Practical Issues in the A nalysis of Univariate GA R C H
Models
Eric Zivot
April 18, 2008
Abstract
This paper gives a tour through the empirical analysis of univariate GARC H
models for nancial time series with stops along the way to discuss various prac-
tical issues associated with model specication, estimation, diagnostic evaluation
and forecasting.
1Introduction
There are many very good surveys covering the mathematical and statistical prop-
erties of GARCH models. See, for example, [9], [14], [74], [76], [27] and [83]. There
are also several comprehensive surveys that focus on the forecasting performance
of GARCH models including [78], [77], and [3]. However, there are relatively few
surveys that focus on the practical econometric issues associated with estimating
GARCH models and forecasting volatility. This paper, which draws heavily from
[88], gives a tour through the empirical analysis of univariate GARCH models for
nancial time series with stops along the way to discuss various practical issues.
Multivariate GARCH models are discussed in the paper by [80]. The plan of this pa-
per is as follows. Section 2 reviews some stylized facts of asset returns using example
data on Microsoft and S&P 500 index returns. Section 3 reviews the basic univariate
GARCH model. Testing for GARCH eects and estimation of GARCH models are
covered in Sections 4 and 5. Asymmetric and non-Gaussian GARCH models are dis-
cussed in Section 6, and long memory GARCH models are briey discussed in Section
7. Section 8 discusses vol atility forecasting, and nal remarks are given Section 9
1
.
1
Asset Mean Med Min Max Std. Dev Skew Kurt JB
Daily Returns
MSFT 0.0016 0.0000 -0.3012 0.1957 0.0253 -0.2457 11.66 13693
S&P 500 0.0004 0.0005 -0.2047 0.0909 0.0113 -1.486 32.59 160848
Monthly Returns
MSFT 0.0336 0.0336 -0.3861 0.4384 0.1145 0.1845 4.004 9.922
S&P 500 0.0082 0.0122 -0.2066 0.1250 0.0459 -0.8377 5.186 65.75
Notes: Sample period is 03/14/86 - 06/30/03 giving 4365 daily observations.
Table 1: Summary Statistics for Daily and Monthly Stock Returns.
2 Some St y lized Facts of Asset Returns
Let P
t
denote the price of an asset at the end of trading day t. The con tinuously
compounded or log return is dened as r
t
=ln(P
t
/P
t1
). Figure 1 plots the daily
log returns, squared returns, and absolute value of returns of Microsoft stock and
the S&P 500 index over the period March 14, 1986 through June 30, 2003. There
is no clear discernible pattern of behavior in the log returns, but there is some per-
sistence indicated in the plots of the squared and absolute returns which represent
the volatility of returns. In particular, the plots show evidence of volatility clus-
tering - low values of volatility followed by low values and high values of volatility
followed by high values. This behavior is conrmed in Figure 2 which shows the
sample autocorrelations of the six series. The log returns show no evidence of serial
correlation, but the squared and absolute returns are positively autocorrelated. Also,
the decay rates of the sample autocorrelations of r
2
t
and |r
t
| appear much slower,
especially for the S&P 500 index, than the exponential rate of a covariance station-
ary autoregressive-moving average (ARMA) process suggesting possible long memory
behavior. Monthly returns, dened as the sum of daily returns over the month, are
illustrated in Figure 3. The monthly returns display much less volatility clustering
than the daily returns.
Table 1 giv es some standard summary statistics along with the Jarque-Bera test
for normality. The latter is computed as
JB =
T
6
Ã
[
skew
2
+
(
d
kurt 3)
2
4
!
, (1)
where
[
skew denotes the sample skewness and
d
kurt denotes the sample kurtosis. Under
the null that the data are iid normal, JB is asymptotically distributed as chi-square
The paper was prepared for the Handbook of Financial Time S eries , edited by T.G. Ande rsen ,
R.A. Davis, J-P Kreiss, and T. M ikosch. Thanks to Saraswata Chaudhuri, R ichard Davis, Ron
Scho enberg and Jiahui Wang for helpful comments a nd suggestions. Financial support from the
Gary Waterman Distinguished Scholarship is greatly appreciated.
Department of Economics, Box 353330, University of Washington. ezivot@u.washington.edu.
1
All of the examples in the pap er were constructed using S-PLUS 8.0 and S+FinMetrics 2.0. Script
les for replicating the examples may be downloaded from http://faculty.washington.edu/ezivot.
2
1986 1990 1994 1998 2002
-0.30 0.10
Microsoft Returns
1986 1990 1994 1998 2002
-0.20 0.05
S & P 500 Returns
1986 1990 1994 1998 2002
0.00 0.08
Microsoft Squared Returns
1986 1990 1994 1998 2002
0.000 0.040
S & P 500 Squared Returns
1986 1990 1994 1998 2002
0.00 0.30
Microsoft Absolute Returns
1986 1990 1994 1998 2002
0.00 0.30
S & P 500 Absolute Returns
Figure 1: Daily returns, squared returns and absolute returns for Microsoft and the
S&P 500 index.
with 2 degrees of freedom. The distribution of daily returns is clearly non-normal
with negative skewness and pronounced excess kurtosis. Part of this non-normality
is caused by some large outliers around the October 1987 stock market crash and
during the bursting of the 2000 tech bubble. However, the distribution of the data
still appears highly non-normal even after the removal of these outliers. Monthly
returns have a distribution that is much closer to the normal than daily returns.
3 The AR CH and GAR CH Model
[33] showed that the serial correlation in squared returns, or conditional heteroskedas-
ticity, can be modeled using an autoregressive conditional heteroskedast icity (ARCH)
modeloftheform
y
t
= E
t1
[y
t
]+
t
, (2)
t
= z
t
σ
t
, (3)
σ
2
t
= a
0
+ a
1
2
t1
+ ···+ a
p
2
tp
, (4)
3
Lag
ACF
0 5 10 15 20
0.0 0.6
Microsoft Returns
Lag
ACF
0 5 10 15 20
0.0 0.6
S&P 500 Returns
Lag
ACF
0 5 10 15 20
0.0 0.6
Microsoft Squared Returns
Lag
ACF
0 5 10 15 20
0.0 0.6
S&P 500 Squared Returns
Lag
ACF
0 5 10 15 20
0.0 0.6
Microsoft Absolute Returns
Lag
ACF
0 5 10 15 20
0.0 0.6
Microsoft Absolute Returns
Figure 2: Sample autocorrelations of r
t
,r
2
t
and |r
t
| for Microsoft and S&P 500 index.
where E
t1
[·] represents expectation conditional on information available at time t1,
and z
t
is a sequence of iid random variables with mean zero and unit variance. In the
basic A RCH model z
t
is assumed to be iid standard normal. The restrictions a
0
> 0
and a
i
0(i =1,...,p) are required for σ
2
t
> 0. The representation (2) - (4) is
convenient for deriving properties of the model as well as for specifying the likelihood
function for estimation. The equation for σ
2
t
can be rewritten as an AR(p) process
for
2
t
2
t
= a
0
+ a
1
2
t1
+ ···+ a
p
2
tp
+ u
t
, (5)
where u
t
=
2
t
σ
2
t
is a martingale dierence sequence (MDS) since E
t1
[u
t
]=0and
it is assumed that E(
2
t
) < .Ifa
1
+ ···+ a
p
< 1 then
t
is covariance stationary,
the persistence of
2
t
and σ
2
t
is measured by a
1
+ ···+ a
p
and ¯σ
2
=var(
t
)=E(
2
t
)=
a
0
/(1 a
1
··· a
p
).
An important extension of the ARCH model proposed by [12] replaces the AR(p)
representation in (4) with an ARMA(p, q) formulation
σ
2
t
= a
0
+
p
X
i=1
a
i
2
ti
+
q
X
j=1
b
j
σ
2
tj
, (6)
where the coecients a
i
(i =0, ··· ,p) and b
j
(j =1, ··· ,q) are all assumed to be
4
1986 1990 1994 1998 2002
-0.3 0.4
Microsoft Returns
1986 1990 1994 1998 2002
-0.20 0.10
S&P 500 Returns
1986 1990 1994 1998 2002
0.02 0.18
Microsoft Squared Returns
1986 1990 1994 1998 2002
0.005
S&P 500 Squared Returns
Lag
ACF
0 5 10 15 20
0.0 0.6
Microsoft Squared Returns
Lag
ACF
0 5 10 15 20
0.0 0.6
S&P 500 Squared Returns
Figure 3: Monthly Returns, Squared Returns and Sample Autocorrelations of
Squared Returns for Microsoft and the S&P 500.
positive to ensure that the conditional variance σ
2
t
is always positive.
2
The model in
(6) together with (2)-(3) is known as the generalized ARCH or GARCH(p, q) model.
The GARCH(p, q) model can be shown to be equivalent to a particular ARCH()
model. When q =0, the GAR CH model reduces to the ARC H model. In order for
the GARCH parameters, b
j
(j =1, ··· ,q), to be identied at least one of the ARCH
coecients a
i
(i>0) must be nonzero. Usually a GARCH(1,1) model with only
three parameters in the conditional variance equation is adequate to obtain a good
model tfornancial time series. Indeed, [49] provided compelling evidence that is
dicult to nd a volatility model that outperforms the simple GARCH(1,1).
Just as an AR CH model can be expressed as an AR model of squared residuals, a
GARCHmodelcanbeexpressedasanARMAmodelofsquaredresiduals.Consider
the GARCH(1,1) model
σ
2
t
= a
0
+ a
1
2
t1
+ b
1
σ
2
t1
. (7)
Since E
t1
(
2
t
)=σ
2
t
, (7) can be rewritten as
2
t
= a
0
+(a
1
+ b
1
)
2
t1
+ u
t
b
1
u
t1
, (8)
2
Positive coecients are sucient but not necessary conditions for the positivity of c onditional
variance. See [72] and [23] for more general conditions.
5
whichisanARMA(1,1)modelwithu
t
=
2
t
E
t1
(
2
t
) being the MDS disturbance
term.
Given the ARMA(1,1) representation of the GARCH(1,1) model, many of its
properties follow easily from those of the corresponding ARMA(1,1) process for
2
t
.
For example, the persistence of σ
2
t
is captured by a
1
+ b
1
and covariance stationarit y
requires that a
1
+ b
1
< 1. The cova riance stationary GARCH(1,1) model has an
AR CH() representation with a
i
= a
1
b
i1
1
, and the unconditional variance of
t
is
¯σ
2
= a
0
/(1 a
1
b
1
).
For the general GARCH(p, q) model (6), the squared residuals
t
behave like an
ARMA(max(p, q),q) process. Covariance stationarity requires
P
p
i=1
a
i
+
P
q
j=1
b
i
< 1
and the unconditional variance of
t
is
¯σ
2
=var(
t
)=
a
0
1
³
P
p
i=1
a
i
+
P
q
j=1
b
i
´
. (9)
3.1 Conditional Mean Specication
Depending on the frequency of the data and the type of asset, the conditional mean
E
t1
[y
t
] is typically specied as a constant or possibly a low order autoregressive-
moving average (ARMA) process to capture autocorrelation caused by market mi-
crostructure eects (e.g., bid-ask bounce) or non-trading eects. If extreme or un-
usual market ev ents have happened during sample period, then dummy variables
associated with these events are often added to th e conditional mean specication to
remove these eects. Therefore, the typical conditional mean specication is of the
form
E
t1
[y
t
]=c +
r
X
i=1
φ
i
y
ti
+
s
X
j=1
θ
j
tj
+
L
X
l=0
β
0
l
x
tl
+
t
, (10)
where x
t
is a k × 1 vector of exogenous explanatory variables.
In nancial investmen t, high risk is often expected to lead to high returns. Al-
though modern capital asset pricing theory does not imply such a simple relationship,
it does suggest that there are some interactions between expected returns and risk as
measured by volatility. Engle, Lilien and Robins (1987) proposed to extend the basic
GARCH model so that the conditional volatility can generate a risk premium which
is part of the expected returns. This extended GARCH model is often referred to
as GARCH-in-the-mean or GARCH-M model. The GARCH-M model extends the
conditional mean equation (10) to include the additional regressor g(σ
t
),whichcan
be an arbitrary function of conditional volatility σ
t
. The most common specications
are g(σ
t
)=σ
2
t
t
, or ln(σ
2
t
).
