7

Estimating GARCH Models by Quasi-Maximum Likelihood

The quasi-maximum likelihood (QML) method is particularly relevant for GARCH models because it provides consistent and asymptotically normal estimators for strictly stationary GARCH processes under mild regularity conditions, but with no moment assumptions on the observed process. By contrast, the least-squares methods of the previous chapter require moments of order 4 at least.

In this chapter, we study in details the conditional QML method (conditional on initial values). We first consider the case when the observed process is pure GARCH. We present an iterative procedure for computing the Gaussian log-likelihood, conditionally on fixed or random initial values. The likelihood is written as if the law of the variables ηt were Gaussian image(0, 1) (we refer to pseudo- or quasi-likelihood), but this assumption is not necessary for the strong consistency of the estimator. In the second part of the chapter, we will study the application of the method to the estimation of ARMA-GARCH models. The asymptotic properties of the quasi-maximum likelihood estimator (QMLE) are established at the end of the chapter.

7.1 Conditional Quasi-Likelihood

Assume that the observations image1, …, imagen constitute a realization (of length n) of a GARCH(p, q) process, more precisely a nonanticipative strictly stationary solution of

(7.1) c07e001_fmt

where (ηt) is a sequence of iid variables of variance 1, ω0 > 0, α0i ≥ 0 (i = 1, …, q), and β0j ≥ 0 (j = 1, …, p). The orders p and q are assumed known. The vector of the parameters

(7.2) c07e002_fmt

belongs to a parameter space of the form

(7.3) c07e003_fmt

The true value of the parameter is unknown, and is denoted by

c07ue001_fmt

To write the likelihood of the model, a distribution must be specified for the iid variables ηt. Here we do not make any assumption on the distribution of these variables, but we work with a function, called the (Gaussian) quasi-likelihood, which, conditionally on some initial values, coincides with the likelihood when the ηt are distributed as standard Gaussian. Given initial values image0, …, image1 − q, image to be specified below, the conditional Gaussian quasi-likelihood is given by

c07ue002_fmt

where the sigma-tilde2t_fmt are recursively defined, for t ≥ 1, by

(7.4) c07e004_fmt

For a given value of θ, under the second-order stationarity assumption, the unconditional variance (corresponding to this value of θ) is a reasonable choice for the unknown initial values:

(7.5) c07e005_fmt

Such initial values are, however, not suitable for IGARCH models, in particular, and more generally when the second-order stationarity is not imposed. Indeed, the constant (7.5) would then take negative values for some values of θ. In such a case, suitable initial values are

(7.6) c07e006_fmt

or

(7.7) c07e007_fmt

A QMLE of θ is defined as any measurable solution imagen of

c07ue003_fmt

Taking the logarithm, it is seen that maximizing the likelihood is equivalent to minimizing, with respect to θ,

(7.8) c07e008_fmt

and image is defined by (7.4). A QMLE is thus a measurable solution of the equation

(7.9) c07e009_fmt

It will be shown that the choice of the initial values is unimportant for the asymptotic properties of the QMLE. However, in practice this choice may be important. Note that other methods are possible for generating the sequence image; for example, by taking image where the ci(θ) are recursively computed (see Berkes, Horváth and Kokoszka, 2003b). Note that for computing image, this procedure involves a number of operations of order n2, whereas the one we propose involves a number of order n. It will be convenient to approximate the sequence (image(θ)) by an ergodic stationary sequence. Assuming that the roots of imageθ(z) are outside the unit disk, the nonanticipative and ergodic strictly stationary sequence (image)t = {image (θ)}t is defined as the solution of

(7.10) c07e010_fmt

Note that image0) = ht.

Likelihood Equations

Likelihood equations are obtained by canceling the derivative of the criterion image with respect to θ, which gives

(7.11) c07e011_fmt

These equations can be interpreted as orthogonality relations, for large n. Indeed, as will be seen in the next section, the left-hand side of equation (7.11) has the same asymptotic behavior as

c07ue004_fmt

the impact of the initial values vanishing as n → ∞.

The innovation of image is image. Thus, under the assumption that the expectation exists, we have

c07ue005_fmt

because image is a measurable function of the imageti, i > 0. This result can be viewed as the asymptotic version of (7.11) at θ0, using the ergodic theorem.

7.1.1 Asymptotic Properties of the QMLE

In this chapter, we will use the matrix norm defined by image A image = Σ | aij | for all matrices A = (aij). The spectral radius of a square matrix A is denoted by ρ(A).

