View original document

The full text on this page is automatically extracted from the file linked above and may contain errors and inconsistencies.

www.clevelandfed.org/research/workpaper/index.cfm

Working Paper 9010

CONSUMPTION AND FRACTIONAL DIFFERENCING:
OLD AND NEW ANOMALIES
by Joseph G. Haubrich

Joseph G. Haubrich is an economic
advisor at the Federal Reserve Bank of
Cleveland. The author would like to
thank Andrew Abel, Angus Deaton, Roger
Kormendi, Andrew Lo, and seminar
participants at the University of
Pennsylvania, the Federal National
Mortgage Association, and the winter
Econometric Society meetings for
stimulating discussions.
Working papers of the Federal Reserve
Bank of Cleveland are preliminary
materials circulated to stimulate
discussion and critical comment. The
views stated herein are those of the
author and not necessarily those of
the Federal Reserve Bank of Cleveland
or of the Board of Governors of the
Federal Reserve System.
September 1990

www.clevelandfed.org/research/workpaper/index.cfm

1. Introduction
Consumption depends on income, so testing theories of consumption involves
testing theories of income. A prominent recent example is the work by
Campbell and Deaton (1989), which uncovers a paradox. They model income as
having a unit root instead of as a fluctuation around a trend, and so they
find that consumption looks too smooth: the permanent-income hypothesis does
not hold. Like some previous researchers, they find that a
difference-stationary process fits the data better than a trend-stationary
process.
The choice between a difference-stationary process and a trend-stationary
process, however, ignores the intermediate class of fractionally differenced
processes.

Since fractional processes exhibit long-term dependence, they are

often classified as having a unit root rather than as trend stationary. This
makes permanent income seem rougher than it really is, while consumption,
which responds to the true, fractional income, looks too smooth. Specifying
consumption correctly removes the paradox.
This paper reviews the techniques of fractionally differenced stochastic
processes, calculates the stochastic properties of consumption when income
follows a fractional stochastic process, and shows how this may explain the
excess-smoothness results.

2. Fractional Methods
Intuition suggests that differencing a time series roughens it, while summing
a time series smooths it. A fractional difference between 0 and 1 can be

www.clevelandfed.org/research/workpaper/index.cfm

described as a filter that roughens a series less than does a first
difference:

The series is rougher than a random walk but smoother than white

noise. This is apparent from the infinite-order moving-average
representation. Let X, follow

(1

-

LldX, =

where

E,

E,,

is white noise, d is the degree of differencing, and L is the lag

operator. If d

=

0, X, is white noise, and if d

- 1, X,

is a random walk.

However, as Granger and Joyeux (1980) and Hosking (1981) show, d need not be
an integer. The binomial theorem provides the relation

defined as

with the binomial coefficient (;)

(,Id

=

d(d

-

l)(d

-

2)...(d
k!

-

k

+

1)

for real d and nonnegative integer k. Using this definition, the
autoregressive (AR) form of X, follows

with the AR coefficient expressed compactly in terms of the gamma function

www.clevelandfed.org/research/workpaper/index.cfm

Manipulating equation (5) yields the corresponding moving average (MA)
representation of X,:

The time-series properties of X, depend crucially on the difference
parameter, d. For example, when d is less than one-half,X, is stationary;
when d is greater than minus one-half,X, is invertible (Granger and Joyeux
[1980], Hosking [1981]).

Likewise, the autocorrelation properties of X,

depend on the parameter d. The MA coefficients, E$,

indicate the effect of

a shock K periods ahead and the extent to which current levels depend on past
values. Using Stirling's approximation, we know that

Comparing this with the decay of an AR(1) process highlights the central
"long-memory" feature of fractional processes: They decay hyperbolically, at
rate kl-d,
rather than at the exponential rate, pk, of an AR(1)

.

For

example, compare in Figure 1 the autocorrelation function of the fractionally
differenced series (~-L)~.~"X,
=

c,

with the AR(l)X,

=

0.9X,-, +

both have first-order autocorrelations of 0.90, the AR(1)'s

c,.