3.2 Ex planat ory Variables in the Co nditional Var iance Equat io n
Just as exogenous variables may be added to the conditional mean equation, exoge-
nous explanatory variables may also be added to the conditional variance formula (6)
6
in a straightforward w ay giving
σ
2
t
= a
0
+
p
X
i=1
a
i
2
ti
+
q
X
j=1
b
j
σ
2
tj
+
K
X
k=1
δ
0
k
z
tk
,
where z
t
is a m × 1 vector of variables, and δ is a m × 1 vector of positive coe-
cients. Variables that have been shown to help predict volatility are trading volume,
macroeconomic news announcements ([58], [43], [17]), implied volatility from option
prices and realized volatility ([82], [11]), overnight returns ([46], [68]), and after hours
realized volatility ([21])
3.3 The GARCH M odel and Stylized Facts of Asset Returns
Previously it was shown that the daily returns on Microsoft and the S&P 500 ex-
hibited the “stylized facts” of volatility clustering as w ell as a non-normal empirical
distribution. Researchers have documented these and many other st ylized facts about
the volatility of economic and nancial time series. [14] gave a complete account of
these facts. Using the ARMA representation of GARCH models shows that the
GARCH model is capable of exp laining many of those stylized facts. The four most
important ones are: volatility clustering, fat tails, volatility mean reversion, and
asymmetry.
To understand volatility clustering, consider the GARCH(1, 1) model in (7). Usu-
ally the GARCH coecient b
1
is found to be around 0.9 for many daily or weekly
nancial time series. Giv en this value of b
1
,itisobviousthatlargevaluesofσ
2
t1
will
be followed b y large values of σ
2
t
, and small values of σ
2
t1
will be followed b y small
values of σ
2
t
. The same reasoning can be obtained from the ARMA represent ation in
(8), where large/small changes in
2
t1
will be followed by large/small changes in
2
t
.
It is well know n that the distribution of many high frequency nancial time series
usually have fatter tails than a normal distribution. That is, extreme values occur
more often than implied by a normal distribution. [12] gave the condition for the
existence of the fourth order moment of a GARCH(1, 1) process. Assuming the
fourth order moment exists, [12] showed that the kurtosis implied by a GARCH(1, 1)
process with normal errors is greater than 3, the kurtosis of a normal distribution.
[51] and [52] extended these results to general GARC H(p, q) models. Thus a GARCH
model with normal errors can replicate some of the fat-tailed behavior observed in
nancial time series. A more thorough discussion of extreme value theory for GARCH
is given by [24]. Most often a GARCH model with a non-normal error distribution
is required to fully capture the observed fat-tailed behavior in returns. These models
are reviewed in sub-Section 6.2.
Although nancial markets may experience excessive volatilit y from time to time,
it appears that volatilit y will ev entually settle down to a long run level. Recall, the
unconditional variance of
t
for the stationary GARCH(1, 1) model is ¯σ
2
= a
0
/(1
a
1
b
1
). To see that the volatilit y is always pulled toward this long run, the ARMA
representation in (8) may be rewritten in mean-adjusted form as:
(
2
t
¯σ
2
)=(a
1
+ b
1
)(
2
t1
¯σ
2
)+u
t
b
1
u
t1
. (11)
7
If the abo ve equation is iterated k times, it follows that
(
2
t+k
¯σ
2
)=(a
1
+ b
1
)
k
(
2
t
¯σ
2
)+η
t+k
,
where η
t
is a moving average process. Since a
1
+ b
1
< 1 for a covariance stationary
GAR CH (1, 1) model, (a
1
+ b
1
)
k
0 as k →∞. Although at time t there may be a
large deviation between
2
t
and the long run variance,
2
t+k
¯σ
2
will approach zero
“on average” as k gets large; i.e., the vo latility “mean reverts” to its long run level
¯σ
2
. The magnitude of a
1
+ b
1
controls the speed of mean reversion. The so-called
half-life of a volatility shock, dened as ln(0.5)/ ln(a
1
+ b
1
), measures the average
time it takes for |
2
t
¯σ
2
| to decrease by one half. Obviously, the closer a
1
+ b
1
is to
one the longer is the half-life of a volatility shock. If a
1
+ b
1
> 1, the GARCH model
is non-stationary and the volatility will eventually explode to innit y as k →∞.
Similar arguments can be easily constructed for a GARCH(p, q) model.
The standard GARCH(p, q) model with Gaussian errors implies a symmetric dis-
tribution for y
t
and so cannot account for the observed asymmetry in the distribution
of returns. Ho wever, as shown in Section 6, asymmetry can easily be built into the
GARCH model by allowing
t
to have an asymmetric distribution or by explicitly
modeling asymmetric behavior in the conditional variance equation (6).
3.4 Temporal Aggregation
Volatility clustering and non-Gaussian behavior in nancial returns is typically seen
in weekly, daily or intraday data. The persistence of conditional volatility tends to
increase with the sampling frequency
3
. However, as shown in [32], for GARCH models
there is no simple aggregation principle that links the parameters of the model at
one sampling frequency to the parameters at another frequency. This occurs because
GAR C H models imply that the squared residual process follows an ARMA t ype
process with MDS innovations which is not closed under temporal aggregation. The
practical result is that GARCH models tend to be t to the frequency at hand. This
strategy, however, may not provide the best out-of-sample volatility forecasts. For
example, [68] showed that a GARCH model t to S&P 500 daily returns produces
better forecasts of weekly and month ly volatility than GARCH models ttoweekly
or monthly returns, respectively.
4 T esting for ARCH/GARCH eects
The stylized fact of volatility clustering in returns manifests itself as autocorrelation
in squared and absolute returns or in the residuals from the estimated conditional
mean equation (10). The signicance of these autocorrelations may be tested using
3
The empirical result that aggregated returns exhib it sm aller GARCH e ects and approach
Gaussian behavior can be explained by the results of [26] who showed that a central limit the-
orem holds for standardized sums of random variables that follow covariance stationary GARCH
pro cesses.
8
the Ljung-Box or modied Q-statistic
MQ(p)=T (T +2)
p
X
j=1
ˆρ
2
j
T j
, (12)
where ˆρ
j
denotes the j-lag sample autocorrelation of the squared or absolute returns.
IfthedataarewhitenoisethentheMQ(p) statistic has an asymptotic chi-square dis-
tribution with p degrees of freedom. A signicant value for MQ(p) provides evidence
for time varying conditional volatility.
To test for autocorrelation in the raw returns when it is suspected that there are
GAR CH eects present, [27] suggested using the following heteroskedasticity robust
version of (12)
MQ
HC
(p)=T(T +2)
p
X
j=1
1
T j
Ã
ˆσ
4
ˆσ
4
γ
j
!
ˆρ
2
j
,
where ˆσ
4
is a consisten t estimate of the squared unconditional variance of returns,
and ˆγ
j
is the sample autocovariance of squared returns.
Since an ARCH model implies an AR model for the squared residuals
2
t
, [33]
showed that a simple Lagrange multiplier (LM) test for ARCH eects can be con-
structed based on the auxiliary regression (5). Under the null hypothesis that there
are no A RCH eects, a
1
= a
2
= ···= a
p
=0, the test statistic
LM = T · R
2
(13)
has an asymptotic chi-square distribution with p degrees of freedom, where T is the
sample size and R
2
is computed from the regression (5) using estimated residuals.
Even though the LM test is constructed from an ARCH model, [61] show that it
also has power against more general GARCH alternatives and so it can be used as a
general specication test for GARCH eects.
[64], however, argued that the LM test (13) may reject if there is general mis-
specication in the conditional mean equation (10). They showed that such misspec-
ication causes the estimated residuals ˆ
t
to be serially correlated which, in turn,
causes ˆ
2
t
to be serially correlated. Therefore, care should be exercised in specifying
the conditional mean equation (10) prior to testing for AR CH eects.
4.1 Testing for ARCH Eects in Daily and M onthly R etur ns
Table 2 shows values of MQ(p) computed from daily and monthly squared returns
and the LM test for ARCH, for various values of p, for Microsoft and the S&P 500.
There is clear evide n ce of volatility clustering in the daily returns, but less evidence
for monthly returns especially for the S&P 500.
5 Estimation of GA RCH M odels
The general GARCH(p, q) model with normal errors is (2), (3) and (6) with z
t
iid N(0, 1). For simplicity, assume that E
t1
[y
t
]=c. Given that
t
follows Gaussian
9
MQ(p) r
2
t
LM
Asset p 15101510
Daily Returns
MSFT
56.81
(0.000)
562.1
(0.000)
206.8
(0.000)
56.76
(0.000)
377.9
(0.000)
416.6
(0.000)
S&P 500
87.59
(0.000)
415.5
(0.000)
456.1
(0.000)
87.52
(0.000)
311.4
(0.000)
329.8
(0.000)
Monthly Returns
MSFT
0.463
(0.496)
17.48
(0.003)
31.59
(0.000)
0.455
(0.496)
16.74
(0.005)
33.34
(0.000)
S&P 500
1.296
(0.255)
2.590
(0.763)
6.344
(0.786)
1.273
(0.259)
2.229
(0.817)
5.931
(0.821)
Notes: p-values are in parentheses.
Table 2: Tests for ARCH Eects in Daily Stock Returns
distribution conditional on past history, the prediction error decomposition of the
log-likelihood function of the GARCH model conditional on initial values is
log L =
T
X
t=1
l
t
=
T
2
log(2π)
1
2
T
X
t=1
log σ
2
t
1
2
T
X
t=1
2
t
σ
2
t
, (14)
where l
t
=
1
2
(log(2π)+logσ
2
t
)
1
2
2
t
σ
2
t
. The conditional loglikelihood (14) is used
in practice since the unconditional distribution of the initial values is not known
in closed form
4
. As discussed in [69] and [20], there are several practical issues to
consider in the maximization of (14). Starting values for the model parameters c, a
i
(i =0, ··· ,p) and b
j
(j =1, ··· ,q) need to be chosen and an initialization of
2
t
and
σ
2
t
must be supplied. The sample mean of y
t
is usually used as the starting value for
c, zero values are often given for the conditional variance parameters other than a
0
and a
1
, and a
0
is set equal to the unconditional variance o f y
t
5
. For the initial values
of σ
2
t
, a popular choice is
σ
2
t
=
2
t
=
1
T
T
X
s=1
2
s
,t 0,
where the initial values for
s
are computed as the residuals from a regression of y
t
on a constant.
Once the log-likelihood is initialized, it can be maximized using numerical op-
timization techniques. The most common method is based on a Newton-Raphson
iteration of the form
ˆ
θ
n+1
=
ˆ
θ
n
λ
n
H(
ˆ
θ
n
)
1
s(
ˆ
θ
n
),
4
[29] gave a computationally intensive numerical pro cedure for approximating the exact log-
likelihood.
5
Setting the starting values for all of the A RCH coecients a
i
(i =1,...,p) to ze ro may create
an ill-behaved likelihood and lead to a local minimum since the remaining GARCH parameters are
not identied.
10
where θ
n
denotes the vector of estimated model parameters at iteration n, λ
n
is a
scalar step-length parameter, and s(θ
n
) and H(θ
n
) denote the gradien t (or score)
vector and Hessian matrix of the log-likelihood at iteration n, respectively. The step
length parameter λ
n
is chosen such that ln L(θ
n+1
) ln L(θ
n
). For GARCH models,
the BHHH algorithm is often used. This algorithm approximates the Hessian matrix
using only rst derivative information
H(θ) B(θ)=
T
X
t=1
∂l
t
θ
∂l
t
θ
0
.
In the application of the Newton-Raphson algorithm, analytic or numerical deriva-
tives may be used. [41] pro vided algorithms for computing analytic derivatives for
GARCH models.
The estimates that maximize the conditional log-likelihood (14) are called the
maximum likelihood (ML) estimates. Under suitable regularity conditions, the ML
estimates are consistent and asymptotically normally distributed and an estimate of
the asymptotic covariance matrix of the ML estimates is constructed from an estimate
of the nal Hessian matrix from the optimization algorithm used. Unfortunately,
verication of the appropriate regularity conditions has only been done for a limited
number of simple GARCH models, see [63], [60], [55], [56] and [81]. In practice, it is
generally assumed that the necessary regularity conditions are satised.