Strong Consistency

Recall that model (7.1) admits a strictly stationary solution if and only if the sequence of matrices A0 = (A0t), where

c07ue006_fmt

admits a strictly negative top Lyapunov exponent, γ (A0) < 0, where

(7.12) c07e012_fmt

Let

c07ue007_fmt

By convention, imageθ (z) = 0 if q = 0 and imageθ (z) = 1 if p = 0. To show strong consistency, the following assumptions are used.

A1: θ0 image Θ and Θ is compact.

A2: γ(A0) < 0 and for all θ image Θ, image.

A3: image has a nondegenerate distribution and E image = 1.

A4: If p > 0, A-theta_fmt(z) and B-theta_fmt(z) have no common roots, A-theta_fmt(1) ≠ 0, and α0q + β0p ≠ 0.

Note that, by Corollary 2.2, the second part of assumption A2 implies that the roots of imageθ(z) are outside the unit disk. Thus, a nonanticipatlve and ergodic strictly stationary sequence (image)t is defined by (7.10). Similarly, define

c07ue008_fmt

Example 7.1 (Parameter space of a GARCH(1, 1) process) In the case of a GARCH(1, 1) process, assumptions Al and A2 hold true when, for instance, the parameter space is of the form

c07ue009_fmt

where δ image (0, 1) Is a constant, small enough so that the true value θ0 = (ω0, α0, β0)′ belongs to Θ. Figure 7.1 displays, in the plane (α, β), the zones of strict stationarity (when ηt is image(0, 1)

Figure 7.1 GARCH(1, 1): zones of strict and second-order stationarity and parameter space image.

c07f001_fmt

distributed) and of second-order stationarity, as well as an example of a parameter space Θ (the gray zone) compatible with assumptions A1 and A2.

The first result states the strong consistency of imagen. The proof of this theorem, and of the next ones, is given in Section 7.4.

Theorem 7.1 (Strong consistency of the QMLE) Let (imagen) be a sequence of QMLEs satisfying (7.9), with initial conditions (7.6) or (7.7). Under assumptions A1–A4, almost surely

c07ue010_fmt

Remark 7.1

1. It Is not assumed that the true value of the parameter θ0 belongs to the interior of Θ. Thus, the theorem allows to handle cases where some coefficients, αi or βj, are null.

2. It Is important to note that the strict stationarlty condition Is only assumed at θ0, not over all Θ. In view of Corollary 2.2, the condition image is weaker than the strict stationarlty condition.

3. Assumption A4 disappears In the ARCH case. In the general case, this assumption allows for an overidentification of either of the two orders, p or q, but not of both. We then consistently estimate the parameters of a GARCH(p − 1, q) (or GARCH(p, q − 1)) process if an overparameterized GARCH(p, q) model is used.

4. When p ≠ 0, assumption A4 precludes the case where all the α0i are zero. In such a case, the strictly stationary solution of the model is the strong white noise, which can be written in multiple forms. For instance, a strong white noise of variance 1 can be written in the GARCH(1, 1) form with image

5. The assumption of absence of a common root, in A4, is restrictive only if p > 1 and q > 1. Indeed if q = 1, the unique root of A-theta_fmt(z) is 0 and we have B-theta_fmt(0) ≠ 0. If p = 1 and β01 ≠ 0, the unique root of B-theta_fmt(z) is 1/β01 > 0 (if β01 = 0, the polynomial does not admit any root). Because the coefficients α0i are positive this value cannot be a zero of A-theta_fmt(z).

6. The assumption Eηt = 0 Is not required for the consistency (and asymptotic normality) of the QMLE of a GARCH. The conditional variance of imaget is thus, in general, only proportional to ht: Var(imaget | imageu, u < t) = {1 − (Eηt)2}ht. The assumption E image is made for identifiability reasons and is not restrictive provided that E image.

Asymptotic Normality

The following additional assumptions are considered.

A5: θ0 image image, where image denotes the interior of Θ.

A6: κη = E image < ∞.

The limiting distribution of imagen is given by the following result.

Theorem 7.2 (Asymptotic normality of the QMLE) Under assumptions A1–A6,

c07ue011_fmt

where

(7.13) c07e013_fmt

is a positive definite matrix.

Remark 7.2

1. Assumption A5 is standard and entails the first-order condition (at least asymptotically). Indeed if imagen is consistent, it also belongs to the interior of Θ, for large n. At this maximum the derivative of the objective function cancels. However, assumption A5 is restrictive because it precludes, for instance, the case α01 = 0.

2. When one or several components of θ0 are null, assumption A5 is not satisfied and the theorem cannot be used. It is clear that, in this case, the asymptotic distribution of image(imagenθ0) cannot be normal because the estimator is constrained. If, for instance, α01 = 0, the distribution of image(image1 − α 01) is concentrated in [0, ∞), for all n, and thus cannot be asymptotically normal. This kind of ‘boundary’ problem is the object of a specific study in Chapter 8.