Although

autocorrelation

function decays much more rapidly. Figure 2A plots the impulse-response
functions of these two processes. At lag 1, the MA coefficients of the
fractionally differenced series and the AR(1) are 0.475 and 0.900,

www.clevelandfed.org/research/workpaper/index.cfm

respectively; at lag 10, they are 0.158 and 0.349, and at lag 100, they are
0.048 and 0.000027. The persistence of the fractionally differenced series is
apparent at the longer lags. Alternatively, we may ask what value of an
AR(1)'s

autoregressive parameter will, for a given lag, yield the same impulse

response as the fractionally differenced series (equation [I]).
simply the k-th root of
when d

- 0.475.

This value is

%, and is plotted in Figure 2B for various lags

For large values of k, this autoregressive parameter must be

very close to unity.
These representations also show how standard econometric methods can fail
to detect fractional processes. Although a high-order ARMA process can mimic
the hyperbolic decay of a fractionally differenced series in finite samples,
the large number of parameters required would give the estimation a poor
rating from the usual Akaike or Schwartz criteria. A n explicitly fractional
process, however, captures that pattern with a single parameter, d. Granger
and Joyeux (1980) and Geweke and Porter-Hudak (1983) provide empirical support
by showing that fractional models often out-predict fitted ARMA models.
The lag polynomials A(L) and B(L) provide a metric for the persistence of

.

Suppose % represents GNP, which falls unexpectedly this year. How

much should this decline change a forecast of future GNP? To address this
issue, define

% as the coefficients of the lag polynomial, C(L), that

satisfies the relation (1
given by equation (1).

-

L)%

=

C(L)e,,

where the process

% is

One measure used by Campbell and Mankiw (1987) is

www.clevelandfed.org/research/workpaper/index.cfm

For large values of k, the value of B, measures the response of Xt+k
to an innovation at time t, a natural metric for persistence. From
equation (7), it is immediate that for 0 < d < 1, C(l)

=

0, and that,

asymptotically, there is no persistence in a fractionally differenced series,
even though the autocorrelations die out very slowly. This holds true not
only for d

- 1/2 (the stationary case), but also for 1/2 < d < 1, when the

process is nonstationary.
From these calculations, it is apparent that the long-run dependence of
fractional processes relates to the slow decay of the autocorrelations, not to
any permanent effect. This distinction is important; for example, an IMA(1,l)
can have small but positive persistence, but the coefficients will never mimic
the slow decay of a fractional process.

3.

Fractional Differencing and the Theory of Consumption

The excess-smoothness paradox can be stated more precisely as follows.
Assuming the standard certainty equivalence framework (for example, quadratic
utility; see Hall [1978], Flavin [1981], and Zeldes [1989]), we can find how
the variance of consumption depends on the income process:

where

- consumption,
r - the real interest rate,

C,

www.clevelandfed.org/research/workpaper/index.cfm

8,

=

the MA coefficients of income Yt,

4t

=

the AR coefficients of Y,,

A
u:

- the difference operator A - (1 - L), and

-

the variance of income shocks.

Hansen and Sargent (1981) show that this formula holds for both stationary
and nonstationary processes. Since consumption is a random walk (more
generally a martingale) in this framework, the variance of the change in
consumption (equation [9]) also represents the variance of innovations to
consumption. Under the traditional assumption that income follows a
trend-stationary process (because the shocks die out), the variance of
innovations to consumption, var(ACt), should be less than the variance of
innovations to income, i. This is what Friedman was trying to explain with
the permanent-incomehypothesis

--

namely, that consumption looks smoother

than income. If, however, income is first-difference stationary, as
researchers since Nelson and Plosser (1982) have claimed, the revision in
permanent income exceeds the revision in actual income. Consumption
imovation should then exceed income innovation, a:.

Deaton (1987)

finds that it does not.

A numerical example based on the data used in this paper illustrates
excess smoothness. Suppose income is a random walk.