In GARCH models for which the distribution of z
t
is symmetric and the parame-
ters of the conditional mean and variance equations are variation free, the information
matrix of the log-lik elihood is block diagonal. The implication of this is that the pa-
rameters of the conditional mean equation can be estimated separately from those
of the conditional variance equation without loss of asymptotic eciency. This can
greatly simplify estimation. An common model for which block diagonality of the
information matrix fails is the GARCH-M model.
5.1 Num erical Accuracy of GARCH Estimates
GARCH estimation is widely available in a number of commercial software packages
(e.g. EVIEWS, GAUSS, MATLAB, Ox, RATS, S-PLUS, TSP) and there are also
a few free open source implementations. [41], [69], and [20] discussed numerical ac-
curacy issues associated with maximizing the GARCH log-likelihood. They found
that starting values, optimization algorithm choice, and use of analytic or numerical
derivatives, and convergence criteria all inuence the resulting numerical estimates
of the GARCH parameters. [69] and [20] studied estimation of a GARCH(1,1) model
from a variety of commercial statistical packages using the exchange rate data of [15]
as a benchmark. They found that it is often dicult to compare competing software
since the exact construction of the GAR CH likelihood is not always adequately de-
scribed. In general, they found that use of analytic derivatives leads to more accurate
estimation than procedures based on purely numerical evaluations.
In practice, the GARCH log-likelihood function is not always well behaved, es-
pecially in complicated models with many parameters, and reaching a global max-
imum of the log-likelihood function is not guaranteed using standard optimization
11
techniques. Also, the positive variance and stationarity constraints are not straight-
forward to implement with common optimization software and are often ignored in
practice. P oor choice of starting values can lead to an ill-behaved log-likelihood and
cause convergence problems. Therefore, it is always a good idea to explore the surface
of the log-likelihood by perturbing the starting values and re-estimating the GARC H
parameters.
In many empirical applications of the GARCH(1,1) model, the estimate of a
1
isclosetozeroandtheestimateofb
1
is close to unity. This situation is of some
concern since the GARCH parameter b
1
becomes unidentied if a
1
=0, and it is
well kno w n that the distribution of ML estimates can become ill-behaved in models
with nearly unidentied parameters. [66] studied the accuracy of ML estimates of
the GARCH parameters a
0
,a
1
and b
1
when a
1
is close to zero. They found that the
estimated standard error for b
1
is spuriously small and that the t-statistics for testing
hypotheses about the true value of b
1
are severely size distorted. They also showed
that the concentrated loglikelihood as a function of b
1
exhibits multiple maxima. To
guard against spurious inference they recommended comparing estimates from pure
AR CH(p) models, which do not suer from the identication problem, with estimates
from the GARCH(1,1). If the volatility dynamics from these models are similar then
the spurious inference problem is not likely to be present.
5.2 Quasi-M axim um Likelihood Estimation
Another practical issue associated with GARCH estimation concerns the correct
choice of the error distribution. In particular, the assumption of conditional normality
is not always appropriate. However, as shown by [86] and [16], even when normal-
ity is inappropriately assumed, maximizing the Gaussian log-likelihood (14) results
in quasi-maximum likelihood estimates (QMLEs) that are consistent and asymptot-
ically normally distributed provided the conditional mean and variance functions of
the GAR CH model are correctly specied. Inaddition,[16]derivedanasymptotic
covariance matrix for the QMLEs that is robust to conditional non-normality. This
matrix is estimated using
H(
ˆ
θ
QML
)
1
B(
ˆ
θ
QML
)H(
ˆ
θ
QML
)
1
, (15)
where
ˆ
θ
QML
denotes the QMLE of θ, and is often called the “sandwich” estima-
tor. The coecient standard errors computed from the square roots of the diagonal
elements of (15) are sometimes called “Bollerslev-Wooldridge” standard errors. Of
course, the QMLEs will be less ecient than the true MLEs based on the correct er-
ror distribution. However, if the normality assumption is correct then the sandwich
covariance is asymptotically equivalent to the inverse of the Hessian. As a result, it
is good practice to routinely use t h e sandwich covariance for inference purposes.
[35] and [16] evaluated the accuracy of the quasi-maximum likelihood estimation
of GARCH(1,1) models. They found that if the distribution of z
t
in (3) is symmetric,
then QMLE is often close to the MLE. However, if z
t
has a skewed distribution then
theQMLEcanbequitedierent from the MLE.
12
5.3 Model Selectio n
An important practical problem is the determination of the ARCH order p and the
GARCH order q for a particular series. Since GARCH models can be treated as
ARMA models for squared residuals, traditional model selection criteria such as the
Akaike information criterion (AIC) and the Bayesian information criterion (BIC) can
be used for selecting models. For daily returns, if attention is restricted to pure
AR CH(p) models it is typically found that large values of p are selected b y AIC and
BIC. For GARCH(p, q) models, those with p, q 2 are typically selected by AIC
and BIC. Lo w order GARCH(p,q) models are generally preferred to a high order
AR CH(p) for reasons of parsimon y and better numerical stability of estimation (high
order GARCH(p, q) processes often have many local maxima and minima). For many
applications, it is hard to beat the simple GARC H(1,1) model.
5.4 Evaluat io n of Es timated GARCH models
After a GARCH model has been t to the data, the adequacy of the tcanbe
evaluated using a number of graphical and statistical diagnostics. If the GARCH
model is correctly specied, then the estimated standardized residuals ˆ
t
/ˆσ
t
should
behave like classical regression residuals; i.e., they should not display serial correla-
tion, conditional heteroskedasticity or any type of nonlinear dependence. In addition,
the distribution of the standardized residuals ˆ
t
/ˆσ
t
should match the specied error
distribution used in the estimation.
Graphically, ARCH eects reected by serial correlation in ˆ
2
t
/ˆσ
2
t
can be uncovered
by plotting its SACF. The modied Ljung-Box statistic (12) can be used to test the
null of no autocorrelation up to a specic lag, and Engle’s LM statistic (13) can be
used to test the null of no remaining ARCH eects
6
. If it is assumed that the errors
are Gaussian, then a plot of ˆ
t
/ˆσ
t
against time should have roughly ninety ve percent
of its values between ±2; a normal qq-plot of ˆ
t
/ˆσ
t
should look roughly linear
7
;and
the JB statistic should not be too mu ch larger than six.
5.5 Estima tio n of GA RCH Models for D a ily and M o nthly R etu rn s
Table 3 gives model selection criteria for a variety of GARCH(p, q) tted to the daily
returns on Microsoft and the S&P 500. For pure ARC H(p) models, an ARCH(5)
is chosen by all criteria for both series. For GARCH(p, q) models, AIC picks a
GARCH(2,1) for both series and BIC picks a GARCH(1,1) for both series
8
.
Table4givesQMLEsoftheGARCH(1,1)model assuming normal errors for the
Microsoft and S&P 500 daily returns. For both series, the estimates of a
1
are around
6
These tests should be viewed as ind icative, since the distribution of the tests a re inuenced by
the estimation of the GARCH model. For valid LM tests, the partial derivatives of σ
2
t
with respect
to the conditional volatility parameters should b e added as additional regressors in the auxiliary
regression (5) based on estimated residuals.
7
If an error distribution other than the Gaussian is assumed, then the qq-plot should be con-
structed using the quantiles of the assumed distribution.
8
The low log-likelihood values for the GARCH(2,2) models indicate that a local maximum was
reached.
13
(p, q) Asset AIC BIC Likelihood
(1,0) MSFT -19977 -19958 9992
S&P 500 -27337 -27318 13671
(2,0) MSFT -20086 -20060 10047
S&P 500 -27584 -27558 13796
(3,0) MSFT -20175 -20143 10092
S&P 500 -27713 -27681 13861
(4,0) MSFT -20196 -20158 10104
S&P 500 -27883 -27845 13947
(5,0) MSFT -20211 -20166 10113
S&P 500 -27932 -27887 13973
(1,1) MSFT -20290 -20264 10149
S&P 500 -28134 -28109 14071
(1,2) MSFT -20290 -20258 10150
S&P 500 -28135 -28103 14072
(2,1) MSFT -20292 -20260 10151
S&P 500 -28140 -28108 14075
(2,2) MSFT -20288 -20249 10150
S&P 500 -27858 -27820 13935
Table 3: Model Selection Criteria for Estimated GARCH(p,q) Models.
0.09 and the estimates of b
1
are around 0.9. Using both ML and QML standard er-
rors, these estimates are statistically dierent from zero. Howev er, the QML standard
errors are considerably larger than the ML standard errors. The estimated volatility
persistence, a
1
+ b
1
, is very high for both series and implies half-lives of shocks to
volatility to Microsoft and the S&P 500 of 15.5 days and 76 days, respectively. The
unconditional standard deviation of returns, ¯σ =
p
a
0
/(1 a
1
b
1
), for Microsoft
and the S&P 500 implied by the GARCH(1,1) models are 0.0253 and 0.0138, respec-
tively, and are very close to the sample standard deviations of returns reported in
Table 1.
Estimates of GARCH-M(1,1) models for Microsoft and the S&P 500, where σ
t
is added as a regressor to the mean equation, show small positive coecients on σ
t
and essentially the same estimates for the remaining parameters as the GARCH(1,1)
models.
Figure 4 shows the rst dierences of returns along with the tted one-step-
ahead volatilities, ˆσ
t
, computed from the GARCH(1,1) and ARCH(5) models. The
ARCH(5) and GARCH(1,1) models do a good job of capturing the observed volatil-
ity clustering in returns. The GARCH(1,1) volatilities, however, are smoother and
display more persistence than the ARC H(5) volatilities.
Graphical diagnostics from the tted GARCH(1,1) models are illustrated in Fig-
ure 5. The SACF of ˆ
2
t
/ˆσ
2
t
does not indicate any signicant autocorrelation, but
the normal qq-plot of ˆ
t
/ˆσ
t
shows strong departures from normality. The last three
columns of Table 4 give the standard statistical diagnostics of the tted GARCH
14
GARCH Parameters Residual Diagnostics
Asset a
0
a
1
b
1
MQ(12) LM(12) JB
Daily Returns
MSFT
2.80e
5
(3.42e
6
)
[1.10e
5
]
0.0904
(0.0059)
[0.0245]
0.8658
(0.0102)
[0.0371]
4.787
(0.965)
4.764
(0.965)
1751
(0.000)
S&P 500
1.72e
6
(2.00e
7
)
[1.25e
6
]
0.0919
(0.0029)
[0.0041]
0.8990
(0.0046)
[0.0436]
5.154
(0.953)
5.082
(0.955)
5067
(0.000)
Monthly Returns
MSFT
0.0006
[0.0006]
0.1004
[0.0614]
0.8525
[0.0869]
8.649
(0.733)
6.643
(0.880)
3.587
(0.167)
S&P 500
3.7e
5
[9.6e
5
]
0.0675
[0.0248]
0.9179
[0.0490]
3.594
(0.000)
3.660
(0.988)
72.05
(0.000)
Notes: QML standard errors are in brackets.
Table 4: Estimates of GARCH(1,1) Model with Diagnostics.
models. Consistent with the SACF, the MQ statistic and Engle’s LM statistic do
not indicate remain ing ARCH eects. Furthermore, the extremely large JB statistic
conrms nonnormality.
Table 4 also sho ws estimates of GARCH(1,1) models t to the monthly returns.
The GARCH(1,1) models t to the monthly returns are remarkable similar to those
t to the daily returns. There are, however, some important dierences. The monthly
standardized residuals are much closer to the normal distribution, especially for Mi-
crosoft. Also, the GARCH estimates for the S&P 500 reect some of the character-
istics of spurious GARCH eects as discussed in [66]. In particular, the estimate of
a
1
is close to zero, and has a relatively large QML standard error, and the estimate
of b
1
is close to one and has a very small standard error.
