h
u
U
'
x ) h
x
+ |V
'
(
h
)(h
xx
+ h
xu
U
'
x ) (1.4.18)
Therefore, the steady state version of the system (1.4.17) - (1.4.18) is given by:

0 = t
uu
x
*
,u
*
U
'
x x
*
+ |V " x
*
h
x
x
*
,u
*
( ) () () ( ) e
(1.4.19)
+h
u
x ,u U
x
x
?
h
u x ,u + |V '
x
e
h
ux x ,u + h
uu
x ,u U
x
x
?
(
* *
) () (
'
*
?
* *
) () (
*


* *
) (
* *
) ()

V " x
*
= |V " x
*
h
x
x
*
,u
*
+h
u
x
*
,u
*
U
'
x x
*
? h
x
x
*
,u
*
'
*
?
() () ( ) ( ) () ( )
e ?
(1.4.20)
+|V ' x
*
h
xx
x
*
,u
*
+ h
xu
x
*
,u
*
U
'
x x
*
?
()
e
( ) ( ) ()
?

These equations define a quadratic system for the unknowns V "(x ) and
U
(
x
) .
* ' *
x



14
Going back to the stochastic case, the first order condition with respect to u is given by:

0 = t
u
(
U
( x,
c
)) + | E V '
h
( x,
U
( x,
c
),c z
t
+
1
) h
u
( x,
U
( x,
c
),c z
t
+
1
) | O
t
(1.4.21)
{( ) }
Taking the derivative of the Bellman equation with respect to x yields:

V
'
(x) = | E V
'
h (x,
U
(x,
c
),c z
t
+
1
) h
x
(x,
U
(x,
c
),c z
t
+
1
) | O
t
{( ) } (1.4.22)
In order to obtain a local approximation of the control law around c = 0 , its derivatives

with respect to c must exist and be known. To find these values one needs to

differentiate equations (1.4.21) - (1.4.22) with respect to c , make c = 0 and solve the

resulting system for the values of the derivatives of U with respect to c when c = 0 ,

i.e., for U
c
'
( x
*
,
0
) . Once that value is found, one can compute a Taylor expansion for

U
( x,
c
)
around
( x
*
,
0
) .


If the model requires the addition of an inequality constraint such as (1.2.3) which

could be the representation of a liquidity constraint or a gross investment constraint, the

Bellman equation (1.4.8) becomes:

V (x
t
) = muax
t
(u
t
) + · min (
f
( x
t
, x
t
÷
1
),
0
)3 + | E ?
V
(
h
( x
t
,u
t
,c z
t
)) | O
t
? (1.4.23)
t
{
?
?}
where · is the penalty parameter.




I.4.3. Flavors of Perturbation Methods

Economic modeling problems have used a variety of approximation methods that

may be characterized as perturbation methods. The most common use of perturbation

methods is the method of linearization around the steady state. Such linearization

provides a description on how a dynamical system evolves near its steady state. It has

often been used to compute the reaction of a system to shocks. While the first-order




15
perturbation method exactly corresponds to the solution obtained by standard

linearization of first-order conditions, one well known drawback of such a solution,

especially in the case of asset pricing models, is that it does not take advantage of any

piece of information contained in the distribution of the shocks. Collard and Juillard

(2001) use higher order perturbation methods and apply a fixed-point algorithm, which

they call "bias reduction procedure", to capture the fact that the policy function depends

on the variance of the underlying shocks. Similarly, Schmitt-Grohé and Uribe (2004)

derive a second-order approximation to the policy function of a general class of dynamic,

discrete-time, rational expectations models using a perturbation method that incorporates

a scale parameter for the standard deviations of the exogenous shocks as an argument of

the policy function.





I.4.4. Alternative Local Approximation Methods

There are also certain local approximations techniques used in the literature that

may look like perturbation methods when in fact they are not. One frequently used

approach is to find the deterministic steady state and then to replace the original nonlinear

problem with a linear-quadratic problem that is similar to the original problem. The

linear-quadratic problem can then be solved using standard methods. This method differs

from the perturbation method in that the idea here is to replace the nonlinear problem

with a linear-quadratic problem, whereas the perturbation approach focuses on computing

derivatives of the nonlinear problem. Let me consider again the problem defined by

equations (1.2.1) - (1.2.2). The idea is to approximate the original problem by a





16
combination of a quadratic objective and a linear constraint, which would take the

following form:

max E ?
¿
( ? · |
t
Q +Wu + Ru
2
| O ?
u
t
?t =0
t
t
) 0
?

?
(1.4.24)

s.t. x
t
= Ax
t
÷
1
+ Bu
t
+ Cy
t
+ D (1.4.25)
where Q, R, W , A, B, C and D are scalars.


In order to obtain the new specification, the first step is to compute the steady

state for the deterministic problem (which means z
t
= 0 in equation (1.2.4)). Therefore,

one has to formulate the Lagrangian:

·
L
=
¿ |t {
t
(u
t
) ÷ ì
t
?x
t
÷
h
( x
t
÷
1
,u
t
, y
0
)
?
}
t =0
? ?
(1.4.26)

The first order conditions for (1.4.26) is a system of 3 equations with unknowns

x,u and ì . The solution of the system represents the steady
state,
( x
*
,u
*
, ì
*
) . The next

step is to take the second order Taylor expansion for
t
(u
t
) and first order Taylor

expansion for
h
(x
t
÷
1
,u
t
, y
t
)
around
( x
*
,u
*
, y
0
) . Thus,

t
(
u
) =
t
(
u
) + t
'
(
u
) (u ÷
u
) + t

"
(
u
) (


*2
t * * t * *
u ÷
u
) t
2
(1.4.27)

h
( x
t
÷
1
,u
t
, y
t
) =
h
( x
*
,u
*
, y
0
) + h
'
x ( x
*
,u
*
, y
0
)( x
t
÷
1
÷ x
*
) +
(1.4.28)
+ h
u
'
( x
*
,u
*
, y
0
)(u
t
÷ u
*
) + h
'
y ( x
*
,u
*
, y
0
)( y
t
÷ y
0
)
These expansions allow one to identify the parameters Q, R, W , A, B, C and D .


Specifically,

*2
Q =
t
( u
*
) ÷ t
'
( u
*
) u
*
+ t
"
(
u
*
)
u
2
t
"
(
u
*
)


(1.4.29)
W = t
'
( u
*
) ÷ t
"
( u
*
)
u
*




17
R=
2

A = h
'
x ( x
*
,u
*
, y
0
) B = h
u
'
( x
*
,u
*
, y
0
) C = h
'
y ( x
*
,u
*
, y
0
)
(1.4.30)
D =
h
( x
*
,u
*
, y
0
) ÷ h
'
x ( x
*
,u
*
, y
0
) x
*
÷ h
u
'
( x
*
,u
*
, y
0
)u
*
÷ h
'
y ( x
*
,u
*
, y
0
) y
0

Once the parameters have been identified, the problem can be written in the form

described by (1.4.24) and (1.4.25) which has a quadratic objective function and linear

constraints
12
.

If the model needs to account for an additional inequality constraint such as

(1.2.3), the Lagrangian (1.4.26) becomes

L
=
¿ |t {
t
(u
t
) ÷ ì
t
?x
t
÷
h
( x
t
÷
1
,u
t
, y
0
)? + ·
t f
( x
t
, x
t
÷
1
)
} ·
t =0
? ?
(1.4.31)
and the additional Kuhn-Tucker conditions have to be taken into account.





I.4.5. Notes on Local Approximation Methods

The perturbation methods provide a good alternative for dealing with the major

drawback of the method of linearization around steady state, that is, its lack of accuracy

in the case of high volatility of shocks or high curvature of the objective function. While

the first order perturbation method coincides with the standard linearization, the higher

order perturbation methods offer a much higher accuracy
13
.

Some of the local approximation implementations such as the linear-quadratic

method
14
do fairly well when it comes to modeling movements of quantities, but not as


12



There are some other variations of this approach used in the literature such as
Christiano (1990b).

13

See Collard and Juillard (2001) for a study on the accuracy of perturbation methods in
the case of an asset-pricing model.

14

Dotsey and Mao (1992), Christiano (1990b) and McGrattan (1990) have documented
the quality of some implementations of the macroeconomic linear-quadratic approach.


18
well with asset prices. The reason behind this result is that approximation of quantity

movements depends only on linear-quadratic terms whereas asset-pricing movements are

more likely to involve higher-order terms.




I.5. Discrete State-Space Methods
15



These methods can be applied in several situations. In the case where the state

space of the model is given by a finite set of discrete points these methods may provide

an "exact" solution
16
. In addition, these methods are frequently applied by discretizing an

otherwise continuous state space. The use of discrete state-space methods in models with

a continuous state space is based on the result
17
that the fixed point of a discretized

dynamic programming problem may converge point wise to its continuous equivalent
18
.

The discrete state-space methods sometimes prove to be a useful alternative to

linearization and log-linear approximations to the first order necessary conditions,

especially for certain model specifications.





15

16





This section draws heavily on Burnside (1999) and on Tauchen and Hussey (1991)

This may be the case in models without endogenous state variables, especially when
there is only one state variable that follows a simple finite state process. Examples are
Mehra and Prescott (1985) and Cecchetti, Lam and Mark (1993).

17

As documented in Burnside (1999), Atkinson (1976) and Baker (1977) present
convergence results related to the use of discrete state spaces to solve integral equations.
Results concerning pointwise and absolute convergence of solutions to asset pricing
models obtained using discrete state spaces are presented in Tauchen and Hussey (1991)
and Burnside (1993).

18

The procedure employed by discrete state-space methods in models with a continuous
state space is sometimes referred to as 'brute force discretization'.



19
I.5.1. Example. Discrete State-Space Approximation Using Value-Function Iteration

As before, I consider the following maximization problem:

max E
?
¿ |
tt
(u
t
) | O
0
? ·

subject to
u
t
?
?t =0


x
t
+
1
=
h
(x
t
,u
t
, y
t
)
?
?
(1.5.1)




(1.5.2)

where y
t
is a realization from an n -state Markov chain, u
t
is the control variable and x
t


is the state variable. Let Y
=
{Y
1
, Y
2
,..., Y
n
} be the set of all possible realizations for y
t
.

In order to be able to apply the above mentioned methodology one has to establish a grid

for the state variable. Let the ordered set X
=
{X
1
, X
2
,..., X
k
} be the grid for x
t
.


Assuming that the control variable u
t
can be explicitly determined from equation (1.5.2)

as a function of x
t
, x
t
+
1
and y
t
, then the dynamic programming problem can be

expressed as:

V (x
t
, y
t
) = max
t
( x
t
, x
t
+
1
, y
t
) + | E ?
V
( x
t
+
1
, y
t
+
1
) | O
t
?
x eXt+1
{
?
?
}
(1.5.3)

Let H(x
t
, y
t
) be the Cartesian product of Y and X , that is, the set of all possible

m = n ? k pairs (x
i
, y
j
) . Formally, H(x
t
, y
t
) = (x
i
, y
j
) | x
i
ŒX ¸ ÷
k
and y
j
ŒY ¸ ÷
n
.
{ }

Hence H(x
t
, y
t
) ¸ ÷
k
· ÷
n
= ÷
m
. If equation (1.5.3) is discretized using the grid given

by H(x
t
, y
t
) one can think of function
V
(
?
) as a point in ÷
m
. Similarly, the expression

t
(x
t
, x
t
+
1
, y
t
) + |
E
(
V
(x
t
+
1
, y
t
+
1
) | O
t
) can be thought of as a mapping M from ÷
m
into

÷
m
. In this context
V
(
?
) is a fixed point for M , that is, V =
M
(
V
) . One of the methods


commonly used to solve for the fixed point in these situations is the value function

iteration.


20
In order to solve the maximization problem one can use various algorithms. The

algorithm I am going to present follows, to some degree, Christiano (1990a). Let

S
j
(X
p
, Y
q
) be the value of x
t
+
1
that maximizes
M
(V
j
) for given values of x
t
and y
t
,

( x
t
, y
t
)
=
(X
p
,Y
q
) c H . Formally,

S
t
j+
1
(X
p
, Y
q
) = arg max
t
(X
p
, x
t
+
1
, Y
q
) + | E ?V
j
( x
t
+
1
, y
t
+
1
) | O
t
?
x
t
+
1
eX
{
?
?
}
(1.5.4)

where j represents the iteration. The idea is to go through all the possible values for x
t
+
1
,

that is, the set X , and find the value that maximizes the right hand side of (1.5.4). That

will become the value assigned to S
j
(X
p
, Y
q
) . Then the procedure will be repeated for a


different value of the pair x
t
and y
t
belonging to set H(x
t
, y
t
) and, finally, a global

maximum will be found. The exposition of the algorithm so far implies an exhaustive

search of the grid. The speed of the algorithm can be improved by choosing a starting

point for the search in every iteration and continue the search only until the first decrease

in the value function is encountered
19
. The decision rule for u
t
can then be derived by

substituting S
t
+
1
for x
t
+
1
in the law of motion.



I.5.2. Fredholm Equations and Numerical Quadratures

Let me consider the model specified by (1.2.1) - (1.2.2). Then the Bellman

equation is given by:

V
( x
t
, y
t
) = muax
t
(u
t
) + | E ?
V
( x
t
+
1
, y
t
+
1
) | O
t
?
t
{
?
?
}
(1.5.5)



19



T his change in the algorithm, as presented by Christiano (1990a), is valid only when
the value function is globally concave.



21
If y
t
follows a process such as (1.2.4), one can rewrite the conditional expectation and

consequently the whole equation (1.5.5) as:

V
( x
t
, y
t
) = muax
t
(u
t
) +
|
}
V
( x
t
+
1
, y
t
+
1
)
q
( y
t
+
1
| y
t
) dy
t
+1{t
}


(1.5.6)
In the above equation, the term needing approximation is the integral

}
V
( x t +1

, y
t
+
1
)
q
( y
t
+
1
| y
t
) dy
t
+1

If V (x
t
+
1
, y
t
+
1
) is continuous in y
t
+
1
for every x , the integral can be replaced by an N-

point quadrature approximation. An N-point quadrature method is based on the notion

that one can find some points y
i
,
N
and some weights w
i
,
N
in order to obtain the following

approximation

N
¿
V
( x t +1
i=1
, y
i
.
N
) w
i
,
N
~
}
Y
V
( x
t
+
1
, y
t
+
1
)
q
( y
t
+
1
| y
t
) dy
t
+1
(1.5.7)

where the points y
i
,
N
eY ,i = 1,K, N , are chosen according to some rule, while the weight

given to each point, w
i
,
N
, relates to the density function
q
(
y
) in the neighborhood of


those points. In general, a quadrature method requires a rule for choosing the points, y
i
,
N
,


and a rule for choosing the weights, w
i
,
N
. The abscissa y
i
,
N
and weights w
i
,
N
depend only

on the density
q
(
y
) , and not directly on the function V .


Quadrature methods differ in their choice of nodes and weights. Possible choices

are Newton-Cotes, Gauss, Gauss-Legendre and Gauss-Hermite approximations. For a

classical N-point Gauss rule along the real line, the abscissa y
i
,
N
and weights w
i
,
N
are

determined by forcing the rule to be exact for all polynomials of degree less than or equal

to 2N ÷1.





22
For most rational expectation models, integral equations are a very common

occurrence both in Bellman equations such as (1.5.6), as well as in Euler equations. One

of the most common forms of integral equations mentioned in the literature is the

Fredholm equation
20
. Therefore, in this section I will present an algorithm similar to the

one used by Tauchen and Hussey (1991) for solving such equation.

Now let me assume for a moment that the Euler equation of the model is given by

a Fredholm equation of the second kind:

v( y
t
)
=
v¢ ( y
t
+
1
, y
t
)v( y
t
+
1
)q( y
t
+
1
y
t
)dy
t
+
1
+¸ ( y
t
) (1.5.8)
where y
t
is an n-dimensional vector of variables, E
t
is the conditional expectations

operator based on information available through period t , and
¢
( y
t
, y
t
+
1
) and
¸
( y
t
) are

functions of y
t
and y
t
+
1
that depend upon the specific structure of the economic model,

and where
v
( y
t
) is the solution function of the model. The
process
{y
t
} is characterized


by a conditional density, q( y
t
+
1
y
t
) .

Following Tauchen and Hussey (1991), let the T[?] operator define the integral

term in equation (1.5.8). Then (1.5.8) can be written as:

v = T [v ] + ¸ (1.5.9)

Under regularity conditions, the operator [I ÷ T ]
÷
1
exists, where I denotes the identity

operator, and the exact solution is:

v = [I ÷ T ]
÷
1¸ (1.5.10)





20





One example where this form of integral equation appears is a version of the asset
pricing model. See Tauchen and Hussey (1991) and Burnside (1999) for more details.



23
An approximate solution is obtained using T
N
in place of T , where T
N
is an

approximation of T using quadrature methods for large N . Then [I ÷ T
N
] can be

inverted.

v
N
= [I ÷ T
N
]
÷
1¸ (1.5.11)

In some cases, the function ¸ is of the form ¸ = T[¸
0
] and then the approximate solution

is taken as [I ÷ T
N
]
÷
1
T
N

0
] .




I.5.3. Example. Using Quadrature Approximations

This is an example of discrete state-space approximation using quadrature

approximations and value-function iterations. I consider a similar model to the one

described in section I.5.1 with the difference that y
t
is a Gaussian
AR
(
1
) process as

opposed to a Markov chain. Again, the representative agent solves the following

optimization problem

max
E
?
¿ |
tt
(u
t
) O
0
? ·

subject to
u
t
?
?t =0



x
t
+
1
=
h
(x
t
,u
t
, y
t
)
?
?
(1.5.12)




(1.5.13)

where y
t
is a Gaussian
AR
(
1
) process with the law of motion y
t
= µ y
t
÷
1
+ z
t
where z
t

is i.i.d. N 0,o
2
. I assume that u
t
can be expressed as a function of x , i.e.
( )

u
t
=
g
(x
t
, x
t
+
1
, y
t
) . Then the Bellman equation for the dynamic programming is given by

V
( x
t
, y
t
) = mxax
t
(
g
( x
t
, x
t
+
1
, y
t
)) + |
E
{
V
( x
t
+
1
, y
t
+
1
) O
t
+
1
} t+1


(1.5.14)





24
Writing the expectation term explicitly, equation (1.5.14) becomes:

V
( x
t
, y
t
) = mxax
t
(
g
( x
t
, x
t
+
1
, y
t
)) +
|
}
V
( x
t
+
1
, y
+
1
)
f
( y
t
+
1
y
t
) dy
t
+1t+1


(1.5.15)
where

y
t
+
1
= µ y
t
+ z
t
+1 (1.5.16)
To convert the dynamic programming problem in (1.5.15) to one involving

discrete state spaces one needs first to approximate the law of motion of y
t
using a

discrete state-space process. That is, redefine y
t
to be a process which lies in a set
Y = y
i
,N
{}


N


with y
i
,
N
= o a
i
,
N
, where

{
a
}

N


is the set of quadrature points
i=1
i,N
i=1

corresponding to an N-point rule for a standard normal distribution
21
. Let the probability

that y
t
+
1
= y
i
,
N
conditional on y
t
= y
j
,
N
be given by


f y
i
,
N
y
j
,
N
w
i
,N
p
ji
=
(
f y
i
,
N
0 (
) )
s
j
(1.5.17)
where




N
s
j
=
9



f y
i
,
N
y
j
,N (



)
w




(1.5.18)
i=1
f y
i
,
N
0 ( )
i,N

and {w
i
,
N
}N=
1
are the quadrature weights as described in section I.5.2.. With this i

approximation, the Bellman equation can be written as:


V
( x
t
, y
t
) =
mxax ?

?
t g x , x , y + |N
V
( x ,
y
)
p
?
t+
1
?

(
( t t+1
j
) ¿1t+1iji ?? )
i=
(1.5.19)

given y
t
= y
j
, j = 1,..., N .




21




This is in fact the approach used by Tauchen and Hussey (1991) and Burnside (1999),
among others.



25
The next step is to replace the state space by a discrete domain X from which the

solution is chosen. There is no universal recipe for choosing a discrete domain and

therefore it is usually done on a priori knowledge of possible values of the state

variable
22
. The maximization problem can now be solved by value function iteration as

presented in section I.5.1..





I.5.4. Notes on Discrete State-Space Methods

Discrete state-space methods tend to work well for models with a low number of

state variables. As the number of variables increases, this approach becomes numerically

intractable, suffering from what the literature usually refers to as the curse of

dimensionality. In addition, as pointed out in Baxter et al. (1990), when the method is

used to solve continuous models there are two sources of approximation error. One is due

to forcing a discrete grid on continuous state variables and second from using a discrete

approximation of the true distribution of the underlying shocks. There are also instances

where the use of discrete state-space methods is entirely inappropriate since the

discretization process transforms an infinite state space into a finite one and in the

process is changing the information structure. This may not be an issue in most models,

but it definitely has an impact in models with partially revealing rational expectations

equilibria
23
.






22

23






See Tauchen (1990) for an example.

See Judd (1998) pp. 578-581 for an example.



26
I.6. Projection Methods
24



As opposed to the previously presented numerical methods, the techniques that

are going to be presented in this section have a high degree of generality. Projection

methods appear to be applicable to solving a wide variety of economic problems. In fact,

projection methods can be described as general numerical methods that make use of

global approximation techniques
25
to solve equations involving unknown functions.

The idea is to replace the quantity that needs to be approximated by parameterized

functions with arbitrary coefficients that are to be determined later on
26
, or to represent

the approximate solution to the functional equation as a linear combination of known

basis functions whose coefficients need to be determined
27
. In either case, there are

coefficients to be computed in order to obtain the approximate solution. These

coefficients are found by minimizing some form of a residual function.

Further on, a step by step description of the general projection method is

presented, followed by a discussion of the parameterized expectations approach.







24







I borrow this terminology from Judd (1992, 1998). These methods are also called
weighted residual methods by some authors (for example Rust (1996), McGrattan (1999),
Binder et al. (2000)). In fact, one can argue that weighted residual methods are just a
subset of the projection methods with a given norm and inner product.

25

In some cases local approximations are used on subsets of the original domain and then
they are pieced together to give a global approximation. One such case is the finite
element method.

26

See Marcet and Marshall (1994a), Marcet and Lorenzoni (1999), Wright and Williams
(1982a, 1982b, 1984) and Miranda and Helmberger (1988)

27

See McGrattan (1999)





27
I.6.1. The Concept of Projection Methods

Suppose that the functional equation can be described by:

F (d ) = 0 (1.6.1)

where F is a continuous map, F : C
1
÷ C
2
with C
1
and C
2
complete normed function

spaces and d : D c ?
k
÷ ?
m
is the solution to the optimization problem. More generally,

d is a list of functions that enter in the equations that define the equilibrium of a model,

such as decision rules, value functions, and conditional expectations functions, while the

F operator expresses equilibrium conditions such as Euler equations or Bellman

equations.


I.6.1.1. Defining the Problem

The problem is to find d : D c ?
k
÷ ?
m
that satisfies equation (1.6.1). This

translates into finding an approximation
d
ˆ(x;u ) which depends on a finite-dimensional

vector of parameters u
=
|u
1
,u
2
,K,u
n
| such that F

( x;
u
) is as close as possible to
( )


zero.

