Difference Equations
Difference Equations
Klaus Neusser
October 3, 2016
1
Klaus
c Neusser
i
Preface
There are, of course, excellent books on deterministic difference equations:
for example, Elaydi (2005), Agarwal (2000), or Galor (2007). Colonius and
Kliemann (2014) give a presentation from the perspective of dynamical sys-
tems. These books do, however, not go into the specific problems faced in
economics.
The books makes use of linear algebra. Very good introduction to this
topic is presented by the books of Strang (2003) and Meyer (2000).
Bern,
September 2010 Klaus Neusser
ii
Contents
1 Introduction 1
1.1 Notation and Preliminaries . . . . . . . . . . . . . . . . . . . . 1
iii
iv CONTENTS
v
vi LIST OF FIGURES
vii
viii LIST OF DEFINITIONS
List of Theorems
ix
Chapter 1
Introduction
1
2 CHAPTER 1. INTRODUCTION
The aim of the analysis is to assess the existence and uniqueness of a solution
to a given difference equation; and, in the case of many solutions, to charac-
terize the set of all solutions. In the normal linear case of dimension n = 1,
for example, the set of solutions turns out to be a vector space of dimension
p.
One way to pin down a particular solution is to require that the solution
must satisfy some boundary conditions. In this case, we speak of a boundary
value problem. The simplest way to specify boundary conditions is to require
that the solution X(t) must be equal to p prescribed values x1 , . . . , xp , called
initial conditions, at given time indices t1 , . . . , tp :
k(X)k = sup{kXt k, t Z}
(L)Xt = (I 1 L 2 L2 . . . p Lp )Xt
= Xt 1 Xt1 2 Xt2 . . . p Xtp
Xt = Xt1 + Zt , 6= 0. (2.2)
5
6 CHAPTER 2. LINEAR DIFFERENCE EQUATIONS
equation (2.3)
X1 = x0
X2 = X1 = 2 x0
...
Xt = Xt1 = t x0 .
This suggests to take
Xt = t c (2.4)
as the general solution of the first order linear homogenous difference equa-
tion (2.3). Actually, equation (2.4) provides a whole family of solutions
indexed by the parameter c R. To each value of c, there corresponds a
trajectory (Xt ) = (t c). In order to highlight this dependency, we may write
the solutions as Xt (c). Note that the trajectories of two different solutions
Xt (c1 ) and Xt (c2 ), c1 6= c2 , cannot cross.
The parameter c can be pinned down by using a single boundary condition.
A simple form of such a boundary condition requires, for example, that Xt
takes a particular value xt0 in some period t0 . Thus, we require that Xt0 = xt0
in period t0 . In this case we speak of an initial value problem. The value of
c can then be retrieved by solving the equation xt0 = t0 c for c. This leads
x
to c = tt00 . The solution may then be written as
or equivalently
!
(1) (2)
Xt Xt a1
(1) (2) 6= 0
Xt Xt a2
Since this must hold for any a1 , a2 , not both equal to zero, the determinant
of the Casarotian matrix (see Definition 2.3)
!
(1) (2)
Xt Xt
C(t) = (1) (2)
Xt Xt
8 CHAPTER 2. LINEAR DIFFERENCE EQUATIONS
3. The Superposition Principle (see Theorem 2.2) then delivers the gen-
eral solution of the nonhomogeneous equation as the sum of (X (g) ) and
(X (p) ). However, this solution still depends through (X (g) ) on some
constants. To pin down the solution uniquely and therefore solving
the boundary value problem requires additional conditions. These con-
ditions can come in the form of initial values (starting values) or in
the form of requirements that the solution must obey some qualitative
feature. A typical feature in this context is boundedness, a condition
which usually can be given an economic underpinning.
Before continuing with the theoretical analysis consider the following ba-
sic example.
Amortization of a Loan
One of the simplest settings in economics where a difference equation arises
naturally, is compound interest calculation. Take, for example, the evolution
of debt. Denote by Dt the debt outstanding at the beginning of period t,
then the debt in the subsequent period t + 1, Dt+1 , is obtained by the simple
accounting rule:
Dt+1 = Dt + rDt Zt = (1 + r)Dt Zt (2.6)
where rDt is the interest accruing at the end of period t. Here we are using for
simplicity a constant interest rate r. The debt contract is serviced by paying
the amount Zt at the end of period t. This payment typically includes a
payment for the interest and a repayment of the principal. Equation (2.6)
constitutes a linear nonhomogeneous first order difference equation with =
1 + r.
Given the initial debt at the beginning of period 0, D0 , the amount of
debt outstanding in subsequent periods can be computed recursively using
the accounting rule (2.6):
D1 = (1 + r)D0 Z0
D2 = (1 + r)D1 Z1 = (1 + r)2 D0 (1 + r)Z0 Z1
...
Dt+1 = (1 + r)t+1 D0 Zt (1 + r)Zt1 (1 + r)t Z0
Xt
= (1 + r)t+1 D0 (1 + r)i Zti
i=0
t+1
Note
Pt how Dt+1i is determined as the sum of two parts: (1 + r) D0 and
i=0 (1 + r) Zti . The first expression thereby corresponds to the general
10 CHAPTER 2. LINEAR DIFFERENCE EQUATIONS
economics). Given the difference equation for the evolution of debt, the NPG
condition with constant payment per period is equivalent to:
DT +1 T 1 Z
Z
lim T +1
= lim D0 1 (1 + r) = D0 0
T (1 + r) T r r
which implies that Z rD0 . Thus, the NPG condition holds if the constant
repayments Z are at least as great as the interest.
0.9
0.8 steady
state 450line
0.7
0.6
Xt+1
0.5
0.4
f(x)
0.3
0.2
0.1 steady
state
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
X
t
Figure 2.1: Cobweb diagram with steady states of the logistic function: y =
2.5x(1 x) and x0 = 0.1
This provides an example with two steady states. The steady states are
determined by the equation: X = X (1 X ). This equation has two
solutions which give the corresponding two steady states: X = 0 and X =
1
with 0 < X < 1. f (x) attains a maximum value of /4 at x = 1/2.
Figure 2.1 shows the two steady states and the evolution of Xt starting at
X0 = x0 = 0.1 and taking = 2.5.
Another example consists of a linear first order nonhomogeneous differ-
ence equation with independent or forcing variable Zt constant over time, i.e.
Zt = Z. It is easy to compute the steady state in this simple case:
Z
X = X + Z X = for 6= 1. (2.10)
1
Z
Xt = t c + X = t c + for 6= 1.
1
Xt X = (Xt1 X ).
Xt+1 = f (Xt )
|X1 X | = |f (X0 ) f (X )|
The mean value theorem then implies that there exists , X0 < < X , such
that
|f (X0 ) f (X )| = |f 0 ()| |X0 X | .
Hence we have
|X1 X | M |X0 X | .
This shows that X1 is closer to X and is thus also in J because M < 1. By
induction we therefore conclude that
|Xt X | M t |X0 X | .
For any > 0, let = min{, } then |X0 X | < implies |Xt X | <
for all t 0. X is therefore a stable equilibrium point. In addition, X
is attractive because limt |Xt X | = 0. Thus, X is asymptotically
stable.
X2 a
1 a
Xt+1 = Xt t = Xt + .
2Xt 2 Xt
Starting with X0 > 0, the difference equation converges to a whereas if
X0 < 0 the limit is a. The above
difference equation
has a nice intuitive
interpretation. Suppose Xt > a then a/Xt < a, thus by taking the
arithmetic average between Xt and a/X t one can expect to get closer to a.
The argument holds similarly for Xt < a.
Xt = Xt1 + Zt
Xt = (Xt2 + Zt1 ) + Zt = 2 Xt2 + Zt1 + Zt
...
Xt = t X0 + t1 Z1 + t2 Z2 + + Zt1 + Zt
t1
X
t
= X0 + j Ztj
j=0
2.3. SOLUTIONS OF FIRST ORDER EQUATIONS 17
Taking the absolute value of the difference between Xt and the second term
of the right hand side of the equation leads to:
t1
X
Xt j Ztj = t X0 = |t | |X0 |
j=0
which is just the steady state solution described in equation (2.10) of section
2.2.
The requirement that Zt remains bounded can, for example, be violated
if Zt itself satisfies the homogenous difference equation Zt = Zt1 which
implies that Zt = t c for some c 6= 0. Inserting this into equation (2.13)
then leads to j
(b)
X
j tj t
X
Xt = c= c.
j=0 j=0
The infinite sum converges only if | / | < 1. This shows that besides
the stability condition || < 1, some additional requirements with respect
(b)
to the sequence of the exogenous variable are necessary to render Xt in
equation (2.13) a meaningful particular solution. Usually, we assume that
(Zt ) is bounded, i.e. that (Zt ) ` .
Consider next the case || > 1. In this situation the above iteration is
(b)
no longer successful because Xt in equation (2.13) is not welldefined even
18 CHAPTER 2. LINEAR DIFFERENCE EQUATIONS
Xt = 1 Xt+1 1 Zt+1
= 1 1 Xt+2 1 Zt+2 1 Zt+1 = 2 Xt+2 2 Zt+2 1 Zt+1
...
Xh
= h Xt+h 1 j+1 Zt+j for h 1.
j=1
Taking the absolute value of the difference between Xt and the second term
on the right hand side of the equation leads to:
h
X
1 j+1
Zt+j = h Xt+h = |h | |Xt+h |.
Xt +
j=1
As the economy is expected to live forever, there is no end period and the
forward iteration can be carried out indefinitely into the future. Because
|| > 1, the right hand side of the equation converges to zero as h ,
provided that Xt+h remains bounded. This suggests the following particular
solution:
X
(f ) 1
Xt = j+1 Zt+j , || > 1, (2.14)
j=1
where the superscript (f ) indicates that the solution was obtained by iter-
ating the difference equation forward in time. For this to be a meaningful
choice, the infinite sum must be well-defined. This will be guaranteed if, for
example, Zt remains bounded, i.e. if (Z) ` .
In the case || = 1 neither the backward nor the forward iteration strategy
leads to a sensible solution even when Zt is constant and equal to Z 6=
0. Either an equilibrium point does not exist as in the case = 1 or the
equilibrium point exists as is the case for = 1, but Xt oscillates forever
between X0 and X0 + Z so that the equilibrium point is unstable. Most of
the time, we restrict ourself to the case of hyperbolic situations and exclude
the case || = 1.
To summarize, assuming that (Zt ) is bounded, the first order linear dif-
ference equation (2.2) led us to consider the following two representations of
7
Note however that, if Zt satisfies itself a homogeneous difference equation of the form
Zt = Zt1 , the criterion for convergence is, as in the example above, | / | < 1.
2.4. EXAMPLES OF FIRST ORDER EQUATIONS 19
(b)
Note that these equations imply that cb = X0 X0 , respectively that cf =
(f )
X0 X0 . Depending on the value of , we can distinguish the following
three cases:
|| < 1: the backward solution is asymptotically stable in the sense that Xt
(b) (b)
approaches Xt as t . Any deviation of Xt from Xt vanishes
over time, irrespective of the value chosen for cb . The forward solution,
(f )
usually, makes no sense because Xt does not remain bounded even if
the forcing variable Zt is constant over time.
|| > 1: both solutions have an explosive behavior due to the term t . Even
(b) (f )
small deviations from either Xt or Xt will grow without bounds.
There is, however, one and only one solution which remains bounded.
It is given by cf = 0 which implies that Xt always equals its equilibrium
(f )
value Xt .
|| = 1: neither the backward nor the forward solution converge for constant
Zt 6= 0.