6GARCHModelExtensions
In many cases, the basic GARCH conditional variance equation (6) under normality
provides a reasonably good model for analyzing nancial time series and estimating
conditional volatility. Howev er, in some cases there are aspects of the model which
can be improved so that it can better capture the characteristics and dynamics of a
particular time series. For example, the empirical analysis in the previous Section
showed that for the daily returns on Microsoft and the S&P 500, the normality
assumption may not be appropriate and there is evidence of nonlinear behavior in
the standardized residuals from the tted GARCH(1,1) model. This Section discusses
several extensions to the basic GARCH model that make GARCH modeling more
exible.
15
1986 1990 1994 1998 2002
-0.30 0.10
Microsoft Daily Returns
1986 1990 1994 1998 2002
-0.20 0.05
S&P 500 Daily Returns
1986 1990 1994 1998 2002
0.02 0.10
Conditional Volatility from GARCH(1,1)
1986 1990 1994 1998 2002
0.01 0.06
Conditional Volatility from GARCH(1,1)
1986 1990 1994 1998 2002
0.02 0.10
Conditional Volatility from ARCH(5)
1986 1990 1994 1998 2002
0.01 0.09
Conditional Volatility from ARCH(5)
Figure 4: One-step ahead volatilities from tted ARCH(5) and GARCH(1,1) models
for Microsoft and S&P 500 index.
6.1 Asym metric Leverage Eects and News Impact
In the basic GARCH model (6), since only squared residuals
2
ti
enter the conditional
variance equation, the signs of the residuals or shocks have no eect on conditional
volatility. However, a stylized fact of nancial volatility is that bad news (negative
shocks) tends to have a larger impact on volatility than good news (positiv e shock s).
That is, volatilit y tends to be higher in a falling market than in a rising market. [10]
attributed this eect to the fact that bad news tends to drive down the stock price,
thus increasing the leverage (i.e., the debt-equity ratio) of the stock and causing the
stock to be more volatile. Based on this conjecture, the asymmetric news impact on
volatilit y is commonly referred to as the leverage eect.
6.1.1 Testing for Asymmetric Eects on Conditional Volatility
A simple diagnostic for uncovering possible asymmetric leverage eectsisthesample
correlation between r
2
t
and r
t1
. A negative value of this correlation provides some
evidence for potential leverage eects. Other simple diagnostics, suggested by [39],
16
Lag
ACF
0 102030
0.0 0.4 0.8
Microsoft Squared Residuals
Lag
ACF
0102030
0.0 0.4 0.8
S&P 500 Squared Residuals
Quantiles of Standard Normal
Microsoft Standardized Residuals
-2 0 2
-0.3 -0.1 0.1
Quantiles of Standard Normal
S&P 500 Standardized Residuals
-2 0 2
-0.20 -0.05 0.10
Figure 5: Graphical residual diagnostics from tted GARCH(1,1) models to Microsoft
and S&P 500 returns.
result from estimating the following test regression
ˆε
2
t
= β
0
+ β
1
ˆw
t1
+ ξ
t
,
where ˆε
t
is the estimated residual from the conditional mean equation (10), and ˆw
t1
is a variable constructed from ˆε
t1
and the sign of ˆε
t1
. Asignicant value of β
1
indicates evidence for asymmetric eects on conditional volatility. Let S
t1
denote
a dummy variable equal to unity when ˆε
t1
is negative, and zero otherwi se. Engle
and Ng consider three tests for asymmetry. Setting ˆw
t1
= S
t1
gives the Sign
Bias test; setting ˆw
t1
= S
t1
ˆε
t1
gives the Negative Size Bias test; and setting
ˆw
t1
= S
+
t1
ˆε
t1
gives the Positive Size Bias test.
6.1.2 Asymmetric GARCH Models
The leverage eect can be incorporated into a GARCH model in several ways. [71]
proposed the following exponen tial GARCH (EGARCH) model to allow for leverage
eects
h
t
= a
0
+
p
X
i=1
a
i
|
ti
| + γ
i
ti
σ
ti
+
q
X
j=1
b
j
h
tj
, (16)
17
where h
t
=logσ
2
t
.Notethatwhen
ti
is positive or there is “good news”, the
total eect of
ti
is (1 + γ
i
)|
ti
|; in contrast, when
ti
is negative or there is “bad
news”, the total eect of
ti
is (1 γ
i
)|
ti
|. Bad news can have a larger impact on
volatility, and the value of γ
i
would be expected to be negative. An advantage of the
EGARCH model over the basic GARCH model is that the conditional variance σ
2
t
is
guaranteed to be positive regardless of the values of the coecients in (16), because
the logarithm of σ
2
t
instead of σ
2
t
itself is modeled. Also, the EGARCH is covariance
stationary provided
P
q
j=1
b
j
< 1.
Another GARCH variant that is capable of modeling leverage eectsisthethresh-
old GARCH (TGARCH) model,
9
which has the following form
σ
2
t
= a
0
+
p
X
i=1
a
i
2
ti
+
p
X
i=1
γ
i
S
ti
2
ti
+
q
X
j=1
b
j
σ
2
tj
, (17)
where
S
ti
=
½
1if
ti
< 0
0if
ti
0
.
That is, depending on whether
ti
is above or below the threshold value of zero,
2
ti
has dierent eects on the conditional variance σ
2
t
:when
ti
is positive, the
total eects are given by a
i
2
ti
;when
ti
is negative, the total eects are given by
(a
i
+ γ
i
)
2
ti
. So one would expect γ
i
to be positive for bad news to have larger
impacts.
[31] extended the basic GARCH model to allow for leverage eects. Their power
GAR CH (PGARCH(p, d, q)) model has the form
σ
d
t
= a
0
+
p
X
i=1
a
i
(|
ti
| + γ
i
ti
)
d
+
q
X
j=1
b
j
σ
d
tj
, (18)
where d is a positiv e exponent, and γ
i
denotes the coecient of leverage eects.
When d =2, (18) reduces to the basic GARCH model with leverage eects. When
d =1, the PGARCH model is specied in terms of σ
t
which tends to be less sensitive
to outliers than when d =2. The exponent d may also be estimated as an additional
parameter which increases the exibility of the model. [31] sho wed that the PGARCH
model also includes man y other GARCH variants as special cases.
Many other asymmetric GARCH models have been proposed based on smooth
transition and Markov switching models. See [44] and [83] for excellent surveys of
these models.
6.1.3 News Impact Curve
The GAR CH, EGARC H, TGARCH and PGARCH models are all capable of modeling
leverage eect s. To clearly see the impact of leverage eects in these models, [75],
and [39] advocated the use of the so-called news impact curve. They dened the news
9
The original TGARCH mo del prop osed by [87] models σ
t
instead of σ
2
t
. The TGARCH model
is also known as the GJR m odel because [47] proposed essentially the same m odel.
18
GAR CH(1, 1) σ
2
t
= A + a
1
(|
t1
| + γ
1
t1
)
2
A = a
0
+ b
1
¯σ
2
¯σ
2
= a
0
/[1 a
1
(1 + γ
2
1
) b
1
]
TGARCH(1, 1) σ
2
t
= A +(a
1
+ γ
1
S
t1
)
2
t1
A = a
0
+ b
1
¯σ
2
¯σ
2
= a
0
/[1 (a
1
+ γ
1
/2) b
1
]
PGAR CH(1, 1, 1) σ
2
t
= A +2
Aa
1
(|
t1
| + γ
1
t1
)
+a
2
1
(|
t1
| + γ
1
t1
)
2
, A =(a
0
+ b
1
¯σ)
2
¯σ
2
= a
2
0
/[1 a
1
/
p
2 b
1
]
2
EGAR CH(1, 1) σ
2
t
= A exp{a
1
(|
t1
| + γ
1
t1
)/¯σ}
A σ
2b
1
exp{a
0
}
¯σ
2
=exp{(a
0
+ a
1
p
2)/(1 b
1
)}
Table 5: News impact curves for asymmetric GARCH processes. ¯σ
2
denotes the
unconditional variance.
Asset corr(r
2
t
,r
t1
) Sign Bias Negative Size Bias Positive Size Bias
Microsoft 0.0315
0.4417
(0.6587)
6.816
(0.000)
3.174
(0.001)
S&P 500 0.098
2.457
(0.014)
11.185
(0.000)
1.356
(0.175)
Notes: p-values are in parentheses.
Table 6: Tests for Asymmetric GARCH Eects.
impact curve as the functional relationship between conditional variance at time t
and the shock term (error term) at time t1, holding constant the information dated
t2 and earlier, and with all lagged conditional variance evaluated at the level of the
unconditional variance. Table 5 summarizes the expressions dening the news impact
curves, which include expressions for the unconditional variances, for the asymmetric
GARCH(1,1) models.
6.1.4 Asymmetric GARCH Models for Daily Returns
Table 6 shows diagnostics and tests for asymmetric eects in the daily returns on
Microsoft and the S&P 500. The correlation between r
2
t
and r
t1
is negative and
fairly small for both series indicating weak evidence for asymmetry. However, the
Size Bias tests clearly indicate asymmetric eects with the Negative Size Bias test
giving the most signicant results.
Table 7 gives the estimation results for EGAR CH(1,1), TGARCH(1,1) and PGARCH(1,d,1)
models for d =1, 2. All of the asymmetric models show statistically signicant lever-
19
Model a
0
a
1
b
1
γ
1
BIC
Microsoft
EGARCH
0.7273
[0.4064]
0.2144
[0.0594]
0.9247
[0.0489]
0.2417
[0.0758]
-20265
TGAR CH
3.01e
5
[1.02e
5
]
0.0564
[0.0141]
0.8581
[0.0342]
0.0771
[0.0306]
-20291
PGARCH 2
2.87e
5
[9.27e
6
]
0.0853
[0.0206]
0.8672
[0.0313]
0.2164
[0.0579]
-20290
PGARCH 1
0.0010
[0.0006]
0.0921
[0.0236]
0.8876
[0.0401]
0.2397
[0.0813]
-20268
S&P 500
EGARCH
0.2602
[0.3699]
0.0720
[0.0397]
0.9781
[0.0389]
0.3985
[0.4607]
-28051
TGAR CH
1.7e
6
[7.93e
7
]
0.0157
[0.0081]
0.9169
[0.0239]
0.1056
0.0357
-28200
PGARCH 2
1.78e
6
[8.74e
7
]
0.0578
[0.0165]
0.9138
[0.0253]
0.4783
[0.0910]
-28202
PGARCH 1
0.0002
[2.56e
6
]
0.0723
[0.0003]
0.9251
[8.26e
6
]
0.7290
[0.0020]
-28253
Notes: QML standard errors are in brac kets.
Table 7: Estimates of Asymmetric GARCH(1,1) Models.
age eects, and lower BIC values than the symmetric GARCH models. Model selec-
tion criteria indicate that the TGARCH(1,1) is the best tting model for Microsoft,
and the PGARCH(1,1,1) is the best tting model for the S&P 500.
Figure 6 shows the estimated news impact curves based on these models. In
this plot, the range of
t
is determined by the residuals from the tted models. The
TGARCH and PGARCH(1,2,1) models ha ve very similar NICs and show much larger
responses to negative shocks than to positive shocks. Since the EGARCH(1,1) and
PGARCH(1,1,1) models are more robust to extreme shocks, impacts of small (large)
shocks for these model are larger (smaller) compared to those from the other models
and the leverage eect is less pronounced.
6.2 Non-Gaussian Error Distributions
In all the examples illustrated so far, a normal error distribution has been exclusively
used. However, given the well known fat tails in nancial time series, it may be more
appropriate to use a distribution which has fatter tails than the normal distribution.
The most common fat-tailed error distributions for tting GARCH models are: the
Student’s t distribution; the double exponential distribution; and the generalized
error distribution.
[13] proposed tting a GARCH model with a Student’s t distribution for the
standardized residual. If a random variable u
t
has a Student’s t distribution with ν
degrees of freedom and a scale parameter s
t
, the probability density function (pdf)
20
-0.2 -0.1 0.0 0.1 0.2
0.001 0.004
Asymmetric GARCH(1,1) Models for Microsoft
TGARCH
PGARCH 1
PGARCH 2
EGARCH
-0.2 -0.1 0.0 0.1 0.2
0.0 0.002 0.005
Asymmetric GARCH(1,1) Models for S&P 500
TGARCH
PGARCH 1
PGARCH 2
EGARCH
Figure 6: News impact curves from tted asymmetric GARC H(1,1) models for Mi-
crosoft and S&P 500 index.
of u
t
is given by
f(u
t
)=
Γ[(ν +1)/2]
(πν)
1/2
Γ(ν/2)
s
1/2
t
[1 + u
2
t
/(s
t
ν)]
(ν+1)/2
,
where Γ(·) is the gamma function. The variance of u
t
is given by
var(u
t
)=
s
t
ν
ν 2
,v>2.