I.6.1.1.1. Example
28


Consider the following finite horizon problem where the social planner or a

representative agent maximizes

E
e
9 |
tt
(u
t
) O
0
÷ T

subject to
c
©t=0
?
?
(1.6.2)




28

x
t
=
h
(x
t
÷
1
,u
t
, y
t
)


The example in section I.6.1 draws heavily on Binder et al. (2000)



28


(1.6.3)
with x
0
and x
T
given. y
t
is an
AR
(
1
) process with the law of motion

y
t
= µ y
t
÷
1
+ z
t
(1.6.4)

and z
t
are i.i.d. with z
t
~ N 0,o 2
y
. I assume that u can be expressed as a function of
( )
x , i.e. u
t
=
g
(x
t
÷
1
, x
t
, y
t
) . Then the Euler equation for period T ÷ 1 is given by

0 = t '
g
(x
T
÷
2
, x
T
÷
1
, y
T
÷
1
) ? g '
x

1
(x
T
÷
2
, x
T
÷
1
, y
T
÷
1
)
( )
(1.6.5)
+ | E t '
g
(x
T
÷
1
, x
T
, y
T
) ? g '
x

1
(x
T
÷
1
, x
T
, y
T
) O
T
÷1
{( ) }
Let the optimal decision rule for x
T
÷
1
be given by x
*
÷
1
= d
T
÷
1
(x
T
÷
2
, y
T
÷
1
) where T

d
(
?
) is a smooth function. The projection methodology consists of approximating
d
(
?
)

by d
ˆ
(?,
u
) , where u represents an unknown parameter matrix. The unknown parameters

are computed such that the Euler equation also holds for d
ˆ
(?,
u
) .

Further on in this section I present the necessary steps one needs to take when

applying the projection methods, drawing heavily on the formalization provided by Judd

(1998)
29
. As I mentioned above, the methodology consists of finding an approximation

dˆ(x;u ) such that F d
ˆ
( x;
u
) is as close as possible to zero. It becomes obvious that
( )

there are a few issues that need to be addressed: what form of approximation to choose

for dˆ(x;u ) ; does the operator F need to be approximated; what does one understand by,

or in other words, what is the formal representation of "as close as possible to zero".









29










Judd provides a five step check list for applying the projection methods.



29
I.6.1.2. Finding a Functional Form

The first step comes quite naturally from the need to address the question on how

to represent d (x;u ) . In general d
ˆ
is defined as a finite linear combination of basis

functions, ?
i
(x), i = 0,K, n :

n
dˆ(x;u ) = ?
0
(x)
+
¿u
i
?
i
(x) (1.6.6)
i=1

Therefore, the first step consists of choosing a basis over C
1
.

Functions ?
i
(x), i = 0,K, n are typically simple functions. Standard examples of

basis functions include simple polynomials (such as ?
0
(x) = 1, ?
i
(x) = x
i
), orthogonal

polynomials (for example, Chebyshev polynomials), and piecewise linear functions.

Choosing a basis is not a straightforward task. For example, ordinary polynomials are

sometimes adequate in simple cases where they may provide a good solution with only a

few terms. However, since they are not orthogonal on R
+
and they are all monotonically

increasing and positive for x Œ R
+
, for x big enough, they are almost indistinguishable

and hence they tend to reduce numerical accuracy
30
. Consequently, orthogonal bases are

usually preferred to avoid the shortcomings just mentioned.

One of the more popular orthogonal bases is formed by Chebyshev polynomials.

They constitute a set of orthogonal polynomials with respect to the weight function






30






In order to solve for the unknown coefficients u
i
one needs to solve linear systems of
equations. The accuracy of these solutions depends on the properties of the matrices
involved in the computation, i.e. linear independence of rows and columns. Due to the
properties already mentioned, regular polynomials tend to lead to ill-conditioned
matrices.



30
e(x) = 1 1÷ x
2
, that is,
}
1
p
i
(x) p
j
(x)e(x)dx = 0 for all i = j . Chebyshev polynomials
÷1

are defined on the closed
interval
|÷1,
1
| and can be computed recursively as follows:


p
i
(x) = 2xp
i
÷
1
(x) ÷ p
i
÷
2
(x), i = 2, 3, 4,K (1.6.7)
with p
0
(x) = 1 and p
1
(x) = x or, non-recursively, as:

p
i
(x) = cos (i arccos (
x
)) (1.6.8)
Another set of possible basis functions that can be used to construct a piecewise

linear representation for d
ˆ
is given by:

? x ÷ x
i
÷
1
if x
e
| x ,
x
|
? x
i
÷ x
i
÷1
?
i÷1 i
?
i
(x) = ? x
i
+
1
÷ x if x
e
|x
i
, x
i
+
1
|

? i+1 i
?x ÷ x
(1.6.9)
?
?
?
0 elsewhere

The points x
i
, i = 1,K, n that divide the domain D c ? need not be equally spaced. If,

for example, it is known that the function to be approximated has large gradients or kinks

in certain places then the subdivisions can be smaller and clustered in those regions. On

the other hand, in areas where the function is near-linear the subdivisions can be larger

and hence fewer.

Once the basis is chosen, the next step is to choose how many terms and

consequently how many parameters the functional form will have. In general, if the

choice of the basis is good, the higher the number of terms the better the approximations.

However, due to the fact that the more terms are chosen the more parameters have to be

computed, one should choose the smallest number of terms, n , that yields an acceptable






31
approximation. One possible approach is to begin with a small n and then increase its

value until some approximation threshold is reached.




I.6.1.2.1. Example


Going back to the model defined by equations (1.6.2) and (1.6.3) the next step is

choosing a base. I assume that Chebyshev polynomials are used in constructing the

functional form for dˆ
T
÷
1
(?,
u
) . Then:



T
÷
1
(x
T
÷
2
, y
T
÷
1
;u
T
÷
1
) =

n
x
,T ÷
1
n
y
,T ÷1
9 9u



T ÷1,sq


p
s
÷
1
(x
T
÷
1
) p
q
÷
1
( y
T
÷
1
)
s=1 q=1
% %
(1.6.10)

where u
T
÷1,sq is
the
(s,
q
) element of u
T
÷
1
, p
l
(?) is the l -th order Chebyshev polynomial


as defined in (1.6.7) - (1.6.8), while n
x
,T ÷
1
and n
y
,T ÷
1
are the maximum order of the

Chebyshev polynomials assumed for x
T
÷
1
and y
T
÷
1
respectively. In order to restrict the
% %

domain of the polynomials to the unit interval the following transformation is applied:

%T ÷
1
= 2 x
T
ma÷1x ÷ x
T
m÷i1
n
÷ 1
min
x



%
x
T
÷
1
÷ x
T
÷1

min
%
(1.6.11)
y
T
÷
1
= 2 y
T
ma÷1x ÷ y
T
m÷i1
n
÷1 (1.6.12)
y
T
÷
1
÷ y
T
÷1



I.6.1.3. Choosing a Residual Function

In many cases, computing F (dˆ) may require the use of numerical approximations


such as when F (d ) involves integration of d . In those cases, the F operator has to be

approximated. In addition, once the methodology for approximating d and F has been






32
established, one needs to choose a residual function. Therefore, the third step consists of

defining the residual function and an approximation criterion. Let

R(x;u ) } F (dˆ(?,u ))(x) ˆ


(1.6.13)
be the residual function. At this point, a decision has to be made on how an acceptable

approximation is defined. That is accomplished by choosing an approximation criterion.

One choice is to compute the sum of squared residuals, R(?;u ) } R(?;u ), R(?;u ) and

then determine u such that R(?;u ) is minimized. An alternative would be to choose a

collection of n test functions in C
2
, p
i
: D C R
m
, i = 1,..., n , and for each guess of u to

compute the n projections, P
i
(?) } R(?;u ), p
i
(?)

31
.


It is obvious that this step creates the

projections that will be used to determine the value of the unknown coefficients, u .

Another popular choice in the literature is the weighted residual criterion defined as
32
:

v ¢ (x)R(x;u )dx = 0,
iD


i = 1,K, n


(1.6.14)

where ¢
i
(x), i = 1,K, n are weight functions. Alternatively, the set of equations (1.6.14)

can be written as

v



D


e (x)R(x;u )dx = 0

n


(1.6.15)
where D is the domain for function d , e (x)
=
9e ¢
i
(x) and (1.6.15) must hold for i
i=1

any non-zero weights e
i
, i = 1,K, n . Therefore, the method sets a weighted integral of

R(x;u ) to zero as the criterion for determining u .


31



The choice of the criterion gives the method its name. That is why in the literature the
method appears both under the name "projection method" and "weighted residual
method".

32

See McGrattan (1999).





33
I.6.1.3.1. Example


Going back to the example, recall that Chebyshev polynomials were used in

constructing the functional form for dˆ
T
÷
1
(?,
u
) :



T
÷
1
(x
T
÷
2
, y
T
÷
1
;u
T
÷
1
) =

n
x
,T ÷
1
n
y
,T ÷1
9 9u



T ÷1,sq


p
s
÷
1
( %T÷
1
) p
q
÷
1
( %T÷
1
)
s=1 q=1
x y

As mentioned above, the Euler equation (1.6.5) needs to hold for d
ˆ
(?,
u
) . Therefore, its

right hand side is a prime candidate for defining the residuals function. Let v
T
÷
1
= _
T
÷
2
ˆ
.
x
·
y
T÷1 ˜

With this notation, the residual function is given by:
. +

R
T
÷
1
?v
T
÷
1
; dˆ
T
÷
1
(v
T
÷
1
;u
T
÷
1
)? =
? ?
g v
T
÷
1
, dˆ
T
÷
1
(v
T
÷
1
;u
T
÷
1
), y
T
÷
1
? g '
x

1
v
T
÷
1
, dˆ
T
÷
1
(v
T
÷
1
;u
T
÷
1
), y
T
÷1
( ) ( ) (1.6.16)

+
(t
'
)÷1 E
?
|t ' g dˆ
T
÷
1
(v
T
÷
1
;u
T
÷
1
), x
T
, y
T
? g '
x

1

T
÷
1
(v
T
÷
1
;u
T
÷
1
), x
T
, y
T
?
{
?
( ) (
)
}
?

Then the criterion for computing u%
T
÷
1
is given by the weighted residual integral equation:

v

R
T
÷
1
v
T
÷
1
; dˆ
T
÷
1
(v
T
÷
1
;uˆT ÷
1
)
?
W (v
T
÷
1
)dv
T
÷
1
= 0
v
T
÷1
e ?
(1.6.17)
where W is a weighting function. In the next section it will become clear why the choice

of W is important in the computation of u%
T
÷
1
.




I.6.1.4. Methods Used for Estimating the Parameters

Evidently, the next step is to find u Œ R
n
that minimizes the chosen criterion. In

order to determine the coefficients u
1
,K,u
n
several methods can be used, depending on

the criterion chosen.




34
If the projection criterion is chosen, finding the n components of u means

solving the n equations
R
(x,
u
), p
i
= 0 for some specified collection of test functions,


p
i
. The choice of the test functions p
i
defines the implementation of the projection

method. In the least squares implementation the projection directions are given by the

gradients of the residual function. Therefore, the problem is reduced to solving the

c
R
( x ,
u
)
nonlinear set of equations generated by
R
(x,
u
),
cu
i
= 0 i = 1,..., n .

One alternative is to choose the first n elements of the basis u , that is,

?
i
(x) i = 1,..., n , as the weight functions, ¢
i
(x), i = 1,K, n . In other words, n elements of

the basis used to approximate dˆ(x;u ) are also used as test functions to define the

projection direction, ¢
i
(x) = ¢
i
(x), i = 1,K, n . This technique is known as the Galerkin

method. As a result of this choice, the Galerkin method forces the residual to be

orthogonal to each of the basis functions. Therefore u is chosen to solve the following

set of equations:

P
i
(
u
) =
R
(x,
u
),¢
i
(
x
) = 0 i = 1,..., n (1.6.18)

As long as the basis functions are chosen from a complete set of functions, system

(1.6.18) provides the exact solution, given that enough terms are included. If the basis

consists of monomials, the method is also known as the method of moments. Then u is

the solution to the system:

P
i
(
u
) =
R
(x,
u
), x
i
÷
1
= 0 i = 1,..., n (1.6.19)

The collocation method chooses u so that the functional equation holds exactly at

n fixed points, x
i
, called the collocation points. That is, u is the solution to:



35
R(x
i
;u ) = 0, i = 1,..., n (1.6.20)

where
{x
i
}ni=
1
are n fixed points from D . It is easy to see that this is a special case of the

projection approach, since R(x;u ),o (x ÷ x
i
) = R(x
i
;u ) , where o (x ÷ x
i
) is the Dirac

function at x
i
. If the collocation points x
i
are chosen as the n roots of the n
th
orthogonal

polynomial basis element and the basis elements are orthogonal with respect to the inner

product, the method is called orthogonal collocation. The Chebyshev polynomial basis is

a very popular choice for an orthogonal collocation method.




I.6.1.4.1. Example


Going back to the example, it was established that the criterion for computing

u%
T
÷
1
is given by the following integral equation:

v

R
T
÷
1
v
T
÷
1
; dˆ
T
÷
1
(v
T
÷
1
;uˆT ÷
1
)?W (v
T
÷
1
)dv
T
÷
1
= 0
v
T
÷1
e ?
As discussed in this section, given this criterion, the collocation method is a

sensible choice for computing u%
T
÷
1
. Then the choice for the weighting functions, as used

in Binder et al. (2000), is the n
x
,T ÷
1
, n
y
,T ÷
1
Dirac delta functions o x
T
÷
1
÷ x
i
T ÷
1
, y
T
÷
1
÷ y
i
T ÷
1
,
( )

where x
iT
÷
1
and y
iT
÷
1
are chosen such that x
iT
÷
1
and y
iT
÷
1
are the n
x
,T ÷
1
and n
y
,T ÷
1
zeros of
% %

the Chebyshev polynomials forming the basis of the approximation dˆ
T
÷
1
(v
T
÷
1
;u
T
÷
1
) . The

zeros for the Chebyshev polynomials are given by


?
?
co
s
(2
i

÷
1
)
t


?
?
v%T÷
1
=
? i
2n
x
,T ÷1 ? (1.6.21)
? (2i ÷1)t ?
? cos 2n ?
? y,T ÷1 ?



36
Then the integral equation can be reduced to:

R
T
÷
1
v
ij
÷
1
; dˆ
T
ij÷
1
= 0

for all
(
T
)
(1.6.22)

v
ij
÷
1
= x
i
T ÷
1
, y
T

1
, i = 1, 2,..., n
x
,T ÷
1
,
and
T
( )




T
ij÷
1
= dˆ
T
÷
1

v
ij
÷
1
;uˆT ÷1
j = 1, 2,..., n
y
,T ÷1
(1.6.23)
(
T
)
(1.6.24)

The discrete orthogonality of Chebyshev polynomials implies that:

n
x
,T ÷
1
n
y
,T ÷1
9 9 e
p
(x
%
)
p
(y
%
)??e
p
(x
%
)
p
(y
%
)?? =
0
i=1 j =1 w÷1
i
T ÷1
p÷1
j
T ÷1
s÷1
i
T ÷1
q÷1
j
T ÷1
(1.6.25)
for w t s and /or p t q , and


n
x
,T ÷
1
n
y
,T ÷1
9 9 e
p
(x
%
)
p
(y
%
)??e
p
(x
%
)
p
(y
%
)?? =

c
(n
i=1 j =1 w÷1
i
T ÷1
p÷1
j
T ÷1
s÷1
i
T ÷1
q÷1
j
T ÷1
sq
x,T ÷1
, n
y
,T ÷
1
(1.6.26) )
for w = s and p = q , with




?n
x
,T ÷
1
n
y
,T÷
1
, w = s = p = q = 1
?
? ?w = s = 1 and p = q = 1,
c
sq
(n
x
,T ÷
1
, n
y
,T÷
1
) = ?x,T÷1y,T÷
1
) / 2, ?or
?
(n n
?
(1.6.27)
?
?
??w
= s =
1 and
p =
q

=
1,
?
(n
x
,T ÷
1
n
y
,T÷
1
) / 4, w = s = 1 and p = q = 1 ?
Then u is given by:

n
x
,T ÷
1
n
y
,T ÷1
1
uˆT ÷1,sq =

9 9
p
(x
%
)
p
(y
%
)?e
g
(v
( )
s÷1
i
T ÷1
q÷1
j
T ÷1
T ÷1
, dˆ
T
ij÷
1
, y
T
÷
1
? g '
x

2
v
T
÷
1
, dˆ
T
ij÷
1
,
y
T
÷1
c
sq
n
x
,T ÷
1
, n
y
,T ÷1
i=1 j =1
) ( )
+
(t%
'
)÷1 E |t ' g dˆ
T
ij÷
1
, x
T
, y
T
? g '
x

1

T
ij÷
1
, x
T
, y
T
? v
T
÷1 ?
({
e
( ) ( )
?
}
)
?
?
(1.6.28)

for s = 1, 2,..., n
x
,T ÷
1
, q = 1, 2,..., n
y
,T ÷
1
.





37
The conditional expectation from the above equation needs to be computed

numerically. In order to compute the integral one can use some of the quadrature methods

such as the Gauss quadrature presented in section I.5.2. All that remains is to solve

equation (1.6.28) for u
T
÷1,sq , s = 1, 2,..., n
x
,T ÷
1
, q = 1, 2,..., n
y
,T ÷
1
. Once dˆ
T
÷
1
v
T
÷
1
;uˆT ÷1 (
)


is

computed, one can proceed recursively backwards to period T ÷ 2 . Note that

x
*
÷
1
= dˆ
T
÷
1
v
T
÷
1
;uˆT ÷
1
will be used in the definition of R
T
÷
2
v
ij
÷
2
; dˆ
T
ij÷
2
. The computation
T
( ) (
T
)

of uˆT ÷
2
can now follow the same logic as the computation of uˆT ÷
1
.

So far the flavors of the projection methodology have been categorized either with

respect to the choice of the approximation criterion or with respect to the method

employed for estimating the parameters. The choice of basis functions for the

representation in (1.6.6) can be used to further divide projection methods into two

categories: spectral methods and finite-element methods. Spectral methods use basis

functions that are smooth and non-zero on most of the domain of x such as Chebyshev

polynomials and the same functions are used on all regions of the state space. Finite-

element methods use basis functions that are equal to zero on most of the domain and

non-zero on only a few subdivisions of the domain of x (these are in general piecewise

linear functions such as those defined in (1.6.9)) and they provide different

approximations in different regions of the state space. For problems with many state

variables, there are typically many coefficients to compute and it implies the inversion of

a large, dense matrix. With the finite-element method, however, the same matrix is sparse

and its structure can typically be exploited. For the above-mentioned reasons McGrattan






38
(1996, 1999) argues that a finite-element method is better suited to problems in which the

solution is nonlinear or kinked in certain regions.





I.6.2. Parameterized Expectations

While Marcet (1988) is largely credited in the literature with the introduction of

the parameterized expectations approach, Christiano and Fisher (2000) point out that the

underlying idea of parameterized expectations seems to have surfaced earlier in the work

of Wright and Williams (1982a, 1982b, 1984), and then in the work of Miranda and

Helmberger (1988). Marcet (1988)
33
implemented a variation of that idea and the

approach finally caught on with the publication of Den Haan and Marcet (1990).

In this section, I will concentrate on what Christiano and Fisher (2000) call the

conventional parameterized expectations approach due to Marcet (1988). While one may

argue that this methodology does not belong under the label of projection methods, I

believe that it can be viewed as a special case of projection methods by virtue of its use of

parameterized functions to approximate an unknown quantity, of an implicit choice of a

residual function and an approximation criterion similar to projection methods. In

addition, the techniques used to estimate the parameters are also common to projection

methods. The assumption is that the functional equation has the following form:

g E
t
?
|
(q
t
+
1
,q
t
)? ,q
t
÷
1
,q
t
, z
t
= 0
(
? ?
)
(1.6.29)

where q
t
includes all the endogenous and exogenous variables and z
t
is a vector of

exogenous shocks. As it has been repeatedly asserted in this chapter, the reason why


33


For more information of this variant of the parameterized expectations approach, see
the references cited in Marcet and Marshall (1994b).



39
many dynamic models are difficult to solve is that conditional expectations often appear

in the equilibrium conditions. The assumption under which this methodology operates is

that conditional expectations are a time-invariant function c of some state variables:

c (u
t
) = E
t
||(q
t
+
1
,q
t )
| (1.6.30)

where E
t
||(q
t
+
1
,q
t )
| = E ?|(q
t
+
1
,q
t
) u
t
? is the conditional expectation based on the
? ?

available information at time t , u
t
e R
l
where u
t
is a subset
of
(q
t
÷
1
, z
t
) . As Marcet and

Lorenzoni (1999) point out, a key property of c is that under rational expectations, if

agents use c to form their decisions, the series generated is such that c is precisely the

best predictor of the future variables inside the conditional expectations. So, if c were

known, one could easily simulate the model and check whether this is actually the

conditional expectation.

The basic approach of Marcet and Marshall (1994a) is to substitute the

conditional expectations in equation (1.6.29) by parameterized functions of the state

variables with arbitrary coefficients. Then (1.6.29) is used to generate simulations for u
t


consistent with the parameterized expectations. With these simulations, one can iterate on

the parameterized expectations until they are consistent with the solution they generate.

In this fashion, the process of estimating the parameters is reduced to a fixed-point

problem.





I.6.2.1. Example

Consider again the model specified by (1.6.2) - (1.6.3) with the Euler equation for

period t given by:




40

0 = t '
g
(x
t
, x
t
÷
1
, y
t
) ? g '
x
t
(x
t
, x
t
÷
1
, y
t
)
( )
(1.6.31)
+ | E t '
g
(x
t
+
1
, x
t
, y
t
+
1
) ? g '
x
t
(x
t
+
1
, x
t
, y
t
+
1
) O
t
{( ) }
The idea is to substitute

E
t
t '
g
(x
t
+
1
, x
t
, y
t
+
1
) ? g '
x
t
(x
t
+
1
, x
t
, y
t
+
1
)
{( ) }
by a parameterized function
¢
( x
t
÷
1
, y
t
;
u
) where u is a vector of parameters. For

simplicity, let the function ¢ be given by:

¢
t
( x
t
÷
1
, y
t
;u
1
,u
2
) = u
1
x
t
÷
1
+u
2
y
t
(1.6.32)

The next step is to generate a
series
{z
t
}Tt=
1
as draws from a Gaussian distribution and to

choose starting values for the elements of u , u
i
0
, i = 1, 2 . Then, for uˆi = u
i
0
and assuming


that the initial values for x
t
and y
t
, that is, x
÷
1
and y
0
are given, one can use the

following system

t
'
(
g
(x
t
, x
t
÷
1
, y
t
))? g '
x
(x
t
, x
t
÷
1
, y
t
) +uˆ1x
t
÷
1
+uˆ2 y
t
= 0 for t = 0,...,T ÷1 t
x
t
=
h
(x
t
÷
1
,u
t
, y
t
) for t = 0,...,T , with x
÷
1
given (1.6.33)
y
t
= µ y
t
÷
1
+ z
t
for t = 1,...,T , with y
0
given

to generate
series

tt
j
}
t
=
0
,

t
j
}
t
=
1
and
{u
t
j
}
t
=
0
where j represents the iteration. In order
T T
x y ˆT

to estimate the parameters u , proponents of this methodology run a regression of

?
tj
uˆj = t '
g

t
j
, ˆ
t

1
, ˆ
t
j
) ? g '
x
t

t
j
, ˆ
t

1
, ˆ
t
j
)
() ( xx y ) xx y (1.6.34)

on ¢
t
. Formally, the regression can be written as:

?
tj
uˆj = a
1
ˆ
t

1
+ a
2
ˆ
tj
+ ç
t
() x y

where ç
t
is the error term. The estimates for a
1
and a
2
provide a new set of values for u

for the next iteration. With those values new series will be generated for {ˆ
t
j+
1
}
t
=
0
and T
x




41
{u
ˆ
} j
+1 T
y
T
t t =0
. In this particular case, there is no need to generate new series
for

t
j+
1
}
t
=
1
if the

same vector of shocks {z
t
}Tt=1 is used. In addition, note that a
1
and a
2
are in fact

functions of uˆ . Specifically, for iteration j , the vector of parameters a is a function of

j
, a = G uˆ
j
. Hence the final step is to find the fixed point u =
G
(
u
) . One approach ()

suggested by Marcet and Lorenzoni (1999) is to compute the values of uˆ for iteration
j +1 using the following expression uˆj+
1 =
(1÷
b
)uˆj + bG uˆ
j
where b > 0 . The iteration ()
process should stop when uˆ
j
and G uˆ
j
are sufficiently close. ()




I.6.3. Notes on Projection Methods

As Judd (1992) points out, the advantage of the projection method framework is

that one can easily generate several different implementations by choosing among

different basis, residual functions or methods for estimating the parameters. Obviously,

the many choices also imply some trade-offs among speed, accuracy, and reliability. For

example, the orthogonal collocation method tends to be faster than the Galerkin method,

while the Galerkin method tends to offer more accuracy
34
.