Which solution is appropriate depends on the nature of the economic prob-
lem at hand. In particular, the choice of the boundary condition requires
some additional thoughts and cannot be determined on general grounds. As
the exercises below demonstrate, the nature of the expectations formation
mechanism is sometimes decisive.
Dt = pt , >0 (demand)
St = pet + ut , >0 (supply)
St = Dt (market clearing)
pet = pt1 (expectations formation)
The Figure 2.2 depicts several possible cases depending on the relative
slopes of supply and demand. In the first panel = 0.8 so that we have
an asymptotically stable equilibrium. Starting at p0 , the price approaches
the steady state by oscillating around it. In the second panel = 1 so
that independently of the starting value, the price oscillates forever between
p0 and p1 . In the third panel, we have an unstable equilibrium. Starting at
p0 6= p , pt diverges.
8
The logarithm of the price level is taken to ensure a positive price level.
2.4. EXAMPLES OF FIRST ORDER EQUATIONS 21
1.4
45line 1
1.2
p1
0.9
45line
1 0.8
P
1
0.7
0.8 steady
0.6
Pt+1
steady state
Pt+1
state 0.5
0.6
(/)p + 1 0.4
t
0.4
0.3
P2
0.2 (/)pt + 1
0.2
p 0.1
2
0 0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 0 0.2 0.4 0.6 0.8 1
P P1 p0
0 Pt Pt p1
1
45line
0.9
0.8
p 0.7
1
0.6 steady
state
Pt+1
0.5
0.4 (/)pt + 1
p 0.3
2
0.2
0.1
0
0 0.2 0.4 0.6 p 0.8 1
p0
Pt 1
The constant returns to scale assumption then implies that, for L > 0 fixed,
Y
lim = lim F (1, L/K) = F (1, 0) = 0.
K K K
F (K, 0) = KF (1, 0) = 0.
where f (k) = F (k, 1). Holding K > 0 fixed, LHopitals rule implies
f (k)
lim F (K, L) = K lim = K lim f 0 (k) =
L k0 k k0
where we have used the result that capital is essential (see the previous
Lemma), i.e. that f (0) = 0. The last equality is a consequence of the Inada
conditions. The proof for limK F (K, L) = , L > 0 fixed, is analogous.
Yt = Ct + It . (2.17)
It = sYt . (2.18)
Investment adds to the existing capital stock which depreciates in each period
at a constant rate (0, 1):
Starting in period 0 with some positive capital K0 > 0, the system con-
sisting of the two difference equations (2.19) and (2.20) completely describes
the evolution of the economy over time. A first inspection of the two equa-
tions immediately reveals that both labor and capital tend to infinity. Indeed,
as > 0 labor grows without bound implying according to Lemma 2.2 that
output also grows without bound. This is not very revealing if one is looking
for steady states and is interested in a stability analysis. In such a situation
it is often advisable to look at the ratio of the two variables, in our case at
K/L. This has two main advantages. First, the dimension of the system
is reduced to one and, more importantly, the singularity at infinity is, at
least in the linear case, eliminated.9 Second, these ratios often have a clear
economic meaning making the economic interpretation of the results more
comprehensible.
We apply this device to the Solow model as described by equations (2.19)
and (2.20). Thus, dividing equation (2.19) by Lt+1 and making use of the
constant returns to scale assumption results in the fundamental equation of
the Solow model:
Kt+1 1 s
kt+1 = = kt + f (kt ) = g(kt ) (2.21)
Lt+1 1+ 1+
where kt = K t
Lt
is known as the capital intensity and f (k t ) = F Kt
Lt
, 1 .
The economy starts in period zero with an initial capital intensity k0 > 0.
The nonlinear first order difference equation (2.21) together with the initial
condition uniquely determines the evolution of the capital intensity over time,
and consequently of all other variables in the model. Note that the concavity
of F is inherited by f and thus by g so that we have g 0 > 0 and g 00 < 0.
Moreover, limk0 g(k) = 0 and limk g(k) = .
Proposition 2.1. Given the assumptions of the Solow model, the funda-
mental Solow equation (2.21) has two steady states k = 0 and k > 0.
9
Technically speaking, this induces a new difference equation on the projective space.
In the two dimensional case, the projective space is defined as the set of rays through the
origin. As each ray crosses the unit circle twice, an equivalent definition is given as the
unit circle where opposite points are not distinguished. See Colonius and Kliemann (2014,
chapter 4) for details.
2.4. EXAMPLES OF FIRST ORDER EQUATIONS 25
0.35
steady state
0.3
k*
0.25
capital intensity
g(k)
0.15
45degree line
0.1
0.05
0
0 0.05 0.1 0.15 0.2 0.25 k* 0.3
capital intensity
k = g(k ).
We can therefore study the local behavior of the nonlinear difference equa-
tion (2.21) around the steady state k > 0 by investigating the properties of
the first order homogenous difference equation:
Proof. Note that g 0 (k) > 0 for all k > 0. Concavity of g implies that g(k)
k g 0 (k )(k k ) for all k > 0. Take k < k , then g(k) > k. Thus,
g(k) k < g 0 (k )(g(k) k ) < 0 so that g 0 (k ) < 1.
Starting in period zero with an initial capital intensity k0 > 0, the solution
to this initial value problem is:
kt = k + t (k0 k )
dt + pet+1 pt
r= pet+1 = (1 + r)pt dt (2.25)
pt
2.4. EXAMPLES OF FIRST ORDER EQUATIONS 27
where pet+1 denotes the price expected to prevail in the next period. Assuming
that expectations of the investors are rational which is equivalent to assum-
ing perfect foresight in the context of no uncertainty, the above arbitrage
equation turns into a simple first order difference equation:
where Xt = (1 + r)1
(f ) P j
j=0 (1 + r) dt+j . Note that the forward solution is
only well-defined if the infinite sum converges. A sufficient condition for this
to happen is the existence of a finite index j0 such that |dt+j /(1 + r)j | < M j ,
for j > j0 and some M < 1. This is guaranteed, in particular, by a constant
dividend stream dt+j = d, for all j = 0, 1, 2, . . .
The term (1 + r)t cf is usually called the bubble term because its behavior
(f )
is unrelated to the dividend stream; whereas the term Xt is referred to as
the fundamentals because it is supposed to reflect the intrinsic value of the
share.
Remember that we want to figure out the price of a share. Take period
(f )
0 to be the current period and suppose that cf = p0 X0 > 0. This
means that the current stock price is higher than what can be justified by
the future dividend stream. According to the arbitrage equation (2.25) this
high price (compared to the dividend stream) can only be justified by an
appropriate capital gain, i.e. an appropriate expected price increase in the
next period. This makes the price in the next period even more different
from the fundamentals which must be justified by an even greater capital
gain in the following period, and so on. In the end, the bubble term takes
over and the share price becomes almost unrelated to the dividend stream.
This situation is, however, not sustainable in the long run.10 Therefore, the
10 (f )
A similar argument applies to the case cf = p0 X0 < 0.
28 CHAPTER 2. LINEAR DIFFERENCE EQUATIONS
(f )
only reasonable current share price p0 is X0 which implies that cf = 0. This
effectively eliminates the bubble term and is actually the only nonexplosive
solution. Thus, we have a unique (determinate) rational equilibrium solution.
This solution is
X
(f ) 1
pt = Xt = (1 + r) (1 + r)j dt+j (2.28)
j=0
Thus, the price of a share always equals the present discounted value of the
corresponding dividend stream. Such a solution is reasonable in a situation
with no uncertainty and no information problems. Note this solution implies
that the price immediately responds to any change in the expected dividend
stream. The effect of a change in dt+h , h = 0, 1, 2, . . . on pt is given by
pt
= (1 + r)h1 h = 0, 1, . . .
dt+h
Thus, the effect diminishes the further the change takes place in the fu-
ture. Consider now a permanent change in dividends, i.e. a change where all
dividends increase by some constant amount 4d. The corresponding price
change 4pt equals:
1
X 4d
4pt = (1 + r) (1 + r)j 4d =
j=0
r
1
pt = pt1 + mt = pt1 + Zt (2.29)
1 + 1 +
1
where = 1+ and Zt = 1+ mt .
From our previous discussion we know that the general solution of this
difference equation is given as the sum of the general solution to the ho-
(p)
mogenous equation and a particular solution, pt , to the nonhomogeneous
equation:
(p)
pt = t c + pt
11
See also the analysis in Sargent (1987).
30 CHAPTER 2. LINEAR DIFFERENCE EQUATIONS
p1 = p0 + Z1
p2 = p1 + Z2 = 2 p0 + Z1 + Z2
...
pt = t p0 + t1 Z1 + t2 Z2 + + Zt1 + Zt
t1
X
= t p0 + i Zti
i=0
From a mathematical point of view this expression only makes sense if the
limit of the infinite sum exists. Thus, additional assumptions are required.
Suppose that logged money remained constant, i.e. mt = m < for all
t, then the logic of the model suggests that the logged price
Pleveli should
remain finite as well. In mathematical terms this means that i should
converge. This is, however, a geometric sum so that convergence is achieved
if and only if
|| = <1 (2.30)
1 +
Assuming that this stability condition holds, the general solution of the dif-
ference equation (2.29) implied by the Cagan model is:
X
t
pt = c + i Zti (2.31)
i=0
where the constant c can be computed from an initial value condition. Such
an initial condition arises naturally because the formation of adaptive expec-
tations requires the knowledge of the price from the previous period which
can then serve as an initial condition.
The stability condition therefore has important consequences. First, irre-
spective of the value of c, the first term of the solution (the general solution
2.4. EXAMPLES OF FIRST ORDER EQUATIONS 31
Thus, the price level moves up by the same percentage point. Such a once-
and-for-all change is termed a permanent change. In contrast a transitory
change is a change which occurs only once. The effect of a transitory change
of mt by m in period t on the logged price level in period t + h for some
h 0 is given by
h
1h 1
pt+h = m = m
1 + 1 + 1 +
1 mt
pt+1 = pt + = pt + Zt (2.33)
with Zt = mt /. As = 1
> 1, the stability condition is violated. One
can nevertheless find a meaningful particular solution of the nonhomogeneous
equation by iterating the difference equation forwards in time instead of
32 CHAPTER 2. LINEAR DIFFERENCE EQUATIONS
1.5
0.5
0.5
1.5
2
0 2 4 6 8 10 12 14 16 18 20
h
Figure 2.4: Impulse response function of the Cagan model with adaptive
expectations taking = 0.5 and = 0.9
backwards:
pt = 1 pt+1 1 Zt
= 1 1 pt+2 1 Zt+1 1 Zt = 2 pt+2 2 Zt+1 1 Zt
...
Xh1
h 1
= pt+h i Zt+i for h > 0
i=0
The logged price level in period t, pt , now depends on some expected logged
price level in the future, pt+h , and on the development of logged money
expected to be realized in the future. Because the economy is expected to
live forever, this forward iteration is carried on into the infinite future to
yield:
X
pt = lim h pt+h 1 i Zt+i
h
i=0
As 0 < 1 < 1, the limit and the infinite sum are well defined, provided
that the logged money stock remains bounded. Under the assumption that
the logged money stock is expected to remain bounded, the economic logic
of the model suggests that the logged price level should remain bounded as
2.4. EXAMPLES OF FIRST ORDER EQUATIONS 33
X
(p) 1
t
pt = c + pt t
= c i Zt+i (2.34)
i=0
Due to the term t c, the logged price level grows exponentially without bound
although the logged money stock may be expected to remain bounded, unless
(p)
c = 0. Thus, setting c = 0 or equivalently p0 = p0 guarantees a nonexplosive
rational expectations equilibrium.