In that case, the

variance of the change in consumption should equal the variance of the change
in income, as intuition or equation (9) suggests. In fact, the figure for
consumption is 11.65, while that for income is 61.14.

www.clevelandfed.org/research/workpaper/index.cfm

The key point to note, both in predicting the variance of consumption and
in determining the variance of income innovations, is that we must make some
assumptions or estimates of the income process. By making a different and
better assumption about income

--

fractional differencing - - the paradox can

be resolved.
Another advantage of assuming a fractional-differencing process for income
is that it allows us to retain two assumptions jettisoned by others. First,
the income process is univariate, and consumers have no information about it
that is hidden from the econometrician. West (1988) shows that such hidden
information can spuriously create excess smoothness, because true income
surprises would then be less than measured income surprises. Various methods
that correct for hidden information (Campbell and Deaton [1989], Flavin
[1988]) still show excessive smoothness, however. Second, the
permanent-income hypothesis is maintained throughout. Both Campbell and
Deaton and Flavin show that departures from this can simultaneously produce
both excess smoothness and excess sensitivity.
The remainder of this section attempts to answer two basic questions.
First, does there exist a difference parameter, d, that resolves the
paradox

--

that is, if income follows such a process, consumption will no

longer look too smooth? Second, does actual income follow such a process?
other words, will the fractional parameter that provides a solution fit the
income data that we have?
Using data for the United States, I proceed in four basic steps.l
Section 3.1 reports estimates of the variance of income and consumption

In

www.clevelandfed.org/research/workpaper/index.cfm

changes using both Generalized Method of Moments (GMM) and classical
chi-squared techniques to determine the estimates' precision.

In section

3.2, using the permanent-income hypothesis, I find a range of d in the income
process that will produce the variance of consumption found in the first step.
In section 3.3, I employ a test for fractional differencing in the income
series. Finally, in section 3.4, I use simulations to estimate the
probability that fractional parameters reported in section 3.2 would produce
the value found in section 3.3.

3.1 Distribution of the Sample Variance
I begin by estimating and comparing the variance of income changes and the
variance of consumption changes. Calculating the distribution of the sample
variance depends on assumptions about the underlying process. The classical
approach assumes an i.i.d. sample from a normal distribution and then produces
the familiar result that the scaled sample variance is distributed chi-squared
with degrees of freedom one less than the sample size:

This may be appropriate for consumption, which, according to theory, should
follow a random walk.

It has the advantage of being correct for finite

samples.
The GMM approach allows for heteroskedasticity and autocorrelation.
Designed to handle much more complicated estimation problems (Hansen [1982],
Hansen and Singleton [1982]), it reduces to a fairly simple form when used to

www.clevelandfed.org/research/workpaper/index.cfm

determine the distribution of the sample variance.

(See Ng Lo [I9881 for a

rigorous and clear demonstration of this.) In fact, it reduces to estimating
the covariance matrix. Therefore, I use the Newey-West (1987) covariance
matrix. This provides a positive, definite heteroskedastic and
autocorrelation-consistent covariance matrix. The disadvantage is that it
provides an asymptotic result.
The Newey-West matrix also requires that a choice be made on the number of
lags used to compute the matrix.

The authors suggest using the fourth root of

the sample size, but the convergence results for this small number depend on
mixing conditions, which will generally be violated in the case of long-term
dependence. In more general cases, they suggest employing the cube or square
root, while Chatfield (1984, p. 141) recommends using twice the square root.
With a sample size of 120 for the consumption series and 137 for the two
income series, I use five lags. This follows Ng Lo (1988), who finds that
this choice works well even in larger samples for a variety of series.
Table 1 shows the sample variances for per-capita consumption of
nondurables and services, plus both per-capita income measures used (labor and
disposable).

It also reports the 95 percent confidence bounds obtained using

both the classical and GMM approaches. Since the GMM bounds are broader
(because income shows autocorrelation), they are used in the next part of this
exercise.