The generality of the projection techniques can also be seen from the fact that

even methods that discretize the state space can be thought of as projection methods that

are using step function bases.

While throughout this section I emphasized the wide applicability of projection

methods, there is an aspect that has been overshadowed. Recall that the idea is to replace


34


See Judd (1992) for more details.





4 2
the quantity that needs to be approximated by parameterized functions (basis functions

?
i
(
x
) ) with arbitrary coefficients ( a
i
). In projection methods, the coefficients are chosen

to be the best possible choices relative to the basis ?
i
(
x
) and relative to some criterion.


However, the bases are usually chosen to satisfy some general criteria, such as

smoothness and orthogonality conditions. Such bases may be good but very rarely are

they the best possible for the problem under consideration.

An important advantage of parameterized expectations approach is that, for

specific models, it may implicitly deal with the presence of inequality constraints

eliminating the need to constantly check whether the Kuhn-Tucker conditions are

satisfied
35
.

A key component of the conventional parameterized expectations approach

presented in this section is a cumbersome nonlinear regression step. The regression step

implies simulations involving a huge amount of synthetic data points. The problem with

this approach is that it inefficiently concentrates on a residual amount that is obtained

from visiting only high probability points of the invariant distribution of the model. As

Pointed out by Judd (1992) and Christiano and Fisher (2000), it is important to consider

the tail areas of the distribution as well. Christiano and Fisher (2000) offer a modified

version of the parameterized expectations approach that they call the Chebyshev

parameterized expectations approach, specifically designed to eliminate the shortcoming

discussed above. In fact, Christiano and Fisher (2000) explicitly transform the

parameterized expectations approach into a projection method that they refer to as the

weighted residual parameterized expectations approach. As mentioned above, expressing

35


See Christiano and Fisher (2000) for details.



43
the parameterized expectations approach as a projection method opens the door to a

variety of possible implementations.
36
.



I.7. Comparing Numerical Methods: Accuracy and Computational Burden

It is difficult to define the global criteria of success for numerical methods.

Accuracy is in general at the top of the checklist in defining a good numerical method.

However, it may not always be the most important criterion when choosing a numerical

method. For example, even though a method may not provide the best approximation for

the policy function, it may still be preferred to other methods as long as the loss in

accuracy relative to the policy function does not affect too much the value of the

objective function. In such cases, speed or ease of implementation may take precedence.

There does not seem to be a general agreement in the literature on how to evaluate

the accuracy of numerical methods. Consequently, a number of criteria have been

proposed in order to asses the performance of numerical algorithms.

One widely used strategy for determining accuracy is to test the outcome of a

computational algorithm in a particular case where the model displays an analytical

solution. For example, Collard and Juillard (2001) use an average relative error and a

maximal relative error criterion in order to asses the accuracy of several numerical

methods. While this approach may be useful for certain specifications, the problem is that

for alternative parameterizations of the model the approximation error of the computed

decision and value functions may change substantially. Changes in the curvature of the

objective function and in the discount factor are the usual culprits in influencing

36

In fact, Christiano and Fisher (2000) provide two other modified versions of the
parameterized expectations approach (PEA): PEA Galerkin and PEA collocation.



44
considerably the accuracy of the algorithm. Collard and Juillard (2001) determine that for

an asset pricing model the Galerkin method using fourth order Chebyshev polynomials

clearly outperforms linearization methods as well as lower order perturbation methods.

However, higher order (order four and higher) perturbation methods prove to be quite

accurate.

Another strategy used for analyzing the accuracy of numerical methods is to look

at the residuals of the Euler equation. This seems like a natural choice especially for

approaches that are based on approximating certain terms entering, or the whole, Euler

equation
37
.

A procedure for checking accuracy of numerical solutions based on the Euler

equation residuals was proposed by den Haan and Marcet (1990, 1994). It consists of a

test for the orthogonality of the Euler equation residuals over current and past

information. The idea behind this test is to compute simulated time series for all the

choice and state variables as well as Euler equation residuals, based on a candidate

approximation. Then, using estimated values of the coefficients resulting from regressing

the Euler equation residuals on lagged simulated time series, one can construct measures

of accuracy. As pointed out by Santos (2000), the problem with this approach is that

orthogonal Euler equation residuals may be compatible with large deviations from the

optimal policy. In addition, as referenced by Judd (1992), Klenow (1991) found that the

procedure failed to reject candidate solutions that resulted in relatively high errors for the

choice variable while rejecting solutions resulting in occasional high large errors but

without any discernible pattern.

37

For a detailed discussion on criteria involving Euler equation residuals, please see
Reiter (2000) and Santos (2000).



45
Judd (1992, 1998) suggested an alternative test that consists of computing a one

period optimization error relative to the decision rule. The error is obtained by dividing

the current residual of the Euler equation to the value of next period's decision function.

Subsequently, two different norms are applied to the error term: one gives the average

and the other supplies the maximum.

In a study aimed at comparing various approximation methods, Taylor and Uhlig

(1990) found that performance varies greatly depending on the criterion used for

assessing accuracy. For example, the decision rules indicated that some of the easier to

implement methods such as the linear-quadratic method and the extended-path method

were fairly close to the "exact" decision rule
38
as given by the quadrature-value-function-

grid method of Tauchen (1990) or the Euler-equation grid method of Coleman (1990).

However, neither the linear-quadratic nor the extended-path method performed well

when using the martingale-difference tests for the Euler-equation residual. Not

surprisingly, the parameterized expectations approach performed well when using the den

Haan and Marcet criterion but not as well when measured against the exact decision rule.

While accuracy is very important, computational time may also play an important

role in the eyes of some researchers. While the extended-path method has relatively low

cost when compared to grid methods, it is fair to state that both grid methods and the

extended-path method are computationally quite involved, whereas linear-quadratic

methods are typically quite fast. Most projection methods also fare well in terms of


38



Solutions obtained through discretization methods are sometimes referred to as
"exact". The reason behind this labeling is that models obtained as a result of
discretization may be solved exactly by finite-state dynamic programming methods.
However, one has to keep in mind that reducing a continuous-state problem to a finite-
state problem still involves an approximation error.



46
computational burden when compared to discretization methods or even parameterized

expectations methods. As the state space increases, discretization methods suffer heavily

from the curse of dimensionality.

The fact that none of the methods outperforms the others does not mean that every

method could be applied to any model out there with a good degree of success
39
. One has

to use good judgment when deciding on using a certain numerical method.





I.8. Concluding Remarks

As it has become clear over the course of this chapter, there are quite a few

methodologies available for solving non-linear rational expectations models. However, if

one looks closer, it becomes obvious that all methods share some common elements. For

example, certainty equivalence is at the core of the extended path method but it can also

be used in perturbation methods to find the equilibrium of a (deterministic) system

similar to the one under investigation. The discrete state space approach can be viewed as

a projection method with step functions as a basis. Similarly, the first order perturbation

method is nothing more than a simple linearization around steady state. In addition, the

parameterized expectations approach can be easily transformed into a projection method.

Moreover, since all the functional equations for rational expectations models imply the

existence of some integrals, the quadrature approximation may make an appearance in

almost every methodology.



39




Judd (1998) contains an example of a partially revealing rational expectations problem
which cannot be solved by discretizing the state space, but which can be approximated by
more general projection methods.



47
Several studies have tried to asses the performance of these numerical methods.

However, even for relatively simple models their performance may vary greatly
40
.

Despite all of their sophistication, none of these methods can consistently outperform the

others.

Even comparing the methods is not a walk in the park. Several authors including

Judd (1992), Den Haan and Marcet (1994), Collard and Juillard (2001), Santos (2000)

and Reiter (2000) proposed different criteria for evaluating the performance of numerical

solutions. Unfortunately, each criterion has its caveats and it has to be applied selectively,

based on the specificity of the model under investigation. Therefore, one has to choose

carefully the proper methodology when in need of numerical solutions.




























40




























See the studies by Taylor and Uhlig (1990), Judd (1992), Rust (1997), Christiano and
Fischer (2000), Santos (2000), Collard and Juillard (2001), Fair (2003), Schmitt-Grohé
and Uribe (2004).



48
Chapter II. Using Scenario Aggregation Method to Solve a Finite


Horizon Life Cycle Model of Consumption





II.1. Introduction

Multistage optimization problems are a very common occurrence in the economic

literature. While there exist other approaches to solving such problems, many economic

models involving intertemporal optimizing agents assume that the representative agent

chooses its actions as a result of solving some dynamic programming problem. Lately, an

increasing number of researchers have investigated alternative approaches to modeling

the representative agent, in an attempt to find one that may explain observed facts better

or easier. Following the same line of research, I explore the suitability of scenario

aggregation method as an alternative to describe the decision making process of an

optimizing agent in economic models. The idea is that this methodology offers a different

approach that might be more consistent with the observation that agents are more likely

to behave like chess players, making decisions based only on a subset of all possible

outcomes and using a relatively short horizon
41
. The advantage of scenario aggregation

methodology is that, while it presents attractive features for use in models assuming

bounded rationality, it can also be seen as an alternative numerical method that can be

used for obtaining approximate solutions for rational expectation models. Therefore, I

start by studying in this chapter the viability of the scenario aggregation method, as


41


In the next chapter I will focus more on the length of the span over which the decision
making process takes place.



49
presented by Rockafellar and Wets (1991), to provide a good approximation for the

optimal solution of a simple finite horizon life-cycle model of consumption with

precautionary savings. In the next chapter, I will use scenario aggregation to model the

decision making of the rationally bounded consumer.

The layout of this chapter is as follows. First, I present the setup of a simple life-

cycle consumption model with precautionary saving. Then, I introduce the notion of

scenarios followed by a description of the aggregation method. Next, I introduce the

progressive hedging algorithm followed by its application to a finite horizon life-cycle

consumption model. Then, I present simulation results and conclude the chapter with

final remarks.





II.2. A Simple Life-Cycle Model with Precautionary Saving

I consider the following version of a life-cycle model. Suppose an individual

agent is faced with the following intertemporal optimization problem:

max E

9 |
t
F
t
(c
t
) | I
0
? T
?


(2.2.1)
{c
t
}
T
=0
t
_t=0
e
?
where F
t
is a utility function which has the typical properties assumed in the literature,

i.e. it is twice differentiable, it is increasing with consumption and exhibits negative

second derivative. The information set I
0
contains the level of consumption, assets, labor

income and interest rate for period zero and all previous periods.

Maximization is subject to the following transition equation:

A
t =
(1+ r
t
) A
t
÷
1
+ y
t
÷ c
t
, t = 0,1,...,T ÷1, (2.2.2)

A
t
> ÷b, with A
-1
, A
T
given (2.2.3)



50
where A
t
represents the level of assets at the beginning of period t , y
t
the labor income

at time t , and c
t
represents consumption in period t . The initial and terminal conditions,

A
÷
1
and A
T
, are given. Uncertainty is introduced in the model through the labor income.

The realizations of the labor income are described by the following process:

y
t
= y
t
÷
1
+ ç
t
, t = 1,...,T , with y
o
given (2.2.4)
and ç
t
being drawn from a normal distribution, ç
t
~
N
(0,o2
y
) . For now, I will not make


any particular assumption about the process generating the interest rate, r
t
. Therefore, to

summarize the model, a representative consumer derives utility in period t from

consuming c
t
, discounts future utility at a rate | and wants, in period zero, to maximize

his present discounted value of future utilities for a horizon of T +1 periods. At the

beginning of each period t the consumer receives a stochastic labor income y
t
, and

based on the return on his assets A
t
÷
1
, from the beginning of period t ÷1 to the beginning

of period t , he chooses the consumption level c
t
, and thus determines the level of assets

A
t
according to equation (2.2.2).

Of particular importance in this problem is the random variable ç
t
. In the

standard formulation of the problem, ç
t
is assumed to be distributed normally with mean

zero and some variance o2
y
. Instead of making the standard assumption, if I assume that

ç
t
's sample space has only a few elements, then the optimization problem (2.2.1) -

(2.2.4) is a perfect candidate for being solved using the scenario aggregation method. Let

me assume for the moment that the sample space is given
by
{e
1
,e
2
,...,e
n
} with the

associated probabilities { p
1
, p
2
,..., p
n
} . If S is the set of all scenarios then its cardinal is



51
given by n
T
. It is obvious that as the sample space for the forcing variable increases, the

number of scenarios increases proportional with power T . Therefore, applying the

scenario aggregation method to find an approximate solution for this problem may only

be feasible when T and n are relatively small. In the next chapter, I will present a

solution for T relatively large.





II.3. The Concept of Scenarios


II.3.1. The Problem

In this section, I formally introduce a multistage optimization problem and then,

in the following sections, I will present the idea of scenario aggregation and how it can be

applied to such a problem.

The multistage stochastic optimization problem consists of minimizing an

objective function, F : R
m
C R subject to some constraints, which usually describe the

dynamic links between stages.

The objective function F is time separable and is given by a sum of functions,

T
F
=
9 F
t
with each function F
t
, F
t
: R
m
C R corresponding to stage t of the
t =0

optimization problem. These functions depend on a set of variables u
t
, which in turn

represent the decisions that need to be made at each stage t . For simplicity I assume that

u
t
is a m
u
·1 vector, with m
u
independent of t , that is, the same number of decisions is

to be made at each stage.






52
If U (t) represents the set of all feasible actions at stage t , then u
t
has to be part

of the set U (t) , that is, u
t
¸ U (t), t = 0,...,T , U (t) [ R
m
u . The temporal dimension of

the problem is characterized by stages t and state variables X (t) .

The link between stages is given by:

x
t
+
1
= G
t
(x
t
,u
t
,u
t
+
1
) .
Hence, the problem can be formulated as:

min
E
?
¿ F
t
( x
t
,u
t
) | I
0
? T

subject to:
??
t =0



x
t
+
1
= G
t
(x
t
,u
t
,u
t
+
1

t
)
?
?
(2.3.1)




(2.3.2)

where I
0
is the information set at time t = 0 and ç
t
is the forcing variable.

In the next few sections, I will present the concept of scenarios as well as possible

decomposition methods along with the idea of scenario aggregation.





II.3.2. Scenarios and the Event Tree

In this section, I present an intuitive description for the concept of scenarios. A

formal description is presented in Appendix, section A1. Suppose the world can be

described at each point in time by the vector of state variables x
t
. In the case of a

multistage optimization problem, let u
t
denote the control variable and let ç
t
be the

forcing variable. I assume that an agent makes decisions reflected in the control variable

u
t
. For simplicity let ç
t
be a random variable witch can take two values ç
a
and ç
b
with

probabilities p
a
and 1÷ p
a
.




53

If the horizon has T +1 time periods and {ç
a

b
} is the set of possible

realizations for ç
t
then the sequence

ç
s
=

0
s

1
s
,K,ç
T
s
)
is called a scenario
42
. From now on, for notation simplification, I will refer to a scenario

s simply by ç
s
or by the index s . Given that the set of all realizations for ç
t
is finite,

one can define an event
tree
{N,
A
} characterized by the set of nodes N and the set of

arcs A . In this representation, the nodes of the tree are decision points and the arcs are

realizations of the forcing variables. The arcs join nodes from consecutive levels such

that a node n
t
j
at level t is linked to N
t
+
1
nodes, n
k
+
1
, k = 1,..., N
t
+
1
at level t +1. In t

Figure 1 I represent such a tree for a span of T = 3 periods. As mentioned above, the

forcing variable takes only two
values,

a

b
} and hence the tree has 15 nodes. The arcs

that join nodes from consecutive levels represent realizations of the forcing variable and

are labeled accordingly.

The set of nodes N can be divided into subsets corresponding to each level

(period). Suppose that at time t there are N
t
nodes. For example, for t = 1, there are two

nodes, node2 and node3. The arcs reaching these two nodes belong each to several

scenarios s . The bundle of scenarios that go through one node plays a very important

role in the decomposition as well as in the aggregation process. The term equivalence

class has been used in the literature to describe the set of scenarios going through a

particular node.



42



Other definitions of scenarios can be found in Helgason and Wallace (1991a, 1991b )
and Rosa and Ruszczynski (1994).



54
node1, t=0


çt=ça çt=çb


node2, t=1 node3, t=1

çt=ça çt=çb çt=ça çt=çb


node4, t=2 node5, t=2 node6, t=2 node7, t=2

çt=ça çt=çb çt=ça çt=çb çt=ça çt=çb çt=ça çt=çb


node8, t=3 node9, t=3 node10, t=3 node11, t=3 node12, t=3 node13, t=3 node14, t=3 node15, t=3







Figure 1 Event tree


By definition, an equivalence class at time t is the set of all scenarios having the

first t +1 realizations common. As mentioned in the above description of the event tree,

at time t there are N
t
nodes. Every node is associated with an equivalence class. Then,

the number of distinct equivalence classes at time t is also N
t
.

In Figure 2 one can see that for t = 1 there are two nodes and consequently two

equivalence classes, {s
1
, s
2
, s
3
, s
4
} and {s
5
, s
6
, s
7
, s
8
} . The number of elements of an


equivalence class is given by the number of leaves stemming from the node associated

with it. In this example, the number of leaves stemming from both nodes is four, which is

also the number of scenarios belonging to each class.



55
node1, t=0
(s1,s2,s3,s4,s5,s6,s7,s8)


çt=ça çt=çb


node2, t=1 node3, t=1
(s1,s2,s3,s4) (s5,s6,s7,s8)

çt=ça çt=çb çt=ça çt=çb

node4, t=2 node5, t=2 node6, t=2 node7, t=2
(s1,s2) (s3,s4) (s5,s6) (s7,s8)

çt=ça çt=çb çt=ça çt=çb çt=ça çt=çb çt=ça çt=çb


node8, t=3 node9, t=3 node10, t=3 node11, t=3 node12, t=3 node13, t=3 node14, t=3 node15, t=3
(s1) (s2) (s3) (s4) (s5) (s6) (s7) (s8)







Figure 2 Equivalence classes


The transition from a state at time t to one at time t + 1 is governed by the control

variable u
t
but is also dependent on the realization of the forcing variable, that is, on a

particular scenario s . Since scenarios will be viewed in terms of a stochastic vector ç

with stochastic components ç
0
s

1
s
,K,ç
T
s
, it is natural to attach probabilities to each

scenario. I denote the probability of a particular realization of a scenario, s , with

p(s) = prob(ç
s
) .
Let us consider the case of the event trees represented in Figure 1 and Figure 2 and

assume the probability of realization ç
a
is prob(ç
t
= ç
a
) = p
a
while the probability of

realization ç
b
, is prob(ç
t
= ç
b
) = p
b
, with p
a
+ p
b
= 1. Then, due to independence



56
across time, one can compute the probability of realization for scenario s
1
,

prob(ç
s
= s
1
) = p
3
a . Similarly, the probability of realization for scenario s
2
is

prob(ç
s
= s
2
) = p
2
p
b
, or prob(ç
s
= s
2
) = p
2
(1÷ p
a
) .
a a

Further on, I define
43
the probabilities associated with a scenario conditional upon

belonging to a certain equivalence class at time t . For example, the probability associated

with scenario s
1
, conditional on s
1
belonging to equivalence
class
{s
1
, s
2
, s
3
, s
4
} is given

by prob ( s
1
| s
1 e
{s
1
, s
2
, s
3
, s
4
}) = p
2
a




II.4. Scenario Aggregation
44



In this section, I will show how a solution can be obtained by using special

decomposition methods, which exploit the structure of the problem by splitting it into

manageable pieces, and then aggregate their solutions. In the multistage stochastic

optimization literature, there are two groups of methods that have been discussed: primal

decomposition methods that work with subproblems that are assigned to time stages
45


and dual methods, in which subproblems correspond to scenarios
46
. Most of the methods,

regardless of which group belong to, use the general theory of augmented Lagrangian

decomposition. In this chapter I will concentrate on a methodology that belongs to the

second group and has been derived from the work of Rockafellar and Wets (1991).


43



For a more formal definition, see the Appendix, section A1.

44
Section A2 in the Appendix offers a more formal description of scenario aggregation.

45

46

See the work of Birge (1985), Ruszczynski (1986, 1993), Van Slyke and Wets (1969).

See the work of Mulvey and Ruszczynski (1992), Rockafellar and Wets (1991),
Ruszczynski (1989), Wets (1988).


57
Let us assume for a moment that the original problem can be decomposed into

subproblems, each corresponding to a scenario. Then the subproblems can be described

as:

T

u ŒU [R
mu
min 9 F
t

x
s
t ,u
s
t ,
t t
t=1
( )
s ŒS
(2.4.1)

where u
st
and x
st
are the control and the state variable respectively, conditional on the

realization of scenario s while S is a finite, relatively small set of scenarios. Moreover,

suppose that each individual subproblem can be solved relatively easy. The question then

becomes how to blend the individual solutions into a global optimal solution. Let the

term policy
47
describe a set of chosen control variables for each scenario and indexed by

the time dimension.

The policy function has to satisfy certain constraints if two different scenarios s

and s ' are indistinguishable at time t on information available about them at the time.

Then u
s
t = u
s
t
'
, that is, a policy can not require different actions at time t relative to

scenarios s and s ' if there is no way to tell at time t which of the two scenarios will be

followed. In the literature, this constraint is sometimes referred to as the non-

anticipativity constraint. Going back to Figure 2, for t = 1, if the realization of ç
t
is ç
a
,

the decision maker will find himself at the decision point node2. There are four scenarios

that pass through node2 and the non-anticipativity constraint requires that only one

decision be made at that point since the four scenarios are indistinguishable. A policy is






47






A formal description of the policy function is presented in Appendix.


58
defined as implementable if it satisfies the non-anticipativity constraint, that is, u
t
must

be the same for all scenarios that have common past and present
48
.

In addition, a policy has to be admissible. A policy is admissible if it always

satisfies the constraints imposed by the definition of the problem. It is clear that not all

admissible policies are also implementable.

By definition, a contingent policy
49
is the solution, u
s
, to a scenario subproblem.

It is obvious that a contingent policy is always admissible but not necessarily

implementable. Therefore, the goal is to find a policy that is both admissible and

implementable. Such a policy is referred to as a feasible policy. One way to create a

feasible policy from a set on contingent policies is to assign weights (or probabilities) to

each scenario and then aggregate the contingent policies according to these weights.