To summarize, the Cagan model suggests the following two solutions:
X
t (b) (b)
p t = cb + pt , whereby pt = i Zti
i=0
X
(f ) (f )
pt = t cf + pt , whereby pt = 1 i Zt+i
i=0
distinct roots
In this case we have the following theorem.
Theorem 2.5 (Fundamental Set for equation of order p). If all the roots of
the characteristic equation are distinct, the set {t1 , . . . , tp } forms a funda-
mental set of solutions.
2.5. DIFFERENCE EQUATIONS OF ORDER P 35
Proof. It suffices to show that det C(t) 6= 0 where C(t) is the Casarotian
matrix of {t1 , . . . , tp }.
t
1 t2 ... tp
t+1 t+1 . . . t+1
1 2 p
det C(t) = det ..
.. ... ..
. . .
t+p1 t+p1 t+p1
1 2 . . . p
1 1 ... 1
1 2 . . . p
= t1 t2 . . . tp det .. .. ..
...
. . .
p1
1 p1
2 . . . p1
p
This second
Q matrix is called the Vandermonde matrix whose determinant
equals 1i<jp (j i ) which is different from zero because the roots are
distinct. Thus, det C(t) 6= 0, because the roots are also different from zero.
The above Theorem thus implies that the general solution to the homo-
(g)
geneous equation Xt is given by
(g)
Xt = c1 t1 + c2 t2 + + cp tp . (2.37)
Using the same technique as in the proof of Theorem 2.1, it is easy to demon-
strate that the set of solutions forms a linear space of dimension p.
multiple roots
When the roots of the characteristic equation are not distinct, the situa-
tion becomes more complicated. Denote the r distinct roots by z1 , , zr ,
r < p, and their corresponding multiplicities by m1 , , mr . Writing the
homogeneous difference equation in terms of the lag operator leads to
(1 1 L p Lp ) Xt
= (1 1 L)m1 (1 2 L)m2 (1 r L)mr Xt = 0 (2.38)
where i , 1 i r equals z1i . In order to find the general solution, we will
proceed in several steps. First note if t is a solution to
(1 i L)mi t = 0 (2.39)
it is also a solution to (2.38). Second, Gi = {ti , tti , t2 ti , , tmi 1 ti } is
a fundamental set of solutions for equation (2.39). Before we prove this
statement in Lemma 2.4, we need the following lemma.
36 CHAPTER 2. LINEAR DIFFERENCE EQUATIONS
(1 L)k ts = 0, 0s<k
(g)
where Xt is the general solution to the homogeneous equation given by
(p)
equation (2.40) and Xt is a particular solution to the nonhomogeneous
equation.
In the search for a particular solution, the same ideas as in first order case
can be used. If the nonhomogeneous part is constant, i.e. Zt = Z, the steady
state, if it exists, qualifies for a particular solution to the nonhomogeneous
equation. If the nonhomogeneous part depends on time, a particular solution
can be found by iterating the equation backwards and/or forwards depending
on the location of the roots. This will become clear by analyzing the examples
in section 2.5.4.
Higher order equations will add no new qualitative features. Assuming that
1 1 2 6= 0, the unique fixed point of this homogenous equation is 0. The
corresponding characteristic equation is given by the quadratic equation:
1 1 z 2 z 2 = 0.
Or in terms of = z1 :
p
1 21 + 42
1,2 = . (2.42)
2
To understand the qualitative behavior of Xt , we distinguish three cases:
Xt = rt e(t+) + e(t+)
= 2rt cos(t + )
1. r > 1: both roots are outside the unit circle (i.e. the circle of
radius one and centered in the point (0, 0)). Xt oscillates, but
with ever increasing amplitude. The fixed point zero is unstable.
2. r = 1: both roots are on the unit circle. Xt oscillates, but with
constant amplitude.
3. r < 1: both roots are inside the unit circle. The solution oscillates,
but with monotonically decreasing amplitude and converges to
zero as t . The fixed point zero is asymptotically stable.
3 3 3
2 2 2
Xt 1 1 1
Xt
Xt
0 0 0
1 1 1
2 2 2
3 3 3
0 5 10 0 5 10 0 5 10
case 1: 1 = 1.1 case 2: 1 = 1 case 3: 1 = 0.8
3 3 3
2 2 2
1 1 1
Xt
Xt
Xt
0 0 0
1 1 1
2 2 2
3 3 3
0 5 10 0 5 10 0 5 10
case 4: 1 = 0.8 case 5: 1 = 1 case 6: 1 = 1.1
20
0
Xt
20
40
0 1 2 3 4 5 6 7 8 9 10
case 1: roots outside unit circle (1 = 1+i, 2 = 1i)
1
Xt
0 1 2 3 4 5 6 7 8 9 10
case 2: roots on unit circle (1 = (sqrt(2)/2)(1+i), 2 = (sqrt(2)/2)(1i))
1
Xt
0 1 2 3 4 5 6 7 8 9 10
case 3: roots in unit circle (1 = 0.5(1+i), 2 = 0.5(1i))
Similarly,
p p
2 21 + 42 1 21 + 42
1> > 2 =
2 p 2
2
1 1 + 4 + 41
>
q 2
1 (1 + 2)2
=
2
1 1 2
= = 1
2
If the roots are complex, 1 and 2 are complex conjugate numbers. Their
2 (2 +4 )
squared modulus then equals 1 2 = 1 41 2 = 2 . As 2 < 1, the
modulus of both 1 and 2 is smaller than one.
2.5.4 Examples
Multiplier Accelerator model
A classic economic example of a second order difference equation is the
multiplier-accelerator model originally proposed by Samuelson (1939). It
was designed to demonstrate how the interaction of the multiplier and the
accelerator can generate business cycles. The model is one of a closed econ-
omy and consists of a consumption function, an investment function which
incorporates the accelerator idea and the income identity:
2
explosive explosive
oscillations growth
1
asymptotically
stable
2
0
-1
explosive oscillations
-2
parabola: 21 + 4 2 = 0
-3
-3 -2 -1 0 1 2 3
1
to be between zero and one. The remaining parameters of the model, and
, bear no restriction besides that they have to be positive. Inserting the
consumption and the investment equation into the income identity leads to
the following nonhomogeneous second order difference equation:
not too strong, i.e. if < 1. The steady state Y is therefore asymptotically
stable if one imposes this additional requirement. Yt oscillates around its
steady state if, according to Theorem 2.6, there is no real positive inverse
root of the characteristic equation. The inverse of the characteristic roots
are given by p
( + ) ( + )2 4
1,2 = .
2
If the roots are real, they are both strictly positive and strictly smaller than
one. Thus, Yt can only oscillate around its steady state if and only if the
roots are complex, i.e. if ( + )2 4 < 0. If they are complex, their moduli
are strictly smaller than one.
In the general case where government expenditures are not constant, but
vary over time, we apply the method of undetermined coefficients to find
(p)
a particular solution, Yt , to equation (2.44). This method conjectures a
certain type of solution and then tries to pin down a solution by inserting it
into the difference equation. In the particular case at hand, the roots of the
characteristic function are all outside the unit circle. Thus, we conjecture a
particular solution of the form:
X
(p)
Yt =c+ i Gti
i=0
0 = 1
1 = ( + )0 1 = +
2 = ( + )1 0
j = ( + )j1 j2 , j2
j = d1 j1 + d2 j2 .
The coefficients d1 and d2 can then be determined from the initial conditions:
0 = 1 = d1 + d2
1 = + = d1 1 + d2 2
3
= 4
and = 14 In this case we have a multiple root equal to = 0.5. Ac-
cording to equation (2.40) the impulse response coefficients are there-
fore given by j = (d0 + d1 t)t . The constants d0 and d1 can again
be found by solving the equation system: 0 = 1 = d0 and 1 = 1 =
(d0 + d1 ). The solution is given by d0 = 1 and d1 = 1. The corre-
sponding impulse response coefficients are plotted in Figure 2.8. They
resemble very much to those of the previous case. They even die out
more rapidly.
2
= 3
and = 23 In this case the discriminant is negative so that we have
two complex conjugate roots:
1
1,2 = 2 2
3
1 2
d1 =
2 2
1 2
d2 = +
2 2
1.4
1.2
0.8
0.6
real roots:
= 4/5, = 1/5
0.4 multiple roots:
= 3/4, = 1/4
0.2
0.2
complex roots:
0.4
= 2/3, = 2/3
0.6
0 2 4 6 8 10 12 14 16 18 20
period
Dt = pt , >0 (demand)
St = pet + ut , >0 (supply)
It = (pet+1 pt ), >0 (inventory demand)
St = Dt + (It It1 ), (market clearing)
pet = pt , (perfect foresight)
1 z + z 2 = 0
2.5. DIFFERENCE EQUATIONS OF ORDER P 49
This equation implies that the two roots, z1 and z2 , are given by
p
2 4
z1,2 =
2
First note that because > 2 the roots are real, distinct, and positive.
Second they come in reciprocal pairs as z1 z2 = 1. Thus, one root is smaller
than one whereas the other is necessarily greater than one. Thus, we have
one stable and one explosive root. The solution to the homogenous equation
can therefore be written as
pt = c1 t + c2 t
X
pt = j utj
j=
X X X ut
j ut+1j = j utj j ut1j +
j= j= j=
ut+2 : 1 = 2 3
ut+1 : 0 = 1 2
1
ut : 1 = 0 1 +
ut1 2 = 1 0
ut2 : 3 = 2 1
This shows that the j s follow homogenous second order difference equa-
tions:
j = j1 j2 j1
j = j1 j2 j1
j = d1 j + d2 j
j = e1 j + e2 j
If we impose again the requirement that the price must be finite if the supply
shock has always been constant and is expected to remain constant in the
2.5. DIFFERENCE EQUATIONS OF ORDER P 51
This implies that {pt } is a bounded sequence, i.e. that (pt ) ` . In this
case the price pt is just a function of all past shocks and all expected future
shocks.
Another way to represent this solution is to express pt as
X
1
pt = pt1 j ut+j .
j=0
In this expression the double infinite sum is replaced by a single one. This is
due to the fact that the past evolution of supply is now summarized by pt1
which is supposed to be known in period t. The effect of discounted expected
future supply is just as before.
In order to gain a better understanding of the dynamics, we will analyze
the following numerical example. In this example = 20 9
and the parameters
and are such that = 2.05. This implies that + = 19 . The roots are
then given by = 0.8 and 1 = 1.25. The bounded solution is then given
by
X
pt = 0.8|j| utj
j=
Suppose that the supply shock has been constant forever and is expected
to remain constant at u. The above formula then implies that the logged
price level pt equals 9u and that It = 0. Suppose that an unexpected and
transitory positive supply shock of value 1 hits the market in period 0. Then
according to the first panel in Figure 2.9 the price immediately falls by 1.
At the same time inventories rise because prices are expected to move up in
the future due to the transitory nature of the shock. Here we have a typical
price movement: the price falls, but is expected to increase. After the shock
the market adjusts gradually as prices rise to their old level and by running
down inventories.
Consider now a different gedankenexperiment. Suppose that the shock
is not unexpected, but expected to hit the market only in period 5. In
this case, we see a more interesting evolution of prices and inventories. In
period zero when the positive supply shock for period 5 is announced, market
participants expect the price to fall in the future. They therefore want to
get rid of their inventories by trying to selling them already now.13 As a
13
In our example they actually go short as I0 < 0.