3. Assumption A6 does not concern image, and does not preclude the IGARCH case. Only a fourth-order moment assumption on ηt is required. This assumption is clearly necessary for the existence of the variance of the score vector ∂2113_fmtt0)/∂θ. In the proof of this theorem, it is shown that

c07ue012_fmt

4. In the ARCH case (p = 0), the asymptotic variance of the QMLE reduces to that of the FGLS estimator (see Theorem 6.3). Indeed, in this case we have image. Theorem 6.3 requires, however, the existence of a fourth-order moment for the observed process, whereas there is no moment assumption for the asymptotic normality of the QMLE. Moreover, Theorem 6.4 shows that the QMLE of an ARCH(q) is asymptotically more accurate than that of the OLS estimator.

7.1.2 The ARCH(l) Case: Numerical Evaluation of the Asymptotic Variance

Consider the ARCH(l) model

c07ue013_fmt

with ω0 > 0 and α0 > 0, and suppose that the variables ηt satisfy assumption A3. The parameter is θ = (ω, α)′. In view of (2.10), the strict stationarity constraint A2 is written as

c07ue014_fmt

Assumption Al holds true if, for instance, the parameter space is of the form Θ = [δ, 1/δ] × [0, 1/δ], where δ > 0 is a constant, chosen sufficiently small so that θ0 = (ω0, ρ0)′ belongs to Θ. By Theorem 7.1, the QMLE of θ is then strongly consistent. Since image, the QMLE image is characterized by the normal equation

c07ue015_fmt

with, for instance, image. This estimator does not have an explicit form and must be obtained numerically. Theorem 7.2, which provides the asymptotic distribution of the estimator, only requires the extra assumption that θ0 belongs to image = (δ, 1/δ) × (0, 1/δ). Thus, if α0 = 0 (that is, if the model is conditionally homoscedastic), the estimator remains consistent but is no longer asymptotically normal. Matrix J takes the form

c07ue016_fmt

and the asymptotic variance of image(imagenθ0) is

c07ue017_fmt

Table 7.1 displays numerical evaluations of this matrix. An estimation of J is obtained by replacing the expectations by empirical means, obtained from simulations of length 10 000, when ηt is image(0, 1) distributed. This experiment is repeated 1000 times to obtain the results presented in the table.

In order to assess, in finite samples, the quality of the asymptotic approximation of the variance of the estimator, the following Monte Carlo experiment is conducted. For the value θ0 of the parameter, and for a given length n, N samples are simulated, leading to N estimations image of

Table 7.1. Asymptotic variance for the QMLE of an ARCH(l) process with ηt ~ (0, 1).

c07t001_fmt

Table 7.2. Comparison of the empirical and theoretical asymptotic variances, for the QMLE of the parameter α0 = 0.9 of an ARCH(l), when ηt ~ image(0, 1).

c07t002_fmt

θ0, i = 1, … N. We denote by image their empirical mean. The root mean squared error (RMSE) of estimation of α is denoted by

c07ue018_fmt

and can be compared to image, the latter quantity being evaluated independently, by simulation. A similar comparison can obviously be made for the parameter ω. For θ0 = (0.2, 0.9)′ and N = 1000, Table 7.2 displays the results, for different sample length n.

The similarity between columns 3 and 4 is quite satisfactory, even for moderate sample sizes. The last column gives the empirical probability (that is, the relative frequency within the N samples) that imagen is greater than 1 (which is the limiting value for second-order stationarity). These results show that, even if the mean of the estimations is close to the true value for large n, the variability of the estimator remains high. Finally, note that the length n = 1000 remains realistic for financial series.

7.1.3 The Nonstationary ARCH(l)

When the strict stationarity constraint is not satisfied in the ARCH(l) case, that is, when

(7.14) c07e014_fmt

one can define an ARCH(l) process starting with initial values. For a given value image0, we define

(7.15) c07e015_fmt

where ω0 > 0 and α0 > 0, with the usual assumptions on the sequence (ηt). As already noted, image converges to infinity almost surely when

(7.16) c07e016_fmt

and only in probability when the inequality (7.14) is an equality (see Corollary 2.1 and Remark 2.3 following it). Is it possible to estimate the coefficients of such a model? The answer is only partly positive: it is possible to consistently estimate the coefficient α0, but the coefficient ω0 cannot be consistently estimated. The practical impact of this result thus appears to be limited, but because of its theoretical interest, the problem of estimating coefficients of nonstationary models deserves attention. Consider the QMLE of an ARCH(l), that is to say a measurable solution of

(7.17) c07e017_fmt

where θ = (ω, α), Θ is a compact set of (0, ∞)2, and image for t = 1, …, n (starting with a given initial value for image). The almost sure convergence of image to infinity will be used to show the strong consistency of the QMLE of α0. The following lemma completes Corollary 2.1 and gives the rate of convergence of image to infinity under (7.16).