The question that the scenario aggregation methodology answers is how to obtain

the optimal solution U from a collection of implementable policies U . In this chapter, I ˆ


will present a version of the progressive hedging algorithm originally developed by

Rockafellar and Wets (1991).













48













For certain problems the non-anticipativity constraint can also be defined in terms of
the state variable, that is, x
t
(e ) must be the same for all scenarios that have common past
and present.

49

I borrow this term from Rockafeller and Wets (1991).



59
II.5. The Progressive Hedging Algorithm

The algorithm is based on the principle of progressive hedging
50
which consists of

starting with an implementable policy and creating sequences of improved policies in an

attempt to reach the optimal policy.

Let us go back to the definition of an implementable policy. By computing


u
s
t =
ˆ

¿
}

p s
?
{s
t
}
i
u
s
t
'
= E u
s
t
'
| {s
t
}
i
s?e
{s
t
i
( )
( )
for all s
e
{s
t
}
i
(2.5.1)
for all scenarios s e S and all periods t = 1,...,T , one creates a starting collection of


implementable policies, denoted by U
0
. In equation (2.5.1) E represents the expectation ˆ


operator. Therefore, in order to obtain an initial collection of implementable policies one

should first compute some contingent policies for each scenario and then apply the

expectation operator for each period t and each scenario s conditional on it belonging to

the corresponding equivalence class, {s
t
}
i
.


The progressive hedging algorithm finds a path from U
0
, the set of ˆ


implementable policies, to U , the set of optimal policies, by solving a sequence of

problems in which the scenarios subproblems are not the original ones, but a modified

version of those by including some penalty terms. The algorithm is an iterative process

starting from U
0
and computing at each iteration k a collection of contingent policies ˆ

U
k
which are then aggregated into a collection of implementable policies U
k
that are ˆ


supposed to converge to the optimal solution U . The contingent policies U
k
are found as

optimal solutions to the modified scenario subproblems:


50


This term was coined by Rockafellar and Wets (1991). The idea is based on the theory
of the proximal point algorithm in nonlinear programming.



60

min F
s
( x
s
,u
s
) + w
s
u
s
+
1
µ u
s
÷ u
s
ˆ
2
2
(2.5.2)
where ? is the ordinary Euclidian norm, µ is a penalty parameter and w
s
is an


information price
51
. The use of µ is justified by the fact that the new contingent policy

should not depart too much from the implementable policy found in the previous

iteration. The modified scenario subproblems (2.5.2) have the form of an augmented

Lagrangian.

In the next subsection, I present a detailed description of the progressive hedging

algorithm, which uses subproblems in the form of an augmented Lagrangian as shown

above.





II.5.1. Description of the Progressive Hedging Algorithm


The optimal solution of the problem described by equations (2.3.1) - (2.3.2), U ,

represents the best response an optimizing agent can come up with in the presence of

uncertainty. An advantage of this algorithm is that one does not necessarily need to solve

subproblems (2.5.2) exactly. A good approximation
52
of the solution is enough in

allowing one to solve for the solution of the global problem.

Let U
k
denote a collection of admissible policies and W
k
a collection of

information prices corresponding to iteration k . The progressive hedging algorithm, as

designed by Rockafellar and Wets (1991), consists of the following steps:



51

52



I borrow this term from Rockafellar and Wets (1991).

One can envision transforming the scenario subproblems into quadratic problems by
using second order Taylor approximations.



61
Step 0. Choose a value for µ , W
0
and for U
0
. The value of µ may remain

constant throughout the algorithm but it can also be adjusted from iteration to iteration
53
.

Changing the value of µ may improve the speed of convergence. Throughout this

chapter, I will consider µ as being constant. U
0
can be composed of the contingent

policies u
s
(0
)
= u
1
s(0
)
,u
1
s(0
)
,...,u
T
s(0
)
obtained from solving all the scenarios subproblems,
( )

whether modified or not. W
0
can be initialized to zero, W
0
= 0 . Calculate the collection

of implementable policies, U
0
= JU
0
, where J is the aggregation operator
54
. ˆ

Step 1. For every scenario s e S , solve the subproblem:

min ¿
?
F
t
s
( x
s
t ,u
s
t ) + w
s
t u
s
t +
1
µ u
s
t ÷ u
s
t ? T
?
t=1 ?
2
ˆ 2?
?
(2.5.3)

For iteration k +1 , let u
s
(k+1
)
= u
1
s(k+1
)
,u
s
2(k+1
)
,...,u
T
s(k+1
)
denote the solution to the
( )

subproblem corresponding to scenario s . This contingent policy is admissible but not

necessarily implementable. Let U
k
+
1
be the collection of all contingent policies u
s
(k+1
)
.

Step2. Calculate the collection of implementable policies, U
k
+
1
= JU
k
+
1
. While ˆ

these policies are implementable, they are not necessarily admissible in some cases
55
. If

the policies obtained are deemed a good approximation, the algorithm can stop. A

stopping criterion should be employed in this step.

53

See Rockafeller and Wets (1991) and Helgason and Wallace (1991a, 1991b) for a
discussion on the values of µ . Rosa and Ruszczynski (1994) also provide some
algorithm for updating similar penalty parameters.

54

55

See the appendix for more details on the aggregation operator.

Contingent policies are always admissible. If the domain of admissible policies is
convex then any linear combination of the contingent policies will also belong to that
domain. As noted above, by definition, the aggregation operator is linear. Therefore, for a
convex problem the implementable policies computed in step 1 are also admissible.



62
Step3. Update the collection of information prices W
k
+
1
by the following rule:


W
k
+
1
= W
k
+ µ U
k
÷U
k
(
For each scenario s e S rule (2.5.4) translates into:
ˆ
)
(2.5.4)

w
s
t (k+1
)
= w
s
t (k
)
+ µ u
s
t (k
)
÷ u
s
t(k)
(
ˆ
)
for t = 1,...,T
(2.5.5)
This updating rule is derived from the augmented Lagrangian theory. In principle, the

rule can be changed with something else as long as the decomposition properties are not

altered.

Step 4. Reassign k := k +1 and go back to step one.

Next, I investigate how this methodology can be applied to a type of dynamic

programming problem closed to what is often employed by economists for their models.





II.6. Using Scenario Aggregation to Solve a Finite Horizon Life Cycle Model

In this section, I will take a closer look at the viability of scenario aggregation in

approximating a rational expectations model. I choose a standard finite horizon life cycle

model that has an analytical solution, which will be used as a benchmark for the

performance of the scenario aggregation method.

I start by presenting an algorithm for solving the problem given by (2.2.1) -

(2.2.4) under the assumption that the length of the horizon, T , and the number of

realizations of the forcing variable, n , are relatively small. The algorithm used is similar

to that developed by Rockefeller and Wets (1991). As mentioned above, the idea is to

split the problem into many smaller problems based on scenario decomposition and solve

those problems iteratively imposing the non-anticipativity constraint. For computational

convenience, I will reformulate the problem (2.2.1) - (2.2.4) as a minimization rather than


63
maximization. Hence, for each scenario s e S , represented by the sequence of

realizations y
s
= y
s
0 , y
1
s
,K, y
T
s
, the problem becomes:
( )

min
?
¿ |
t
?÷F
t
(c
s
t ) + w
s
t c
s
t +
1
µ (c
s
t ÷ c
s
t )
?
?
?
T
?
2
??
(2.6.1)

subject to
c
t
? t =0 ?
2
??

A
ts
= 1+ r
t
s
A
t

1
+ y
s
t ÷ c
s
t , t = 0,1,...,T
( ) (2.6.2)


Expressing c
st
and c
st
as a function of A
t
s
and A
t
s
, the augmented Lagrangian function,

for a fixed scenario s , becomes:

L
=
¿ |
t
÷F
t
?
(1+ r
t
s
) A
t

1
+ y
s
t ÷ A
ts
? + w
s
t
?
(1+ r
t
s
) A
t

1
+ y
s
t ÷ A
ts
? +
T

t =0
{
? ? ? ? (2.6.3)
+
1
µ ?(1+ r
t
s
) A
t

1
+ y
s
t ÷ A
ts
÷
(1+ r
t
s
) A
t

1
+ y
s
t ÷ A
ts
? ?
2?
( )
2
??
?
All the underlined variables in the above equations represent implementable policies or

states derived from applying implementable policies.

Before going through the steps of the algorithm, I will make a few assumptions

about the functional form of the utility function as well as about the interest rate. First, it

is assumed that preferences are described by a negative exponential utility function.

Hence:

F
t
(c
t
) = ÷ 1 exp (÷u c
t
)
u


(2.6.4)

where u is the risk aversion coefficient. Secondly, the interest rate, r
t
, is taken to be

constant. Finally, the distribution of the forcing variable is approximated by a discrete

counterpart. The realizations as well as the associated probabilities are obtained using a

Gauss-Hermite quadrature and matching the moments up to order two. The number of

points used to approximate the original distribution determines the number of scenarios.


64
By decomposing the original problem into scenarios, the subproblems become

deterministic versions of the original model.


II.6.1. The Algorithm

Given the assumptions made in the previous section, problem (2.6.1) becomes:


min
?
¿
?
? T |
t
?
1 exp ÷u c
s
+ w
s
c
s
+

1
µ c
s
÷ c
s
2 ??
c
t
? t =0
?u
(
t
) t
t
2 ( t
t
)
?
???
(2.6.5)
Consequently the Lagrangian for scenario s is:

L
=
¿ |
t
? 1 exp ?÷
u
((1+
r
) A
t

1
+ y
s
t ÷ A
t
s
)? + w
s
t
?
(1+
r
) A
t

1
+ y
s
t ÷ A
ts
? + T
t =0
?
?
u
? ? ? ? (2.6.6)
+
1
µ ?(1+
r
) A
t

1
+ y
s
t ÷ A
ts
÷
((1+
r
) A
t
÷
1
+ y
s
t ÷ A
t
)? ? 2
2? ??
?
Since the consumption variable was replaced by a function of the asset level, the

algorithm will be presented in terms of solving for the level of assets.

Step 0. Initialization: Set w
st
= 0 for all stages t and scenarios s . Choose a value

for µ that remains constant throughout the algorithm, let it be µ = 5 . Later on, in this

chapter, I will discuss the impact the value of µ has on the convergence process. At this

point, one needs a first set of policies. The convergence process, and implicitly the speed

of the algorithm, is impacted by the choice of the first set of policies.

One suggestion made in the literature by Helgason and Wallace (1991a, 1991b) is

to use the solution to the deterministic version of the model. This would amount to using

the certainty equivalence solution in this case. I will first implement the algorithm using

as starting point the certainty equivalence solution and then I will take advantage of the

fact that for certain specifications of the model each scenario subproblem has an exact

solution. I will then compare the convergence properties of the algorithm in these two

cases.


65

Let
{c
ceq
}
t
=
0
denote the solution to the deterministic problem. Then, using the T
t


transition equation (2.6.2) one can compute the level of assets for each scenario s ,

A
s
(0
)
= A
0
s(0
)
, A
1
s(0
)
,..., A
T
s(÷01
)
. Next, it becomes possible to compute the implementable
{ }
states A
(
0
)
= A
(
00
)
, A
1
(0
)
,...,
A
T
(0÷)1{
}


as a weighted average of A
0
s(0
)
corresponding to all


scenarios s , using as weights the probabilities of realization for each scenario.

Alternatively, one can compute the first set of contingent policies by solving a

deterministic life cycle consumption model for each scenario s :

mi
sn
¿ |
t
1 exp ÷u
?
(1+
r
) A
t

1
+ y
s
t ÷ A
ts
?
A
t
T

t =0
u
{
?
?
}
(2.6.7)

with A
(
÷s1
)
and A
T
(s
)
given. As before, let A
s
(0
)
= A
0
s(0
)
, A
1
s(0
)
,..., A
T
s(÷01
)
denote the solution
{ }

to this problem. This solution is admissible but not implementable. The implementable

solution for each period t , A
t
0
, is computed as the weighted average of all the contingent

solutions for period t , A
t
s(0
)
, with the weights being given by the probability of

realization for each particular scenario s .

Step 1. For every scenario s e S , solve the subproblem:

mi
sn
¿ |
t
? 1 exp ÷u
?
(1+
r
) A
t

1
+ y
s
t ÷ A
ts
? +
A
t
T

t =0
?
u
?
{
?
?}
+W
ts ?
(1+
r
) A
t

1
+ y
s
t ÷ A
ts
? +
? ?
(2.6.8)
+
1
µ (1+
r
) A
t

1
+ y
s
t ÷ A
ts
÷
?
(1+
r
) A
t
÷
1
+ y
s
t ÷ A
t
? ?
2
{
?
2
??
}
?
A detailed description of how the solution is computed can be found in the Appendix.

The advantage of the scenario aggregation method is that the solution to problem (2.6.8)

does not have to be computed exactly.




66

Let A
s
(k
)
= A
0
s(k
)
, A
1
s(k
)
,...,
A
T
s(÷k1) {

} denote the contingent solution to this problem,


where k denotes the iteration. Based on this solution I also compute the consumption

path for each scenario, c
s
(k
)
. This solution is admissible but not implementable and

therefore the next step is to compute the implementable solution based on the contingent

solutions A
s
(k
)
.

Step 2. First, compute the implementable states A
k
. As it was mentioned in step


0, A
t
k
is computed as the weighted average of all the contingent solutions for period t ,

A
t
s(k
)
, with the weights being given by the probability of realization for each particular

scenario s . Since the space of the solutions for the problem being solved is convex, the

implementable solution is also admissible. At this point, if solution A
k
is considered


good enough, the algorithm can stop and A
k
becomes officially the solution of the


problem described by (2.2.1) - (2.2.4). In order to make a decision on the viability of A
k


as the optimal solution, one needs to define a stopping criterion. Based on the value of

A
k
I compute the implementable consumption path c
k
and then use the following error

sequence
56
:

c(k
)
=
¿ |t
?
( c
(
tk
)
÷ c
(
tk÷1) )2 + A
(
tk
)
÷ A
(
tk÷1) ?
T

t =0
?
?
(
)
?
2
?
(2.6.9)

where k is the iteration number. The termination criterion is c (k
)
< d where d is

arbitrarily chosen. In the next section, I will discuss the importance of the stopping

criterion in determining the accuracy of the method.


56


This is similar to what Helgason and Wallace (1990a) proposed. Later on in this
chapter we will discuss the impact the choice of the value for d has on the results.



67
Step 3. For t = 0,1,...,T and all scenarios s update the information prices:


w
s
t (k+1
)
= w
s
t (k
)
+ µ ?(1+
r
) A
t
s÷(k
)
÷ A
t
(÷k1
)
÷ A
t
s(k
)
÷ A
t
(k
)
? for t = 1,...,T
?
(
1
)( )
?
Step 4. Reassign k := k +1 and go back to step one.




II.6.2. Simulation Results

In this section, I present a brief picture of the results obtained by the

implementation of the scenario aggregation method compared to the analytical solution.

These results show that the numerical approximation obtained through scenario

aggregation is close to the analytical solution for certain parameterizations of the model.

In order to asses the accuracy of the scenario aggregation method I will use several

criteria put forward in the literature. First, I compare the decision rule, i.e. the

consumption path obtained through scenario aggregation with the values obtained from

the analytical solution. In this context, I use two relative criteria similar to what Collard

and Juillard (2001) use. One, E
a
, gives the average departure from the analytical solution R

and is defined as:


a


T +
1
¿
c
*t1
T

c
*
t ÷ c
t
E= R
t =0
(2.6.10)


The other, E
m
, represents the maximal relative error and is defined as: R


E = max ? c
t
÷*c
t
m
?*
?
?
T
(2.6.11)
R
?c
?
? ?
t
?
t
=0 ?

where c
*
t is the analytical solution and c
t
is the value obtained through scenario

aggregation. Alternatively, since the problem is ultimately solved in terms of the level of

assets, the two criteria could also be expressed using the level of assets:


68

E
=
¿t*a

?
t
??T
÷1
R
1
T
÷
1
A
*
÷ A
t
, E
m
= max ? A
*
÷ A
t
T
t
=
0
A
t
R
? A
*
?
?
?
t
?
t
=0 ?


where A
*
is given by the analytical solution and A
t
by the scenario aggregation. Even t

though the scenario aggregation methodology does not use the Euler equation in

obtaining the solution, I will use the Euler equation based criteria proposed by Judd

(1998) as an alternative for determining the accuracy of the approximation. The criterion

is defined as a one period optimization error relative to the decision rule. The measure is

obtained by dividing the current residual of the Euler equation to the value of next

period's decision function. Subsequently, two different norms are applied to the error

term: one, E
a
, gives the average and the other, E
m
, supplies the maximum. Judd (1998)
E E

labeled these criteria as measures of bounded rationality.

The simulations were done using the following common set of parameter values:

the discount factor | = 0.96 ; the initial and terminal values for the level of assets

A
÷
1
= 500 and A
T
= 1000 ; the income generating process has a starting value of

y
0
= 200 . In addition, the interest rate is assumed deterministic. I used two values for the

interest rate, r = 0.04 and r = 0.06 . The distribution of the forcing variable was

approximated by a 3 point discrete distribution. As I mentioned in the description of the

progressive hedging algorithm, a few factors can influence the performance of the

scenario aggregation method. Let us first look at how the starting values and stopping

criterion influence the results.









69
II.6.2.1. Starting Values and Stopping Criterion

As I mentioned above, the starting values and the stopping criterion are very

important elements in the implementation of the algorithm. I consider for the moment

that the starting values are given by the certainty equivalence solution of the life cycle

consumption model. I analyze the case where the value for the coefficient of risk aversion

is u = 0.01, the variance for the income process is o2
y
= 100 and the interest rate is

r = 0.06 . The stopping criterion is given by the sequence c
(
k
)
as defined in (2.6.9) and I

arbitrarily choose d = 0.004 . Therefore when c
(
k
)
becomes smaller than d = 0.004 I stop

and declare the solution obtained in iteration k as the solution to the problem described

by (2.2.1) - (2.2.4). In Table 1 I provide the values for the accuracy measures discussed

above, using the level of assets, as opposed to the level of consumption. One can see that

the approximation to the analytical solution obtained by stopping when c
(
k
)
is smaller

than the arbitrarily chosen d is very good.

Table 1. Accuracy measures for d=.004
u

0.01
o
2
y

E
a
R

E
m
R

E
a
E

E
m
E


100 0.001445515 0.002392885 0.000005019 0.000008735















70
The results presented in Table 1 are obtained after 159 iterations. Next, I will look

at the behavior of the sequence c
(
k ) for the case presented above.



3.5

3

2.5

2

1.5

1

0.5

0


Evolution for the c(k)sequence



5

4

3

2

1

0



x 10
-3


Evolution for the c(k)sequence (zoom)
0 50 100 150 200 250 300 350 150 200 250 300
Iteration k Iteration k

Evolution of the value for Evolution of the value for
the objective function the objective function (zoom)
-111.5

-112

-112.5

-113

-113.5

-114

-114.5
-111.876

-111.878

-111.88

-111.882

-111.884

-111.886

-111.888

-115 -111.89
0 50 100 150 200 250 300 350 150 200 250 300 350
Iteration k Iteration k



Figure 3. Evolution of the c
(
k ) sequence and the value of the objective for u = .01 and o2
y
= 100



One can see in Figure 3 that the value for sequence c
(
k
)
continues to decrease

until iteration 250 when it attains the minimum value. At the same time, the value of the

objective continues to increase until iteration 266 when it attains its maximum. It is worth

noting that the value of the objective is computed as in equation (2.6.12). Based on these

observations one may elect to choose as stopping criterion the point where c
(
k
)
attains its

minimum or when the objective function attains its maximum as opposed to an arbitrary

value d . Next, I look at how close is the approximation to the analytical solution when



71
v
a
l
u
e

o
f

c
( k
) s
e
q
u
e
n
c
e


v
a
l
u
e

o
f

c
( k
) s
e
q
u
e
n
c
e


V
a
l
u
e

o
f

t
h
e

o
b
j
e
c
t
i
v
e

f
u
n
c
t
i
o
n


V
a
l
u
e

o
f

t
h
e

o
b
j
e
c
t
i
v
e

f
u
n
c
t
i
o
n


using these criteria. In Table 2 one can see that there is not much difference between the

last two criteria when compared to the analytical solution. The only difference is that the

value of the expected utility is marginally higher in the second case.

Table 2. Accuracy measures for various stopping criteria
u

0.01
o
2
y

E
a
R

E
m
R

E
a
E

E
m
E


Stopping criterion

100 0.001445515 0.002392885 0.000005019 0.000008735 Arbitrary d = 0.004

100 0.002137894 0.002691210 0.000007190 0.000013733 Minimum of c
(
k)

100 0.002137894 0.002691210 0.000007190 0.000013733 Maximum objective



A somewhat interesting result is that the ad-hoc stopping criterion d = 0.004

leads to a better approximation of the analytical solution. This is explained by the fact

that the progressive hedging algorithm leads to the solution that would be obtained

through the aggregation of the exact solutions for every scenario. Here the starting point

is the certainty equivalent solution and the path to convergence, at some point, is very

close to the analytical solution.





II.6.3. The Role of the Penalty Parameter

In the implementation of the progressive hedging algorithm, I chose the penalty

parameter to be constant. Its role is to keep the contingent solution for each iteration close

to the previous implementable policy. However, its value also has an impact on the speed

of convergence. I will now consider the previous parameterization of the model and I am



72
going to change the value of the penalty parameter to see how it changes the speed of

convergence. In Figure 4 one can see that as µ increases so does the number of iterations

needed to achieve convergence. While a higher value of the penalty parameter helps the

convergence of contingent policies to the implementable policy, it also slows the global

convergence process, requiring more iterations.

Evolution for the c(k)sequence Evolution for the c(k)sequence
-3 for µ = 0.1 -5x 10 x 10 for µ = 0.5
5 3

2.54

2
3
1.5
2
1

10.5

0 0
150 200 250 300 1000 1200 1400 1600 1800 2000 2200
Iteration k Iteration k

Evolution for the c(k)sequence Evolution for the c(k)sequence
x 10
-9
for µ = 2 for µ = 5 x 10
-11

4 8
3.5 7
3 6
2.5 5
2 4
1.5 3
1 2
0.5 1
0 0
7000 7500 8000 8500 9000 9500 2 2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8
Iteration k Iteration k x 10
4




Figure 4. Convergence for different values of the penalty parameter.


For µ = 0.1 , 250 iterations are needed to achieve convergence, while for µ = 0.5 ,

1780 iterations are needed. For higher values, such as µ = 5 , the number of iterations

needed to achieve convergence increases to over 25000 iterations.







73
v
a
l
u
e

o
f

c
( k
) s
e
q
u
e
n
c
e


v
a
l
u
e

o
f

c
( k
) s
e
q
u
e
n
c
e


V
a
l
u
e

o
f

t
h
e

o
b
j
e
c
t
i
v
e

f
u
n
c
t
i
o
n


V
a
l
u
e

o
f

t
h
e

o
b
j
e
c
t
i
v
e

f
u
n
c
t
i
o
n


II.6.4. More simulations

In this section I investigate how close the scenario aggregation solution is to the

analytical solution for various parameters. Table 3 shows the values for the four criteria

enumerated above for different values of the coefficient of risk aversion and of the

variance of the random variable entering the income process. All the simulations whose

results are presented in Table 3 were done using a three point approximation of the

distribution of the random variable entering the income process. The relative measures

are computed using the level of assets.