52 CHAPTER 2. LINEAR DIFFERENCE EQUATIONS
1
0 2 4 6 8 10 12 14 16 18 20
period
expected positive transitory supply shock in period 5
0.5
logged price level / inventory
inventories
0
0.5
logged price level
1
0 2 4 6 8 10 12 14 16 18 20
period
result, the price and the inventories start to fall already before the supply
shock actually takes place. In period 5 when the supply shock finally hits the
market, market participants expect the price to move up again in the future
which leads to a buildup of inventories. Note that this buildup is done when
the price is low. From period 5 on, the market adjusts like in the previous
case because the supply shock is again assumed to be transitory in nature.
Taylor model
In this example we analyze a simple deterministic version of Taylors stag-
gered wage contract model which also has a backward and forward component
(see Taylor (1980) and Ashenfelter and Card (1982)).14 In this model, half
of the wages have to be contracted in each period for two periods. Thus, in
each period half of the wages are renegotiated taking the wages of the other
group as given. Assuming that the two groups are of equal size, wages are
set according to the following rule:
Thus, wage setting in period t takes into account the wages of contracts still
in force, wt1 , and the expected wage contract in the next period, wt+1 . As
14
The model could equally well be applied to analyze staggered price setting behavior.
2.5. DIFFERENCE EQUATIONS OF ORDER P 53
the two groups are of equal size and power, we weight them equally by 0.5.
In addition wages depend on the state of the economy over the length of the
contract, here represented by aggregate demand averaged over the current
and next period. The aggregate wage in period t, Wt , is then simply the
average over all existing individual contract wages in place in period t:
1
Wt = (wt + wt1 ) (2.49)
2
The model is closed by adding a quantity theoretic aggregate demand equa-
tion relating Wt and yt :
yt = Wt + vt , < 0. (2.50)
The negative sign of reflects the fact that in the absence of full accom-
modation by the monetary authority, higher average nominal wages reduce
aggregate demand. vt represents a shock to aggregate demand.
Putting equations (2.48), (2.50), and (2.49) together one arrives at a
linear difference equation of order 2:
or equivalently
wt+1 wt + wt1 = Zt (2.51)
(1h) 2h
with = 2 (1+h) and Zt = (1+h) (vt + vt+1 ). The characteristic equation
for this difference equation is
1 z + z 2 = 0.
The symmetric nature of the polynomial coefficients implies that the roots
appear in pairs such that one root is the inverse of the other.15 This means
that one root, say 1 , is smaller than one whereas the other one is greater
than one, i.e. 2 = 1/1 . To see this note first that the discriminant is equal
to 4 = h > 0. Thus, the roots are real and second that 1 2 = 1. If we
denote 1 by then 2 = 1/ and we have = + 1 .
Applying the superposition principle, the solution becomes
(p)
wt = c1 t + c2 t + wt (2.52)
(p)
where the coefficients c1 and c2 and a particular solution wt have yet to be
determined. In order to eliminate explosive solutions, we set c2 = 0. The
15
This conclusion extends to contracts longer than two periods (see Ashenfelter and
Card, 1982).
54 CHAPTER 2. LINEAR DIFFERENCE EQUATIONS
and insert this solution into the difference equation (2.52) and perform a
comparison of coefficients as in the previous exercise. This leads again to
two homogeneous difference equations for the coefficients (j ) and (j ),
j 1 with solutions
j = d1 j + d2 j
j = e1 j + e2 j
3.1 Introduction
This chapter treats systems of linear difference equations. For each variable
X1t , , Xnt , n 1, we are given a linear nonhomogeneous difference equa-
tion of order p where each variable can, in principle, depend on all other
variables with a lag. Writing each difference equation separately, the system
is given by
(1) (1) (1)
X1t = 11 X1,t1 + 12 X2,t1 + + 1n Xn,t1
(p) (p) (p)
+ + 11 X1,tp + 12 X2,tp + + 1n Xn,tp + Z1t
(1) (1) (1)
X2t = 21 X1,t1 + 22 X2,t1 + + 2n Xn,t1
(p) (p) (p)
+ + 21 X1,tp + 22 X2,tp + + 2n Xn,tp + Z2t
(1) (1)
Xnt = n1 X1,t1 + n2 X2,t1 + + (1)
nn Xn,t1
(p) (p)
+ + n1 X1,tp + n2 X2,tp + + (p)
nn Xn,tp + Znt
Using matrix notation this equation system can be written more com-
pactly as
55
56 CHAPTER 3. SYSTEMS OF DIFFERENCE EQUATIONS
(k)
1, 2, . . . , p, denote the matrices k = i,j for k = 1, 2, . . . , p. The
i,j=1,2,...,n
solution of this difference equation is based again on the same principles as
in the univariate case (see page 8). Before doing so we show how to reduce
this p-th order system to a first order system.
Any system of order p can be rewritten as a system of order 1. In order to
see this, define a new variable Yt as the stacked vectors Xt , Xt1 , , Xtp+1 .
This new variable then satisfies the following first order system:
Xt Xt1 Zt
Xt1 1 2 3 . . . p1 p
Xt2 0
Xt2 In 0 0 . . . 0 0
Xt3 0
Yt = .. = 0 In 0 . . . 0 0 .. + ..
. .. .. .. . . .. .. . .
. . . . . .
Xtp+2 Xtp+1 0
0 0 0 . . . In 0
Xtp+1 Xtp 0
= Yt1 + Zt (3.2)
0
where Zt is redefined to be Zt0 0 0 . . . 0 0 . In denotes the identity
matrix of dimension n. The matrix is an np np matrix called the com-
panion matrix of (3.1).1 Thus, multiplying out the equation system (3.2) one
can see that the first equation gives again the original equation (3.1) whereas
the remaining p 1 equations are just identities. The study of a p-th order
system can therefore always be reduced to a first order system.
The Casarotian matrix is closely related to the issue whether or not the
sequences are independent.
Lemma 3.1. If det C(t) of n sequences (X (i) ), 1 i n, is different from
zero for at least one t0 0, then (X (i) ), 1 i n, are linearly independent
for t 0.
Proof. Suppose that (X (i) ), 1 i n, are linearly dependent. Thus, there
exists a nonzero vector c such that C(t)c = 0 for all t 0. In particular,
C(t0 )c = 0. This stands, however, in contradiction with the assumption
det C(t0 ) 6= 0.
Note that the converse
is not true
as can be seen from the following
(1) 1 (2) t
example: Xt = and Xt = 2 . These two sequences are linearly
t t
independent, but det C(t) = 0 for all t 0. The converse of Lemma 3.1 is
true if the sequences are solutions to the homogenous equation (3.5).
Lemma 3.2. If (X (i) ), 1 i n, are n linearly independent solutions of
the homogenous system (3.5), then det C(t) 6= 0 for all t 0.
Proof. Suppose there exists a t0 such det C(t0 ) = 0. This implies that there
(i)
exists a nonzero vector c such that C(t0 )c = ni=1 ci Xt = 0. Because the
P
(i) (i)
Xt are solutions so is the linear combination Yt = ni=1 ci Xt . For this
P
solution Yt0 = 0 thus Yt = 0 for all t because the uniqueness of the solution.
As the solutions are, however, linearly independent c must be equal to 0
which stands in contradiction to c 6= 0.
We can combine the two Lemmas to obtain the following theorem.
Theorem 3.1. The solutions (X (i) ), 1 i n, of the homogenous system
(3.5) are linearly independent for t 0 if and only if there exists t0 0 such
that det C(t0 ) 6= 0.
(1) (n)
The above Theorem implies that the n solutions Ut , , Ut of the
homogenous system (3.5) which satisfy the initial conditions
0
(i) (i) 0 0 1 0 0
U0 = e = |{z} , 1 i n, (3.6)
i-th element
(3.5) satisfying the initial condition (3.6). As the solutions are uniquely
determined, we have thus shown that the space of all solutions to the ho-
mogenous system (3.5) is a linear space of dimension n. Thus, any solution
can be written as
Xn
(i)
Xt = Xi,0 Ut = U(t)X0 (3.7)
i=1
Definition 3.3. Any nn matrix U(t) which is nonsingular for all t 0 and
which satisfies the homogenous matrix system (3.8) is called a fundamental
matrix. If in addition the matrix satisfies U(0) = In then it is called a
principal fundamental matrix.
Note that if V(t) is any fundamental matrix then U(t) = V(t)V 1 (0)
is a principal fundamental matrix. Note also if a V(t) is a fundamental
matrix then V(t)C is also a fundamental matrix where C is any nonsingular
matrix. This implies that there are infinitely many fundamental matrices for
a given homogenous matrix system. There is, however, only one principal
fundamental matrix because the matrix difference equation (3.8) uniquely
determines all subsequent matrices once an initial matrix is given. In the
case of a principal fundamental matrix this initial matrix is the identity
matrix. In this monograph we will not pursue the concept of the fundamental
matrix further because it does not payoff in the context of constant coefficient
systems.3 Only note that U(t) = t where U(t) is a principal fundamental
matrix. Thus, any solution to the homogenous system (3.5) has the form:
Xt = U(t)c = t c (3.9)
Distinct Eigenvalues
If all the eigenvalues, 1 , , n of are distinct, then is diagonalizable,
i.e. similar to a diagonal matrix. Thus, there exists a nonsingular matrix
Q such that Q1 Q = where = diag(1 , , n ). The columns of Q
consist of the eigenvectors of . With this similarity transformation in mind
it is easy to compute t :
t = QQ1 QQ1 QQ1 = Qt Q1
| {z }
t times
t
1 0 0 t1 0 0
0 2 0 0 t 0
1 2 1
= Q .. Q = Q .. Q
.. . . .. .. .. ..
. . . . . . . .
0 0 n 0 0 tn
In the case of distinct eigenvalues, it is easy to proof the following theorem.
Theorem 3.2. If the spectrum of , () = {1 , , n }, consists of n
distinct eigenvalues with corresponding eigenvectors qi , i = 1, . . . , n, then
the set
(i)
Xt = qi ti , i = 1, . . . , n,
represents a fundamental set of solutions to the homogenous system (3.5).
Proof. As qi is an eigenvector of corresponding to i , we have
(i) (i)
Xt+1 = qi t+1
i = i qi ti = qi ti = Xt , t 0.
(i)
The third equality follows from Q = Q. Thus, the Xt , i = 1, , n
are solutions to the homogenous system (3.5). In addition, we have that the
determinant of the corresponding
Casarotian matrix evaluated at t = 0 is
det C(0) = det q1 , , qn = det Q 6= 0 because Q consists of n linearly
independent eigenvectors and is therefore nonsingular. Thus, according to
Theorem 3.1 these solutions are linearly independent.
4
Recommended books on linear algebra are among others Meyer (2000) and Strang
(2003).
62 CHAPTER 3. SYSTEMS OF DIFFERENCE EQUATIONS
Thus, the general solution to the homogenous system (3.5) can be written
as n
X
Xt = ci qi ti , t0 (3.10)
i=1
for some constants c1 , , cn .
Another way to understand the result of Theorem 3.2 is to observe that
the similarity transformation actually decomposes the interrelated system in
Xt into n unrelated univariate first order difference equations. This decom-
position is achieved by the variable transformation Yt = Q1 Xt :
Yt+1 = Q1 Xt+1 = Q1 QQ1 Xt = Yt
Y1,t+1 1 0 Y1,t
.. .. . . . .
. = . . .. ..