Lemma 7.1 Define the ARCH(l) model by (7.15) with any initial condition image ≥ 0. The nonstationarity condition (7.16) is assumed. Then, almost surely, as n → ∞,

c07ue019_fmt

for any constant ρ such that

(7.18) c07e018_fmt

This result entails the strong consistency and asymptotic normality of the QMLE of α0.

Theorem 7.3 Consider the assumptions of Lemma 7.1 and the QMLE defined by (7.17) where θ0 = (ω0, α0) image Θ. Then

(7.19) c07e019_fmt

and when θ0 belongs to the interior of Θ,

(7.20) c07e020_fmt

as n → ∞.

In the proof of this theorem, it is shown that the score vector satisfies

c07ue020_fmt

In the standard statistical inference framework, the variance J of the score vector is (proportional to) the Fisher information. According to the usual interpretation, the form of the matrix J shows that, asymptotically and for almost all observations, the variations of the log-likelihood image log 2113_fmtt(θ) are insignificant when θ varies from (ω0, α0) to (ω0 + h, α0) for small h. In other words, the limiting log-likelihood is flat at the point (ω0, α0) in the direction of variation of ω0. Thus, minimizing this limiting function does not allow θ0 to be found. This leads us to think that the QML of ω0 is likely to be inconsistent when the strict stationarity condition is not satisfied. Figure 7.2 displays numerical results illustrating the performance of the QMLE in finite samples. For different values of the parameters, 100 replications of the ARCH(1) model have been generated, for the sample sizes n = 200 and n = 4000. The top panels of the figure correspond to a second-order stationary ARCH(1), with parameter θ0 = (1, 0.95). The panels in the middle correspond to a strictly stationary ARCH(1) of infinite variance, with θ0 = (1, 1.5). The results obtained for these two cases are similar, confirming that second-order stationarity is not necessary for estimating an ARCH. The bottom panels, corresponding to the explosive ARCH(1) with parameter θ0 = (1, 4), confirm the asymptotic results concerning the estimation of α0. They also illustrate the failure of the QML to estimate ω0 under the nonstationarity assumption (7.16). The results even deteriorate when the sample size increases.

Figure 7.2 Box-plots of the QML estimation errors for the parameters ω0 and α0 of an ARCH(l) process, with ηt ~ image (0, 1).

c07f002_fmt

7.2 Estimation of ARMA-GARCH Models by Quasi-Maximum Likelihood

In this section, the previous results are extended to cover the situation where the GARCH process is not directly observed, but constitutes the innovation of an observed ARMA process. This framework is relevant because, even for financial series, it is restrictive to assume that the observed series is the realization of a noise. From a theoretical point of view, it will be seen that the extension to the ARMA-GARCH case is far from trivial. Assume that the observations X1, …, Xn are generated by a strictly stationary nonanticipative solution of the ARMA(P, Q)-GARCH(p, q) model

(7.21) c07e021_fmt

where (ηt) and the coefficients ω0, α0i and β0j are defined as in (7.1). The orders P, Q, p, q are assumed known. The vector of the parameters is denoted by

c07ue021_fmt

where θ is defined as previously (see (7.2)). The parameter space is

c07ue022_fmt

The true value of the parameter is denoted by

c07ue023_fmt

We still employ a Gaussian quasi-likelihood conditional on initial values. If qQ, the initial values are

c07ue024_fmt

These values (the last p of which are positive) may depend on the parameter and/or on the observations. For any image, the values of imaget(image), for t = −q + Q + 1, …, n, and then, for any θ, the values of image (image), for t = 1, …, n, can thus be computed from

(7.22) c07e022_fmt

When q < Q, the fixed initial values are

c07ue025_fmt

Conditionally on these initial values, the Gaussian log-likelihood is given by

c07ue026_fmt

A QMLE is defined as a measurable solution of the equation

c07ue027_fmt

Strong Consistency

Let image and image. Standard assumptions are made on these AR and MA polynomials, and assumption A1 is modified as follows:

A7: image0 image Ф and Ф is compact.

A8: For all image image Ф, aimage(z)bimage(z) = 0 implies |z| > 1.

A9: a(z) and b(z) have no common roots, a0P ≠ 0 or b0Q ≠ 0.