Table 3. Accuracy measures for various parameters when interest rate r=0.04
u

o2y 0.01 0.05 0.1

E
a
R

E
m
R

E
a
E

E
m
E

E
a
R

E
m
R

E
a
E

E
m
E

E
a
R

E
m
R

E
a
E

E
m
E


1 .0000 .0000 .0000 .0000 .0001 .0001 .0000 .0000 .0002 .0003 .0000 .0000

4 .0000 .0001 .0000 .0000 .0004 .0005 .0000 .0000 .0009 .0011 .0000 .0000

25 .0005 .0007 .0000 .0000 .0029 .0037 .0000 .0000 .0058 .0074 .0000 .0000

100 .0023 .0029 .0000 .0000 .0116 .0147 .0000 .0000 .0230 .0290 .0000 .0000




For lower values of the coefficient of risk aversion the approximation is relatively good.

As the coefficient of risk aversion increases in tandem with the variance of the income

process, the accuracy suffers when looking at relative measures. The Euler equation

measure still indicates a very good approximation.

Let us now look at how this approximation affects the value of the original

objective, i.e. the expected discounted utility over the lifetime horizon. Table 4 shows the





74
ratio of the expected utilities for the whole horizon with the scenario aggregation as the

as the denominator and the analytical solution as the numerator.


Table 4. The ratio of lifetime expected utilities

F
as

F
sc
u

o2y 0.01 0.03 0.05 0.1


1 1.00000 1.00000 1.00000 1.00003

4 1.00000 1.00001 1.00003 1.00058

25 1.00000 1.00002 1.00141 1.02027

100 1.00003 1.00051 1.02273 1.39364




The discounted utilities are computed as in the original formulation of the problem:

F
sc
= 1


N
¿1 ?
?
¿
0
??
u
(
t
)
?
??? ?÷
T |
t
? 1 exp ÷u c
i
??


(2.6.12)

and
N
i= t=

F
as
= 1
N

?÷T |
t
? 1 exp ÷u c
i
* ?? (2.6.13)
N
¿1 ?
?
¿
0
??
u
(
t
)
?
???
i= t=

where N is the number of simulations, F
sc
is the discounted utility obtained with

scenario aggregation and


F
as
is the discounted utility obtained with the analytical

solution. In this formulation, both quantities are negative so their ratio is positive. Note

however that the initial formulation of the problem using the objective function specified

in (2.6.12) and (2.6.13) was a maximization. Therefore, higher ratio in Table 4 means that

the solution obtained through scenario aggregation leads to higher discounted lifetime

utility than the analytical solution. I simulate 2000 realizations of the income process and

then I average the discounted utilities over this sample. The result shows that the solution


75
obtained through scenario aggregation leads to higher overall expected utility as the

coefficient of risk aversion increases. This is explained by the fact that the level of

consumption in the first few periods is higher in the case of scenario aggregation. In the

context of a short horizon, this leads to higher levels of discounted utility.





II.7. Final Remarks

The results show that scenario aggregation can be used to provide a good

approximation to the solution of a life-cycle model for certain values of the parameters.

There are a few remarks to be made regarding the convergence. As pointed out earlier in

this chapter the value of µ has an impact on the speed of convergence. Higher values of

µ lead to faster convergence of the contingent policies towards an implementable policy

but that also means that the overall convergence is slower and hence it impacts the

accuracy if an ad-hoc stopping criterion is used. Therefore, one needs to choose carefully

the values of the ad-hoc parameters. On the other hand, if the scenario problems have an

exact solution then the final implementable policy can be obtained through a simple

weighted average with the weights being the probabilities of realization for each scenario.



















76
Chapter III. Impact of Bounded Rationality
57
on the Magnitude of


Precautionary Saving



III.1. Introduction

It is fair to say that nowadays the assumption of rational expectations has become

routine in most economic models. Recently, however, there has been an increasing

number of papers, such as Gali et al. (2004), Allen and Carroll (2001), Krusell and Smith

(1996), that have modeled consumers using assumptions that depart from the standard

rational expectations paradigm. Although they are not explicitly identified as modeling

bounded rationality, these assumptions clearly take a bite from the unbounded rationality,

which is the standard endowment of the representative agent. The practice of imposing

limits on the rationality of agents in economic models is part of the attempts made in the

literature to circumvent some of the limitations associated with the rational expectations

assumption. Aware of its shortcomings, even some of the most ardent supporters
58
of the

rational expectations paradigm have been looking for possible alterations of the standard

set of assumptions. As a result, a growing literature in macroeconomics is tweaking the

unbounded rationality assumption resulting in alternative approaches that are usually

presented under the umbrella of bounded rationality.




57





The concept of bounded rationality in this chapter should be understood as a set of
assumptions that departs from the usual rational expectation paradigm. Its meaning will
become clear later in the chapter when the underlying assumptions are spelled out.

58

Sargent (1993) for example, identifies several areas in which bounded rationality can
potentially help, such as equilibrium selection in the case of multiple possible equilibria
and behavior under "regime changes".



77
One may ask why is there a need to even consider bounded rationality. First,

individual rationality tests led various researchers to "hypothesize that subjects make

systematic errors by using ... rules of thumb which fail to accommodate the full logic of a

decision" (J. Conlisk, 1996). Secondly, some models assuming rational expectations fail

to explain observed facts, or their results may not match empirical evidence. Since most

of the time models include other hypotheses besides the unbounded rationality

assumption, the inability of such models to explain certain observed facts could not be

blamed solely on rational expectations. Yet, it is worth investigating whether bounded

rationality plays an important role in such cases. Finally, as Allen and Carroll (2001)

point out, even when results of models assuming rational expectations match the data, it

is still worth asking the question of how can an average individual find the solution to

complex optimization problems that until recently economists could not solve. To

summarize, the main idea behind this literature is to investigate what happens if one

changes the assumption that agents being modeled have a deeper understanding of the

economy than researchers do, as most rational expectations theories assume. Therefore,

instead of using rational expectations, it is assumed that economic agents make decisions

behaving in a rational manner but being constrained by the availability of data and their

ability to process the available information.

While the vast literature on bounded rationality continues to grow, there is yet to

be found an agreed upon approach to modeling rationally bounded economic agents.

Among the myriad of methods being used, one can identify decision theory, simulation-

based models, artificial intelligence based methodologies such as neural networks and

genetic algorithms, evolutionary models drawing their roots from biology, behavioral




78
models, learning models and so on. Since there is no standard approach to modeling

bounded rationality, most of the current research focuses on investigating the importance

of imposing limits on rationality, as well as on choosing the methods to be used in a

particular context. When modeling consumers, the method of choice so far seems to be

the assumption that they follow some rules of thumb
59
. Instead of imposing some rules of

thumb, my approach in modeling bounded rationality focuses on the decision making

process. I borrow the idea of scenario aggregation from the multistage optimization

literature and I adapt it to fit, what I believe to be, a reasonable description of the decision

making process for a representative consumer. Besides the decision making process per

se, I also add a few other elements of bounded rationality that have to do with the ability

to gather and process information.

In the previous chapter, the method of scenario aggregation was introduced as an

alternative method for solving non-linear rational expectation models. Even though it

performs well in certain circumstances, the real advantage of the scenario aggregation

lays in a different area. Its structure presents itself as a natural way to describe the

process through which a rationally bounded agent, faced with uncertainty, makes his

decision. In this chapter, I consider several versions of a life-cycle consumption model

with the purpose of investigating how the magnitude of precautionary saving changes

with the underlying assumptions on the (bounded) rationality of the consumer.








59








Some of the examples are Gali et al. (2004), Allen and Carroll (2001), Lettau and
Uhlig (1999) and Ingram (1990).



79
III.2. Empirical Results on Precautionary Saving


There seems to be little agreement in the empirical literature on precautionary

saving, especially when it comes to its relationship to uncertainty. Skinner (1988) found

that saving was lower than average for certain groups
60
of households that are perceived

to have higher than average income uncertainty. In the same camp, Guiso, Jappelli and

Terlizzese (1992), using data from the 1989 Italian Survey of Household Income and

Wealth, found little correlation between the level of future income uncertainty and the

level of consumption
61
. In addition, Dynan (1993), using data from the Consumer

Expenditure Survey, estimated the coefficient of relative prudence and found it to be "too

small to be consistent with widely accepted beliefs about risk aversion".

On the other hand, Dardanoni (1991) basing his analysis on the 1984 cross-

section of the UK FES (Family Expenditure Survey) suggested that the majority of

saving in the sample arises for precautionary motives. He found that average

consumption across occupation and industry groups was negatively related to the within

group variance of income. Carroll (1994) found that income uncertainty was statistically

important in regressions of current consumption on current income, future income and

uncertainty. Using UK FES data, Merrigan and Normandin (1996) estimated a model

where expected consumption growth is a function of expected squared consumption

growth and demographic variables and their results, based on the period 1968-1986,


60

61



Specifically, the groups identified were farmers and self-employed.

In fact the study on Italian consumers did find that consumption was marginally lower
while wealth was marginally higher for those who were facing higher income uncertainty in
the near future.





80
indicate that precautionary saving is an important part of household behavior. Miles

(1997), using several years of cross-sections of the UK micro data and regressing

consumption on several proxies for permanent income and uncertainty, found that, for

each cross-section, the latter variable played a statistically significant role in determining

consumption. In a study trying to measure the impact of income uncertainty on household

wealth, Carroll and Samwick (1997), using the Panel Study of Income Dynamics, found

that about a third of the wealth is attributable to greater uncertainty. Later on, Banks et al.

(2001), exploiting not only the cross-sectional, but also the time-series dimension of their

data set, find that section specific income uncertainty as opposed to aggregate income

uncertainty plays a role in precautionary saving. Finally, Guariglia (2001) finds that

various measures of income uncertainty have a statistically significant effect on savings

decisions.

In this chapter, I am going to show that, by introducing bounded rationality in a

standard life cycle model, one can increase the richness of the possible results. Even if

the setup of the model would imply the existence of precautionary savings, under certain

parameter values and rules followed by consumers, the precautionary saving is apparently

almost inexistent. As opposed to most of the literature
62
studying precautionary savings, I

introduce uncertainty in the interest rate, beside income uncertainty. In this context, the

size of precautionary saving no longer depends exclusively on income uncertainty.









62










A notable exception is Binder et al. (2000).



81
III.3. The Model


I start this section by presenting the formulation of a standard finite horizon life-

cycle consumption model. Then I will introduce a form of bounded rationality
63
and

investigate the path for consumption and savings.

Consider the finite-horizon life-cycle model under negative exponential utility.

Suppose an individual agent is faced with the following intertemporal optimization

problem:

ma
T
x E
?
¿ ÷|
t
1 exp (÷u c
t
) | I
0
? T

subject to
{c
t
}
t
=0
??
t =0
u ?
?
(3.3.1)

A
t =
(1+ r
t
) A
t
÷
1
+ y
t
÷ c
t
, t = 0,1,...,T ÷1, (3.3.2)

A
t
> ÷b, with A
-1
, A
T
given (3.3.3)

where u is the coefficient of risk aversion, A
t
represents the level of assets at the

beginning of period t , y
t
the labor income at time t , and c
t
represents consumption in

period t . The initial and terminal conditions, A
÷
1
and A
T
are given. The information set

I
0
contains the level of consumption, assets, labor income and interest rate for period

zero and all previous periods. The labor income is assumed to follow an arithmetic

random walk:



63




As it was already mentioned above, the approach in defining bounded rationality in this
chapter has some similarities to the approach followed by Lettau and Uhlig (1999) in the
sense that several rules are used to account for the inability of the boundedly rational
agent to optimize over long horizons.



82
y
t
= y
t
÷
1
+ ç
t
, t = 1,...,T , with y
o
given (3.3.4)

and ç
t
being drawn from a normal distribution, ç
t
~
N
(0,o2
y
) . When the interest rate is

deterministic, this problem has an analytical solution
64
. However, if the interest rate is

stochastic, the solution of this finite horizon life cycle model becomes more complicated

and it can not be computed analytically. For now, I will not make any particular

assumption about the process generating the interest rate. Therefore, to summarize the

model, a representative consumer derives utility in period t from consuming c
t
,

discounts future utility at a rate | and wants, in period zero, to maximize his present

discounted value of future utilities for a horizon of T +1 periods. At the beginning of

each period t the consumer receives a stochastic labor income y
t
, finds out the return r
t


on his assets A
t
÷
1
, from the beginning of period t ÷1 to the beginning of period t , and, by

choosing c
t
, determines the level of assets A
t
according to equation (3.3.2).

Now, I introduce a rationally bounded agent in the following way. First, I assume

that the agent does not have either the resources or the sophistication to be able to

optimize over a long horizon. For example, if the agent enters the labor force at time zero

and faces the problem described by (3.3.1) - (3.3.4) over a time span extending until his

retirement, let it be period T , the assumption is that the agent does not have the ability to

optimally choose, at time zero, a consumption plan over that span. Instead, he focuses on

choosing a consumption plan over a shorter horizon, let it be T
h
+1 periods.

Secondly, because of his limited ability to process large amounts of information

he repeats this process every period in order to take advantage of any new available


64



See the appendix for a detailed description of the analytical solution.



83
information. This idea of a shorter and shifting optimization horizon is similar to the

approach taken by Prucha and Nadiri
65
(1984, 1986, and 1991). Now, the question is how

an individual who lacks sophistication, can optimally
66
choose a consumption plan even

for a short time span. In order to model the decision process I make use of the scenario

aggregation method. Under this assumption, the agent evaluates several possible paths

based on the realization of the forcing variables specified in the model. By assigning

probabilities to each of the possible paths, the agent is in the position to aggregate the

scenarios (paths), i.e., to compute the expected value for his decision.

In order to be able to use the scenario aggregation method, the forcing variables

need to have a discrete distribution but in the model presented above, they are described

as being drawn from a normal distribution. This leads to the third element that can be

brought under the umbrella of bounded rationality. Since the agent has limited

computational ability, the distribution of the forcing variable is approximated by a

discrete distribution with the same mean and variance as the original distribution. This

approximation does not necessarily have to be viewed as a bounded rationality element

since similar approaches have been employed repeatedly in numerical solutions using

state space discretization
67
.

Given the assumptions made about the abilities of the rationally bounded

representative agent, I will now go through the details of solving the problem described



65



In their work, a finite and shifting optimization horizon is used to approximate an
infinite horizon model.

66

67

Optimality here means the best possible solution given the level of ability.

Tauchen, among others, used this kind of approximation on various occasions, such as
Tauchen (1990), Tauchen and Hussey (1991).



84
by equations (3.3.1) - (3.3.4). Hence, at every point in time, t , the agent solves the

problem:

ma}xh
E
?
¿ ÷|
t
1 exp (÷u c
t
+
t
) | I
t
?
for
t = 0,1,...,T ÷ T
h
T
h

or
{c
t
+tT=0t
?
?
t
=0
u ?
?
(3.3.5)

ma}x
÷
t
E
?
¿ ÷|
t
1 exp (÷u c
t
+
t
) | I
t
? for t = T ÷ T
h
+1,...,T ÷1 T ÷t

subject to
{c
t
+t
T
=0t
??
t
=
0
u ?
?
(3.3.6)

A
t
+
t
=
(1+ r
t
+
t
) A
t
+t÷
1
+ y
t
+
t
÷ c
t
+
t
,
(3.3.7)
t = 0,1,...,T ÷1, t = 0,..., min (T
h
,T ÷
t
)

with A
-1
, A
t
-1, A
t
+T
h
and A
T
given (3.3.8)

where A
t
+
t
represents the level of assets at the beginning of period t +t , y
t
+
t
the labor

income at time t +t , and c
t
+
t
represents consumption in period t +t . The initial and

terminal conditions, A
-1
, A
t
÷
1
, A
t
+T
h
and A
T
are given. The information set I
t
contains the

level of consumption, assets, labor income and interest rate for period t and all previous

periods. The labor income is assumed to follow an arithmetic random walk:

y
t
+
t
= y
t
+t ÷
1
+ ç
t
b+
t
,

t = 1,...,T , t = 0,..., min (T
h
,T ÷
t
)
with y
0
given


(3.3.9)

ç
t
b+
t
being drawn from a discrete distribution,
D
(0,o2
y
) with a small number of


realizations.

In making the above assumptions, the belief is that they would better describe the

way individuals make decisions in real life. It is often the case that plans are made for

shorter horizons, but not entirely forgetting about the big picture.






85
Recalling the results of Skinner (1988) who found that saving was lower than

average for farmers and self employed, groups that are otherwise perceived to have

higher than average income uncertainty, one can assume that planning for those groups

does not follow the recipe given by the standard life cycle model. Given the high level of

uncertainty, I believe it would be more appropriate to model these consumers as if they

plan their consumption path only for a short period of time and then reevaluate. This

would be consistent with the fact that farmers change their crop on a cycle of several

years and may be influenced by the fluctuations in the commodities markets and other

government regulations. Similarly, some among the self employed are likely to have

short term contracts and are more prone to reevaluate their strategy on a high frequency

basis. Therefore, the model above seems like a good description on how the decision

making process works. The only detail that remains to be decided is how the consumer

chooses the short horizon terminal condition, that is, the level of assets, or the wealth. For

this purpose, I propose three different rules and I investigate their effect on the saving

behavior.

So far, no assumption has been made about the process governing the realizations

of the interest rate. From now on, I assume that the interest rate is also described by an

arithmetic random walk:

r
t
= r
t
÷
1
+u
t
, t = 1,...,T , with r
o
given (3.3.10)
Since in this formulation the problem does not have an analytical solution, the classical

approach would be to employ numerical methods in order to describe the path of

consumption, even for a very short horizon. In order to find the solution corresponding to

the model incorporating the bounded rationality assumption I will use the scenario




86
aggregation
68
methodology. Then I will compare this solution with the numerical

solution
69
that would result from the rational expectation version of the model when

optimizing over the whole T period horizon.


III.3.1. Rule 1

Under rule 1, the consumer considers several possible scenarios for a short

horizon and assumes that for later periods certainty equivalence holds. In this context, he

makes a decision for the current period and moves on to the next period when he

observes the realization of the forcing variables. Then he repeats the process by making a

decision based on considering all the relevant scenarios for the near future and assuming

certainty equivalence for the distant future. Hence, the decision making process takes

place every period. More precisely, when optimizing in period t , the consumer considers

all the scenarios in the tree event determined by the realizations of the forcing variable

for the first T
h
periods. From period t + T
h
he considers that certainty equivalence holds

for the remaining T ÷ t ÷ T
h
periods. This translates specifically to considering that

income and interest rate are frozen for each existing scenario for the remaining T ÷ t ÷ T
h


periods. To be more specific, for time t = 0 , the consumer considers all the scenarios

available in the event tree for the first T
h
periods and assumes certainty equivalence for

68

Since an analytical solution can be obtained when income follows an arithmetic
random walk and interest rate is deterministic, it is not necessary to discretize both
forcing variables, but only the interest rate. This approach reduces considerably the
computational burden. A short description on the methodology used along with the
solution for one scenario with deterministic, interest rate is presented in the appendix.
More details on the scenario aggregation methodology can be found in the second
chapter.

69

The numerical solution is obtained using projection methods and is due to Binder et al.
(2000).



87
the remaining T ÷ T
h
periods. When it advances to period t = 1, he optimizes again

considering all the scenarios available in the tree event for periods 1, 2,...,T
h
+1 and

assumes certainty equivalence for the remaining T ÷ T
h
÷1 periods.

In fact, this rule can be considered as an extension to the scenario aggregation

method in order to avoid the dimensionality curse. One may recall that due to its

structure, the number of scenarios in the scenario aggregation method increases

exponentially with the number of periods. In effect, this rule is limiting the number of

scenarios considered and it is consistent with a rationally bounded decision maker who

can only consider a limited and, most likely, low number of possible scenarios.

Following are some graphical representations of the simulations for rule 1. Each

graph contains the values for the coefficient of risk aversion, u . The graphs also contain

the numerical solution and, for comparison purposes, the evolution of assets if the

solution were computed in the case of certainty equivalence. I first consider a group of 12

cases varying certain parameters of the model. For all simulations in this group, the total

number of periods considered is T = 40 and the optimizing horizon is T
h
= 6 . The

starting level of income is y
0
= 200 , the initial level of assets is A
÷
1
= 500 while the

terminal value is A
T
= 1000 . The discount factor is | = 0.96 , the starting value for the

interest rate, r
0
= 0.06 while the standard deviation for the interest process is given by

o
r
= 0.0025. I use a discrete distribution with three possible realizations to approximate

the original distribution of the forcing variable and that implies that in each period t , for

t s T ÷ T
h
= 34 , the optimization process goes over 3
T
h = 729 scenarios. For periods

34 = T ÷ T
h
< t s T ÷1 = 39 the number of scenarios considered decreases to 3
T
÷
t
. The



88
parameters that are changing in the simulations are the variance for the income process

and the coefficient of risk aversion. I consider all cases obtained combining three values

for the standard deviation of income, o
y e
{1, 5, 10} and four values for the coefficient

of risk aversion, u
e
{0.005, 0.01, 0.05,
0.1
} . The results presented in this section as well


as for the rest of the chapter are based on 1000 simulations. This means that for both the

income generating process and the interest rate generating process, I consider 1000

realizations for each period. The decision to use only 1000 realizations was based on the

observation that the sample drawn provided a good representation of the arithmetic

random walk process assumed in the model. Specifically, both the mean and the standard

deviation of the sample were close to their theoretical values.

Some general results have emerged from all these simulations. First, the path for

the level of assets for the solution obtained in the bounded rationality case always lies

below the path for the level of assets for the numeric al solution obtained in the rational

expectation case. Consequently, the consumption path in the bounded rationality case

starts with values of consumption higher than in the rational expectations case.

Eventually the paths cross and the consumption level in the rational expectations case

ends up being higher toward the end of the horizon.

















89


360
340
320
300
280
260
240
220
200
180
160
ConsumptionPathfor u =0.005
numerical solution
boundedrationality


360
340
320
300
280
260
240
220
200
180
160
ConsumptionPathfor u =0.01
numerical solutionbounded
rationality

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40



360
340
320
300
280
260
240
220
200
180
160
Time t (periods)

ConsumptionPathfor u =0.05
numerical solution
boundedrationality



360
340
320
300
280
260
240
220
200
180
160
Time t (periods)

ConsumptionPathfor u =0.1
numerical solutionbounded
rationality

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
Time t (periods) Time t (periods)


Figure 5. Consumption paths for o
y
= 1, r
0
= 0.06 and o
r
= 0.0025.

One can see in Figure 5 that consumption is increasing over time for both

solutions, with the steepest path corresponding to the lowest value of the coefficient of

risk aversion.

When looking at the asset path for the same value of the standard deviation of the

income process, one notices in Figure 6 that the level of saving in the certainty

equivalence case is mostly higher than the level of saving obtained in the bounded

rationality case as well as under the rational expectations assumption.











90
C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l



1600

1400

1200

1000

800

600
Level of Assets for u =0.005
1600

1400

1200

1000

800

600
Level of Assets for u =0.01

400

numerical solution
boundedrationality
certaintyequivalence

400

numerical solution
boundedrationality
certaintyequivalence

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40



1600

1400

1200

1000

800

600
Time t (periods)

Level of Assets for u =0.05



1600

1400

1200

1000

800

600
Time t (periods)

Level of Assets for u =0.1

400

numerical solution
boundedrationality
certaintyequivalence

400

numerical solution
boundedrationality
certaintyequivalence

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
Time t (periods) Time t (periods)


Figure 6. Asset paths for o
y
= 1, r
0
= 0.06 and o
r
= 0.0025.