Yn,t+1 0 n Yn,t
Thus, through this transformation we have obtained n unrelated univariate
first order difference equations in Yi,t , i = 1, . . . , n:
Y1,t+1 = 1 Y1,t
Yn,t+1 = n Yn,t
These equations can be solved one-by-one by the methods discussed in chap-
ter 2. The general solutions to these univariate first order homogenous equa-
tions are therefore Yi,t = ci ti , i = 1, , n. Transforming the system in Yt
back to the original system by multiplying Yt from the left with Q yields
exactly the solution in equation (3.10).
Repeated Eigenvalues
The situation with repeated eigenvalues is more complicated. Before dealing
with the general case, note that even with repeated eigenvalues the matrix
can be diagonalizable. This is, for example, the case for normal matrices, i.e.
matrices for which 0 = 0 . Examples of normal matrices include sym-
metric matrices ( = 0 ), skew symmetric matrices ( = 0 ) and unitary
or orthogonal matrices (0 = 0 = I). More generally, is diagonalizable
if and only if, for each eigenvalue, the algebraic multiplicity (the multiplicity
as a root of the characteristic polynomial) equals the geometric multiplicity
(the maximum number of linearly independent eigenvectors). It this case the
eigenvalue is called semisimple. If the maximum number of linearly indepen-
dent eigenvectors corresponding to some eigenvalue is strictly less than its
algebraic multiplicity, the matrix is called defective.
3.2. FIRST ORDER SYSTEM OF DIFFERENCE EQUATIONS 63
where
0 1 0 00
0
0 1 0
0
N = ... .. .. ..
.. .
. . .
.
0 0 0 0 1
0 0 0 0 0
5
A detailed treatment of the Jordan canonical form can be found, for example, in
Meyer (2000). The current exposition uses the complex Jordan form. There is, however,
an equivalent presentation based on the real Jordan form (see Colonius and Kliemann,
2014).
64 CHAPTER 3. SYSTEMS OF DIFFERENCE EQUATIONS
Jj (i )t = (i I + N )t
t t1 t t2 2 t
t
= i I + N+ N + + tk+1 N k1
1 i 2 i k1 i
t t t1 t t2 t
tk+1
i 1 i
2 i
k1 tk+2
i
t t t1 t
0
i 1
i k2
i
= ... .. ... ... .. (3.12)
. .
t
t1
i
1
t
0 0 0 i
where N is the nilpotent matrix of size corresponding to the Jordan block.
Xt = Xt1 + Zt . (3.13)
Theorem 3.3. Every solution (Xt ) to the first order nonhomogeneous sys-
tem (3.13) can be represented as the sum of the general solution to the
(g)
homogeneous system (3.5), (Xt ), and a particular solution to the nonho-
(p)
mogeneous system (3.13), (Xt ):
(g) (p) (p)
Xt = Xt + Xt = t c + Xt . (3.14)
Xt = Xt1 + Zt
Xt = (Xt2 + Zt1 ) + Zt = 2 Xt2 + Zt1 + Zt
...
Xt = t X0 + t1 Z1 + t2 Z2 + + Zt1 + Zt
t1
X
t
= X0 + j Ztj
j=0
Lemma 3.3. The zero solution of the homogeneous system (3.5) is stable if
and only if there exists M > 0 such that
t
M for all t 0. (3.20)
Proof. Suppose that the inequality (3.20) is satisfied then kXt k M kX0 k.
Thus, for > 0, let = M . Then kX0 k < implies kXt k < so that the
zero point is stable. Conversely, suppose that the zero point is stable. Then
for all kX0 k <
= sup
t
= 1 sup
t X0
= M
t
kk1 kX0 k
where the first inequality is just the definition of the matrix norm correspond-
ing to the norm in Rn . The second equality is a consequence of kX0 k .
The inequality above follows from the assumption that the zero point is a
stable equilibrium.
The condition given in equation (3.20) is equivalent to the condition that all
solutions are bounded.
Theorem 3.4. For the homogeneous system (3.5) the following statements
are true:
(i) The zero solution is stable if and only if () 1 and the eigenvalues
on the unit circle are semisimple.
(ii) The zero solution is asymptotically stable if and only if () < 1. In
this case, the solution is even exponentially stable.
Proof. According to previous lemma 3.3 we have to prove that kt k M for
some M > 0. Using the Jordan canonical form = QJQ1 this amounts to
kt k = kQJ t Q1 k M . But this is equivalent to the existence of a M > 0
M
such that kJ t k M . M may then be taken as M = kQkkQ 1 k . Now the power
The elements in this matrix become unbounded if |i | > 1. They also become
unbounded if |i | = 1 and Jj (i ) is not a 1 1 matrix. If, however, for all
eigenvalues with |i | = 1 the largest Jordan blocks are 1 1, then the Jordan
segment corresponding to i with |i | = 1, J (i ), is just a diagonal matrix
with ones in the diagonal and is therefore obviously bounded. The Jordan
segments, J (i ), corresponding to eigenvalues |i | < 1, converge to zero, i.e.
limt J (i )t = 0 because tk ti 0 as t by lHopitals rule.
9
The spectrum of a matrix is given by the set of its distinct eigenvalues.
70 CHAPTER 3. SYSTEMS OF DIFFERENCE EQUATIONS
Proof. The proof follows Elaydi (2005, theorem 4.14). Let X(t, x0 ) be a
solution with x0 W s . The definition of the (generalized) eigenvector corre-
sponding to some eigenvalue implies that E = E where E denotes
the space generated by the (generalized) eigenvectors corresponding to .
Hence W s = W s and s s
PrX(t, x0 ) W for all t 0. Given x0 W , we can
represent x0 as x0 = j=1 j ej where r is the number of eigenvalues, distinct
or not, strictly smaller than one and ej denotes the (generalized) eigenvec-
tors corresponding to those eigenvalues. Let J = Q1 Q be the Jordan form
of . Next, partition the Jordan matrix into blocks of stable, respectively
unstable eigenvalues accordingly:
Js 0
J=
0 Ju
The last equality follows from the fact that the Q1 ej s are of the form
(1j , . . . , rj , 0, . . . , 0)0 which is a consequence of the ej s being (generalized)
eigenvectors. This implies that X(t, x0 ) 0 as t because Jst 0 as
t .
The proof of (ii) is analogous.
Linearization
As in the case of one dimensional difference equations (see Theorem 2.4), we
can analyze the stability properties of nonlinear systems via a linear approx-
imation around the steady state. Consider for this purpose the nonlinear
homogeneous system Xt+1 = F (Xt ) with F : Rn Rn and fixed point
3.3. STABILITY THEORY 71
Xt+1 X = A(Xt X )
Definition 3.8 (Saddle Point). The zero solution of the hyperbolic linear
difference equation Xt+1 = AXt is called a saddle point if there exist at least
two eigenvalues of A, u and s , such that |u | > 1 and |s | < 1.
10
This means that the function is r-times continuously differentiable and that it is a
homeomorphism (i.e. one-to-one and onto, continuous with continuous inverse).
72 CHAPTER 3. SYSTEMS OF DIFFERENCE EQUATIONS
Squaring the second inequality in the first line and simplifying gives: tr() <
1 + det . Squaring the first inequality in the second line gives 1 det <
tr(). Combining both results gives the first part of the stability condition
(3.21). The second part follows from the observation that det = 1 2 and
the assumption that |1,2 | < 1. If the roots are complex, they are conjugate
complex, so that the second part of the stability (3.21) results from det =
1 2 = |1 | |2 | < 1. The first part follows from tr2 () 4 det < 0 which
is equivalent to 0 < tr2 () < 4 det . This can be used to show that
4 (1 + det tr()) > 4 + tr2 () 4tr() = (2 tr())2 > 0.
which is the required inequality.
Conversely, if the stability condition (3.21) is satisfied and if the roots are
real, we have
p p
2 + tr2 () 4 det tr() + tr2 () 4 det
1 < < 1 =
2 p 2
2
tr() + tr () + 4 4tr()
<
q 2
tr() + (2 tr())2
= < 1.
2
3.4. TWO-DIMENSIONAL SYSTEMS 73
Similarly, for 2 . If the roots are complex, they are conjugate complex and we
2 2
have |1 |2 = |2 |2 = 1 2 = tr ()tr 4()+4 det = det < 1. This completes
the proof.
It is clear that (0, 0)0 is an equilibrium for this system. In order to understand
the dynamics of the system, we can draw two lines in a (X1 , X2 )-diagram.
The first line is given by all points such that the first variable does not change,
i.e. where X1,t+1 = X1,t . From equation (3.23), these points are represented
by a line with equation (11 1)X1,t + 12 X2,t = 0. Similarly, the points
where the second variable does not change is, from equation (3.24), the line
with equation 21 X1,t + (22 1)X2,t = 0. These two lines divide the R R-
plane into four regions I, II, III, and IV as in figure 3.1. In this example both
lines have positive slopes.
The dynamics of the system in each of the four regions can be figured
out from the signs of the coefficients as follows. Suppose we start at a point
on the X1,t+1 X1,t = (11 1)X1,t + 12 X2,t = 0 schedule then we know
that the first variable does not change. Now increase X2,t a little bit. This
moves us into region I or IV. If 12 is positive, this implies that X1,t+1
X1,t > 0 so that the first variable increases. Thus, we know that above the
X1,t+1 X1,t = 0 line, X1,t increases and that below this line X1,t decreases.
We can indicate this result in figure 3.1 by arrows from left to right in regions
I and IV and arrows from right to left in regions II and III. If 12 is negative,
we obtain, of course, the opposite result. Similarly, consider the schedule
X2,t+1 X2,t = 21 X1,t + (22 1)X2,t = 0 where the second variable does
not change. Consider now an increase in X1,t . This moves us into region III
or IV. If the coefficient 21 is positive, this implies that X2,t+1 X2,t > 0 so
11
We may also interpret the systems as written in deviations from steady state.
74 CHAPTER 3. SYSTEMS OF DIFFERENCE EQUATIONS
X2 t
schedule: X1 t+1
, - X1 t = 0
region I region II
schedule:
steady X2 t+1
, - X2 t = 0
state
X1 t
region III
region IV
that the second variable must increase. As before we can infer that below the
X2,t+1 X2,t = 0 line X2,t increases whereas above this line X2,t decreases.
We can again indicate this behavior by arrows: upward arrows in regions
III and IV and downward arrows in regions I and II. If the sign of 21 is
negative, the opposite result is obtained. This type of analysis gives us a
phase diagram as in figure 3.1. This diagram shows us the dynamics of the
system starting in every possible point. In this example, the arrows indicate
that wherever we are, we will move closer to the steady state. But this
corresponds exactly to the definition of an asymptotically stable equilibrium
point (see definition 3.18). Thus, we conclude from this diagram that the
steady state is an asymptotically stable equilibrium point.
Of course the situation depicted in figure 3.1 is not the only possible
one. In order to analyze all possible situations which can arise in a two-
dimensional system, we make a variable transformation using the Jordan
canonical form of . If has the Jordan canonical form = QJQ1 then
we make the variable transformation Yt = Q1 Xt . This results in a new first
order homogenous difference equation system:
Note that the steady state is not affected by this variable transformation. It
is still the point (0, 0)0 . Let us treat these three cases separately.