3.2

Implied Variance

www.clevelandfed.org/research/workpaper/index.cfm

The variance of income and consumption depends on an unobservable (to the
econometrician) variable:

shocks to income. If income follows a fractional

process with parameter d, we have from Hosking (1981) that

Likewise, the variance-of-consumptionformula (equation [9]) specializes in
this case to

where C, is consumption, Bt are the MA coefficients of income Y,, and

A is the difference operator, A

=

(1 - L).

The estimates for income and

consumption variance give estimates of the shock variance, a:.
Notice that the implied shock variance changes with different assumptions
about the income process, that is, with changes in the differencing parameter,
d.

Inverting equations (11) and (12) yields the variance of income shocks as

a function of d. Then, comparing the implied shock variances across income
and consumption yields the d values that make the income process consistent
with observed consumption behavior.
Implementing the above procedure requires choosing an interest rate. I
use three different quarterly rates: r

=

0.2 percent, which corresponds to

the long-run average rate used in Mehra and Prescott (1985); r

=

1 percent, a

www.clevelandfed.org/research/workpaper/index.cfm

high interest rate; and r

=

0.05 percent, a low interest rate. Using these

numbers made a noticeable, if not dramatic, difference in the variance
estimates.
Tables 2A and 2B report the results of this investigation and make clear
the choice of bounds on d used: 0.79 and 0.95 for labor income, and 0.72 and
0.96 for disposable income.

3.3

Testing for Fractional Differencing

The next step ascertains whether the d values obtained above are consistent
with the observed income process. This section tests for fractional
differencing using the modified rescaled range (R/S) statistic developed by Lo
([forthcoming] and Haubrich and Lo [1989]).

In section 3.4, I use simulations

to determine the probability that the values obtained from the test could come
from distributions with a d parameter in the range calculated above.
The modified R/S statistic tests whether a process X, shows long-term
dependence, (It is based on a statistic originally developed by Hurst [I9511
and popularized by Mandelbrot [1972].) More formally, consider a process
defined as

X,

=

p+ct,

where p is an arbitrary but fixed constant. For the null hypothesis H,
assume that the disturbances
(c1)

E(E~) = 0 for all t,

(E~)

satisfy the conditions

www.clevelandfed.org/research/workpaper/index.cfm

sup E [JE~~'] <

(C2)

w

for some 8 > 2,

t

[A[: f

$ = l in+a
mE

(c3)

2

Ej]

j=l

(c4)

(E,)

]

exists and u2 > 0, and

is strong-mixing,with mixing coefficients

,

that satisfy

Conditions (C2) through (C4) allow dependence and heteroskedasticity, but
prevent them from being too large. Thus, short-term dependent processes, such
as finite-order ARMA models, are included in the null hypothesis, as are
models with conditional heteroskedasticity. Unlike the statistic used by
Mandelbrot, the modified R/S statistic used here is robust to short-term
dependence. A more in-depth discussion of these conditions appears in
Phillips (1987), Haubrich and Lo (1989), and Lo (forthcoming).
To construct the modified R/S statistic, take a sample XI, 3 ,
X,,

with sample mean

a = an(g)
where

[

max
e
l
n

En,choose q
k

- -

X (Xj - X,)
j=1

lags, and calculate:
min

k
j=1

...

www.clevelandfed.org/research/workpaper/index.cfm

Intuitively, the numerator in equation (14) measures the memory in the process
via the partial sums. White noise does not stay long above the mean:
Positive values are soon offset by negative values. A random walk will remain
above or below zero for a long time, and the partial sums (positive or
negative) will grow quickly, making the range large. Fractional processes
fall in between. Mandelbrot (1972) refers to this as the "Joseph Effect" - seven fat and seven lean years. The denominator normalizes not only by the
variance, but by a weighted average of autoco~ariances.~This innovation
over Hurst's R/S statistic provides the robustness to short-term dependence.
The partial sums of white noise constitute a random walk, so a(q) grows
without bound as n increases. A further normalization makes the statistic
easier to work with and interpret:
Vn(q)

- QJq)/*j(n).