While for lower levels of the coefficient of risk aversion u
e
{0.005,
0.01
} , the


asset path obtained assuming certainty equivalence crosses under the other two paths in

the later part of the horizon, the same is not true for higher values of the coefficient of

risk aversion, u
e
{0.05,
0.1
} .


It is not only the relative position of the three paths that changes in the context of

an increasing coefficient of risk aversion, but also the absolute size of the level of

savings. Moreover, the shape of the paths for both the rational expectation and bounded

rationality case changes from concave to convex.

I present now a new set of simulations with the standard deviation of income

being increased to o
y
= 5 . One can see in Figure 7 that the consumption paths for



91
A
s
s
e
t

l
e
v
e
l


A
s
s
e
t

l
e
v
e
l


A
s
s
e
t

l
e
v
e
l


A
s
s
e
t

l
e
v
e
l


u
e
{0.005,
0.01
} are not much different from those presented in Figure 5 while for

higher values of the risk aversion coefficient, u
e
{0.05,
0.1
} , the consumption paths are


steeper than in the previous case.

Looking now at the level of savings, one notices in Figure 8 a similar change to

that observed in the case of consumption. While not much has changed for the lower

values of the coefficient for risk aversion, the asset paths for higher values of the risk

aversion coefficient, u
e
{0.05,
0.1
} , have changed, effectively becoming concave, as


opposed to convex in the previous case. Besides the concavity change, one can observe

that for u = 0.1 the level of assets resulting from the numerical approximation of the

rational expectations model is higher than in the case of certainty equivalence for the

bigger part of the lifetime horizon.



360
340
320
300
280
260
240
220
200
180
160


Consumption Path for u=0.005
numerical solution
bounded rationality



360
340
320
300
280
260
240
220
200
180
160


Consumption Path for u=0.01
numerical solution
bounded rationality

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40



360
340
320
300
280
260
240
220
200
180
160
Time t (periods)

Consumption Path for u=0.05
numerical solution
bounded rationality



360
340
320
300
280
260
240
220
200
180
160
Time t (periods)

Consumption Path for u=0.1
numerical solution
bounded rationality

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
Time t (periods) Time t (periods)


Figure 7. Consumption paths for o
y
= 5 , r
0
= 0.06 and o
r
= 0.0025.




92
C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l



1600

1400

1200

1000

800

600
Level of Assets for u=0.005
1600

1400

1200

1000

800

600
Level of Assets for u=0.01

numerical solution numerical solution
400 bounded rationality 400 bounded rationality
certainty equivalence certainty equivalence

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40



1600

1400

1200

1000

800

600
Time t (periods)

Level of Assets for u=0.05



1600

1400

1200

1000

800

600
Time t (periods)

Level of Assets for u=0.1

numerical solution numerical solution
400 bounded rationality 400 bounded rationality
certainty equivalence certainty equivalence

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
Time t (periods) Time t (periods)


Figure 8. Asset paths for o
y
= 5 , r
0
= 0.06 and o
r
= 0.0025.

By raising the variance of the income again, one can see in Figure 9 that the path

for consumption becomes a lot steeper for u
e
{0.05,
0.1
} . On the other hand, there

seems to be little change in the consumption pattern for u = 0.005 .

On the savings front, the level of precautionary saving increases tremendously for

the highest coefficient of risk aversion, u = 0.1, and quite substantially for u = 0.05 .

Consequently, in these two cases, the level of savings for the rational expectation model,

as well as the bounded rationality version, becomes noticeably higher than what certainty

equivalence produces. Yet, the level of savings continues to be higher for the much lower

coefficient of risk aversion, u = 0.005 , when compared with the savings pattern for

u = 0.01 and u = 0.05 .







93
A
s
s
e
t

l
e
v
e
l


A
s
s
e
t

l
e
v
e
l


A
s
s
e
t

l
e
v
e
l


A
s
s
e
t

l
e
v
e
l



400

350

300

250

200

150
Consumption Path for u=0.005
numerical solution
boundedrationality

400

350

300

250

200

150
Consumption Path for u=0.01
numerical solution bounded
rationality

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40



400

350

300

250

200

150
Time t (periods)

Consumption Path for u=0.05
numerical solution
boundedrationality



400

350

300

250

200

150
Time t (periods)

Consumption Path for u=0.1
numerical solution bounded
rationality

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
Time t (periods) Time t (periods)


Figure 9. Consumption paths for o
y
= 10 , r
0
= 0.06 and o
r
= 0.0025.

Another interesting observation is that if one compares the level of savings from

the panel corresponding to u = 0.05 and o
y
= 10 in Figure 10, to the level of savings

from the panel corresponding to u = 0.005 and o
y
= 1 in Figure 6, the two are almost the


same, if not the later higher. This is to say that for values of coefficient of risk aversion

and of standard deviation for income ten times as high as the ones in Figure 6, the level

of precautionary saving is almost unchanged.















94
C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


Level of Assets for u =0.005 Level of Assets for u =0.01

1800 1800

1600 1600

1400 1400

1200 1200

1000 1000

800 800

600 600numerical
solution numerical
solution
boundedrationality boundedrationality400 400
certaintyequivalence certaintyequivalence

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
Time t (periods) Time t (periods)

Level of Assets for u =0.05 Level of Assets for u =0.1

1800 1800

1600 1600

1400 1400

1200 1200

1000 1000

800 800

600 600
numerical solution numerical solution
400 boundedrationality 400 boundedrationality
certaintyequivalence certaintyequivalence

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
Time t (periods) Time t (periods)


Figure 10. Asset paths for o
y
= 10 , r
0
= 0.06 and o
r
= 0.0025.

As a general observation, it seems that the level of precautionary saving derived

from the rational expectation model is consistently higher, even if not by high margins,

than the level of savings obtained in the case of bounded rationality. For consumption,

the paths can be steeper or flatter but the general allure remains the same. The rationally

bounded consumer tends to start with a higher consumption while after a few periods the

unboundedly rational consumer tends to take over and continue to consume more until

the end of the horizon.











95
A
s
s
e
t

l
e
v
e
l


A
s
s
e
t

l
e
v
e
l


A
s
s
e
t

l
e
v
e
l


A
s
s
e
t

l
e
v
e
l


III.3.2. Rule 2

Under rule 2, the consumer considers all the relevant scenarios for the immediate

short horizon and then, for the later periods, he only takes in account what I call the

extreme cases. Rule 2 is similar to rule 1 in the way the decision maker emphasizes the

importance of scenarios only for the short term horizon. The difference is that under rule

2, rather than assuming certainty equivalence for the later periods, the consumer

considers the extreme case scenarios as a way of hedging against uncertainty in the

distant future. More precisely, when optimizing in period t , the consumer considers all

the scenarios in the event tree determined by the realizations of the forcing variable for

the first T
h
periods but then he becomes selective and only considers the extreme cases
70


for the remaining T ÷ t ÷ T
h
periods. To be more specific, for time t = 0 , the consumer

considers all the scenarios available in the event tree for the first T
h
periods and only the

extreme cases for the remaining T ÷ T
h
periods. When it advances to period t = 1, he

optimizes again considering all the scenarios available in the tree event for periods

1, 2,...,T
h
+1 and only the extreme cases for the remaining T ÷ T
h
÷1 periods.

In fact, this rule can also be considered as an extension to the scenario

aggregation method in an attempt order to avoid the dimensionality curse. One may recall

that due to its structure, the number of scenarios in the scenario aggregation method

increases exponentially with the number of periods. This rule is in fact limiting the

number of scenarios considered by trying to keep intact the possible variation in the

forcing variable. As opposed to rule 1 where from time t + T
h
the assumption is that the

70


The notion of extreme cases covers scenarios for which the realization of the forcing
variable remains the same. For more details see section 0 in the appendix.



96
forcing variable keeps its unconditional mean value, that is, zero, until the end of the

horizon, this rule expands the number of scenarios by adding all the extreme case

scenarios stemming from the nodes existent at time t + T
h
. This expansion can also be

seen as the equivalent of placing more weight on the tails of the original distribution of

the forcing variable. This rule is consistent with a rationally bounded decision maker who

can only consider a limited and, most likely, low number of possible scenarios but wants

to account for the variance in the forcing variable in the later periods of the optimization

horizon.

Following are some graphical representations of the simulations for rule 2. The

graphs depicting the consumption paths contain the bounded rationality solution as well

as the numerical solution. For comparison purposes, the graph panels containing the

evolution of assets display the savings pattern resulting from the solution obtained in the

case of certainty equivalence on top of the solutions for the rational expectations and the

bounded rationality models.

As in the case of rule 1, one can see in Figure 11 that consumption is increasing

over time for both solutions, with the steepest path corresponding to the lowest value of

the coefficient of risk aversion.

As opposed to the previous rule, the rationally bounded consumer does not always

start with a higher level of consumption. In fact, in this panel, for u = 0.05 and u = 0.1,

the solution of the rational expectations model has higher starting values for

consumption.








97


360
340
320
300
280
260
240
220
200
180
160
ConsumptionPathfor u =0.005
numerical solution
boundedrationality


360
340
320
300
280
260
240
220
200
180
160
ConsumptionPathfor u =0.01
numerical solution
boundedrationality

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40



360
340
320
300
280
260
240
220
200
180
160
Time t (periods)

ConsumptionPathfor u =0.05
numerical solution
boundedrationality



360
340
320
300
280
260
240
220
200
180
160
Time t (periods)

ConsumptionPathfor u =0.1
numerical solution
boundedrationality

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
Time t (periods) Time t (periods)


Figure 11. Consumption paths for o
y
= 1, r
0
= 0.06 and o
r
= 0.0025.

Looking at the asset paths for the same value of the standard deviation of the

income process, one can notice in Figure 12 that the level of saving in the certainty

equivalence case is mostly higher than the level of saving obtained in the bounded

rationality case as well as under the rational expectations assumption. While for lower

levels of the coefficient of risk aversion u
e
{0.005,
0.01
} , the asset path obtained


assuming certainty equivalence crosses under the other two paths in the later part of the

horizon, the same is not true for higher values of the coefficient of risk aversion. For

u
e
{0.05,
0.1
} there is only one period, the one next to last, when the level of savings


under certainty equivalence is lower than in the other two cases.





98
C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l




1600

1400

1200

1000

800

600
Level of Assets for u =0.005


1600

1400

1200

1000

800

600
Level of Assets for u =0.01

numerical solution numerical solution
400 boundedrationality 400 boundedrationality
certaintyequivalence certaintyequivalence

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40



1600

1400

1200

1000

800

600
Time t (periods)

Level of Assets for u =0.05



1600

1400

1200

1000

800

600
Time t (periods)

Level of Assets for u =0.1

numerical solution numerical solution
400 boundedrationality 400 boundedrationality
certaintyequivalence certaintyequivalence

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
Time t (periods) Time t (periods)


Figure 12. Asset paths for o
y
= 1, r
0
= 0.06 and o
r
= 0.0025.

As it was the case with rule 1, an increase in the coefficient of risk aversion

results in a decrease of the absolute size of the level of savings. Moreover, the shape of

the paths for both the rational expectation and bounded rationality cases changes from

concave to convex. As opposed to rule 1, for u
e
{0.05,
0.1
} the level of savings under


bounded rationality is higher than under rational expectations.

The next set of simulations has the standard deviation of income increased to

o
y
= 5 . The consumption paths for u
e
{0.005,
0.01
} in Figure 13 are not much different


from those presented in Figure 11 while for higher values of the risk aversion coefficient,

u
e
{0.05,
0.1
} , the consumption paths are steeper than in the previous case.





99
A
s
s
e
t

l
e
v
e
l


A
s
s
e
t

l
e
v
e
l


A
s
s
e
t

l
e
v
e
l


A
s
s
e
t

l
e
v
e
l




360
340
320
300
280
260
240
220
200
180
160
ConsumptionPathfor u =0.005
numerical solution
boundedrationality


360
340
320
300
280
260
240
220
200
180
160
ConsumptionPathfor u =0.01
numerical solutionbounded
rationality

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40




360
340
320
300
280
260
240
220
200
180
160
Time t (periods)

ConsumptionPathfor u =0.05
numerical solution
boundedrationality




360
340
320
300
280
260
240
220
200
180
160
Time t (periods)

ConsumptionPathfor u =0.1
numerical solutionbounded
rationality

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
Time t (periods) Time t (periods)


Figure 13. Consumption paths for o
y
= 5 , r
0
= 0.06 and o
r
= 0.0025.

For the level of savings, the change is similar to that observed in the case of

consumption. In Figure 14 one can see that, while not much has changed for the lower

values of the coefficient for risk aversion, the asset paths for higher values of the risk

aversion coefficient, u
e
{0.05,
0.1
} , have changed, effectively becoming concave, as


opposed to convex in the previous case. Besides the concavity change, one can observe

that for u = 0.1 the level of assets resulting from the numerical approximation of the

rational expectations model is higher than in the case of certainty equivalence for the

bigger part of the lifetime horizon.








100
C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l



1600

1400

1200

1000

800

600
Level of Assets for u =0.005

1600

1400

1200

1000

800

600
Level of Assets for u =0.01
numerical solution numerical solution
400 boundedrationality 400 boundedrationality
certaintyequivalence certaintyequivalence

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40



1600

1400

1200

1000

800

600
Time t (periods)

Level of Assets for u =0.05



1600

1400

1200

1000

800

600
Time t (periods)

Level of Assets for u =0.1

numerical solution numerical solution
400 boundedrationality 400 boundedrationality
certaintyequivalence certaintyequivalence

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
Time t (periods) Time t (periods)


Figure 14. Asset paths for o
y
= 5 , r
0
= 0.06 and o
r
= 0.0025.

For a yet higher variance of income, one can notice in Figure 15 that the path for

consumption becomes a lot steeper for u
e
{0.05,
0.1
} . On the other hand, there seems to

be little change in the consumption pattern for u = 0.005 . On the savings front, the level

of precautionary saving increases tremendously for the highest value of the coefficient of

risk aversion considered here, u = 0.1, and quite substantially for u = 0.05 . As it can be

easily seen in Figure 16, in these two cases, the level of savings for the rational

expectation model, as well as the bounded rationality version, becomes noticeably higher

than what certainty equivalence produces.










101
A
s
s
e
t

l
e
v
e
l


A
s
s
e
t

l
e
v
e
l


A
s
s
e
t

l
e
v
e
l


A
s
s
e
t

l
e
v
e
l



400

350

300

250

200

150
Consumption Path for u=0.005
numerical solution
bounded rationality

400

350

300

250

200

150
Consumption Path for u=0.01
numerical solution
bounded rationality

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40



400

350

300

250

200

150
Time t (periods)

Consumption Path for u=0.05
numerical solution
bounded rationality



400

350

300

250

200

150
Time t (periods)

Consumption Path for u=0.1
numerical solution
bounded rationality

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
Time t (periods) Time t (periods)


Figure 15. Consumption paths for o
y
= 10 , r
0
= 0.06 and o
r
= 0.0025.

Yet, the level of savings continues to be higher for the much lower coefficient of

risk aversion, u = 0.005 , when compared with the savings pattern for u = 0.01 and

u = 0.05 .

As in the case of rule 1, comparing the level of savings from the panel

corresponding to u = 0.05 and o
y
= 10 in Figure 16, to the level of savings from the

panel corresponding to u = 0.005 and o
y
= 1 in Figure 12, leads to the observation that

the two are almost the same. This is to say that for values of the coefficient of risk

aversion and of standard deviation for income ten times as high as the ones in Figure 12,

the level of precautionary saving is almost unchanged.









102
C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


Level of Assets for u =0.005 Level of Assets for u =0.01

1800 1800

1600 1600
1400 1400

1200 1200
1000 1000

800 800
600 600numerical
solution numerical
solution
400 boundedrationality 400 boundedrationality
certaintyequivalence certaintyequivalence

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
Time t (periods) Time t (periods)

Level of Assets for u =0.05 Level of Assets for u =0.1
1800 1800

1600 1600
1400 1400

1200 1200
1000 1000

800 800
600 600
numerical solution numerical solution
400 boundedrationality 400 boundedrationality
certaintyequivalence certaintyequivalence

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
Time t (periods) Time t (periods)


Figure 16. Asset paths for o
y
= 10 , r
0
= 0.06 and o
r
= 0.0025.

As in the case of rule 1 the level of savings under bounded rationality is fairly

close to the level of precautionary saving derived from the rational expectation model.

However, in contrast to rule 1, the relative size depends on the parameters of the model

and hence the level of precautionary saving derived from the rational expectation model

is no longer consistently higher when compared to the level of savings obtained in the

case of bounded rationality. Consequently, the rationally bounded consumer no longer

starts consistently with a higher consumption level.












103
A
s
s
e
t

l
e
v
e
l


A
s
s
e
t

l
e
v
e
l


A
s
s
e
t

l
e
v
e
l


A
s
s
e
t

l
e
v
e
l


III.3.3. Rule 3

In this section, I will consider a simpler rule than the previous two, meaning that

the level of wealth A
t
+T
h
is chosen such that, given the number of periods left until time

T , a constant growth rate would insure that the final level of wealth is A
T
.

Following are some graphical representations of the simulations for rule 3. All the

graphs contain a representation of the numerical solution and, for comparison purposes,

the graphs detailing the evolution for the level of assets also contain the certainty

equivalent solution.

The simulations for rule 3 use the same values of the parameters as in the

previous two sections. Consequently, the numerical solution for the rational expectations

model exhibits the same characteristics as discussed before. Therefore, when presenting

the results in this section I will concentrate on the solution derived from assuming

bounded rationality.

As one can see in Figure 17, the consumption paths have kept their upward slope

but for lower values of the coefficient of risk aversion, the difference between the rational

expectation and bounded rationality solutions is considerably higher than for the previous

two rules. The difference can be clearly seen in the picture, with the rationally bounded

consumer consuming more in the beginning while the unboundedly rational consumers

consumes more from the 12
th
period until the end of the horizon. On the other hand, for

higher values of the coefficient of risk aversion, consumption paths are almost

indistinguishable.








104

360
340
320
300
280
260
240
220
200
180
160
Consumption Path for u=0.005
numerical solution
boundedrationality

360
340
320
300
280
260
240
220
200
180
160
Consumption Path for u=0.01
numerical solution bounded
rationality
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40



360
340
320
300
280
260
240
220
200
180
160
Time t (periods)

Consumption Path for u=0.05
numerical solution
boundedrationality



360
340
320
300
280
260
240
220
200
180
160
Time t (periods)

Consumption Path for u=0.1
numerical solution bounded
rationality

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
Time t (periods) Time t (periods)



Figure 17. Consumption paths for o
y
= 1, r
0
= 0.06 and o
r
= 0.0025.

Looking at the asset paths in Figure 18 one will notice that, for low values of the

coefficient of risk aversion, the bounded rationality assumption leads to much lower

levels of precautionary saving than in the case of rational expectations or certainty

equivalence. However, the surprising result is that for higher values of the coefficient of

risk aversion, there is almost no difference between the level of savings under rational

expectations and bounded rationality.

By increasing the standard deviation of income to o
y
= 5 , one can see in Figure


19 a clear difference between the consumption paths for bounded rationality and rational

expectations for all levels of risk aversion. As before, the two consumption paths have an

upward slope with the rational expectation solution being the steeper one.





105
C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l



1600

1400

1200

1000

800

600
Level of Assets for u =0.005

1600

1400

1200

1000

800

600
Level of Assets for u =0.01

numerical solution numerical solution
400 boundedrationality 400 boundedrationality
certaintyequivalence certaintyequivalence

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40



1600

1400

1200

1000

800

600
Time t (periods)

Level of Assets for u =0.05



1600

1400

1200

1000

800

600
Time t (periods)

Level of Assets for u =0.1

400

numerical solution
boundedrationality
certaintyequivalence

400

numerical solution
boundedrationality
certaintyequivalence

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
Time t (periods) Time t (periods)


Figure 18. Asset paths for o
y
= 1, r
0
= 0.06 and o
r
= 0.0025.



360
340
320
300
280
260
240
220
200
180
160


Consumption Path for u=0.005
numerical solution
bounded rationality



360
340
320
300
280
260
240
220
200
180
160


Consumption Path for u=0.01
numerical solution
bounded rationality

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40



360
340
320
300
280
260
240
220
200
180
160
Time t (periods)

Consumption Path for u=0.05
numerical solution
bounded rationality



360
340
320
300
280
260
240
220
200
180
160
Time t (periods)

Consumption Path for u=0.1
numerical solution
bounded rationality

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
Time t (periods) Time t (periods)


Figure 19. Consumption paths for o
y
= 5 , r
0
= 0.06 and o
r
= 0.0025.


106
A
s
s
e
t

l
e
v
e
l


A
s
s
e
t

l
e
v
e
l


A
s
s
e
t

l
e
v
e
l


A
s
s
e
t

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


The asset paths represented in Figure 20 show clearly a higher level of

precautionary saving in the case of rational expectations. The path corresponding to

certainty equivalence produces higher levels of saving than the bounded rationality path.





1600

1400

1200

1000

800

600





Level of Assets for u =0.005





1600

1400

1200

1000

800

600





Level of Assets for u =0.01

numerical solution numerical solution
400 bounded rationality 400 bounded rationality
certainty equivalence certainty equivalence

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40



1600

1400

1200

1000

800

600
Time t (periods)

Level of Assets for u =0.05



1600

1400

1200

1000

800

600
Time t (periods)

Level of Assets for u =0.1

numerical solution numerical solution
400 bounded rationality 400 bounded rationality
certainty equivalence certainty equivalence

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
Time t (periods) Time t (periods)


Figure 20. Asset paths for o
y
= 5 , r
0
= 0.06 and o
r
= 0.0025.

Increasing again the standard deviation for income to o
y
= 10 , one will notice in


Figure 21 that there is not much change in the paths for consumption at low levels of risk

aversion. However, the slope of consumption for u = 0.1 increases quite a lot.











107
A
s
s
e
t

l
e
v
e
l


A
s
s
e
t

l
e
v
e
l


A
s
s
e
t

l
e
v
e
l


A
s
s
e
t

l
e
v
e
l



400

350

300

250

200

150
Consumption Path for u=0.005
numerical solution
bounded rationality

400

350

300

250

200

150
Consumption Path for u=0.01
numerical solution bounded
rationality

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40



400

350

300

250

200

150
Time t (periods)

Consumption Path for u=0.05
numerical solution
bounded rationality



400

350

300

250

200

150
Time t (periods)

Consumption Path for u=0.1
numerical solution bounded
rationality

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
Time t (periods) Time t (periods)


Figure 21. Consumption paths for o
y
= 10 , r
0
= 0.06 and o
r
= 0.0025.

On the saving side, one can see in Figure 22 that for the highest coefficient of risk

aversion, the rational expectations solution provides a much higher level of savings,

while the rationally bounded consumer still saves less than in the case of certainty

equivalence for u = 0.01.

While the level of precautionary saving depends heavily on the parameter values

of the model for the unboundedly rational consumer, the same can not be said for the

rationally bounded consumer in the case of rule 3. The asset path for the rationally

bounded consumer is barely concave and increasing the variance of income does not

seem to create the same type of changes as the ones observed for the fully rational

consumer. This behavior is the result of optimizing for only short periods of time coupled

with the fact that the intermediary asset level targets are chosen assuming a constant

growth rate.