0.8
0.6
0.4
0.2
Y2,t
0.2
steady
state
0.4
0.6
0.8
1
1 0.8 0.6 0.4 0.2 0 0.2 0.4 0.6 0.8 1
Y1,t
1000
500
Y2,t
steady
state
500
1000
6 4 2 0 2 4 6
Y
1,t
Y1,t too, which in this reduced setting is just Y1,0 = y10 = 0. Thus,
the solution is given by Y1,t = 0 and Y2,t = t2 y2,0 for some initial value
y2,0 . Note that the saddle path, in contrast to the other paths, is a
straight line through the origin. This property is carried over when
the system is transformed back to its original variables. In fact, the
solution becomes X1,t = q12 t2 y2,0 and X2,t = q22 t2 y2,0 where (q12 , q22 )0
is the eigenvector corresponding to 2 . Thus, the ratio of X1,t and X2,t
equals q12 /q22 constant.12 As saddle point equilibria are very promi-
nent in economics, we investigate this case in depth in Section 3.5. In
particular, we will go beyond the two-dimensional systems and analyze
the role of initial values.
When there are multiple eigenvalues with two independent eigenvec-
tors, can again be reduced by a similarity transformation to a di-
agonal matrix. The trajectories are then straight lines leading to the
origin if the eigenvalue is smaller than one as in figure 3.5, and straight
lines leading away from the origin if the eigenvalue is larger than one
in absolute terms.
When one eigenvalue equals one whereas the second eigenvalue is smaller
than one in absolute value, we arrive at a degenerate situation. Whereas
12
If q22 = 0, we take the ratio X2,t /X1,t which again defines a straight line through the
origin.
78 CHAPTER 3. SYSTEMS OF DIFFERENCE EQUATIONS
0.8
0.6
0.4
0.2
2,t
0
Y
0.2
steady
0.4 state
0.6
0.8
1
6 4 2 0 2 4 6
Y
1,t
0.8
0.6
0.4
0.2
2,t
0
Y
0.2
steady
0.4 state
0.6
0.8
1
1 0.8 0.6 0.4 0.2 0 0.2 0.4 0.6 0.8 1
Y1,t
1.5
0.5
Y2,t
steady
state
0.5
1.5
1.5 1 0.5 0 0.5 1 1.5
Y1,t
Y1,t remains at its starting value y10 , Y2,t converges to zero so that the
system converges to (y10 , 0)0 as is exemplified by figure 3.6. Only when
y10 = 0 will there be a convergence to the steady state (0, 0). However,
(0, 0) is not the only steady state because the definition of an eigen-
value implies that A I is singular. Thus, there exists X 6= 0 such
that X = X .
In case that the first eigenvalue equals minus one, there is no conver-
gence as Y1,t will oscillate between y10 and y10 .
0.5
steady
state
Y2,t
schedule:
Y1,t+1Y1,t = 0
0.5
namics in the case of eigenvalues inside the unit circle. One can clearly
discern the oscillatory behavior and the convergence to the steady state.
Figure 3.9 displays a situation with an unstable steady state where all
trajectories move away from the steady state. Finally figure 3.10 dis-
plays a degenerate case where the eigenvalues are on the unit circle. In
such a situation the system moves around its steady state in a circle.
The starting values are (0.25, 0.25), (0.5, 0.5), (1, 1), and (1.5, 1.5).
schedule:
Y1,t+1Y1,t = 0
1
0.8
0.6
0.4
0.2
0
schedule:
Y2,t+1Y2,t = 0
0.2
0.4
0.6
0.8
Figure 3.8: Complex eigenvalues with Stable Steady State (1,2 = 0.7 0.2)
10
schedule: schedule:
Y2,t+1Y2,t = 0 Y1,t+1Y1,t = 0
10
10 5 0 5 10
Figure 3.9: Complex eigenvalues with Unstable Steady State (1,2 = 10.5)
82 CHAPTER 3. SYSTEMS OF DIFFERENCE EQUATIONS
2.5
1.5
0.5
0.5
1.5
2.5
2 1.5 1 0.5 0 0.5 1 1.5 2 2.5
Figure 3.10: Complex eigenvalues on the unit circle (1,2 = cos (/4)
sin (/4))
bounded. In the well-behaved scenario, this will give just enough additional
initial values to determine a unique solution and the model is said to be
determinate (see also Section 1.1). Geometrically, this solution has the form
of a saddle path and the steady state is a saddle point.
A concise treatment of such models in the context of stochastic difference
equations was first given by Blanchard and Kahn (1980). Klein (2000) and
Sims (2001) provide further insights and solution approaches (see Chapter 5
for further details). As the stochastic setting delivers similar conclusions
with respect to uniqueness, we adopt the Blanchard-Kahn setup to the de-
terministic case by replacing rational expectations by perfect foresight. This
framework delivers first order linear nonautonomous difference equations:
x0 = RX0 (3.27)
13
The case k = 0 can be treated in a similar manner (See the example in section 4.3).
If k = n there are just enough initial conditions and the solution procedure of Section 3.2
can be directly applied.
3.5. BOUNDARY VALUE PROBLEM 83
or in matrix form
X10
...
x10 1 ... 0 0 ... 0
X
x0 = ... = ... . . . ... ... . . . ... k0 = (Ik , 0knk )X0
Xk+1,0
xk0 0 ... 1 0 ... 0 ...
Xn0
greater than one. We focus on the case where the zero solution is a saddle
point, meaning that there exists at least two eigenvalues 1 and 2 such that
|1 | < 1 and |2 | > 1 (see Definition 3.8).14 With this notation, we propose
the following particular solution:
X X
(p)
Xt = Qj1 Q1 Ztj Qj 1
2 Q Zt+j (3.28)
j=0 j=1
The reader is invited to verify that this is indeed a solution to equation (3.26).
The solution proposed in equation (3.28) has the property that variables
corresponding to eigenvalues smaller than one are iterated backwards whereas
those corresponding to eigenvalues larger than one are iterated forwards. This
(p)
ensures that {Xt } remains bounded whenever {Zt } is. The general solution
therefore is of the form
X X
t 1 t 1
Xt = Q1 Q c + Q2 Q c + j 1
Q1 Q Ztj Qj 1
2 Q Zt+j . (3.29)
j=0 j=1
The final step consists in finding the constant c. There are two types of
restrictions: the first one are the initial values given by equation (3.27); the
second one are the requirement that we are only interested in non-exploding
solutions. Whereas the first type delivers k restrictions because rank(R) =
k, the second type delivers n2 restrictions. Thus, c must be determined
according to the following equation system:
(p)
initial values: R c = x0 RX0
no explosive solutions: Q(2) c = Q(21) c1 + Q(22) c2 = 0 (3.30)
(11)
Q Q(12) .
where Q1 = (21) (22) , Q
(2)
= Q(21) .. Q(22) and c = (c01 , c02 )0 and
Q Q
where the partitioning of Q1 and c conforms to the partitioning of the
eigenvalues. Note that Q(2) is a n2 n matrix. Depending on whether the
number of independent restrictions is greater, smaller or equal to n, several
situations arise.
Theorem 3.8 (Blanchard-Kahn). Let GL(n) be hyperbolic then the
nonhomogeneous difference equation (3.26) with initial values given by (3.27)
has a unique nonexplosive solution if and only if
R
rank = n. (3.31)
Q(2)
The solution is given by equation (3.29) with c uniquely determined from (3.30).
14
If n1 or n2 equal zeros then the terms corresponding to the matrices 1 and 2 are
omitted.
3.5. BOUNDARY VALUE PROBLEM 85
If k = rank(R) > n1 then there are too many restrictions and there is no
nonexplosive solution. We may, however, soften the boundedness condition
and accept explosive solutions in this situation.
If k = rank(R) < n1 there are not enough initial conditions so that it
is not possible to pin down c uniquely. We thus have an infinite amount
of solutions and we call such a situation indeterminate. The multiplicity of
equilibria indeterminacy offers the possibility of sunspot equilibria.15 Sunspot
equilibria have been introduced by Cass and Shell (1983), Azariadis (1981),
and Azariadis and Guesnerie (1986) (see also Azariadis (1993) and Farmer
(1993)).
15
Sunspot equilibria explore the idea that extraneous beliefs about the state of nature
influence economic activity. The disturbing feature of sunspot equilibria is that economic
activity may change across states although nothing fundamental has changed.
86 CHAPTER 3. SYSTEMS OF DIFFERENCE EQUATIONS
Chapter 4
Examples of Linear
Deterministic Systems of
Difference Equations
where the parameters , , , , and are all positive. The variables are
all expressed in logarithms. The IS-equation represents the dependence of
aggregate demand ytd on the relative price of foreign to home goods and on
the real interest rate. A devaluation of the exchange rate2 , an increase in the
1
Rogoff (Rogoff (2002)) provides an appraisal of this influential paper.
2
The exchange rate et is quoted as the price of a unit of foreign currency in terms of
the domestic currency. An increase in et therefore corresponds to a devaluation of the
home currency.
87
88 CHAPTER 4. EXAMPLES: LINEAR SYSTEMS
foreign price level, p , or a decrease in the domestic price level pt all leads to an
increase in aggregate demand. On the other hand, an increase in the domestic
nominal interest rate, rt , or a decrease in the expected inflation rate, pt+1 pt ,
lead to a reduction in aggregate demand. A crucial feature of the Dornbusch
model is that prices are sticky and adjust only slowly. In particular, the price
adjustment pt+1 pt is proportional to the deviation of aggregate demand
from potential output y. If aggregate demand is higher than potential output,
prices increase whereas, if aggregate demand is below potential output, prices
decrease. In particular, the price level is treated as predetermined variable
whose value is fixed in the current period. The LM-equation represents
the equilibrium on the money market. The demand for real balances, m
pt , depends positively on potential output3 and negatively on the domestic
nominal interest rate. The model is closed by assuming that the uncovered
interest parity holds where r denotes the foreign nominal interest rate. In
contrast to the price level, the exchange rate is not predetermined. It can
immediately adjust within the current period to any shock that may occur.
For simplicity, the exogenous variables y, m, r , and p are assumed to remain
constant.
This system can be reduced to a two-dimensional system in the exchange
rate and the price level:
1
et+1 et = (y + pt m) r (4.1)
h i
pt+1 pt = (et + p pt ) y (y m + pt ) (4.2)
1
The first equation was obtained by combining the LM-equation with the
UIP. The second equation was obtained by inserting the IS-equation into
the price adjustment equation, replacing the nominal interest rate using the
LM-equation and then solving for pt+1 pt .
The steady state of this system is obtained by setting et = ess and pt = pss
for all t and solving for this two variables:
pss = r + m y (4.3)
1
ess = pss p + (y + r ) (4.4)
In the steady state, UIP and the price adjustment imply rt = r and ytd = y.
The system can be further reduced by writing it in terms of deviations from
3
This represents a simplification because money demand should depend on aggregate
demand, ytd , and not on potential output, y. However, we adopt this simplified version for
expositional purposes.
4.1. EXCHANGE RATE OVERSHOOTING 89
steady state:
1
ss
1 ss
et+1 e
et e
= ( + )
pt+1 pss 1 pt pss
1 1
ss
e e
= t (4.5)
pt pss
P(1) > 0 finally implies that one eigenvalue, say 1 , is larger than 1
whereas the second eigenvalue, 2 , lies between 1 and 1.
Because the eigenvalues are distinct, we can diagonalize as = QQ1
where is a diagonal matrix with 1 and 2 on its diagonal. The column of
the matrix Q consist of the eigenvectors of . Multiplying the system (4.5)
by Q1 , we obtain the transformed system:
ss
ss
et+1 1 et+1 e 1 1 et e
= Q = Q QQ
pt+1 pt+1 pss pt pss
1 0 et
= (4.6)
0 2 pt
et = c1 t1
pt = c2 t2
and 12 > 0, q12 and q22 must be of opposite sign. The same conclusion is
reached using the second equation. This reasoning also shows that q22 6= 0.