If the error term
t
in a GARCH model follows a Student’s t distribution with ν
degrees of freedom and var
t1
(
t
)=σ
2
t
, the scale parameter s
t
should be chosen to
be
s
t
=
σ
2
t
(ν 2)
ν
.
Thus the log-likelihood function of a GARCH model with Student’s t distributed
errors can be easily constructed based on the a bo ve pdf.
[71] proposed to use the generalized error distribution (GED) to capture the fat
tails usually observed in the distribution of nancial time series. If a random variable
21
u
t
has a GED with mean zero and unit variance, the pdf of u
t
is given by
f(u
t
)=
ν exp[(1/2)|u
t
|
ν
]
λ · 2
(ν+1)
Γ(1)
,
where
λ =
"
2
2
Γ(1)
Γ(3)
#
1/2
,
and ν is a positive parameter governing the thickness of the tail behavior of the
distribution. When ν =2the above pdf reduces to the standard normal pdf; when
ν<2, the density has thicker tails than the normal density; when ν>2, the density
has thinner tails than the normal density.
When the tail thickness parameter ν =1, the pdf of GED reduces to the pdf of
double exponential distribution:
f(u
t
)=
1
2
e
2|u
t
|
.
Based on the above pdf, the log-likelihood function of GARCH models with GED or
double exponential distributed errors can be easily constructed. See to [48] for an
example.
Several other non-Gaussian error distribution have been proposed. [42] in troduced
the asymmetric Studen t’s t distribution to capture both skewness and excess kurtosis
in the standardized residuals. [85] proposed the normal inverse Gaussian distribution.
[45] provided a very exible seminonparametric innovation distribution based on a
Hermite expansion of a Gaussian density. Their expansion is capable of capturing
general shape departures from Gaussian behavior in the standardized residuals of the
GARCH model.
6.2.1 Non-Gaussian GARCH Models for Daily Returns
Table 8 gives estimates of the GARCH(1,1) and best tting asymmetric GARCH(1,1)
models using Student’s t innovations for the Microsoft and S&P 500 returns. Model
selection criteria indicated that models using the Student’s t distribution t better
than the models using the GED distribution. The estimated degrees of freedom for
Microsoft is about 7, and for the S&P 500 about 6. The use of t-distributed errors
clearly improves the t of the GARCH(1,1) models. Indeed, the BIC values are even
lower than the values for the asymmetric GARCH(1,1) models based on Gaussian
errors (see Table 7). Ov erall, the asymmetric GAR CH(1,1) models with t-distributed
errors are the best tting models. The qq-plots in Figure 7 shows that the Student’s t
distribution adequately captures the fat-tailed behavio r in the standardized residuals
for Microsoft but not for the S&P 500 index.
7 Long Mem ory GARCH Models
If returns follow a GARCH(p, q) model, then the autocorrelations of the squared and
absolute returns should decay exponentially. However, the SACF of r
2
t
and |r
t
| for
22
Model a
0
a
1
b
1
γ
1
v BIC
Microsoft
GARCH
3.39e
5
[1.52e
5
]
0.0939
[0.0241]
0.8506
[0.0468]
6.856
[0.7121
-20504
TGAR CH
3.44e
5
[1.20e
5
]
0.0613
[0.0143]
0.8454
[0.0380]
0.0769
[0.0241]
7.070
[0.7023]
-20511
S&P 500
GARCH
5.41e
7
[2.15e
7
]
0.0540
[0.0095]
0.0943
[0.0097]
5.677
[0.5571]
-28463
PGAR CH
d =1
0.0001
[0.0002]
0.0624
[0.0459]
0.9408
[0.0564]
0.7035
[0.0793]
6.214
[0.6369]
-28540
Notes: QML standard errors are in brackets.
Table 8: Estimates of Non Gaussian GARCH(1,1) Models.
Microsoft and the S&P 500 in Figure 2 appear to decay much more slowly. This is
evidence of so-called long memory behavior. Formally, a stationary process has long
memory or long range dependence if its autocorrelation function behaves like
ρ(k) C
ρ
k
2d1
as k →∞,
where C
ρ
is a positive constant, and d is a real number between 0 and
1
2
. Thus the
autocorrelation function of a long memory process decays slowly at a hyperbolic rate.
In fact, it decays so slowly that the autocorrelations are not summable:
X
k=−∞
ρ(k)=.
It is important to note that the scaling property of the autocorrelation function does
not dictate the general behavior of the autocorrelation function. Instead, it only
species the asymptotic behavior when k →∞. What this means is that for a long
memory process, it is not necessary for the autocorrelation to remain signicant at
large lags as long as the autocorrelation function decays slowly. [8] gives an example
to illustrate this property.
The following subSections describe testing for long memory and GARCH models
that can capture long memory behavior in volatility. Explicit long memory GARCH
models are discussed in [83].
7.1 Testing for Long Mem ory
One of the best-known and easiest to use tests for long memory or long range de-
pendence is the rescaled range (R/S) statistic, which was originally proposed by [53],
and later rened by [67] and his coauthors. The R/S statistic is the range of partial
sums of deviations of a time series from its mean, rescaled by its standard deviation.
Specically, consider a time series y
t
,fort =1, ··· ,T. The R/S statistic is dened
23
-10
-5
0
5
-5 0 5
Microsoft
-5 0 5
S&P 500
Figure 7: QQ-plots of Standardized Residuals from Asymmetric GARCH(1,1) models
with Student’s t errors.
as
Q
T
=
1
s
T
max
1kT
k
X
j=1
(y
j
¯y) min
1kT
k
X
j=1
(y
j
¯y)
, (19)
where ¯y =1/T
P
T
i=1
y
i
and s
T
=
q
1/T
P
T
i=1
(y
i
¯y)
2
.Ify
t
is iid with nite variance,
then
1
T
Q
T
V,
where denotes weak convergence and V is the range of a Brownian bridge on the
unit interval. [62] gives selected quantiles of V .
[62] pointed out that the R/S statistic is not robust to short range dependence. In
particular, if y
t
is autocorrelated (has short memory) then the limiting distribution
of Q
T
/
T is V scaledbythesquarerootofthelongrunvarianceofy
t
. To allow for
short range dependence in y
t
, [62] modied the R/S statistic as follows
˜
Q
T
=
1
ˆσ
T
(q)
max
1kT
k
X
j=1
(y
j
¯y) min
1kT
k
X
j=1
(y
j
¯y)
, (20)
24
where the sample standard deviation is replaced by the square root of the Newey-
West ([73]) estimate of the long run variance with bandwidth q.
10
[62] showed that if
there is short memory but no long memory in y
t
,
˜
Q
T
also converges to V , the range
of a Brownian bridge. [18] found that (20) is eective for detecting long memory
behavior in asset return volatility.
7.2 Two Componen t G ARCH Model
In the covariance stationary GARCH model the conditional volatility will always
mean revert to its long run level unconditional value. Recall the mean reverting
form of the basic GARCH(1, 1) model in (11). In many empirical applications, the
estimated mean reverting rate ˆa
1
+
ˆ
b
1
isoftenverycloseto1. For example, the
estimated value of a
1
+ b
1
from the GAR CH(1,1) model for the S&P 500 index is
0.99 and the half life of a volatility shock implied by this mean reverting rate is
ln(0.5)/ ln(0.956) = 76.5 days. So the tted GARCH(1,1) model implies that the
conditional volatility is very persistent.
[37] suggested that the high persistence and long memory in volatility may be due
toatime-varyinglongrunvolatilitylevel. In particular, they suggested decomposing
conditional variance into two components
σ
2
t
= q
t
+ s
t
, (21)
where q
t
is a highly persistent long run component, and s
t
is a transitory short run
component. Long memory behavior can often be well approximated by a sum of two
such components. A general form of the two componen ts model that is based on a
modied version of the PGARCH(1,d,1) is
σ
d
t
= q
d
t
+ s
d
t
, (22)
q
d
t
= α
1
|
t1
|
d
+ β
1
q
d
t1
, (23)
s
d
t
= a
0
+ α
2
|
t1
|
d
+ β
2
s
d
t1
. (24)
Here, the long run component q
t
follows a highly persistent PGARCH(1,d,1) model
and the transitory componen t s
t
follow s another PGARCH(1 ,d,1) model. For the tw o
components to be separately identied the parameters should satisfy 1 < (α
1
+β
1
) <
(α
2
+ β
2
). It can be shown that the reduced form of the two components model is
σ
d
t
= a
0
+(α
1
+ α
2
)|
t1
|
d
(α
1
β
2
+ α
2
β
1
)|
t2
|
d
+(β
1
+ β
2
)σ
d
t1
β
1
β
2
σ
d
t2
,
whichisintheformofaconstrainedPGARCH(2,d,2) model. However, the two
components model is not fully equivalent to the PGARCH(2,d,2) model because not
all PGARCH(2,d,2) models have the component structure. Since the two compo-
nents model is a constrained version of the PGARCH(2,d,2) model, the estimation
of a two components model is often numerically more stable than the estimation of
an unconstrained PGARCH(2,d,2) model.
10
The long-run variance is the asymptotic variance of
T y μ).
25
Asset
˜
Q
T
r
2
t
|r
t
|
Microsoft 2.3916 3.4557
S&P 500 2.3982 5.1232
Table 9: Modied R/S Tests for Long Memory.
a
0
α
1
β
1
α
2
β
2
v BIC
Microsoft
2.86e
6
[1.65e
6
]
0.0182
[0.0102]
0.9494
[0.0188]
0.0985
[0.0344]
0.7025
[0.2017]
-20262
1.75e
6
5.11e
7
0.0121
[0.0039]
0.9624
[0.0098]
0.0963
[0.0172]
0.7416
[0.0526]
6.924
[0.6975]
-20501
S&P 500
3.2e
8
[1.14e
8
]
0.0059
[0.0013]
0.9848
[0.0000]
0.1014
[0.0221]
0.8076
[0.0001]
28113
1.06e
8
[1.26e
8
]
0.0055
[0.0060]
0.9846
[0.0106]
0.0599
[0.0109]
0.8987
[0.0375]
5.787
[0.5329]
28457
Notes: QML standard errors are in brackets.
Table 10: Estimates of Two Component GARCH(1,1) Models.
7.3 Integrated GARCH Model
The high persistence often observed in tted GARCH(1,1) models suggests that
volatility might be nonstationary implying that a
1
+ b
1
=1, in which case the
GARCH(1,1) model becomes the integrated GARCH(1,1) or IGARCH(1,1) model.
In the IGARCH(1,1) model the unconditional variance is not nite and so the model
does not exhibit volatility mean reversion. However, it can be shown that the model is
strictly stationary pro v ided E[ln(a
1
z
2
t
+b
1
)] < 0. If the IGARCH(1,1) model is strictly
stationary then the parameters of the model can still be consistently estimated by
MLE.
[27] argued against the IGARCH specication for modeling highly persistent
volatility processes for two reasons. First, they argue that the observed convergence
toward normality of aggregated returns is inconsisten t with the IGARCH model. Sec-
ond, they argue that observed IGARCH behavior may result from misspecication
of the conditional variance function. For example, a two components structure or
ignored structural breaks in the unconditional variance ([58] and [70]) can result in
IGARCH behavior.
7.4 Long Memory GA RCH M odels for Daily Returns
Table 9 gives Lo’s modied R/S statistic (20) applied to r
2
t
and |r
t
| for Microsoft
and the S&P 500. The 1% right tailed critical value for the test is 2.098 ([62] Table
5.2) and so the modied R/S statistics are signicant at the 1% level for both series
26
providing evidence for long memory behavior in volatility.
Table 10 shows estimates of the two component GARCH(1,1) with d =2, using
Gaussian and Student’s t errors, for the daily returns on Microsoft and the S&P 500.