108
C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


C
o
n
s
u
m
p
t
i
o
n

l
e
v
e
l


Level of Assets for u=0.005 Level of Assets for u=0.01

1800 1800
1600 1600
1400 1400
1200 1200
1000 1000
800 800
600 600
numerical solution numerical solution
400 boundedrationality 400 boundedrationality
certaintyequivalence certaintyequivalence

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40



1800
1600
1400

1200
1000
800
600
Timet (periods)

Level of Assets for u=0.05



1800
1600
1400

1200
1000
800
600
Timet (periods)

Level of Assets for u=0.1
numerical solution numerical solution
400 boundedrationality 400 boundedrationality
certaintyequivalence certaintyequivalence

0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
Timet (periods) Timet (periods)


Figure 22. Asset paths for o
y
= 10 , r
0
= 0.06 and o
r
= 0.0025.

In conclusion, in the case of rule 3, the rule employed by the rationally bounded

consumer for the accumulation of assets is overshadowing the precautionary motives

embedded in the functional specification of the model.






















109
A
s
s
e
t

l
e
v
e
l


A
s
s
e
t

l
e
v
e
l


A
s
s
e
t

l
e
v
e
l


A
s
s
e
t

l
e
v
e
l


III.4. Final Remarks


The level of precautionary saving under bounded rationality depends quite heavily

on the behavior assumptions. While in many of the simulations presented in this chapter

the level of precautionary saving chosen on average by the rationally bounded consumer

is below that resulting from a rational expectations model, there a few parameterizations

of the model, under rule 2, for which the rationally bounded consumer saves more.

The simulations also show that for low coefficients of risk aversion, variation in

income uncertainty does not affect much the level of saving. If one adds to this

observation the possibility that self selection exists (individuals with high risk aversion

choose occupations with low income uncertainty), it is easy to see why some empirical

studies would find relatively low levels of precautionary saving.

Another interesting result is that under rule 3, where the rationally bounded

consumer follows some form of financial planning, there is not much difference for asset

paths across various levels of risk aversion and income uncertainty. This result is

consistent with the observation made by Lusardi (1997) that the saving rates do not

change much across occupations.

Most of the studies looking to asses the importance of precautionary saving, or the

impact of income uncertainty on precautionary saving, have assumed that interest rate

uncertainty does not play an important role in the decision making process. For the model

discussed in this chapter, the assumption of a constant interest rate would result in an

asset path that is constant regardless of the realizations for the income process. By

introducing uncertainty in the interest rate process, that is no longer the case. The

dynamic of the asset path is especially influenced by the realization of the interest rate




110
process for lower levels of risk aversion. Therefore, the empirical literature should also

consider the impact of interest rate uncertainty when studying the importance of

precautionary motives on the level of saving.

While the results presented in this chapter point to an important role for the

bounded rationality in the decision making process, it would be difficult to test the

model's validity in a standard empirical setting. The problem is that the results depend

heavily on the rules adopted as well as on the parameterization of the model and it would

be difficult to distinguish between the effects of the general assumptions corresponding to

bounded rationality and those specific to a particular rule. Therefore, a more appropriate

framework for testing the validity of the model would be an experimental setting. In such

a framework, one can potentially "calibrate" the model by identifying the level of risk

aversion and the level of patience for each subject. Once these parameters are determined

it becomes easier to test hypotheses regarding the decision making process. There have

been several studies in the field of experimental economics investigating consumption

behavior under uncertainty (Hey and Dardanoni (1988), Ballinger et al. (2003) and

Carbone and Hey (2004)) that concluded that actual behavior differs significantly from

what is considered optimal. While these studies provide some insights in the decision

making process, they do not test for any particular alternative to the optimal behavior

corresponding to an unboundedly rational individual. Therefore a future area of research

is the design of an experimental framework that could test the hypotheses regarding the

decision making process advanced in this chapter.









111
Appendices




Appendix A. Technical notes to chapter 2





Appendix A1. Definitions for Scenarios, Equivalence Classes and Associated Probabilities

Suppose the world that can be described at each point in time by the vector of

state variables x
t
, and let u
t
denote the control variable while ç
t
is the forcing variable.

Suppose ç
t
is a random variable, with the underlying probability space
71
(O, E,
P
) . ç
t
is

defined as ç
t
: O C R where O is countable and finite. If the horizon has T +1 time

periods and ç
t
(
e
) is a realization of ç
t
for the event e ŒO in time period t , then the

sequence

ç
s
(
e
)
=

0
s
(
e
),ç
1
s
(
e
),K,ç
T
s
(
e
))

is called a scenario
72
. From now on, for notation simplification, I will refer to a scenario

s simply by ç
s
or by the index s and, in vector form, by ç
s
= ç
0
s

1
s
,K,ç
T
s
.
( )

Let S (e ) denote the set of all scenarios. Given that O is finite, the set S (e ) is

also finite. Therefore, one can define an event tree {N,
A
} characterized by the set of

nodes N and the set of arcs A . In this representation, the nodes of the tree are decision

points and the arcs are realizations of the forcing variables. The arcs join nodes from


71

72


O is the sample space, E is the sigma field and P is the probability measure.

Other definitions of scenarios can be found in Helgason and Wallace (1991a, 1991b)
and Rosa and Ruszczynski (1994).



112
consecutive levels such that a node n
it
at level t is linked to N
t
+
1
nodes n
k
+
1
, k = 1,..., N
t
+1 t

at level t +1.

The set of nodes N can be divided into subsets corresponding to each level

(period). Suppose that at time t there are N
t
nodes. The arcs reaching the nodes

n
it
, i = 1,K, N
t
belong each to several scenarios ç
q
(e ), q = 1,..., L
t
where L
t
represents

the number of leaves stemming from a node at level t . The bundle of scenarios that go

through one node plays a very important role in the decomposition as well as in the

aggregation process. The term equivalence class has been used in the literature to

describe the set of scenarios going through a particular node.

By definition, the equivalence class {s
t
}
i
, i = 1,K, N
t
is the set of all scenarios

having the first t +1 coordinates, ç
0
,K,ç
t
common. This means that for two scenarios

ç
j
=

0
j

1
j
,K,ç
t

1

t
j
,...,ç
T
j
) and ç
k
=

0
k

1
k
,K,ç
t

1

t
k
,...ç
T
k
) that belong to the


equivalence class

{
s
} , i =1,K, N t
i


t


the first t +1 elements are common, that is,

ç
lj
= ç
l
k for l = 0,...,t . Formally,

{
s
}
=

t
i


k



lk
= ç
l
i


for l = 0,...,t }

As mentioned in the above description of the event tree, at time t there are N
t


nodes. Then, the number of distinct equivalence classes

{
s
}
t



i


is also N
t
, that is,

i = 1,K, N
t
. Every node n
it
, i = 1,K, N
t
is associated with an equivalence
class
{s
t
}
i
.

The number of elements of the
set
{s
t
}
i
is given by the number of leaves stemming from

node i , level (stage) t .



113
Since scenarios are viewed in terms of a stochastic vector ç with stochastic

components ç
0
s

1
s
,K,ç
T
s
, it is natural to attach probabilities to each scenario. I denote

the probability of a particular realization of a scenario, s , with

p(s) = prob(ç
s
) .

These probabilities are non-negative numbers and sum to one. Formally,

p(s) > 0 and

9 p(s) =1. I assume that for each scenario ç
sŒS

s


the stochastic components

ç
0
s

1
s
,K,ç
T
s
are independent. Then




(




)




T
p(s) = prob ç
s
(
e
)
=
÷ prob ç
t
s
(
e
) ( ) (A.1.1)
t =0

Further on, I define the probabilities associated with a scenario conditional upon

belonging to a certain equivalence
class
{s
t
}
i
at time t :


p s s Œ s
t
( {
}
) =
prob
(ç i
s
ç
s
Œ
{s
t
}
i
= )

p
(p{(ss)
}
), t
i
where p s
t
({
}
) i


is the probability mass of all scenarios belonging to the class


t

{
s
} .
t
i
Under the assumptions outlined above, p s
t s
({
}
)
=
÷
prob
(ç t
(
e
)) .
Therefore, the


conditional probability is easily computed as
i
t=0


p
s
{s
t
} = prob ç
s
ç
s
e
{s
t
}
i
=
(
i
)
( )
T
[
prob

t
(
e
)) s
t =t +1

The transition from the state at time t to that at time t + 1 is governed by the control

variable u
t
but is also dependent on the realization of the forcing variable, that is, on a

particular scenario s .




114
Appendix A2. Description of the Scenario Aggregation Theory

The idea is to show how a solution can be obtained by using special

decomposition methods that exploit the structure of the problem by splitting it into

manageable pieces and coordinate their solution.

Let us assume for a moment that the original problem can be decomposed into

subproblems, each corresponding to a scenario. Then the subproblems can be described

as:

T
min 9 F
t
x
s
t ,u
s
t ,
u ŒU [R
mu
( )
s ŒS
(A.2.1)
t t
t=1

where u
st
and x
st
are the control and the state variable respectively, conditional on the

realization of scenario s while S is a finite, relatively small set of scenarios.

Formally, by definition, a policy is a function or a mapping U : S ÷ R
m
assigning

to each scenario s e S a sequence of controls U (s)
=
(u
s
0 ,u
1
s
,K,u
s
t ,K,u
T
s
) , where u
st



denotes the decision to be made at time t if the scenario happens to be s . Similarly, the

state variable at each stage is associated with a particular scenario s . I use the notation

x
st
to show the link between the state variable and scenario s at time t . One can think of


T
the mappings U : S ÷ R
m
as a set of time linked mappings U
t
: S ÷ R
m
t with m
=
¿ m
t
.
t =1

The policy function has to satisfy certain constraints if two different scenarios s

and s ' are indistinguishable at time t on information available about them at time t .

Then u
s
t = u
s
t
'
, that is, a policy can not require different actions at time t relative to

scenarios s and s ' if there is no way to tell at time t which of the two scenarios will be


115
followed. This constraint is referred to as the non-anticipativity constraint. One way to

model this constraint is to introduce an information structure by bundling scenarios into

equivalence classes
73
as defined above. In this way, the scenario set S is partitioned at

each time t into a finite number of disjoint
sets,
{s
t
}
i
. Let the collection of all scenario

equivalence classes at time t be denoted by B
t
, where B
t
=
U{s
t
}
i
. In most cases i

partition B
t
+
1
is a refinement of partition B
t
, that is, every equivalence
class
{s
t
}
i
e B
t
is


a union of some equivalence classes

{
s
} t
+1



j


e B
t
+
1
. Formally,

{
s
}
t
i


=

U
{
s
} t
+1

j =1...m
i



j


.


Looking back to the event tree representation discussed in the previous section, m
i


represents the number of nodes n
t
j+
1
at level t +1 that are linked to the same node n
it
.

A policy is defined as implementable if it satisfies the non-anticipativity

constraint, that is, u
t
(e ) must be the same for all scenarios that have common past and

present
74
. In other words, a policy is implementable if for all t = 0,K,T the t
th
element

is common to all scenarios in the same class

{
s
} , i.e. if t
i

u
t

i
) = u
t

k
) whenever

{
s
}
=
{
s
}
t i t k
.

Let E be the space of all mappings U : S ÷ R
n
with components U
t
: S ÷ R
n
t .

Then the subspace




73

74




Some authors, such as Rockaffeler and Wets (1991), use the term scenario bundle.

For certain problems the non-anticipativity constraint can also be defined in terms of
the state variable, that is, x
t
(e ) must be the same for all scenarios that have common past
and present.



116

H = U e E |U
t
is constant on each class {s
t
}
i
e B
t
, for t = 1,...,T {


identifies the policies that meet the non-anticipativity constraint.
}

A policy is admissible if it always satisfies the constraints imposed by the

definition of the problem. It is clear that not all admissible policies are also

implementable. By definition, a contingent policy is the solution, u
s
, to a scenario

subproblem. It is obvious that a contingent policy is always admissible but not

necessarily implementable. Therefore, the goal is to find a policy that is both admissible

and implementable. Such a policy is referred to as a feasible policy.

One way to create a feasible policy from a set on contingent policies is to assign

weights (or probabilities) to each scenario and then blend the contingent policies

according to these weights. Specifically, if the probabilities associated with each scenario

are defined as in (A.2.1), one calculates for every period t and for every equivalence

class
{s
t
}
i
e B
t
the new policy u
t
by computing the expected value:


u
t

({
s
}
)
=
¿
}
p
(s
?
{
s
}
)u (s?)
t i
s?e
{
s
t
i
t i t
(A.2.2)

Then one defines the new policy for all scenarios s that belong to the equivalence class

{
s
} e B
t
i


t


as:


u
s
t = u
t
ˆ

({
s
}
) for all s
e
{
s
}
t i t i
(A.2.3)


Based on its definition, u
st
is implementable. The operator J :U ÷ U defined by (A.2.2)
ˆ


and (A.2.3) is called the aggregation operator.

Let us
rewrite equation
(2.4.1) as:
ˆ





117

min F
s
( x
s
,u
s
), seS (A.2.4)
u
t
eU
t
_R
mu


T
by defining the functional F
s
( x
s
,u
s
)
=
¿ F
t
( x
t
(s),u
t
(s
)
) .
t =1

Then the overall problem can be reformulated as:

min ¿ p
s
F
s
( x
s
,u
s
) over all U e E I H
seS






(A.2.5)


Let us assume for a moment that u
s
is an implementable policy obtained as in (A.2.3) ˆ


from contingent policies u
s
and u
s
is the optimal policy for the particular scenario s of

the problem described by (A.2.5). Let U and U be the collections of policies u
s
and
ˆ ˆ


u
s
respectively. One can easily see that U represents the optimal policy for the problem

described by (A.2.5). The question that the scenario aggregation methodology answers is

how to obtain the optimal solution U from a collection of implementable policies U . ˆ







Appendix A3. Solution to a Scenario Subproblem

In order to take advantage of the fact that scenario aggregation does not require

the computation of an exact solution for each scenario, I transform the Lagrangian (2.6.8)

by replacing the utility function with a first order Taylor series expansion around the

solution obtained in the previous iteration. Hence:

e
÷
uc
t
= e
÷
uc
t
s
s(k÷1)
?1÷u c
st
÷ c
s
t (k÷1
)
?
?
( )
?

From the transition equation, consumption can be expressed as:

c
s
t
=
(1+
r
) A
t

1
+ y
s
t ÷ A
t
s



118

Then

e
÷
uc
t
= e
÷
uc
t
s
s(k÷1)
{1÷u ??(1+
r
)
(A
s
and
t ÷1
÷ A
t
s÷(k÷1
)
÷ A
ts
÷ A
t
s(k÷1
)
? .For iteration

scenario s the Lagrangian becomes:
1
)(
)
}
?
(
k
)


min ¿ |
?
T
t
?e
÷
uC
s
t(k÷1) ?
?
u ?1÷
u
(1+
r
) A
t

1
÷ A
t
s÷(k÷1
)
+u A
ts
÷ A
t
s(k÷1
)
? +W
ts ?
(1+
r
) A
t

1
+ y
s
t ÷ A
ts
? +
t =0
?
?
(
1
)( )
? ? ?

+
1
µ ?(1+
r
) A
t

1
÷ A
(
t÷k1÷1
)
÷ A
ts
÷ A
(
tk÷1
)
?
2?
( )( )
?
2
}
Then, the first order condition with respect to A
t
s
is given by:


|
t
e
÷
uc
(
{
s k÷1)
t
÷W
t
s(k
)
÷ µ ?(1+
r
) A
t

1
÷ A
(
t÷k1÷1
)
÷ A
ts
÷ A
(
tk÷1
)
? +
?
( )(
)
} ?
|t+
1 ÷
(1+
r
) e
÷
uc{

s(k÷1)
t+1
+
(1+
r
)W
t
s+(1k
)
+
µ
(1+
r
)
?
(1+
r
) A
ts
÷ A
(
tk÷1
)
÷ A
t
s+
1
÷ A
(
t+k1
)
? = 0


Rearranging the terms leads to:
?
( )(
1
?
)
}


1 ?e
÷
uc
s
t(k÷1)
÷
(1+
r
) | e
÷
uc
s
t(+k1÷1) ÷W
s
(k
)
+ (1+
r
) |W
s
(k
)
? +
µ
?
?
t
t +
1
? ?
+
(1+
r
) A
(
t÷k1÷1
)
÷ A
(
tk÷1
)
÷ (1+
r
)2 | A
(
tk÷1
)
+ (1+
r
) | A
(
t+k1÷1
)
= (A.3.1)
=
(1+
r
) A
t

1
÷ A
ts ÷
(1+
r
)2 | A
ts +
(1+
r
) | A
t
s+1

Let

I
s
t (k
)
= 1 ?e
÷
uc
t
÷
(1+
r
) | e
÷
uc
t
+
1
÷W
t
s(k
)
+ (1+
r
) |W
t
s+(1k
)
?
s(k÷1) s(k÷1)
µ
?
?


Then the first order condition with respect to A
t
s
can be written as:
?
?
(A.3.2)

I
s
t (k
)
+ (1+
r
) A
(
t÷k1÷1
)
÷ ?1
+
(1+
r
)2 | ? A
(
tk÷1
)
+ (1+
r
) | A
(
t+k1÷1
)
=
? ?
=
(1+
r
) A
t

1
÷ ?1
+
(1+
r
)2 | ? A
ts +
(1+
r
) | A
t
s+1
? ?

For t = T ÷1 the first order condition becomes:

I
s
T(÷k1
)
+ (1+
r
) A
T
(k÷÷21
)
÷ ?1
+
(1+
r
)2 | ? A
T
(k÷÷11
)
+ (1+
r
) | A
T
(k÷1
)
=
? ?
(A.3.3)
=
(1+
r
) A
T

2
÷ ?1
+
(1+
r
)2 | ? A
T

1
+
(1+
r
) | A
T
? ?




119
Noting that A
T
(k÷1
)
= A
T
= A
T
equation (A.3.3) can be written as:


I
s
T(÷k1
)
+ (1+
r
) A
T
(k÷÷21
)
÷ ?1
+
(1+
r
)2 | ? A
T
(k÷÷11
)
= (1+
r
) A
T

2
÷ ?1
+
(1+
r
)2 | ? A
T
s÷1
?

Similarly, for t = 0 one obtains:
? ? ?

I
s
0(k
)
+ (1+
r
) A
(
÷k1÷1
)
÷ ?1
+
(1+
r
)2 | ? A
(
0k÷1
)
+ (1+
r
) | A
1
(k÷1
)
=
? ?
(A.3.4)
(1+
r
) A
÷
s
1
÷ ??1
+
(1+
r
)2 | ?? A
0s +
(1+
r
) | A
1
s

Again, noting that A
÷
1
is given, A
(
÷k1÷1
)
= A
÷
s
1
so equation (A.3.4) becomes:


I
s
0(k
)
÷ ?1
+
(1+
r
)2 | ? A
(
0k÷1
)
+ (1+
r
) | A
1
(k÷1
)
= ÷ ?1
+
(1+
r
)2 | ? A
0s +
(1+
r
) | A
1
s
? ? ? ?

Rewriting the system of equations in matrix form, leads to:

?÷?1
+
(1+
r
)2 |?
??
?
?
(1+
r
)
?
(1+
r
)|

÷?1
+
(1+
r
)2 |?

0 0

0
K
0

0
0

0
?
?
s

?? A
0
?
?
?
0
?
(1+
r
)
?
(1+
r
)|

÷?1
+
(1+
r
)2 |? (1+
r
) |
K
0
?? A
s
1 ?
?? s ?
? ? ? K
0
?? A
2
? =
?
M M M M M M M
?? M ?
?? s ?
?
?
??A
T
÷
1
? ?
?
?
?
0 0 0 0
K
(1+
r
)

÷?1
+
(1+
r
) |
?
?
2

??
?

I
s
0(k
)
÷?1
+
(1+
r
)2 |?
A
(
0k÷1
)
+(1+
r
) |A
(
1k÷1)
?
? ?
? ? ? ?
?I
s
(k
)
+(1+
r
) A
(
k÷1
)
÷?1
+
(1+
r
)2 |? A
(
k÷1
)
+(1+
r
) |A
(
k÷1
)
? ?1
=?
?
?
0
?

M
?1
2
?
? ? ?
?
I
T
÷
1
+
(1+
r
) A
T
÷
2
?
( )
?
T÷1
?
s(k)

(k÷1
)
÷?1+ 1+r 2 |? A
(
k÷1)
?
?











120
Appendix B. Technical notes to chapter 3





Appendix B1. Analytical Solution for a Scenario with Deterministic Interest Rate

Consider the problem described by (3.3.1) - (3.3.4). Solving the period-by-period

budget constraint (3.3.2) for c
t
, t = T ÷ 1 and t = T , and substituting back into the utility

function, the period T ÷ 1 optimization problem is given by:

e exp ÷u

(1+ r
T
÷
1
) A
T
÷
2
+ y
T
÷
1
÷ A
T
÷
1
?
max ¹÷
c
{
e
}

A
T
÷1
©
¹
u

_ exp ÷u

(1+ r
T
) A
T
÷
1
+ y
T
÷
A
T
?
(B.1.1)
{
e
}
¹
E·| u
?

I
ˆ÷ T÷
1
˜ ?

subject to
·
.



A
T
÷
1
> ÷b

+?
˜¹




(B.1.2)

Taking derivatives with respect to A
T
÷
1
, the Euler equation for (B.1.1) is given by:

exp ÷
u
(1+ r
T
÷
1
) A
T
÷
2
÷u y
T
÷
1
+u A
T
÷
1
?
e
e
¹
?
exp ÷
u
(1+ r
T
÷
1
) A
T
÷
2
÷u y
T
÷
1
÷u b? ,
= max c
e ? ÷
¹
?
(B.1.3)
¹
|
(1+ r
T
) E exp ÷
u
(1+ r
T
) A
T
÷
1
÷u y
T
+u A
T
? I
T
÷1 ©
{
e ?
}
? u
2
o
2
y
?
¹
?
Note that y
T
= y
T
÷
1
+ ç
T
while E ?
exp
(÷uç
T
) | I
T
÷
1
? = exp ?
? ?
? 2 ? and hence solving ?
? ?

(B.1.3) for the optimal wealth level at the beginning of period T ÷ 1 yields:

e*
A
= max c
¹÷b
,

(1+ r
T
÷
1
) A
T
÷
2
+ I*
T
+ A
T
?
÷
.
T ÷1
e

?
(B.1.4)
©
¹
(2 + r
T
)
¹?

where I*
T
= I + log
|
(1+ r
T
)? /u , and I = uo 2
y
/ 2 .
{
e
?}

Going now to period T ÷ 2 , the optimization problem is given by



121

? exp ÷u
?
(1+ r
T
÷
2
) A
T
÷
3
+ y
T
÷
2
÷ A
T
÷
2
?
{
?
}
{
*


? ÷ | E ? exp ÷u
?
(1+ r
T
÷
1
) A
T
÷
2
+ y
T
÷
1
÷ A
T
÷
1
? +
max ?

? ? ?
}
A
T
÷2
?
?
u

exp ?÷
u
((1+ r
T
) A
*
÷
1
+ y
T
÷ A
T
)? T
?
?
u
+| ?
?
I ?? ???



subject to
u
T ÷2
?
?
?
?


(B.1.5)

A
T
÷
2
> ÷b (B.1.6)

Taking derivatives with respect to A
T
÷
2
, and noting that

E exp ÷u A
*
÷
1
I
T
÷
2
? = exp ÷u A
*
÷
1
,
e
(
T
)
?
(
T
)

the Euler equation for (B.1.5) is given by:

exp ?÷
u
(1+ r
T
÷
2
) A
T
÷
3
÷u y
T
÷
2
+u A
T
÷
2
?
?
?
?
exp ?÷
u
(1+ r
T
÷
2
) A
T
÷
3
÷u y
T
÷
2
÷ub? ,
=
max ? ?
?