Haubrich and Lo derive the asymptotic distribution of V, calculating a mean
and standard deviation of approximately 1.25 and 0.27. Tables 3A and

www.clevelandfed.org/research/workpaper/index.cfm

3B present fractiles of the distribution of V and confidence intervals about
the mean.

Figure 3 plots the distribution and density. Note that the

distribution is skewed, with most of its mass between three-fourths and two.
Table 4 reports the results of the modified R/S statistic applied to first
differences of labor income and disposable income. Note that none are
significantly different from the mean at the 5 percent level.

3.4 Simulation Results
Although the modified R/S statistic provides a good test (in terms of size and
power) for detecting long-term dependence, it does not directly provide the d
parameter. To better assess the chances that a d parameter from the correct
range will fit the data, I use simulation methodology.
Simulations employed here ran as follows. I used a Vax Fortran program (a
modification of one written by Lo) to generate 10,000 series of length 135
(not quite matching the data-series length of 136, to compare this study to
other papers).

The series were generated to have fractional differencing

parameter d for several d. I then computed the modified R/S statistic for
each series and counted the number of times that this value fell below the
value obtained from the income data above (Table 4).

This gives the

percentage of times the statistic would be that low if the income series
actually had that d parameter. I emphasize low because in first-difference
form the relevant d would be negative, which should show up as a low R/S
statistic. Table 5 reports these results and also answers the question: If

www.clevelandfed.org/research/workpaper/index.cfm

the process is really fractionally differenced with a particular d, what is
the probability that we would see the V,(q)

number found in the data, or

even a lower number? Of course, subtracting these numbers from one gives the
probability of obtaining a higher R/S statistic. The reader may draw
different conclusions from Table 5, but I think that the results provide mild
support for the belief that fractional processes can explain the
excess-smoothness problem. It seems unlikely that the actual d for either
income process is smaller than the lower bounds obtained above; we would
expect to see much lower numbers than those in Table 4. That is, Table 5
tells us that the probability of seeing that number or a lower one is very
high for such a process with a d of -0.21 or -0.28. On the other hand, the
chance of d

=

-0.04or -0.05 producing such a number is more reasonable.

Earlier in this section, we saw what range d could fall into and still
resolve the Deaton paradox. Now we see, in a general way, how likely it is
that d could be in that range. The chance remains that d is too close to
zero to resolve the paradox by invoking fractional methods. I submit that
Table 5 opens the very real possibility that d falls into the relevant range.
4. Conclusion
Judging the smoothness of consumption depends on the estimate of permanent
income, which in turn depends on our estimate of income. Paradoxes under one
specification

--

--

excess smoothness when income is assumed to have a unit root

do not arise when income is fractional.
The explanation that I propose leaves intact two similar problems in the

consumption literature. First, panel studies have found excess sensitivity of

www.clevelandfed.org/research/workpaper/index.cfm

precisely the opposite type Campbell and Deaton find in aggregate data.
Consumption variance is too high given the estimates for income. Flavin
finds a different type of excess sensitivity, namely, that consumption depends
on past income; it is not a martingale (the expected future value equals
today's value), as the permanent-income hypothesis predicts. Campbell and
Deaton refer to this as the "nonorthogonality" problem.
Nonetheless , without dropping either the permanent- income hypothesis or
the univariate representation of income, fractional processes resolve the
Deaton paradox. Theoretically, a fractional-income process matches the
observed variance of both income and consumption. Empirically, on the basis
of a new statistic and simulations, the evidence supports income following
such a process.

www.clevelandfed.org/research/workpaper/index.cfm

References

Auerbach, Alan J., and Kevin Hassett, "Corporate Savings and a Shareholder
Consumption," Working Paper No. 2994, National Bureau of Economic
Research, June 1989.
Campbell, John, and Angus Deaton, "Why is Consumption So Smooth?" Review
of Economic Studies, 56, 1989, pp. 357-373.