Notice that the BIC values are smaller than the BIC values for the unconstrained
GARCH(2,2) models given in Table 3, which conrms the better n u merical stability
of the two component model. For both series, the two components are present and
satisfy 1 < (α
1
+ β
1
) < (α
2
+ β
2
). For Microsoft, the half-lives of the two components
from the Gaussian (Student’s t) models are 21 (26.8) days and 3.1 (3.9) da ys, re-
spectively. For the S&P 500, the half-lives of the two components from the Gaussian
(Student’s t) models are 75 (69.9) days and 7.3 (16.4) days, respectively.
8 GA RCH Model Prediction
An important task of modeling conditional volatility is to generate accurate forecasts
for both the future value of a nancial time series as well as its conditional volatility.
Volatilit y forecasts are used for risk management, option pricing, portfolio allocation,
trading strategies and model evaluation. Since the conditional mean of the general
GARCH model (10) assumes a traditional ARMA form, forecasts of future values
of the underlying time series can be obtained following the traditional approach for
ARMA prediction. However, by also allowing for a time varying conditional variance,
GARCH models can generate accurate forecasts of future volatility, especially over
short horizons. This Section illustrates how to forecast v olatility using GARC H
models.
8.1 GA RCH and Forecasts for the Conditional M ean
Suppose one is interested in forecasting future values of y
T
in the standard GARCH
model described by (2), (3) and (6). For simplicity assume that E
T
[y
T +1
]=c. Then
the minim um mean squared error h step ahead forecast of y
T +h
is just c, which
does not depend on the GARCH paramet ers, and the corresponding forecast error is
T +h
= y
T +h
E
T
[y
T +h
].
The conditional variance of this forecast error is then
var
T
(
T +h
)=E
T
[σ
2
T +h
],
which does depend on the GARCH parameters. Therefore, in order to produce
condence bands for the hstep ahead forecast the hstep ahead volatility forecast
E
T
[σ
2
T +h
] is needed.
8.2 Forecasts from the GARCH(1,1) M odel
For simplicity, consider the basic GARCH(1, 1) model (7) where
t
= z
t
σ
t
such that
z
t
iid (0, 1) and has a symmetric distribution. Assume the model is to be estimated
overthetimeperiodt =1, 2, ··· ,T. The optimal, in terms of mean-squared error,
27
forecast of σ
2
T +k
given information at time T is E
T
[σ
2
T +k
] and can be computed using
a simple recursion. For k =1,
E
T
[σ
2
T +1
]=a
0
+ a
1
E
T
[
2
T
]+b
1
E
T
[σ
2
T
] (25)
= a
0
+ a
1
2
T
+ b
1
σ
2
T
,
where it assumed that
2
T
and σ
2
T
are known
11
. Similarly, for k =2
E
T
[σ
2
T +2
]=a
0
+ a
1
E
T
[
2
T +1
]+b
1
E
T
[σ
2
T +1
]
= a
0
+(a
1
+ b
1
)E
T
[σ
2
T +1
].
since E
T
[
2
T +1
]=E
T
[z
2
T +1
σ
2
T +1
]=E
T
[σ
2
T +1
]. In general, for k 2
E
T
[σ
2
T +k
]=a
0
+(a
1
+ b
1
)E
T
[σ
2
T +k1
]
= a
0
k1
X
i=0
(a
1
+ b
1
)
i
+(a
1
+ b
1
)
k1
(a
1
2
T
+ b
1
σ
2
T
). (26)
An alternative representation of the forecasting equation (26) starts with the mean-
adjusted form
σ
2
T +1
¯σ
2
= a
1
(
2
T
¯σ
2
)+b
1
(σ
2
T
¯σ
2
),
where ¯σ
2
= a
0
/(1 a
1
b
1
) is the unconditional variance. Then by recursive substi-
tution
E
T
[σ
2
T +k
] ¯σ
2
=(a
1
+ b
1
)
k1
(E[σ
2
T +1
] ¯σ
2
). (27)
Notice that as k →∞, the volatility forecast in (26) approaches ¯σ
2
if the GARCH
process is covariance stationary and the speed at which the forecasts approaches ¯σ
2
is captured by a
1
+ b
1
.
The forecasting algorithm (26) produces forecasts for the conditional variance
σ
2
T +k
. The forecast for the conditional volatility, σ
T +k
, is usually dened as the
square root of the forecast for σ
2
T +k
.
The GARCH(1,1) forecasting algorithm (25) is closely related to an exponentially
weighted mo ving average (EWMA) of past values of
2
t
. This typ e of forecast is
commonly used by RiskMetrics ([54]). The EWMA forecast of σ
2
T +1
has the form
σ
2
T +1,EW MA
=(1 λ)
X
s=0
λ
s
2
ts
(28)
for λ (0, 1). In (28), the w eights sum to one, the rstweightis1λ, and the remain-
ing weights decline exponen tially. To relate the EWMA forecast to the GARCH(1,1)
formula (25), (28) may be re-expressed as
σ
2
T +1,EW MA
=(1 λ)
2
T
+ λσ
2
T,EWMA
=
2
T
+ λ(σ
2
T,EWMA
2
T
),
11
In practice, a
0
,a
1
,b
1
,
T
and σ
2
T
are t he tted values computed from the estimated GARCH(1,1)
mo de l instead of the unobserved “true” values.
28
whic h is of the form (25) with a
0
=0,a
1
=1 λ and b
1
= λ. Therefore, the
EWMA forecast is equivalen t to the forecast from a restricted IGARCH(1,1) model.
It follows that for any h>0
2
T +h,EW MA
= σ
2
T,EWMA
. As a result, unlike the
GARCH(1,1) forecast, the EWMA forecast does not exhibit mean reversion to a
long-run unconditional variance.
8.3 Forecasts from Asym metric GA RCH(1,1) M odels
To illustrate the asymmetric eects of leverage on forecasting, consider the TGARCH(1,1)
model (17) at time T
σ
2
T
= a
0
+ a
1
2
T 1
+ γ
1
S
T 1
2
T 1
+ b
1
σ
2
T 1
.
Assume that
t
has a symmetric distribution about zero. The forecast for T +1 based
on information at time T is
E
T
[σ
2
T +1
]=a
0
+ a
1
2
T
+ γ
1
S
T
2
T
+ b
1
σ
2
T
,
whereitassumedthat
2
T
,S
T
and σ
2
T
are known. Hence, the TGARCH(1,1) forecast
for T +1will be dierent than the GARCH(1,1) forecast if S
T
=1(
T
< 0). The
forecast at T +2is
E
T
[σ
2
T +2
]=a
0
+ a
1
E
T
[
2
T +1
]+γ
1
E
T
[S
T +1
2
T +1
]+b
1
E
T
[σ
2
T +1
]
= a
0
+
³
γ
1
2
+ a
1
+ b
1
´
E
T
[σ
2
T +1
],
which follows since E
T
[S
T +1
2
T +1
]=E
T
[S
T +1
]E
T
[
2
T +1
]=
1
2
E
T
[σ
2
T +1
]. Notice that the
asymmetric impact of leverage is present even if S
T
=0. By recursive substitution
for the forecast at T + h is
E
T
[σ
2
T +h
]=a
0
+
³
γ
1
2
+ a
1
+ b
1
´
h1
E
T
[σ
2
T +1
], (29)
which is similar to the GARCH(1,1) forecast (26). The mean reverting form (29) is
E
T
[σ
2
T +h
] ¯σ
2
=
³
γ
1
2
+ a
1
+ b
1
´
h1
¡
E
T
[σ
2
T +h
] ¯σ
2
¢
where ¯σ
2
= a
0
/(1
γ
1
2
a
1
b
1
) is the long run variance.
Forecasting algorithms for σ
d
T +h
in the PGARCH(1,d,1)andforln σ
2
T +h
in the
EGARCH(1,1) follow in a similar manner and the reader is referred to [31], and [71]
for further details.
8.4 Simulation-Based Forecas ts
The forecasted volatility can be used together with forecasted series values to generate
condence intervals of the forecasted series values. In many cases, the forecasted
volatility is of central interest, and condence intervals for the forecasted volatility can
be obtained as we ll. However, analytic formulas for condence intervals of forecasted
29
volatility are only known for some special cases (see [6]). In models for which analytic
formulas for condence intervals are not known, a simulation-based method can be
used to obtain condence intervals for forecasted volatility from any GARCH that
can be simulated. To obtain volatility forecasts from a tted GARCH model, simply
simulate σ
2
T +k
from the last observation of the tted model. This process can be
repeated many times to obtain an “ensemble” of volatility forecasts. The point
forecast of σ
2
T +k
may then be computed by averaging over the simulations, and a
95% condence interval may be computed using the 2.5% and 97.5% quantiles of the
simulation distribution, respectively.
8.5 Foreca stin g the Volatilit y of Mu ltiperiod R e tur ns
In many situations, a GARCH model is t to daily continuously compounded returns
r
t
=ln(P
t
) ln(P
t1
), where P
t
denotes the closing price on da y t. The resulting
GARCH forecasts are for daily volatility at dierent horizons. For risk management
and option pricing with stochastic volatility, volatility forecasts are needed for multi-
period returns. With continuously compounded returns, the hday return between
days T and T + h is simply the sum of h single da y returns
r
T +h
(h)=
h
X
j=1
r
T +j
.
Assuming returns are uncorrelated, the conditional variance of the hperiod return
is then
var
T
(r
T +h
(h)) = σ
2
T
(h)=
h
X
j=1
var
T
(r
T +j
)=E
T
[σ
2
T +1
]+···+ E
T
[σ
2
T +h
]. (30)
If returns have constant variance ¯σ
2
, then σ
2
T
(h)=h¯σ
2
and σ
T
(h)=
h¯σ. This
is known as the “square root of time” rule as the hda y volatility scales with
h. In
this case, the hday variance per day, σ
2
T
(h)/h, is constant. If returns are described
by a GARCH model then the square root of time rule does not necessarily apply. To
see this, suppose returns follow a GARCH(1,1) model. Plugging the GARCH(1,1)
model forecasts (27) for E
T
[σ
2
T +1
],...,E
T
[σ
2
T +h
] into (30) gives
σ
2
T
(h)=h¯σ
2
+(E[σ
2
T +1
] ¯σ
2
)
1 (a
1
+ b
1
)
h
1 (a
1
+ b
1
)
¸
For the GARCH(1,1) process the square root of time rule only holds if E[σ
2
T +1
]=¯σ
2
.
Whether σ
2
T
(h) is larger or smaller than h¯σ
2
depends on whether E[σ
2
T +1
] is larger
or smaller than ¯σ
2
.
8.6 Evaluating Volatilit y Pred ic tions
GARCH models are often judged b y their out-of-sample forecasting ability, see [22]
for an overview. This forecasting ability can be measured using traditional forecast
30
error metrics as well as with specic economic considerations suc h as value-at-risk
violations, option pricing accuracy, or portfolio performance. Out-of-sample forecasts
for use in model comparison are typically computed using one of two methods. The
rst method produces recursive forecasts. An initial sample using data from t =
1,...,T is used to estimate the models, and hstep ahead out-of-sample forecasts are
produced starting at time T. Then the sample is increased by one, the models are re-
estimated, and hstep ahead forecasts are produced starting at T +1. This process is
repeated until no more hstep ahead forecasts can be computed. The second method
produces rolling forecasts. An initial sample using data from t =1,...,T is used to
determine a window width T, to estimate the models, and to form hstep ahead out-
of-sample forecasts starting at time T. Th en the window is moved ahead one time
period, the models are re-estimated using data from t =2,...,T +1, and hstep
ahead out-of-sample forecasts are produced starting at time T +1. This process is
repeated until no more hstep ahead forecasts can be computed.
8.6.1 Traditional Forecast Evaluation Statistics
Let E
i,T
[σ
2
T +h
] denote the hstep ahead forecast of σ
2
T +h
at time T from GARCH
model i using either recursive or rolling methods. Dene the corresponding forecast
error as e
i,T +h|T
= E
i,T
[σ
2
T +h
] σ
2
T +h
. Common forecast evaluation statistics based
on N out-of-sample forecasts from T = T +1,...,T + N are
MSE
i
=
1
N
T +N
X
j=T +1
e
2
i,j+h|j
,
MAE
i
=
1
N
T +N
X
j=T +1
¯
¯
e
i,j+h|j
¯
¯
,
MAPE
i
=
1
N
T +N
X
j=T +1
¯
¯
e
i,j+h|j
¯
¯
σ
j+h
.