*
?
?
?
?
(
B.
1.
7)
?
?|
(1+ r
T
÷
1
) exp ?÷
u
(1+ r
T
÷
1
) A
T
÷
2
+u A
T
÷
1
? E ?exp (÷u y
T
÷
1
) I
T
÷
2
?? ?
? ? ??

Since y
T
÷
1
= y
T
÷
2
+ ç
T
÷
1
, (B.1.7) can be rewritten as:

exp ?÷
u
(1+ r
T
÷
2
) A
T
÷
3
÷u y
T
÷
2
+u A
T
÷
2
?
?
?
?
exp ?÷
u
(1+ r
T
÷
2
) A
T
÷
3
÷u y
T
÷
2
÷ub? ,
= max ?
? ?
?
?
?
*
? ?
?|
(1+ r
T
÷
1
) exp ?÷
u
(1+ r
T
÷
1
) A
T
÷
2
÷u y
T
÷
2
+u A
T
÷
1
? E ?exp (÷uç
T
÷
1
) I
T
÷
2
??
? ? ? ??

Assuming that liquidity constraint is not binding, solving (B.1.7) for A
T
÷
2
yields:

ln
|
(1+ r
T
÷
1
)? uo 2y
(1+ r
T
÷
2
) A
T
÷
3
÷ A
T
÷
2
= ÷
e u ?
+
(1+ r
T
÷
1
) A
T
÷
2
÷
A
*
÷
1
÷ T
2
(B.1.8)
Using the notation from above, equation (B.1.8) can be written as:

I*T ÷
1
= ÷ A
*
÷
1
+
(2 + r
T
÷
1
) A
T
÷
2
÷
(1+ r
T
÷
2
) A
T
÷3 T


Similarly, for period t , the equivalent of equation (B.1.9) is given by:



122


(B.1.9)

I*t+
1
= ÷ A
*
+
1
+
(2 + r
t
+
1
) A
t
÷
(1+ r
t
) A
t
÷1 t
(B.1.10)
It is clear that the optimal wealth level at the beginning of period t does not depend on

labor income received at the beginning of the period. This result is not general, but is

rather specific to the life-cycle model with a negative exponential utility function and

labor income following an arithmetic random walk process.

Solving for the beginning-of-period wealth levels from t = 0 to t = T ÷1 means

solving the system of linear equations:

_ A
0
ˆ
_
(1+ r
0
) A
÷
1
+ I*
1
ˆ
· A
1
˜ ·
I*
2
˜
· ˜ · ˜
· A
2
˜ · I*3 ˜


·
M
˜ =
·
˜·
· A
T
÷
3
˜ ·
M
I*T ÷2
˜
˜
˜
(B.1.11)
·A ˜ ·
˜
· T ÷
2
˜ · I
T
÷1
*
˜
· A
T
÷1 ˜ ·
. +. A
T
+ I*
T


where D is a tridiagonal coefficient matrix,

_ 2 + r
1
÷1 0 L 0 0 0ˆ
·
÷
(1+ r
1
) 2 + r
2
÷1 L 0 0 0
˜

· ˜
D= · M M M M M M˜ (B.1.12)
·
0 0 0
· L
÷
(1+ r
T
÷
2
) 2 + r
T
÷1 ÷1
˜
˜
·0 .
0 0
L
0
÷
(1+ r
T
÷
1
)
2 + r
T
˜ +

Once the values for wealth levels are computed, the consumption levels follow.

The solution presented in this section is in fact the solution for a scenario obtained by

discretizing the distribution of the forcing variable for the interest rate. Since an

analytical solution can be obtained when income follows an arithmetic random walk and

interest rate is deterministic, it is no longer necessary to discretize both forcing variables,

but only the interest rate. This approach reduces considerably the computational burden.




123
For different labor income processes, a dual discretization is necessary, that is, for both

forcing variables.





Appendix B2. Details on the Assumptions in Rule 1

In period t the consumer wants to solve the optimization problem given by:


ma
T
x
E
_9
T ÷| t ÷
t
_ 1 ˆ

exp
÷u c |
I
?

subject to
{c
t
}
t
=t
et=t
·
u
˜
.+
(
t
) t ??
(B.2.1)

A
t =
(1+ r
t
) A
t
÷
1
+ y
t
÷ c
t
, t = t,t +1,...,T ,
(B.2.2)
with A
t
÷
1
, A
T
given t = 0,1,...,T ÷1,

y
t
= y
t
÷
1
+ ç
t
, t = t +1,...,T ,
(B.2.3)
with y
t
given t = 0,1,...,T ÷1,

r
t
= r
t
÷
1
+u
t
,
with r
t
given
t = t +1,...,T ,
t = 0,1,...,T ÷1,


(B.2.4)

The assumption is that the forcing variable u
t
has three possible realizations,

{u
a
,u
b
,u
c
} . The set of its realizations determines the event tree and consequently the set


of scenarios. For T
h
periods the number of all scenarios is 3
T
h . The consumer considers

all the possible scenarios from period t to period t + T
h
. From there on it assumes that

for every leaf the scenario will be determined by u
t
taking its unconditional mean, that

is, zero. For example, if the short optimizing horizon is given by T
h
= 4 and the sequence

of realizations for u
t
up to period t + 4 , for a particular scenario,
is
{u
a
,u
c
,u
b
,u
c
} , the

assumption made by consumer is that for this particular scenario the realizations of u
t
for

the rest of the periods will be 0 , that is, the whole scenario
is
{u
a
,u
c
,u
b
,u
c
, 0, 0,...,
0
} .



124
This process is repeated as the consumer advances to period t +1 and goes again

through the optimization procedure. The number of scenarios considered remains the

same unless T ÷ t < T
h
, which is to say that there are fewer than T
h
periods left until the

terminal period.





Appendix B3. Details on the Assumptions in Rule 2

In period t the consumer wants to solve the optimization problem given by:

ma
T
x
E

9 ÷|t÷
t
_ 1 ˆ
exp
(÷u c
t
) | I
t
? T
?

subject to
{c
t
}
t
=t
e_
t=t
·
u
˜
.+
?
(B.3.1)

A
t =
(1+ r
t
) A
t
÷
1
+ y
t
÷ c
t
, t = t,t +1,...,T ,
(B.3.2)
with A
t
÷
1
, A
T
given t = 0,1,...,T ÷1,

y
t
= y
t
÷
1
+ ç
t
, t = t +1,...,T ,
(B.3.3)
with y
t
given t = 0,1,...,T ÷1,

r
t
= r
t
÷
1
+u
t
,
with r
t
given
t = t +1,...,T ,
t = 0,1,...,T ÷1,


(B.3.4)

The assumption is that the forcing variable u
t
has three possible realizations,

{u
a
,u
b
,u
c
} . The set of its realizations determines the event tree and consequently the set


of scenarios. For T
h
periods the number of all scenarios is 3
T
h . The consumer considers

all the possible scenarios from period t to period t + T
h
. From there on it assumes that

for every leaf only three more scenarios emerge, with u
t
taking only one of the three

values
{u
a
,u
b
,u
c
} every period until the end of the horizon. For example, if the short

optimizing horizon is given by T
h
= 4 and the sequence of realizations for u
t
up to



125
period t + 4 , for a particular scenario, is {u
a
,u
c
,u
b
,u
c
} , the assumption made by


consumer is that only three more scenarios will stem from the leaf corresponding to

scenario
{u
a
,u
c
,u
b
,u
c
} . These three scenarios are given
by
{u
a
,u
c
,u
b
,u
c
,u
a
,u
a
,...,u
a
} ,

{u
a
,u
c
,u
b
,u
c
,u
b
,u
b
,...,u
b
}
and
{u
a
,u
c
,u
b
,u
c
,u
c
,u
c
,...,u
c
} . Effectively, the total number


of scenarios considered is 3
T
h +
1
as opposed to 3
T
÷
t
which would represent the total

number of scenarios for the horizon from period t to period T .

This whole process is repeated as the consumer advances to period t +1 and goes

again through the optimization procedure. The number of scenarios considered remains

the same unless T ÷ t < T
h
, which is to say that there are fewer than T
h
periods left until

the terminal period.
































126
Bibliography




Allen, T. W., Carroll, C. D. (2001). Individual Learning About Consumption.
Macroeconomic dynamics, 5, 255-271.

Anderson, G., and Moore, G. (1986). An Efficient Procedure for Solving Nonlinear
Perfect Foresight Models. Working Paper, January 1986.

Atkinson, K.E. (1976). A Survey of Numerical Methods for the Solution of Fredholm
Integral Equations of the Second Kind. Society for Industrial and Applied Mathematics,
Philadelphia.

Baker, C. T. H. (1977). The Numerical Treatment of Integral Equations. Clarendon Press,
Oxford.

Ballinger, T. P., Palumbo, M. G., and Wilcox, N. T. (2003). Precautionary Saving and
Social Learning across Generations: An Experiment. The Economic Journal, 113, 920-
947.

Banks, J., Blundell, R., and Brugiavini, A. (2001). Risk Pooling, Precautionary Saving
and Consumption Growth. Review of Economic Studies, 68, 757-779.

Baxter M., Crucini, M., and Rouwenhorst, K. G. (1990). Solving the Stochastic Growth
Model by a Discrete-State-Space, Euler Equation Approach. Journal of Business and
Economic Statistics, 8, 19-21.

Binder, M., Pesaran, M. H., Samiei, S. H. (2000). Solution of Nonlinear Rational
Expectations Models with Applications to Finite-Horizon Life-Cycle Models of
Consumption. Computational Economics, 15, 25-57.

Birge, J. R. (1985). Decomposition and Partitioning Methods for Multistage Stochastic
Linear Programs. Operations Research, 33, 989-1007.

Bitros, G. C., and Kelejian, H. H. (1976). A Stochastic Control Approach to Factor
Demand. International Economic Review, 17, 701-717.

Burnside, C. (1993). Consistency of a Method of Moments Estimator Based on
Numerical Solutions to Asset Pricing Models. Econometric Theory, 9, 602-632.

Burnside, C. (1999). Discrete State-Space Methods for the Study of Dynamics
Economies. Computational methods for the study of dynamic economies, ed. by R.
Marimon and A. Scott. Oxford University Press, Oxford and New York, 95-113.

Carbone, E., and Hey, J. D. (2004). The Effect of Unemployment on Consumption: An
Experimental Analysis. The Economic Journal, 114, 660-683.



127
Carroll, C. D. (1994). How Does Future Income Affect Current Consumption? Quarterly
Journal of Economics, 109, 111-147.

Carroll, C. D., and Samwick, A. (1997). The Nature of Precautionary Wealth. Journal of
Monetary Economics, 40, 41-71.

Cecchetti, S. G., Lam, P. -S., and Mark, N. C. (1993). The Equity Premium and the Risk-
Free Rate: Matching the Moments. Journal of Monetary Economics, 32, 21-45.

Christiano, L. J. (1990a). Solving the Stochastic Growth Model by Linear-Quadratic
Approximation and by Value-Function Iteration. Journal of Business and Economic
Statistics, 8, 23-26.

Christiano, L. J. (1990b). Linear-Quadratic Approximation and Value-function Iteration:
A Comparison. Journal of Business and Economic Statistics, 8, 99-113.

Christiano, L. J. and Fischer, J. D. M. (2000). Algorithms for Solving Dynamic Models
with Occasionally Binding Constraints. Journal of Economic Dynamics and Control, 24,
1179-1232.

Chow, G. C. (1973). Effect of Uncertainty on Optimal Control Policies. International
Economic Review, 14, 632-645.

Chow, G. C. (1976). The Control of Nonlinear Econometric Systems with Unknown
Parameters. Econometrica, 44, 685-695.

Coleman, W.J., II (1990). Solving the Stochastic Growth Model by Policy-Function
Iteration. Journal of Business and Economic Statistics, 8, 27-29.

Collard, F. and Juillard, M. (2001). Accuracy of Stochastic Perturbation Methods: The
Case of Asset Pricing Models. Journal of Economic Dynamics and Control, 25, 979-99.

Conlisk, J. (1996). Why Bounded Rationality? Journal of Economic Literature, 34, 669-
700.

Dardanoni, V. (1991). Precautionary Savings under Income Uncertainty: A Cross-
Sectional Analysis. Applied Economics, 23, 153-160.

Deaton, A. and Laroque, G. (1992). On the Behavior of Commodity Prices. Review of
Economic Studies, 59, 1-23.

Den Haan, W. J. and Marcet, A. (1990). Solving the Stochastic Growth Model by
Parameterizing Expectations. Journal of Business and Economic Statistics, 8, 31-34.

Den Haan, W. J., and Marcet, A. (1994). Accuracy in Simulations. Review of Economic
Studies, 61, 3-17.





128
Dotsey, M. and Mao, C. S. (1992). How Well Do Linear Approximation Methods Work?
The Production Tax Case. Journal of Monetary Economics, 29, 25-58.

Dynan, K. (1993). How Prudent Are Consumers? Journal of Political Economy, 101,
1104-1113.

Fair, R. C. (2003). Optimal Control and Stochastic Simulation of Large Nonlinear
Models with Rational Expectations. Journal of Economic Dynamics and Control, 21, 245-
256.

Fair, R. C., and Taylor, J.B. (1983). Solution and Maximum Likelihood Estimation of
Dynamic Nonlinear Rational Expectations Models. Econometrica, 51, 1169-1185.

Fuhrer, J. C., and Bleakley, C. H. (1996). Computationally Efficient Solution and
Maximum Likelihood Estimation of Nonlinear Rational Expectations Models. Mimeo,
Federal Reserve Bank of Boston.

Gagnon, J. E. (1990). Solving the Stochastic Growth Model by Deterministic Extended
Path. Journal of Business and Economic Statistics, 8, 35-36.

Gali, J., Lopez-Salido, J. D., Valles, J. (2004). Rule-of-Thumb Consumers and the Design
of Interest Rate Rules. NBER Working Paper Series. Working Paper 10392.

Guariglia, A. (2001). Saving Behaviour and Earnings Uncertainty: Evidence from the
British Household Panel Survey. Journal of Population Economics, 14, 619-634.

Guiso, L., Jappelli, T., Terlizzese, D. (1992). Earnings Uncertainty and Precautionary
Saving. Journal of Monetary Economics, 30, 307-337.

Helgason, T., and Wallace, S. W. (1991a). Approximate Scenario Solutions in the
Progressive Hedging Algorithm. Annals of Operation Research, 31, 437-444.

Helgason, T., and Wallace, S. W. (1991b). Structural Properties of the Progressive
Hedging Algorithm. Annals of Operation Research, 31, 445-456.

Hey, J. D., and Dardanoni, V. (1988). Optimal Consumption under Uncertainty: An
Experimental Investigation. The Economic Journal, 98, 105-116.

Ingram, B. F. (1990). Equilibrium Modeling of Asset Prices: Rationality versus Rules of
Thumb. Journal of Business and Economic Statistics, 8, 115-125.

Judd, K.L. (1992). Projection Methods for Solving Aggregate Growth Models. Journal of
Economic Theory, 58, 410-452.

Judd, K. L. (1998). Numerical Methods in Economics. MIT Press, Cambridge, MA.

Klenow, P.J. (1991). Externalities and Business Cycles. Ph.D. thesis, Department of
Economics, Stanford University.



129
Krusell, P., Smith A. A. Jr. (1996). Rules of Thumb in Macroeconomic Equilibrium - A
Quantitative Analysis. Journal of Economic Dynamics and Control, 20, 527-558.

Kydland, F., and Prescott, E. (1982). Time to Build and Aggregate Fluctuations.
Econometrica, 50, 1345-1370.

Lettau, M., and Uhlig, H. (1999). Rules of Thumb versus Dynamic Programming. The
American Economic Review, 89, 141-172.

Lusardi, A. (1997). Precautionary Saving and Subjective Earnings Variance. Economic
Letters, 57, 319-326.

Marcet, A. (1988). Solving Nonlinear Stochastic Growth Models by Parameterizing
Expectations. Carnegie-Mellon University manuscript.

Marcet, A. (1994). Simulation Analysis of Dynamic Stochastic Models: Application to
Theory and Estimation. Advances in Econometrics, Sixth World Congress, Vol. II, ed. by
C. Sims. Cambridge University Press, Cambridge U.K., 91-118.

Marcet, A., and Lorenzoni, G. (1999). The Parameterized Expectations Approach; Some
Practical Issues. Computational methods for the study of dynamic economies, ed. by R.
Marimon and A. Scott. Oxford University Press, Oxford and New York, 143-171.

Marcet, A. and Marshall, D. A. (1994a). Convergence of Approximate Model Solutions
to Rational Expectations Equilibria using the Method of Parameterized Expectations.
Working Paper No. 73, Department of Finance, Kellogg Graduate School of
Management, Northwestern University.

Marcet, A. and Marshall, D. A. (1994b). Solving Non-linear Rational Expectations
Models by Parameterized Expectations: Convergence to Stationary Solutions. Federal
Reserve Bank of Chicago, Working Paper 94-20.

McGrattan, E. R. (1990). Solving the Stochastic Growth Model by Linear-Quadratic
Approximation. Journal of Business and Economic Statistics, 8, 41-44.

McGrattan, E. R. (1996). Solving the Stochastic Growth Model with a Finite-Element
Method. Journal of Economic Dynamics and Control, 20, 19-42.

McGrattan, E. R. (1999). Application of Weighted Residual Methods to Dynamic
Economic Models. Computational methods for the study of dynamic economies, ed. by R.
Marimon and A. Scott. Oxford University Press, Oxford and New York, 114-142.

Mehra, R. and Prescott, E. C. (1985). The Equity Premium: A Puzzle. Journal of
Monetary Economics, 15, 145-161.

Merrigan, P., Normandin, M. (1996). Precautionary Saving Motives: An Assessment
from UK Time Series of Cross-Sections. Economic Journal, 106, 1193-1208.




130
Miles, D. (1997). A Household Level Study of the Determinants of Income and
Consumption. Economic Journal, 107, 1-25.

Miranda, M.J., and Helmberger, P.G. (1988). The Effects of Commodity Price
Stabilization Programs. American Economic Review, 78, 46-58.

Miranda, M. J., and Rui, X. (1997). Maximum Likelihood Estimation of the Nonlinear
Rational Expectations Asset Pricing Model. Journal of Economic Dynamics and Control,
21, 1493-1510.

Mulvey, J. M., and Ruszczynski, A. (1992). A diagonal quadratic approximation method
for large scale linear programs. Operations Research Letters, 12, 205-215.

Novales, A. et al. (1999). Solving Nonlinear Rational Expectations Models by
Eigenvalue-Eigenvector Decomposition. Computational methods for the study of
dynamic economies, ed. by R. Marimon and A. Scott. Oxford University Press, Oxford
and New York, 62-92.

Prucha, I. R., and Nadiri, M. I. (1984). Formulation and Estimation of Dynamic Factor
Demand Equations under Non-Static Expectations: A Finite Horizon Model, NBER
Working Paper Series. Revised technical working paper no. 26.

Prucha, I. R., and Nadiri, M. I. (1986). A Comparison of Alternative Methods for the
Estimation of Dynamic Factor Demand Models under Non-Static Expectations. Journal
of Econometrics 33, 187-211.

Prucha, I. R., and Nadiri, M. I. (1991). On the Specification of Accelerator Coefficients in
Dynamic Factor Demand Models. Economic Letters 35, 123-129.

Reiter, M. (2000). Estimating the Accuracy of Numerical Solutions to Dynamic
Optimization Problems. Mimeo.

Rockafellar, R. T., and Wets, R. J. -B. (1991). Scenarios and Policy Aggregation in
Optimization under Uncertainty. Mathematics of Operations Research, 16, 1-23.

Rosa, C., and Ruszczynski, A. (1994). On Augmented Lagrangian Decomposition
Methods for Multistage Stochastic Programming. International Institute for Applied
Analysis, Working Paper WP-94-125.

Rust, J. (1996). Numerical Dynamic Programming in Economics. Handbook of
Computational Economics, Vol. I, ed. by H. Amman, D. Kendrick and J. Rust.
Amsterdam: North-Holland, 619-729.

Rust, J. (1997). A Comparison of Policy Iteration Methods for Solving Continuous-state,
Infinite-horizon Markovian Decision Problems Using Random, Quasi-Random, and
Deterministic Discretizations. Manuscript, Yale University.





131
Ruszczynski, A. (1986). A Regularized Decomposition Method for Minimizing a Sum of
Polyhedral Functions. Mathematical Programming, 35, 309-333.

Ruszczynski, A. (1989). An Augmented Lagrangian Decomposition Method for Block
Diagonal Linear Programming Problems. Operations Research Letters, 8, 287-294.

Ruszczynski, A. (1993). Parallel Decomposition of Multistage Stochastic Programs.
Mathematical Programming, 58, 201-228.

Santos, M. S. (2000). Accuracy of Numerical Solutions Using the Euler Equation
Residuals. Econometrica, 68, 1377-1402.

Santos, M. S. and Vigo, J. (1998). Analysis of a Numerical Dynamic Programming
Algorithm Applied to Economic Models. Econometrica, 66, 409-426.

Sargent, T. J. (1993). Bounded Rationality in Macroeconomics. Oxford University Press.

Schmitt-Grohé, S. and Uribe, M. (2004). Solving Dynamic General Equilibrium Models
Using a Second-Order Approximation to the Policy Function. Journal of Economic
Dynamics and Control, 28, 755-775.

Skinner, J. (1988). Risky Income, Life Cycle Consumption, and Precautionary Savings.
Journal of Monetary Economics, 22, 237-255.

Tauchen, G. (1990). Solving the Stochastic Growth Model by Using Quadrature Methods
and Value-Function Iterations. Journal of Business and Economic Statistics, 8, 49-51.

Tauchen, G. and Hussey, R. (1991). Quadrature-Based Methods for Obtaining
Approximate Solutions to Nonlinear Asset Pricing Models. Econometrica, 59, 371-396.

Taylor, J. B. and Uhlig, H. (1990). Solving Nonlinear Stochastic Growth Models: A
Comparison of Alternative Solution Methods. Journal of Business and Economic
Statistics, 8, 1-17.

Uhlig, H. (1999). A Toolkit for Analyzing Nonlinear Dynamic Stochastic Models Easily.
Computational methods for the study of dynamic economies, ed. by R. Marimon and A.
Scott. Oxford University Press, Oxford and New York, 30-61.

Van Slyke, R., and Wets, R. J. -B. (1969). L-Shaped Linear Programs with Applications
to Optimal Control and Stochastic Programming. SIAM Journal on Applied Mathematics,
17, 638-663.

Wets, R. J. -B. (1988). Large Scale Linear Programming. Numerical methods in
stochastic programming, ed. by Yu Ermoliev and R. J. -B. Wets. Springer Verlag, Berlin,
65-94.

Wright, B.D. and Willam, J.C. (1982a). The Economic Role of Commodity Storage.
Economic Journal, 92, 596-614.



132
Wright, B.D. and Willam, J.C. (1982b). The Roles of Public and Private Storage in
Managing Oil Import Disruptions. Bell Journal of Economics, 13, 341-353.

Wright, B.D. and Willam, J.C. (1984). The Welfare Effects of the Introduction of
Storage. Quarterly Journal of Economics, 99, 169-182.


















































133

doc_731940249.docx
 

Attachments

Back
Top