Under assumptions A2 and A8, (Xt) is supposed to be the unique strictly stationary nonanticipative solution of (7.21). Let imageimage and image, where image Is the nonantlcipative and ergodic strictly stationary solution of (7.10). Note that et = imaget (image0) and image. The following result extends Theorem 7.1.

Theorem 7.4 (Consistency of the QMLE) Let image be a sequence of QMLEs satisfying (7.2). Assume that Eηt = 0. Then, under assumptions A2–A4 and A7–A9, almost surely

c07ue028_fmt

Remark 7.3

1. As in the pure GARCH case, the theorem does not impose a finite variance for et (and thus for Xt). In the pure ARMA case, where et = ηt admits a finite variance, this theorem reduces to a standard result concerning ARMA models with iid errors (see Brockwell and Davis, 1991, p. 384).

2. Apart from the condition Eηt =0, the conditions required for the strong consistency of the QMLE are not stronger than in the pure GARCH case.

Asymptotic Normality When the Moment of Order 4 Exists

So far, the asymptotic results of the QMLE (consistency and asymptotic normality in the pure GARCH case, consistency in the ARMA-GARCH case) have not required any moment assumption on the observed process (for the asymptotic normality in the pure GARCH case, a moment of order 4 is assumed for the iid process, not for imaget). One might think that this will be the same for establishing the asymptotic normality in the ARMA-GARCH case. The following example shows that this is not the case.

Example 7.2 (Nonexistence of J without moment assumption) Consider the AR(1)-ARCH(1) model

(7.23) c07e023_fmt

where |α01| < 1, ω0 > 0, α0 ≥ 0, and the distribution of the iid sequence (ηt) is defined, for a > 1, by

c07ue029_fmt

Then the process (Xt) is always stationary, for any value of α0 (because exp image see the strict stationarity constraint (2.10)). By contrast, Xt does not admit a moment of order 2 when α0 ≥ 1 (see Theorem 2.2). The first component of the (normalized) score vector is

c07ue030_fmt

We have

c07ue031_fmt

since, first, ηt−1 = 0 entails imaget−1 = 0 and Xt−1 = a01 Xt−2, and second, ηt−1 and Xt−2 are independent. Consequently, if Eimage and a01 ≠ 0, the score vector does not admit a variance.

This example shows that it is not possible to extend the result of asymptotic normality obtained in the GARCH case to the ARMA-GARCH models without additional moment assumptions. This is not surprising because for ARMA models (which can be viewed as limits of ARMA-GARCH models when the coefficients α0i and β0j tend to 0) the asymptotic normality of the QMLE is shown with second-order moment assumptions. For an ARMA with infinite variance innovations, the consistency of the estimators may be faster than in the standard case and the asymptotic distribution is stable, but non-Gaussian in general. We show the asymptotic normality with a moment assumption of order 4. Recall that, by Theorem 2.9, this assumption is equivalent to ρ {E(A0t image A0t)} < 1. We make the following assumptions:

A10: ρ {E(A0t image A0t)} < 1 and, for all image

A11: image0 image image, where image denotes the interior of Ф.

A12: There exists no set Λ of cardinality 2 such that ℙ(ηt image Λ) = 1.

Assumption A10 implies that κη = (image) < ∞ and makes assumption A2 superfluous. The identiliability assumption A12 is slightly stronger than the first part of assumption A3 when the distribution of ηt is not symmetric. We are now in a position to state conditions ensuring the asymptotic normality of the QMLE of an ARMA-GARCH model.

Theorem 7.5 (Asymptotic normality of the QMLE) Assume thatt = 0 and that assumptions A3, A4 and A8–A12 hold true. Then

c07ue032_fmt

where image,

c07ue033_fmt

If, in addition, the distribution of ηt is symmetric, we have

c07ue034_fmt

with

c07ue035_fmt

Remark 7 4

1. It is interesting to note that if ηt has a symmetric law, then the asymptotic variance Σ is block-diagonal, which is interpreted as an asymptotic independence between the estimators of the ARMA coefficients and those of the GARCH coefficients. The asymptotic distribution of the estimators of the ARMA coefficients depends, however, on the GARCH coefficients (in view of the form of the matrices I1 and J1 involving the derivatives of image). On the other hand, still when the distribution of ηt is symmetric, the asymptotic accuracy of the estimation of the GARCH parameters is not affected by the ARMA part: the lower left block image of Σ depends only on the GARCH coefficients. The block-diagonal form of Σ may also be of interest for testing problems of joint assumptions on the ARMA and GARCH parameters.

2. Assumption A11 imposes the strict positivity of the GARCH coefficients and it is easy to see that this assumption constrains only the GARCH coefficients. For any value of image0, the restriction of Ф to its first P + Q + 1 coordinates can be chosen sufficiently large so that its interior contains image0 and assumption A8 is satisfied.