, and N. Gregory Mankiw, "Are Output Fluctuations
Transitory?" Quarterly Journal of Economics, 102, 1987, pp. 857-880.
Chatfield, C., The Analysis of Time Series: An Introduction, 3rd ed.
New York: Chapman and Hall, 1984.
Deaton, Angus, "Life Cycle Models of Consumption: Is the Evidence Consistent
with the Theory?" in Tnunan Bewley, ed., Advances in Econometrics:
Fifth World Congress, vol. 2, New York: Cambridge University Press,
1987.
Diebold, Francis X., and Glenn D. Rudebusch, "Is Consumption Too Smooth?
Long Memory and the Deaton Paradox." Washington, D.C.: Board of Governors of
the Federal Reserve System, March 1989.
Flavin, Marjorie, "The Adjustment of Consumption to Changing Expectations
about Future Income,"Journal of Political Economy, 89, 1981, pp.
974-1009.

, "The Excess Smoothness of Consumption: Identification and
Interpretation," University of Virginia Working Paper, November 1988.
Geweke, John, and Susan Porter-Hudak, "The Estimation and Application of Long
Memory Time Series Models," Journal of Time Series Analysis, 4, 1983,
pp. 221-238.
Granger, Clive, and Roselyne Joyeux, "An Introduction to Long-Memory Time
Series Models and Fractional Differencing," Journal of Time Series
Analysis, 1, 1980, pp. 14-29.
Hall, Robert E., "Stochastic Implications of the Life Cycle-Permanent Income
Hypothesis: Theory and Evidence," Journal of Political Economy, 86,
1978, pp. 971-987.

www.clevelandfed.org/research/workpaper/index.cfm

Hansen, Lars P., "Large Sample Properties of Generalized Method of Moments
Estimators," Econometrica, 50, 1982, pp. 1029-1054.

, and Thomas J. Sargent, "A Note on Wiener-Kolmogorov
Prediction Formulas for Rational Expectations Models," Economics
Letters, 8, 1981, pp. 255-260.
, and Kenneth J. Singleton, "Generalized Instrumental Variables
Estimation of Nonlinear Rational Expectations Models," Econometrica,
50, 1982, pp. 1269-1286.
Haubrich, Joseph G., and Andrew W. Lo, "The Sources and Nature of Long-Term
Memory in the Business Cycle," Working Paper No. 2951, National Bureau of
Economic Research, April 1989.
Hosking, J. R. M., "Fractional Differencing," Biometrica, 68, 1981, pp.
165- 176.
Hurst, Harold E., "Long Term Storage Capacity of Reservoirs," Transactions
of the American Society of Civil Engineers, 116, 1951, pp. 770-799.
Lo, Andrew W., "Long-Term Memory in Stock Market Prices," Econometrica
(forthcoming) .
Mandelbrot, Benoit, "Statistical Methodology for Non-Periodic Cycles: From
the Covariance to R/S Analysis," Analysis of Economic and Social
Measurement, 1, 1972, pp. 259-290.
Mehra, Rajnish, and Edward C. Prescott, "The Equity Premium: A Puzzle,"
Journal of Monetary Economics, 15, 1985, pp. 145-161.
Nelson, Charles R., and Charles I. Plosser, "Trends and Random Walks in
Macroeconomic Time Series: Some Evidence and Implications," Journal
of Monetary Economics, 10, 1982, pp. 139-162.
Newey, Whitney K., and Kenneth D. West, "A Simple, Positive Semi-Definite
Heteroskedasticity and Autocorrelation Consistent Covariance Matrix,"
Econornetrica, 55, 1987, pp. 703-708.
Ng Lo, Nancy, "An Econometric Analysis of the Role of the Price Discovery
in Futures Markets." Ph.D. dissertation. Wharton School, University of
Pennsylvania, 1988.
Phillips, Peter C. B., "Time Series Regression With a Unit Root,"
Econornetrica, 55, 1987, pp. 277-301.
Quah, Danny, "Permanent and Transitory Movements in Labor Income: An