The m odel which produces the smallest values of the forecast evaluation statistics is
judged to be the best model. Of course, the forecast evaluation statistics are random
variables and a formal statistical procedure should be used to determine if one model
exhibits superior predictive performance.
[28] proposed a simple procedure to test the null hypothesis that one model has
superior predictive performance over another model based on traditional forecast
evaluation statistics. Let {e
1,j+h|j
}
T +N
T +1
, and {e
2,j+h|j
}
T +N
T +1
denote forecast errors
from two dierent GARCH models. The accuracy of each forecast is measured by
a particular loss function L(e
i,T +h|T
),i=1, 2. Common c hoices are the squared
error loss function L(e
i,T +h|T
)=
¡
e
i,T +h|T
¢
2
and the absolute error loss function
L(e
i,T +h|T
)=
¯
¯
e
i,T +h|T
¯
¯
. The Diebold-Mariano (D M) test is based on the loss dier-
ential
d
T +h
= L(e
1,T+h|T
) L(e
2,T+h|T
).
31
The null of equal predictive accuracy is H
0
: E[d
T +h
]=0.The DM test statistic is
S=
¯
d
¡
dav ar(
¯
d)
¢
1/2
, (31)
where
¯
d = N
1
P
T +N
j=T +1
d
j+h
, and davar(
¯
d) is a consistent estimate of the asymptotic
variance of
N
¯
d. [28] recommend using the Newey-West estimate for davar(
¯
d) because
thesampleoflossdieren tials {d
j+h
}
T +N
T +1
are serially correlated for h>1. Under the
null of equal predictive accuracy, S has an asymptotic standard normal distribution.
Hence, the DM statistic can be used to test if a given forecast evaluation statistic (e.g.
MSE
1
) for one model is statistically dierent from the forecast evaluation statistic
for another model (e.g. MSE
2
).
Forecasts are also often judged using the forecasting regression
σ
2
T +h
= α + βE
i,T
[σ
2
T +h
]+e
i,T +h
. (32)
Unbiased forecasts have α =0and β =1, and accurate forecasts have high regression
R
2
values. In practice, the forecasting regression suers from an errors-in-variables
problem when estimated GARCH parameters are used to form E
i,T
[σ
2
T +h
] and this
creates a downward bias in the estimate of β. As a result, attention is more often
focused on the R
2
from (32).
An important practical problem with applying forecast evaluations to volatility
models is that the hstep ahead volatility σ
2
T +h
is not directly observable. Typ-
ically,
2
T +h
(or just the squared return) is used to proxy σ
2
T +h
since E
T
[
2
T +h
]=
E
T
[z
2
T +h
σ
2
T +h
]=E
T
[σ
2
T +h
]. However,
2
T +h
is a very noisy proxy for σ
2
T +h
since
var(
2
T +h
)=E[σ
4
T +h
](κ 1), where κ is the fourth moment of z
t
, and this causes
problems for the in terpretation of the forecast evaluation metrics.
Many empirical papers have evaluated the forecasting accuracy of competing
GARCH models using
2
T +h
as a proxy for σ
2
T +h
. [77] gave a comprehensive sur-
vey. The typical ndings are that the forecasting evaluation statistics tend to be
large, the forecasting regressions tend to be slightly biased, and the regression R
2
values tend to be very low (typically below 0.1). In general, asymmetric GARCH
models tend to hav e the lowest forecast evaluation statistics. The overall conclusion,
however, is that GAR CH models do not forecast very well.
[2] provided an explanation for the apparent poor forecasting performance of
GARCH models when
2
T +h
is used as a proxy for σ
2
T +h
in (32). For the GARCH(1,1)
model in which z
t
has nite kurtosis κ, they showed that the population R
2
value in
(32) with h =1is equal to
R
2
=
a
2
1
1 b
2
1
2a
1
b
1
,
and is bounded from above by 1/κ. Assuming z
t
N(0, 1), this upper bound is 1/3.
With a fat-tailed distribution for z
t
the upper bound is smaller. Hence, very low R
2
values are to be expected even if the true model is a GARCH(1,1). Moreover, [49]
found that the substitution of
2
T +h
for σ
2
T +h
in the evaluation of GARCH models
using the DM statistic (31) can result in inferior models being chosen as the best
32
Error pdf GAR CH TGARCH PGARC H
MSFT
Gaussian
Student’s t
0.0253
0.0247
0.0257
0.0253
0.0256
0.0250
S&P 500
Gaussian
Student’s t
0.0138
0.0138
0.0122
0.0128
0.0108
0.0111
Table 11: Unconditional Volatilities from Estimated GARCH(1,1) Models.
with probability one. These results indicate that extreme care must be used when
interpreting forecast evaluation statistics and tests based on
2
T +h
.
If high frequency intraday data are available, then instead of using
2
T +h
to proxy
σ
2
T +h
[2] suggested using the so-called realized variance
RV
m
t+h
=
m
X
j=1
r
2
t+h,j
,
where {r
T +h,1
,...,r
T +h,m
} denote the squared intraday returns at sampling fre-
quency 1/m for day T + h. For example, if prices are sampled every 5 minutes and
trading takes place 24 hours per day then there are m =2885-minu te intervals per
trading day. Under certain conditions (see [4]), RV
m
t+h
is a consistent estimate of
σ
2
T +h
as m →∞. As a result, RV
m
t+h
is a much less noisy estimate of σ
2
T +h
than
2
T +h
and so forecast evaluations based on RV
m
t+h
are expected to be much more ac-
curate than those based on
2
T +h
. For example, in evaluating GARCH(1,1) forecasts
for the Deutschemark-US daily exchange rate, [2] reported R
2
values from (32) of
0.047, 0.331 and 0.479 using
2
T +1
,RV
24
T +1
and RV
288
T +1
, respectively.
8.7 Forec asting the Volatility of Micr os o ft an d the S&P 500
Figure 8 shows hday ahead volatility predictions (h =1,...,250) from the tted
GARCH(1,1) models with normal errors for the daily returns on Microsoft and the
S&P 500. The horizontal line in the gures represents the estimated unconditional
standard deviation from the tted models. At the beginning of the forecast period,
ˆσ
T
<
b
¯σ for both series and so the forecasts revert upward toward the unconditional
volatility. The speed of volatility mean reversion is clearly shown by the forecast
proles. The forecasts for Microsoft revert to the unconditional level after about four
months, whereas the forecasts for the S&P 500 take over one year.
Figure 8 shows the volatility forecasts from the asymmetric and long memory
GARCH(1,1) models, and Table 11 gives the unconditional volatility from the esti-
mated models. For Microsoft, the forecasts and unconditional volatilities from the
dierent models are similar. For the S&P 500, in contrast, the forecasts and uncon-
ditional v olatilities dier considerably across the models.
33
Predicted Volatility from GARCH(1,1) for Microsoft
Mar Apr May Jun Jul Aug Sep Oct Nov Dec Jan Feb Mar Apr May Jun
2003 2004
0.018 0.022 0.026
Predicted Volatility from GARCH(1,1) for S&P 500
Mar Apr May Jun Jul Aug Sep Oct Nov Dec Jan Feb Mar Apr May Jun
2003 2004
0.010 0.014 0.018
Figure 8: Predicted Volatility from GARCH(1,1) Models
9FinalRemarks
This paper surveyed some of the practical issues associated with estimating univariate
GARCH models and forecasting volatility. Some practical issues associated with the
estimation of multivariate GARCH models and forecasting of conditional covariances
are given in [80].
References
[1] Alexander, C. (2001). Market Models: A Guide to Financial Data Analysis.John
Wiley & Sons, Chichester, UK.
[2] Andersen, T., and Bollerslev, T. (1998). Answering the Skeptics: Yes, Stan-
dard Volatility Models Do Provide Accurate Forecasts. International Economic
Review, 39(4), 885-905.
[3] Andersen, T., and Bollerslev, T., Christoersen, P.K., and Diebold, F.X. (2006).
Volatility Forecasting. In G. Elliott, C.W.J.Granger, and A. Timmermann, edi-
tors, Handbook of Economic Forecasting,Amsterdam,NorthHolland.
34
Jul Sep Nov Jan Mar May
2003 2004
0.019 0.023
GARCH
PGARCH(2)
TGARCH
Microsoft: Asymmetric Models
Jul Sep Nov Jan Mar May
2003 2004
0.019 0.023
GARCH
2 COMP
Microsoft: Long Memory Models
Jul Sep Nov Jan Mar May
2003 2004
0.0105 0.0125
GARCH
PGARCH(1)
TGARCH
S&P 500: Asymmetric Models
Jul Sep Nov Jan Mar May
2003 2004
0.0105 0.0125
GARCH
2 COMP
S&P 500: Long Memory Models
Figure 9: Predicted Volatility from Competing GARCH Models.
[4] Andersen, T, Bollerslev, T., Diebold, F.X., and Labys, P. (2003). The Distribu-
tion of Exchange Rate Volatility. Journal of the A merican Statistical Association,
96, 42-55.
[5] Andreou, E. and Ghysels, E. (2002). Rolling Volatility Estimators: Some new
Theoretical, Simulation and Empirical Results. Journal of Business and Ec o-
nomic Statistics, 20(3), 363-376.
[6] Baillie, R.T., and Bollerslev, T. (1992). Prediction in Dynamic Models with Time
Dependent Conditional Variances. Journal of Econometrics, 52, 91-113.
[7] Baillie, R. T., Bollerslev, T., and Mikkelsen, H. O. (1996). Fractionally Inte-
grated Generalized Autoregressiv e Conditional Heteroskedasticity. Journal of
Econometrics, 74, 3-30.
[8] Beran, J. (1994). Statistics for Long Memory Processes. Chapman and Hall, New
York.
[9] Bera, A.K. and Higgins, M.L. (1995). On ARCH Models: Properties, Estimation
and Testing. Journal of Economic Surveys, 7, 305-362.
35
[10] Black, F. (1976). Studies in Stock Price Volatility Changes. Proceedings of the
1976 Business Meeting of the Business and Economics Statistics Section,Amer-
ican Statistical Association, 177-181.
[11] Blair, B., Poon, S.-H., and Taylor, S.J. (2001). Forecasting S&P 100 Volatility:
The Incremental Information Con tent of Implied Volatilities and High Frequency
Index Returns. Journal of Econometrics, 105, 5-26.
[12] Bollerslev, T. (1986). Generalized Autoregressive Conditional Heteroskedastic-
it y. Journal of Econometrics, 31, 307-327.
[13] Bollerslev, T. (1987). A Conditional Heteroskedastic Time Series Model for Spec-
ulative Prices Rates of Return. Review of Economics and Statistics, 69, 542-547.
[14] Bollerslev, T., Engle, R. F., and Nelson, D. B. (1994). ARCH Models. In R. F.
Engle and D. L. McFadden, editors, Handbook of Econometrics,Vol.4,Elsevier
Science B. V., Amsterdam.
[15] Bollerslev, T., and Ghysels, E. (1996). Periodic Autoregressive Conditional Het-
eroscedasticity. Journal of Business and Economic Statistics, 14, 139-151
[16] Bollerslev, T., and Wooldridge, T. (1992). Quasi-Maximum Likelihood Estima-
tion and Inference in Dynamic Models with Time-varying Covariances. Econo-
metric Reviews, 11, 143-172.
[17] Bomm, A.N. (2003). Pre-Announcemen t Eects, News Eects, and Volatility:
Monetary Policy and the Stock Market. Journal of Banking and Finance, 27,
133-151.
[18] Breidt, F.J., Crato, N. and de Lima, P. (1998). The Detection and Estimation
of Long Memory in Stochastic Volatility. Journal of Econometrics, 73, 325-348.
[19] Brooks, C. (1997). GARCH Modeling in Finance: A Review of the Software
Options. Economic Journal, 107(443), 1271-1276.
[20] Brooks, C., Burke, S., and Pesand, G. (2001). Benchmarks and the Accuracy of
GARCH Model Estimation. International Journal of Forecasting, 17, 45-56.