Suppose that q22 = 0 then q12 must also be zero because (11 2 ) > 0.
This, however, contradicts the assumption that (q12 , q22 )0 is an eigenvector.
Note that although the eigenvector is not uniquely determined, its direction
and thus the slope of the saddle path is.
The Dornbusch model is most easily analyzed in terms of a phase diagram
representing the price level and the exchange rate as in figure 4.1. The graph
consists of two schedules: pt+1 pt = 0 and et+1 et = 0. Their intersection
determines the steady state denoted by S. These two schedules correspond
to the equations (4.4) and (4.3). The et+1 et = 0 schedule does not depend
on the exchange rate and is therefore horizontal intersecting the price axis
at pss . Above this schedule the exchange rate depreciates whereas below
this schedule the exchange appreciates, according to equation (4.5). This
is indicated by arrows pointing to the right, respectively to the left. The
pt+1 pt = 0 schedule is upward sloping. To its left, prices are decreasing
whereas to its right prices are increasing, according to equation (4.5). The
two schedules divide the e-p-quadrant into four regions: I, II, III, and IV. In
each region the movement of e and p is indicated by arrows. In the Dornbusch
model the price level is sticky and considered to be a predetermined variable.
Suppose that in period 0 its level is given by p0 . The exchange rate in this
period is not given, but endogenous and has to be determined by the model.
Suppose that the exchange rate in period 0 is at a level corresponding to
point A. This point is to the left of the pt+1 pt = 0 schedule and above
the et+1 et = 0 schedule and therefore in region I. This implies that the
price level has to fall and the exchange rate to increase. The path of e and
p will continue in this direction until they hit the et+1 et = 0 schedule.
At this time the system enters region IV and the direction is changed: both
the price level and the exchange rate decrease. They will so forever. We are
therefore on an unstable path. Consider now an exchange rate in period 0
corresponding to point B. Like A, this point is also in region I so that the
exchange rate increases and the price level decreases. However, in contrast to
the previous case, the path starting in B will hit the pt+1 pt = 0 schedule and
move into region II. In this region, both the price level and the exchange rate
increase forever. Again this cannot be a stable path. Thus, there must be an
exchange rate smaller than the one corresponding to point B, but higher than
the one corresponding to point A, which sets the system on a path leading
to the steady steady. This is exactly the exchange rate which corresponds
to the saddle path given by equation (4.9). In this way the exchange rate in
period 0 is pinned down uniquely by the requirement that the path of (et , pt )0
converges.
92 CHAPTER 4. EXAMPLES: LINEAR SYSTEMS
schedule:
p pt+1 - pt = 0
I
saddle path
A B
p0
II
S
ss schedule:
p et+1 - et = 0
IV
III
saddle path
e
ss e
p schedule:
pt+1 - pt = 0
saddle path
Snew
ss
pnew
schedule:
et+1 - et = 0
ss
pold
saddle path
ss ss
eold enew e0 e
The constant is called the subjective discount rate. The period utility func-
tion U : R+ R is continuously differentiable, increasing, strictly concave,
and, in order to avoid corner solutions, fulfills limc0 U 0 (c) = .5
The rest of the specification is exactly the same as for the Solow model
(see section 2.4.2): Output is produced according to a neoclassical aggregate
production satisfying the Inada conditions. Recognizing that investment in
period t, It equals It = Kt+1 (1 )Kt the national accounting identity
becomes:
Ct + It = Ct + Kt+1 (1 )Kt = F (Kt , Lt )
respectively,
The first order condition for the optimum is given by the Euler-equation,
sometimes also called the Keynes-Ramsey rule:
U 0 (ct ) = h0 (kt+1 )U 0 (ct+1 ) (4.12)
Thus, the Euler-equation equates the marginal rate of transformation, 1/h0 (kt+1 ),
to the marginal rate of substitution, U 0 (ct+1 )/U 0 (ct ).
The equation system consisting of the transition equation (4.11) and the
Euler-equation (4.12) constitutes a nonlinear difference equation system. The
analysis of this system proceeds in the usual manner. First, we compute
the steady state(s). Then we linearize the system around the steady state.
This gives a linear homogeneous difference equation in terms of deviations
from steady state. We find the solution of this difference equation using
the superposition principle. Finally, we select, if possible, one solution using
initial conditions and boundedness arguments.
saddle path
c Dc=0
c
* E
Dk=0
c0
0 k0 k
*
k
**
k
max
k
Figure 4.3: Phase diagram of the optimal growth model
4.2. OPTIMAL GROWTH MODEL 97
From this equation we can deduce that ct+1 ct when kt+1 < k and vice
versa. Thus, to left of the (c = 0)-schedule consumption rises whereas to
the right consumption falls. Similarly, the transition equation (4.11) implies
that kt+1 kt when ct is lower than the corresponding c implied by the
(k = 0)-schedule. Thus, the two schedules divide the nonnegative orthant
of the (k, c)-plane in four regions. The dynamics in these four regions is
indicated in figure 4.3 by orthogonal arrows. In the region to the left of
the (c = 0)-schedule and above the (k = 0)-schedule, i.e. the north-
west region, consumption would increase whereas the capital intensity would
decrease. This dynamics would continue until the c-axis is hit. When this
happens, the economy has no capital left and therefore produces nothing but
consumes a positive amount. Such a situation is clearly infeasible. Paths with
this property have therefore to be excluded. Starting at k0 , there is, however,
one path, the saddle path, where the forces which lead to explosive paths,
respectively infeasible paths, just offset each other and lead the economy to
the steady state. This is the red line in figure 4.3.
The algebraic analysis requires the linearization of the nonlinear equation
system (4.11) and (4.12). This leads to:
kt+1 k
1
kt k
1+ 0 1
=
U 0 (c )h00 (k ) U 00 (c ) ct+1 c 0 U 00 (c ) ct c
where we used the fact that h0 (k ) = 1. Given that > 0 and U 00 (c ) < 0,
the matrix on the left hand side is invertible. This then leads to the following
linear first order homogenous system:
kt+1 k 1 kt k
1 1
= 1 1
ct+1 c 1 + RA (c )h00 (k ) RA (c )h00 (k ) + (1 + ) ct c
kt k
= (4.16)
ct c
98 CHAPTER 4. EXAMPLES: LINEAR SYSTEMS
with
1 00
tr = 1 + 2 = 1 + [(1 + )]1 RA (c )h (k )/(1 + ) > 2
1
det = 1 2 = >1
(1 + )
4 = (tr)2 4 det = [1 ((1 + ))1 ]2
1 00 1 00
RA (c )h (k ) RA (c )h (k ) 2
+ 2 >0
1+ 1+ (1 + )
1 00
RA (c )h (k )
P(1) = 1 tr + det = <0
1+
where 4 denotes the discriminant of the quadratic equation and where we
used the fact that h00 < 0. The above inequalities have the following impli-
cations for the two eigenvalues:
P(0) = det > 1 finally implies that one eigenvalue, say 1 , is larger
than 1 whereas the second eigenvalue, 2 , lies between 0 and 1.
Next, we show that the saddle path is upward sloping, i.e. that qq12 22
> 0 as
shown by the red line in Figure 4.3. This can be verified by manipulating the
defining equations for the eigenvector corresponding to the second eigenvalue
2 . They are given by (11 2 )q12 +12 q22 = 0 and 21 q12 +(22 2 )q22 = 0.
Because (11 2 ) > 0 and 12 < 0, q12 and q22 are of the same sign. The same
conclusion is reached using the second equation. This argument also shows
that q12 6= 0. Suppose that q12 = 0 then q22 must also be zero because 12 =
1
1+
< 0. This, however, contradicts the assumption that (q12 , q22 )0 is an
eigenvector. Note that although the eigenvector is not uniquely determined,
its direction and thus the slope of saddle path is.
Taxation of Capital
Suppose that the government levies a proportional tax on the gross return
to capital. For simplicity, we assume that the revenues from the tax are just
wasted. Therefore only the Euler equation (4.12) is affected. The new Euler
equation then becomes
where is the tax rate with 0 < < 1. The transition equation for capital
is not altered. The new c = 0 schedule then is
c = 0 : (1 )h0 (k ) = (1 )(f 0 (k ) + 1 ) = 1.
100 CHAPTER 4. EXAMPLES: LINEAR SYSTEMS
new Dc=0
c schedule saddle path corresponding
to new steady state
c0
*
c old Eold Dk=0
*
c
Enew
0
*
k new k0 k
max
k
Figure 4.4: Phase diagram of the optimal growth model with distortionary
taxation of capital
This implies that an increase in the tax rate will lower the steady state
capital stock. The tax is thus distortionary.
k = 0 : c = f (k ) ( + )k g.
unstable path
* E0 old: Dk=0
c0
c1 g
c0
* new: Dk=0
c1
E1
0 k1 k
*
k
Figure 4.5: Phase diagram of the optimal growth model with government
expenditures
4.3. THE NEW KEYNESIAN MODEL 103
where yt , t , and it denote income, inflation and the nominal interest rate, all
measured as deviations from the steady state. ut is an exogenous cost-push
shock. Furthermore, we assume that > 0, > 0, and 0 < < 1. In
addition, we take an aggressive central bank, i.e. > 1.
This system can be solved for (yt+1 , t+1 )0 by inserting the Taylor-rule
and Phillips-curve into the IS-equation:
t+1 1 1 t ut /
Xt+1 = = +
yt+1 ( 1)/ + / yt ut /()
= Xt + Zt+1 (4.22)
with
1
tr = 1 + 2 = 1 + + >2
1
det = 1 2 = + >1
2
2 1 2
4 = (tr) 4 det = 1 + + 2 + 4
P(1) = (1 1 )(1 2 ) = ( 1) > 0 if > 1
First assume that is high such that 4 < 0. In this case we have two complex
conjugate roots. Because det > 1, they are located outside the unit circle.8
Alternatively assume that is small enough such that 4 > 0. In this case
both eigenvalues are real. Using the assumption > 1, P(1) > 0. Thus,
both roots are either greater or smaller than one. They cannot be smaller
than one because tr > 2. Thus, in both cases we reach the conclusion
that the eigenvalues are outside the unit circle. As both variables are non-
predetermined, the boundedness condition (3.30), Qc = 0 which is equivalent
to c = 0, then determines the unique solution:
j
X 1 0 1 ut1+j /
Xt = Q Q
j=1
0 j2 ut1+j /()
e = 1 + 2 = 1 + 1 + > 2
tr
1
det e = 1 2 = > 1
2
2 1 2
4 = (tr) 4 det = 1
e + +2+ >0
P(1) = (1 1 )(1 2 ) = < 0.
The discriminant is now unambiguously positive so that both eigenvalues are
real. Moreover, P(1) < 0 so that one eigenvalue is smaller than one and the
other bigger than one. Thus, the boundedness condition does not determine
a unique solution. Instead there is a continuum of solutions indexed by c1
and we are faced with the case of indeterminacy. The implications of this
indeterminacy for monetary policy and possible remedies are discussed in
Gal (2011).
8
Another way to reach this conclusion is by observing that the real part of the roots is
tr
2 > 1.
Chapter 5
1 Et Xt+1 = 0 Xt + Zt , t = 0, 1, 2 . . . , (5.1)
where {Xt } and {Zt } are real-valued n-dimensional stochastic processes de-
fined on some probability space (, F, P). The random variables Xt and
Zt are measurable with respect to the -algebra Ft = {(Xs , Zs ) : s t}.