Explanation for 'Excess Smoothness' in Consumption," Journal of
Political Economy, 98, 1990, pp. 449-475.

www.clevelandfed.org/research/workpaper/index.cfm

West, Kenneth D., "The Insensitivity of Consumption to News about Income,"
Journal of Monetary Economics, 21, 1988, pp. 17-34.
Zeldes, Stephen P., "Optimal Consumption with Stochastic Income: Deviations
from Certainty Equivalence," Quarterly Journal of Economics, 114,
1989, pp. 275-298.

www.clevelandfed.org/research/workpaper/index.cfm

Table 1
Sample Variances

95% Confidence
Bounds

Variance

Consumption

11.65

GMM

Classical

6.37
9.17

16.93
15.30

Labor income

65.35

GMM

36.43
Classical 52.18

94.28
84.24

Disposable
income

61.14

GMM
28.90
Classical 48.82

93.38
78.81

Consumption

=

First difference of real per-capita consumption of nondurables and services,
1989:IQ-1989:IIQ (quarterly data, seasonally
adjusted). Source: National Income and Product Accounts.

Population

=

U.S. total resident population, including
armed forces. Source: National Income and
Product Accounts.

Labor
income

=

First difference of quarterly real percapita labor income, 1952:IQ-1986:IQ.
Sources: Auerbach and Hassett (1989) and
National Income and Product Accounts.

Disposable
income

=

As above. Source: Auerbach and Hassett (1989).

www.clevelandfed.org/research/workpaper/index.cfm

Table 2A
Implied Income Innovation Variances
Labor Income
Implied variance from consumption
Lower bound
Upper bound

d
Interest rate

=

0.05%

Interest rate

=

1%

Interest rate

=

0.2%

Source:

See table 1.

Implied variance
from income

www.clevelandfed.org/research/workpaper/index.cfm

Table 2B
Implied Income Innovation Variances
Labor Income
Implied variance from consumption
Lower bound
Upper bound

d
Interest rate

=

0.05%

Interest rate

=

1%

Interest rate

=

0.2%

Source:

See table 1.

Imp1ied variance
from income

www.clevelandfed.org/research/workpaper/index.cfm

Note, Tables 2A and 2B

Approximations:

Closed-form solutions f o r the i n f i n i t e sums used i n these
calculations do not e x i s t . An upper bound on t h e f i n i t e
sum of N terms and the i n f i n i t e sum is ;(l/l+r)'.

The approximation i s i n f a c t b e t t e r . 10,000 terms were
used f o r t h e i n t e r e s t r a t e s r = 0.01 and r = 0.002, leading
t o e r r o r s of l e s s than 1 x lo-' and 1.05 x
20,000 terms used f o r r = 0.0005 give an e r r o r of l e s s than
0.09.

www.clevelandfed.org/research/workpaper/index.cfm

Table 3A
Fractiles of the Distribution F,(v)

Source: Haubrich and Lo (1989).

Table 3B
Symmetric Confidence Intervals About the Mean
7

Source: Haubrich and Lo (1989).

www.clevelandfed.org/research/workpaper/index.cfm

Table 4
R/S Analysis of Income

Labor Income
Disposable Income

1.193
1.261

1.310
1.268

1.140
1.245

1.062
1.176

1.018
1.170

Note: Both series per capita.
Sources: See table 1.

Table 5
Probability of Observing R/S Statistic

Probability of
LAG ( 1)

Labor Income
d=-0.21
d=-0.05

Source: Author's simulations.

IVn(q)

Disposable Income
d=-0.04
d=-0.28

www.clevelandfed.org/research/workpaper/index.cfm

26
Footnotes

1. For an estimate of income with a view to explaining consumption anomalies
in the spirit of this section, see the interesting (independent) work of
Diebold and Rudebusch (1989). Quah (1990) explains the paradox using
permanent and temporary movements in income.

2. These weights define the Bartlett window. Newey and West (1987) enumerate
the advantages of this specification.

www.clevelandfed.org/research/workpaper/index.cfm

27

Figure 1

-

9

1 .4 7 5 ~t

pj f o r (
. . .
-. . . . . . . ... . . .. . .. . ... . . .. . . . . . . .
. . . . . ..