[21] Chen, F., Yu, W.-C., and Zivot, E. (2008). Predicting Stock Volatility Using
After-Hours Information. Unpublished manuscript, Department of Economics,
University of Washington.
[22] Clements, M.P. (2005). Evaluating Econometric Forecasts of Economic and Fi-
nancial Variables. Palgrave Texts in Econometrics, Palgrave Macmillan, Hound-
mills, UK.
[23] Conrad, C., and Haag, B.R. (2006). Inequalit y Constraints in the Fractionally
Integrated GARCH Model. Journal of Financial Econometrics, 4(3), 413-449.
36
[24] Davis, R., and T. Mikosch (2008). Extreme Value Theory for GARCH Processes.
In T.G. Andersen, R.A. Davis, J-P Kreiss, and T. Mikosch, editors, Handbook
of Financial Time Series, New York, Springer.
[25] de Lima, P.J.F. (1996). Nuisance Parameter Free Properties of Correlation In-
tegral Based Statistics. Econometric Reviews, 15, 237-259.
[26] Diebold, F.X. (1988). Empirical Modeling of Exchange Rate Behavior. Springer-
Verlag, New York.
[27] Diebold, F.X. and Lopez, J.A. (1996). Modeling Volatilit y Dynamics. In
K. Hoover, editor, Macroeconomics: Developments, Tensions and Prospects.
Klu wer, Boston.
[28] Diebold, F.X. and R.S. Mariano (1995). Comparing Predictive Accuracy, Journal
of Business and Economic Statistics. 13, 253-263.
[29] Diebold, F.X., and Sc huermann, T. (1993). Exact Maximum Likelihood Esti-
mation of ARCH Models. Unpublished manuscript, Department of Economics,
University of Pennsylvania.
[30] Ding, Z., and Granger, C. W. J. (1996). Modeling Volatility Persistence of Spec-
ulative Returns: A New Approach. Journal of Econometrics, 73, 185-215.
[31] Ding, Z., Granger, C. W. J., and Engle, R. F. (1993). A Long Memory Property
of Stock Market Returns and a New Model. Journal of Empirical Finance,1,
83-106.
[32] Drost, F.C. and Nijman, T.E. (1993). Temporal Aggregation of GARCH
Processes. Econometrica, 61, 909-927.
[33] Engle, R.F. (1982). Autoregressive Conditional Heteroskedasticity with Esti-
mates of the Variance of United Kingdom Ination. Econometrica, 50, 987-1007.
[34] Engle, R.F. (2001). GARCH 101: The Use of ARCH/GARCH Models in Applied
Economics. Journal of Economic Perspectives, 15, 157-168.
[35] Engle, R.F. and González-Rivera, G. (1991). Semiparametric ARCH models.
Journal of Business and Economic Statistics, 9, 345-60.
[36] Engle, R.F. and Kroner, K. (1995). Multivariate Simultaneous Generalised
AR CH. Econometric Theory 11, 122-150.
[37] Engle, R. F., and Lee, G. J. (1999). A Long-Run and Short-Run Component
ModelofStockReturnVolatility.InR.F.EngleandH.White,editors,Cointe-
gration, Causality, and Forecasting.OxfordUniversityPress,Oxford.
[38] Engle, R.F., and Mezrich, J. (1995). Grappling with GARCH. RISK, 8(9), 112-
117.
37
[39] Engle, R.F., and Ng, V. (1993). Measuring and Testing the Impact of News on
Volatility. Journal of Finance, 48, 1749-78.
[40] Engle, R.F., and Patton, A. (2001). What Good is a Volatility Model?, Quanti-
tative Finance, 1, 237-245.
[41] Fiorentini, G., Calzolari, G., and Panattoni, L. (1996) Analytic Derivatives and
the Computation of GARCH Estimates. Journal of Applied Econometrics 11,
399-417.
[42] Fernandez, C., and Steel, M. (1998). On Bayesian Modeling of Fat Tails and
Skewness. Journal of the American Statistical Association, 93, 359 -371.
[43] Flannery, M. and Protopapadakis, A. (2002). Macroeconomic Factors Do Inu-
ence Aggregate Stock Returns. The Review of Financial Studies, 15, 751-782.
[44] Franses, P.H. and D. van Dijk (2000). Non-Linear Time Series Models in Em-
pirical Finance. Cambridge University Press, Cambridge.
[45] Gallant, A.R. and Tauchen, G. (2001). SNP: A Program for Nonparametric Time
Series Analysis, Version 8.8, User’s Guide. Unpublished manuscrip, University
of North Carolina at Chapel Hill.
[46] Gallo, G.M., and Pacini, B. (1997). Early News is Good News: The Eect
of Market Opening on Market Volatility. Studies in Nonlinear Dynamics and
Econometrics, 2(4), 115-131.
[47] Glosten, L. R., Jagannathan, R., and Runkle, D. E. (1993). On the Relation
Between the Expected Value and the Volatilit y of the Nominal Excess Return
on Stocks. Journal of Finance, 48(5), 1779-1801.
[48] Hamilton, J. D. (1994). Time Series Analysis. Princeton University Press,
Princeton, NJ.
[49] Hansen, P., and Lunde, A. (2004). A Forecast Comparison of Volatility Models:
Does Anything Beat a GARCH(1,1) Model? Journal of Applied Econometrics,
20, 873-88 9.
[50] Hansen, P., and Lunde, A. (2006). Consistent Ranking of Volatility Models.
Journal of Econometrics, 131, 97-121.
[51] He, C., Teräsvirta, T. (1999a). Properties of Moments of a Family of GARCH
Processes. Journal of Econometrics, 92, 173-192.
[52] He, C., Teräsvirta, T. (1999b). Fourth Moment Structure of the GARCH(p, q)
Process. Econometric Theory, 15, 824-846.
[53] Hurst, H. E. (1951). Long Term Storage Capacity of Reservoirs. Transactions of
the American Society of Civil Engineers, 116, 770-799.
38
[54] J.P. Morgan (1997). RiskMetrics, Technical Documents, 4th Edition.NewYork.
[55] Jensen, T., and Rah bek, A. (2004). Asymptotic Normalit y of the QML Estimator
of ARCH in the Nonstationary Case. Econometrica, 72(2), 641-646.
[56] Kristensen, D., and Rahbek, A. (2005). Asymptotics of the QMLE for a Class
of ARCH(q) Models. Econometric Theory, 21, 946-961.
[57] Laurent, S., and Peters, J.-P. (2002). G@RCH 2.0: An Ox Package for Estimat-
ing and Forecasting Various ARCH Models. Journal of Economic Surveys 16,
pp.447-485.
[58] Lamoureux, C.G., and Lastrapes, W.D. (1990a). Heteroskedasticity in Stock
Return Data: Volume versus GARCH Eects. The Journal of Finance, 45, 221-
229.
[59] Lamoureux, C.G., and Lastrapes, W.D. (1990b). Persistence in Variance, Struc-
tural Change and the GARCH Model. Journal of Business and Economic Sta-
tistics, 8, 225-234.
[60] Lee, and Hansen, B.E. (1993). Asymptotic Theory for the GARCH(1,1) Quasi-
Maximum Likelihood Estimator. Econometric The ory, 10, 29-52.
[61] Lee, J.H.H., and King, M.L. (1993). A Locally Most Powerful Based Score Test
for ARCH and GARCH Regression Disturbances. Journal of Business and Eco-
nomic Statistics, 7, 259-279.
[62] Lo, A. W . (1991). Long Term Memory in Stock Market Prices. Econometrica,
59, 1279-1 313.
[63] Lumsdaine, R. L, (1992). Consistency and Asymptotic Normality of the Quasi-
Maximum Likelihood Estimator in IGARCH(1,1) and Covariance Stationary
GARCH(1,1) Models. Econometrica, 64, 575-596.
[64] Lumsdaine, R.L., and Ng, S. (1999). Testing for ARCH in the Presence of a
Possibly Misspecied Conditional Mean. Journal of Econometrics, 93, 257-279.
[65] Lundbergh, S., and Teräsvirta, T. (2002). Evaluating GARCH Models. Journal
of Econometrics, Vol. 110, 417-435
[66] Ma, J., Nelson, C.R., and Startz, R. (2006). Spurious Inference in the
GARCH(1,1) Model When It Is Weakly Identied. Studies in Nonlinear Dy-
namics and Econometrics, 11(1), Article 1.
[67] Mandelbrot, B. B. (1975). Limit Theorems on the Self-Normalized Range for
Weakly and Strongly Dependent Processes. Zeitschrift für Wahr- scheinlichkeit-
stheorie und verwandte Gebiete, 31, 271-285.
[68] Martens, M. (2002). Measuring and Forecasting S&P 500 Index-Futures Volatil-
ity Using High Frequency Data. The Journal of Futures Markets, 22, 497-518.
39
[69] McCullough, B.D., and Renfro, C.G. (1999). Benchmarks and Software Stan-
dards: A Case Study of GARCH Procedures. Journal of Economic and Social
Measurement, 25, 59-71.
[70] Mikosch, T., and Starica, C. (2004). Non-stationarities in Financial Time Series,
the Long-Range Dependence and the IGARCH Eects. Review of Economics
and Statistics, 86, 378-384.
[71] Nelson, D. B. (1991). Conditional Heteroskedasticity in Asset Returns: a New
Approac h. Econometrica, 59(2), 347-370.
[72] Nelson, D. B., and Cao, C. Q. (1992). Inequality Constrain ts in the Univariate
GARCH Model. Journal of Business and Economic Statistics, 10(2), 229-235.
[73] Newey, W.K. and West, K.D. (1987). A Simple Positive Semidenite Het-
eroskedasticity and Autocorrelation Consisten t Covariance Matrix. Economet-
rica, 55, 703-708.
[74] Pagan, A. (1996). The Econometrics of Financial Markets. Journal of Empirical
Finance, 3, 15-10 2.
[75] Pagan, A., and Schwert G.W. (1990). Alternative Models for Conditional Volatil-
it y. Journal of Econometrics, 45, 267-290.
[76] Palm, F.C. (1996). GARCH models of Volatility. In G.S. Maddala and C.R.
Rao, editors, Handbook of Statistics (Vol 14), Statistical Methods in Finance,
pp. 209-240. North Holland.
[77] Poon, S.-H. (2005). A Practical Guide to Forecasting Financial Market Volatility.
John Wiley & Sons, New York.
[78] Poon, S.-H., and Granger, C. (2003). Forecasting Financial Market Volatility: A
Review. Journal of Economic Literature, 41(2), 478-539.
[79] Poon, S.-H., and Granger, C. (2005). Practical Issues in Forecasting Volatility.
Financial Analysts Journal, 61(1), 45-56.
[80] Silvennoinen, A., and Teräsvirta, T. (2008). Multivariate GARCH. In T.G. An-
dersen,R.A.Davis,J-PKreiss,andT.Mikosch,editors,Handbook of Financial
Time Series, New York, Springer.
[81] Straumann, D. (2005). Estimation in Conditionally Heteroskedastic Time Series
Models. L ecture Notes in Statistics 181, Springer, Berlin.
[82] Taylor, S.J., and Xu, X. (1997). The Incremental Volatilty Information in One
Million Foreign Exchange Quotations. Journal of Empirical Finance, 4, 317-340.
[83] Teräsvirta, T. (2008). An Introduction to Univariate GARCH Models. In T.G.
Andersen, R.A. Davis, J-P Kreiss, and T. Mikosch, editors, Handbook of Finan-
cial Time Series, New York, Springer.
40
[84] Tsay, R.S. (2001). Analysis of Financial Time Series.JohnWiley& Sons,New
York.
[85] Venter, J.H., and de Jongh, P.J. (2002). Risk Estimation Using the Normal
Inverse Gausian Distribution. Journal of Risk,4,1-23.
[86] Weiss, A.A. (1986). Asymptotic Theory for ARCH Models: Estimation and
Testing. Econometric Theory, 2, 107-131.
[87] Zakoian, M. (1994). Threshold Heteroscedastic Models. Journal of Economic
Dynamics and Control, 18, 931-955.
[88] Zivot, E., and Wang, J. (2005). Modeling Financial Time Series with S-PLUS,
Second Edition. Springer-Verlag, New York.
41