This makes the sequence {Ft } a filtration adapted to {Xt } and {Zt }. In eco-
nomics Ft is also called the information set. Et then denotes the conditional
expectation with respect to Ft . As expectations are based on current and
past Xt s and Zt s only and not on extraneous variables, we have eliminated
the possibility of sunspot solutions.
Expectational difference equations of the type (5.1) arise typically in the
context of rational expectations models. Starting with the seminal paper by
Blanchard and Kahn (1980), an extensive literature developed which analyzes
the existence and nature of its solutions. The most influential papers, at least
for the present exposition, are Gourieroux et al. (1982), Klein (2000), Sims
(2001), among others. There is no loss of generality involved by confining
the analysis to first order equations as higher order equations can be reduced
to first order ones by inflating the dimension of the process (see Binder and
Peseran, 1994).
In the following, 1 is not necessarily invertible. Thus, we allow for the
possibility that some equations do not involve expectational terms. The
105
106 CHAPTER 5. STOCHASTIC DIFFERENCE EQUATION
stochastic theory developed below is therefore more general than its deter-
ministic counterpart.
To make the problem tangible, we consider only a certain class of so-
lutions. In particular, we require that the exogenous input process {Zt } is
bounded in Lp , p > 1.1 This means that supt kZt kp < . This assumption
implicitly restricts the solution processes to be bounded as well.
Remark 5.1. The class of stationary processes is the prime example of such
processes as the expected value and the variance remain constant. In many
applications, {Zt } is specified as an ARMA-process.2
Throughout the analysis we follow King and Watson (1998) and assume
that the linear matrix pencil 1 z + 0 is regular:3
As was already pointed out by Blanchard and Kahn (1980), the notion of
a predetermined variable or process is key for understanding the nature of the
solution.4 Following Klein (p.1412, 2000), we adopt the following definition.
|| < 1: In this case, the representation (5.4) implies that the solution fol-
lows an autoregressive process of order one. This representation admits
a causal representation with respect to the
P expectational errors. This
representation is given by Xt = t X0 + t1 j=0 j
tj and is bounded.
However, if {Xt } is not predetermined, there is no starting value X0
and we are faced with a situation of indeterminacy because any mar-
tingale difference sequence defined with respect to Ft would satisfy the
difference equation.
one obtains:
Xt = 1 Et Xt+1 1 Zt
= 1 Et 1 Et+1 Xt+2 1 Zt+1 1 Zt
= 2 Et Xt+2 1 Zt 2 Et Zt+1
= ...
k
X
k1 1
= Et Xt+k+1 j Et Zt+j
j=0
As we are looking for solutions which remain bounded, this suggests to take
X
1
Xt = j Et Zt+j (5.6)
j=0
This solution, however, makes only sense when {Xt } is predetermined so that
X0 is given.
1 mt
Et pt+1 = pt + , < 0, (5.7)
110 CHAPTER 5. STOCHASTIC DIFFERENCE EQUATION
where {mt } is now a stochastic process and where the expected price level is
replaced by the conditional expectation of the logged price level.
Given that < 0, the coefficient of pt , = ( 1)/, is positive and
strictly greater than one. Therefore the only bounded (stationary) solution
(g)
to the homogeneous equation is {Xt } = 0 and a particular solution can be
found by forward iteration. Thus the solution is:
j
1 X
pt = Et mt+j . (5.8)
1 j=0 1
This shows that the relation between the price level and money supply de-
pends on the conduct of monetary policy, i.e. it depends on the autoregressive
coefficient a. Thus whenever the monetary authority changes its rule, it af-
fects the relation between pt and mt . This cross-equation restriction is viewed
by Hansen and Sargent (1980) to be the hallmark of rational expectations.
It also illustrates that a simple regression of pt on mt can not be considered a
structural equation, i.e. cannot uncover the true structural coefficients ( in
our case), and is therefore subject to the so-called Lucas-critique (see Lucas
(1976)).
A similar conclusion is reached if money supply follows a moving average
process of order one instead of an autoregressive process of order one:
t = mt mt1 + 2 mt2 . . .
and
1 invertible
The invertibility of 1 implies that we can rewrite equation (5.1) as
Et Xt+1 = Xt + Zt , t = 0, 1, 2 . . .
where = 1 1
1 0 and Zt = 1 Zt . Let us further assume that is diagonal-
izable with = QQ1 , diagonal. As in the discussion of the deterministic
case in Section 3.5, we partition as
1 0
=
0 2
112 CHAPTER 5. STOCHASTIC DIFFERENCE EQUATION
such that the eigenvalues in 1 are strictly inside the unit circle whereas those
in 2 are strictly outside the unit circle. We disregard the case of eigenvalues
on the unit circle.
We make the following assumption with respect to the dimension of 1
and 2 .
Following the logic of the discussion in Section 5.2, the unique bounded
(2)
solution for {Yt } is
X
X
(2)
j
(2)
Yt = 1
2
1
2 Et Zt+j = 2 j
2 Et Q(21) (1)
Z t+j + Q(22) (2)
Z t+j
j=0 j=0
X
= 1 j
2 2 Q(21) Q(22) Et Zt+j . (5.9)
j=0
5.3. THE MULTIVARIATE CASE 113
Turn next to the first part of the decoupled equation. Note that the
(1)
predetermined variable Xt satisfies the identity:
(1) (1) (1) (1) (2) (2)
Xt+1 Et Xt+1 = Q11 Yt+1 Et Yt+1 + Q12 Yt+1 Et Yt+1 = t+1 .
(1)
Inserting this equation into the equation for Yt+1 , we get
(1) (1) (1) (1)
Yt+1 = Et Yt+1 + Yt+1 Et Yt+1
(1) (2) (2)
= 1 Yt + Zt (1) + Q1
11 t+1 Q 1
11 Q 12 Yt+1 Et Y t+1
(1)
= 1 Yt + Zt (1) + Q1 1
11 t+1 Q11 Q12 t+1 (5.10)
(2) (2)
where t+1 = Yt+1 Et Yt+1 is an exogenous martingale difference process.
Thus, equation (5.10) is a first order autoregressive scheme with starting
value given by
(1) (k) (2)
Y0 = Q1 11 X 0 Q 12 Y 0 .
Equations (5.10) and (5.9) determine the solution for Yt . This step in the
derivation is only valid if Q11 is invertible. Otherwise, we could not determine
(1) (1)
the initial values of Yt from those of Xt and there would be a lack of initial
values for Yt .6 Hence, Assumption 5.3 is not sufficient for the uniqueness of
the solution. In addition, we need the following assumption.
Assumption 5.4. Q11 is nonsingular.
Finally, the solution for Yt can be turned back into a solution for Xt by
multiplying Yt by Q.
We can get further insights into the nature of the solution by assuming
that {Zt } is a causal autoregressive process of order one:
Zt+1 = AZt + ut+1 , ut WN(0, 2 ) and kAk < 1
where {ut } is exogenous. This specification implies that Et Zt+j = Aj Zt ,
j = 1, 2, . . . Inserting this into equation (5.9) we find that
X
(2)
1 j Q(21) Q(22) Aj Zt = M Zt
Yt = 2 2
j=0
1
P j (21) (22)
j (1)
where M = 2 j=0 2 Q Q A . The solution to Yt then can
be written as
(1) (1)
Yt+1 = 1 Yt + Zt (1) + Q1 1
11 t+1 Q11 Q12 M ut+1 .
6
See Klein (2000, section 5.3.1) and King and Watson (2002) for details and examples.
114 CHAPTER 5. STOCHASTIC DIFFERENCE EQUATION
Remark 5.3. The above derivation remains valid even if the matrix is not
diagonalizable. In this case, we will have to work with the Jordan canonical
form instead (see Section 3.2.2).
Remark 5.4. The derivation excluded the possibility of roots on the unit
circle.
1 singular
In many practical applications, 1 is not invertible so that the procedure just
outlined is not immediately applicable. This, for example, is the case when
a particular equation contains no expectations at all which translates into a
corresponding row of zeros in 1 . One way to deal with this problem is to
take a generalized inverse of 1 and proceed as explained above.
The most appropriate type of generalized inverse in the context of dif-
ference equations is the Drazin-inverse (see Campbell and Meyer, 1979, for
a comprehensive exposition). This generalized inverse can be obtained for
any n n matrix A in the following manner. Denote by IndA the smallest
nonnegative integer k such that rankAk = rankAk+1 . This number is called
the index of A. Then the following Theorem holds (see Theorem 7.2.1 in
Campbell and Meyer, 1979).
With this Theorem in mind, we can now decouple the system in two
parts. The first part will be similar to the case when 1 is invertible. The
second one will correspond to the singular part and will in some sense solve
5.3. THE MULTIVARIATE CASE 115
out the expectations. Multiply for this purpose equation (5.1) from the left
by (z1 + 0 )1 where assumption 5.2 guarantees that the inverse exists for
some number z. Thus, we get
Applying the law of iterated expectations and multiplying the equation from
the left by N k1 gives
e (2) = (In2 zN )2 N k2 Et X
0 = N k Et X e (2)
t+k t+k2
3 k3
= (In2 zN ) N Et X e (2)
t+k3
= ...
et(2)
= (In2 zN )k Et X
et(2) .
= (In2 zN )k X
Complex Numbers
These two operations will turn C into a field where (0, 0) and (1, 0) play
the role of 0 and 1.1 The real numbers R are embedded into C because we
identify any a R with (a, 0) C.
The number = (0, 1) is of special interest. It solves the equation x2 +1 =
0, i.e. 2 = 1. The other solution being = (0, 1). Thus any complex
number (a, b) may be written as (a, b) = a + b where a, b are arbitrary real
numbers.2
1
Substraction and division can be defined accordingly:
2
A more detailed introduction of complex numbers can be found in Rudin (1976) or
any other mathematics textbook.
117
118 APPENDIX A. COMPLEX NUMBERS
z=a+ib
b
i
1
r
imaginary part
1 a
0
1
1
i
b
unit circle: z=aib
a2 + b2 = 1
2
2 1 0 1 2
real part
z = a + b Cartesian coordinates
= re = r(cos + sin ) polar coordinates.
e + 1 = (cos + sin ) + 1 = 1 + 1 = 0.
119
e + e a
cos = = ,
2 r
e e b
sin = = .
2 r
Further implications are de Moivres formula and the Pythagoras theorem
(see Figure A.1):
n
de Moivres formula re = rn en = rn (cos n + sin n)
Pythagoras theorem 1 = e e = (cos + sin )(cos sin )
= cos2 + sin2
(z) = 1 1 1 1 1
1 z 2 z . . . 1 p z .
3
The notation with j z j instead of j z j was chosen to conform to the notation
of AR-models.
120 APPENDIX A. COMPLEX NUMBERS
Appendix B
Matrix Norm
Thus the induced matrix norm is the maximum amount a vector on the unit
sphere can be stretched. The matrix norm induced by the Euclidian vector
norm is given by: p
kAk = max kAxk = (A0 A)
kxk=1
where (A0 A) is the spectral radius of A0 A, i.e. (A0 A) = max{|| : is an eigenvalue of A0 A}.
Another convenient matrix norm is the Frobenius norm, sometimes also
called the Hilbert-Schmidt or the Schur norm. It is defined as follows:
n
X n
X
kAk2 = |aij |2 = tr(A0 A) = i
i,j i
where i are the eigenvalues of A0 A. Thus the Frobenius norm stakes the
columns of A into a long n2 -dimensional vector and takes its Euclidian norm.
The matrix norm has the following properties:
121
122 APPENDIX B. MATRIX NORM
Bibliography
123
124 BIBLIOGRAPHY