...

.

.

.

.

.

.

.

.

-

.

. . . . . . . . . . . . . .

....

..;... . . . . ... . . . .. . . . :.......: . . . . . . . . .: ............ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . ... .
. . -.. . . .. . . . . .:. . . . . . . . . . . . . . . . . .
-. .........-... - . .. - . . . . .. . .. . . . . . . .. . . .:.
.
.
.
..1.
..:..
i
.
.
i
.
.
:
.
.
.
:
.
.
. . . . . . .. . . . .. . . . . . . . . .. .
+
. .. .. . .
.
.
.
.
.
.\..:
.............
.........
.;
...........................
...!.
......................
>...
..i.... ..... .:.. ..... ...,
4
--lo-... . . .
.
. . . . ..' " ' . . . . . . . .. . . . .. . . . . . . .. .
. \
. . . . . . . . .
..;..W
. .. . . . . . . . . . . . . ... . .. :..: . . . . . . . . . . . . . . . .
:
.: . . .. . .. . .. . .
. .
- ..!.
. . : ..:.. .:.. . . '
: . ..:. . . . : .
.
1 . .. . . .:...:..
;...
.
.
.
.
.
.
\ ............1 ...... :.. ......... ..; ..... . . . . . . i . ....: : . . . . :.. . : . .: ...;....:
.
.
. . .
. . . . . .. . . . .. . .. . . . .. . .. .
.......
0 . .../ :. . .; . . :. . :
- . . .. . . .. . . . . . . ' . . . . : . . , . . :. . . . .; .1 . . .. . . . .. . .. .
- . . 1...:. . . .
.. . ,. .
.
.
.. , .
4

-\ . . .. . . ... . . ... . . . ... . . . . . . ..

0

*

i

:

-

:...

.

"

e
e

-

:.

,' 9

-

f

"-

0

..

:\ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
.

0

- . . . . . . \. .
-. . .
. . ,
; ., . ,

,

O

0.

30

.

.

.

.

.

.

.

.

.

.

..

.

-

. . . . . . .

. . . . . . . . ..

,

-

.

.

. . .
,

60

....

. . .

.
.
\ _ . .. . . :. . . . . . . . . .

90

.

.

I

.

120

LAG

Autocorrelation functions of an AR(1) with coefficient 0.90 [dashed line] and a fractionally
differenced series Xt = (1 - L ) - ~ @with differencing parameter d = 0.475 [solid lie]. Although both processes have a &&order autocorrelationof 0.90, the fractionallydifferenced
process decays much more slowly.

Source: Haubrich and Lo (1989).

www.clevelandfed.org/research/workpaper/index.cfm

28

F i g u r e 2A

Impulse response function [solid line] of the fractiindly differenced time series Xt =
(1 ~
) for ~differencing
e
~parameter d = 0.475. For comparisiin, the imp&-function of an AR(1) with autoregressiveparameter 0.90 is slso plotted [dashed lines].

-

Source: Haubrich and Lo (1989).

www.clevelandfed.org/research/workpaper/index.cfm

Equivalent p of AR(I)

..

... . . . . .
:

LAG

Values of aa AR(1)kr a u t o ~ parameta
e
required to generate the same k-th order
autocorrelation as the &actionally diflaenced suies Xt = (1 L)'~Q for diff-~ing
parameter d = 0.475 [wUd he]. Formdo, th& ia himply the k-th root of the frrctionally
differenced eeries' hpubmpomw function [dashed line]. For large k, the autmegrasipe
parameter must be very close to unity.

-

Source: Haubrich and Lo (1989).

www.clevelandfed.org/research/workpaper/index.cfm

30
Figure 3

Distribution and density function of the range V of a Brownian bridge. Dashed curves are
the n o d distribution and density functions with mean and variance equal to t h e ofV.

Source: Haubrich and Lo (1989).