3. In the proof of the theorem, the symmetry of the iid process distribution is used to show the following result, which is of independent interest.

If the distribution of ηt is symmetric then,

(7.24) c07e024_fmt

provided this expectation exists (see Exercise 7.1).

Example 73 (Numerical evaluation of the asymptotic variance) Consider the AR(1)-ARCH(l) model defined by (7.23). In the case where ηt follows the image(0, 1) law, condition A10 for the existence of a moment of order 4 is written as image, that is, α0 < 0.577 (see (2.54)). In the case where ηt follows the χ2(1) distribution, normalized in such a way that Eηt = 0 and E image = 1, this condition is written as image, that is, α0 < 0.258. To simplify the computation, assume that ω0 = 1 is known. Table 7.3 provides a numerical evaluation of the asymptotic variance Σ, for these two distributions and for different values of the parameters a0 and ρ0. It is clear that the asymptotic variance of the two parameters strongly depends on the distribution of the iid process. These experiments confirm the independence of the asymptotic distributions of the AR and ARCH parameters in the case where the distribution of ηt is symmetric. They reveal that the independence does not hold when this assumption is relaxed. Note the strong impact of the ARCH coefficient on the asymptotic variance of the AR coefficient. On the other hand, the simulations confirm that in the case where the distribution is symmetric, the AR coefficient has no impact on the asymptotic accuracy of the ARCH coefficient. When the distribution is not

Table 7.3. Matrices Σ of asymptotic variance of the estimator of (a0, α0) for an AR(1)-ARCH(l), when ω0 = 1 is known and the distribution of ηt is image(0, 1) or normalized χ2(l).

c07t003_fmt

symmetric, the impact, if there is any, is very weak. For the computation of the expectations involved in the matrix Σ, see Exercise 7.8. In particular, the values corresponding to α0 = 0 (AR(1) without ARCH effect) can be analytically computed. Note also that the results obtained for the asymptotic variance of the estimator of the ARCH coefficient in the case a0 = 0 do not coincide with those of Table 7.2. This is not surprising because in this table ω0 is not supposed to be known.

7.3 Application to Real Data

In this section, we employ the QML method to estimate GARCH(1, 1) models on daily returns of 11 stock market indices, namely the CAC, DAX, DJA, DJI, DJT, DJU, FTSE, Nasdaq, Nikkei, SMI and S&P 500 indices. The observations cover the period from January 2, 1990 to January 22, 20091 (except for those indices for which the first observation is after 1990). The GARCH(1, 1) model has been chosen because It constitutes the reference model, by far the most commonly used in empirical studies. However, in Chapter 8 we will see that it can be worth considering models with higher orders p and q.

Table 7.4 displays the estimators of the parameters ω, α, β, together with their estimated standard deviations. The last column gives estimates of image, obtained by replacing the unknown parameters by their estimates and image by the empirical mean of the fourth-order moment of the standardized residuals. We have image if and only If ρ4 < 1. The

Table 7.4. GARCH(1, 1) models estimated by QML for 11 indices. The estimated standard deviations are given In parentheses. image.

c07t004_fmt

estimates of the GARCH coefficients are quite homogenous over all the series, and are similar to those usually obtained in empirical studies of daily returns. The coefficients α are close to 0.1, and the coefficients β are close to 0.9, which indicates a strong persistence of the shocks on the volatility. The sum α + β is greater than 0.98 for 10 of the 11 series, and greater than 0.96 for all the series. Since α + β < 1, the assumption of second-order statlonarity cannot be rejected, for any series (see Section 8.1). A fortiori, by Remark 2.6 the strict statlonarity cannot be rejected. Note that the strict statlonarity assumption, E log (α1image + ²1) < 0, seems difficult to test directly because it not only relies on the GARCH coefficients but also involves the unknown distribution of ηt. The existence of moments of order 4, image < ∞, is questionable for all the series because image is extremely close to 1. Recall, however, that the asymptotic properties of the QML do not require any moment on the observed process but do require strict stationarity.

7.4 Proofs of the Asymptotic Results*

We denote by K and p generic constants whose values can vary from line to line. As an example, one can write for 0 < ρ1 < 1 and 0 < ρ2 < 1, i1 ≥ 0, i2 ≥ 0,

c07ue036_fmt

Proof of Theorem 7.1

The proof is based on a vectorial autoregressive representation of order 1 of the vector image analogous to that used for the study of stationarity. Assumption A2 allows us to write image as a series depending on the infinite past of the variable image. It can be shown that the initial values are not Important asymptotically, using the fact that, under the strict stationarity assumption, image necessarily admits a moment order s, with s > 0. This property also allows us to show that the expectation of 2113_fmtt0) is well defined in image and that E_theta-0_fmt (2113_fmtt(θ)) − E_theta-0_fmt(2113_fmtt0)) ≥ 0, which guarantees that the limit criterion is minimized at the true value. The difficulty is that image can be equal to +∞. Assumptions A3 and A4 are crucial to establishing the identifiability: the former assumption precludes the existence of a constant linear combination of the image, j > 0. The assumption of absence of common root is also used. The ergodicity of 2113_fmtt(θ) and a compactness argument conclude the proof.

It will be convenient to rewrite (7.10) in matrix form. We have

(7.25) c07e025_fmt

where

(7.26) c07e026_fmt

We will establish the following intermediate results.

(a) limn→∞ supθimageΘ |In(θ) − imagen(θ)| = 0, a.s.

(b) (image t image image such that image a.s.) ⇒ θ = θ0.

(c) image, and if image.

(d) For any θ ≠ θ0, there exists a neighborhood V(θ) such that

c07ue037_fmt

(a) Asymptotic irrelevance of the initial values. In view of Corollary 2.2, the condition image of assumption A2 implies that ρ(B) < 1. The compactness of Θ implies that

(7.27) c07e027_fmt

Iterating (7.25), we thus obtain

(7.28) c07e028_fmt

Let image be the vector obtained by replacing image by image in image, and let image be the vector obtained by replacing image by the initial values (7.6) or (7.7). We have

(7.29) c07e029_fmt

From (7.27), it follows that almost surely

(7.30) c07e030_fmt

For x > 0 we have log xx − 1. It follows that, for image. We thus have almost surely, using (7.30),

(7.31) c07e031_fmt

The existence of a moment of order s > 0 for image, deduced from assumption Al and Corollary 2.3, allows us to show that image a.s. (see Exercise 7.2). Using Cesàro’s lemma, point (a) follows.

(b) Identifiability of the parameter. Assume that image a.s. By Corollary 2.2, the polynomial image θ(B) is invertible under assumption A2. Using (7.10), we obtain

c07ue038_fmt

If the operator in B between braces were not null, then there would exist a constant linear combination of the image, j > 0. Thus the linear innovation of the process (image) would be equal to zero. Since the distribution of image is nondegenerate, in view of assumption A3,

c07ue039_fmt

We thus have

(7.32) c07e032_fmt

Under assumption A4 (absence of common root), it follows that Aθ(z) = A-theta_fmt(z), Bθ(z) = B-theta_fmt(z) and ω = ω0. We have thus shown (b).

(c) The limit criterion is minimized at the true value. The limit criterion is not integrable at any point, but image is well defined in image image {+ ∞} because, with the notation x = max(−x, 0) and x+ = max(x, 0),

c07ue040_fmt2

It is, however, possible to have image for some values of θ. This occurs, for instance, when θ = (ω, 0, …, 0) and (imaget) is an IGARCH such that image. We will see that this cannot occur at θ0, meaning that the criterion is integrable at θ0. To establish this result, we have to show that image. Using Jensen’s inequality and, once again, the existence of a moment of order s > 0 for image, we obtain

c07ue041_fmt

because

c07ue042_fmt

Thus

c07ue043_fmt

Having already established that image, it follows that image is well defined in image. Since for all x > 0, log xx − 1 with equality if and only if x = 1, we have

(7.33) c07e033_fmt

with equality if and only if image0)/image(θ)image-a.s., that is, in view of (b), if and only if θ = θ0.3

(d) Compactness of Θ and ergodicity of (2113_fmtt(θ)). For all θ image Θ and any positive integer k, let Vk(θ) be the open ball of center θ and radius l/k. Because of (a), we have

c07ue044_fmt

To obtain the convergence of this empirical mean, the standard ergodic theorem cannot be applied (see Theorem A.2) because we have seen that 2113_fmtt(θ*) is not necessarily integrable, except at θ0. We thus use a modified version of this theorem, which allows for an ergodic and strictly stationary sequence of variables admitting an expectation in image image {+∞)} (see Exercise 7.3). This version of the ergodic theorem can be applied to {2113_fmtt (θ*)}, and thus to {image} (see Exercise 7.4), which allows us to conclude that

c07ue045_fmt

By Beppo Levi’s theorem, image increases to image as k → ∞. Given (7.33), we have shown (d).

The conclusion of the proof uses a compactness argument. First note that for any neighborhood V0) of θ0,

(7.34) c07e034_fmt

The compact set Θ is covered by the union of an arbitrary neighborhood V0) of θ0 and the set of the neighborhoods V(θ) satisfying (d), θ image Θ V0). Thus, there exists a finite subcover of Θ of the form V(θ0), V1), …, Vk), where, for i = 1, …, k, Vi) satisfies (d). It follows that

c07ue046_fmt

The relations (d) and (7.34) show that, almost surely, imagen belongs to V0) for n large enough. Since this is true for any neighborhood V0), the proof is complete.      Box_fmt

Proof of Theorem 7.2

The proof of this theorem is based on a standard Taylor expansion of criterion (7.8) at θ0. Since imagen converges to θ0, which lies in the interior of the parameter space by assumption A5, the derivative of the criterion is equal to zero at image n. We thus have

(7.35) c07e035_fmt

where the image are between imagen and θ0. It will be shown that

(7.36) c07e036_fmt

and that

(7.37) c07e037_fmt

The proof of the theorem immediately follows. We will split the proof of (7.36) and (7.37) into several parts:

(a) image

(b) J is invertible and image.

(c) There exists a neighborhood image0) of θ0 such that, for all i, j, k image {1, …, p + q + 1},

c07ue047_fmt

(d) image and image tend in probability to 0 as n → ∞.

(e) image.

(f) image.

(a) Integrability of the derivatives of the criterion at θ0. Since image, we have

(7.38) c07e038_fmt

(7.39) c07e039_fmt

At θ = θ0, the variable image/image = image is independent of image and its derivatives. To show (a), it thus suffices to show that

(7.40) c07e040_fmt

In view of (7.28), we have

(7.41) c07e041_fmt

(7.42) c07e042_fmt

where image, and B(j) is a p × p matrix with 1 in position (1, j) and zeros elsewhere. Note that, in view of the positivity of the coefficients and (7.41)–(7.42), the derivatives of image are positive or null In view of (7.41), it is clear that ∂image/ ∂ ωis bounded. Since image, the variable image is also bounded. This variable thus possesses moments of all orders. In view of the second equality in (7.41) and of the positivity of all the terms involved in the sums, we have

c07ue048_fmt

It follows that

(7.43) c07e043_fmt

The variable image thus admits moments of all orders at θ = θ0. In view of (7.42) and βjB(j)B, we have

(7.44) c07e044_fmt

Using (7.27), we have image Bk imageKρk for all k. Moreover, image having a moment of order s image (0, 1), the variable image has the same moment.4 Using in addition (7.44), the inequality image and the relation x/(1 + x) ≤ xs for all x ≥ 0,5 we obtain

(7.45) c07e045_fmt

Under assumption A5 we have β0j for all j which entails that the first expectation in (7.40) exists.

We now turn to the higher-order derivatives of image. In view of the first equality of (7.41), we have

(7.46) c07e046_fmt

We thus have

c07ue049_fmt

which is a vector of finite constants (since ρ(B) < 1). It follows that image is bounded, and thus admits moments of all orders. It is of course the same for image. The second equality of (7.41) gives

(7.47) c07e047_fmt

The arguments used for (7.45) then show that

c07ue050_fmt

This entails that image is integrable. Differentiating relation (7.42) with respect to βj, we obtain

(7.48) c07e048_fmt

because βjB(j)B. As for (7.45), it follows that

c07ue051_fmt

and the existence of the second expectation in (7.40) is proven.

Since image is bounded, and since by (7.43) the variables image are bounded at θ0, it is clear that

c07ue052_fmt

for i = 1, …, q + 1. With the notation and arguments already used to show (7.45), and using the elementary inequality x/(1 + x) ≤ xs/2 for all x ≥ 0, Minkowski’s inequality implies that

c07ue053_fmt

Finally, the Cauchy–Schwarz inequality entails that the third expectation of (7.40) exists.

(b) Invertibility of J and connection with the variance of the criterion derivative. Using (a), and once again the independence between image and image and its derivatives, we have by (7.38),

c07ue054_fmt

Moreover, in view of (7.40), J exists and satisfies (7.13). We also have

(7.49) c07e049_fmt

Assume now that J is singular. Then there exists a nonzero vector λ in imagep + q + 1 such that image a.s.6 In view of (7.10) and the stationarity of image, we have

c07ue055_fmt

Let λ = (λ0, λ1, …, λq + p)′. It is clear that λ1 = 0, otherwise image would be measurable with respect to the σ-field generated by {ηu, u < t − 1}. For the same reason, we have λ2 = … = λ2+1 = 0 if λq+1 = … = λq+i = 0. Consequently, λ ≠ 0 implies the existence of a GARCH(p − 1, q − 1) representation. By the arguments used to show (7.32), assumption A4 entails that this is impossible. It follows that λ′Jλ = 0 implies λ = 0, which completes the proof of (b).

(c) Uniform integrability of the third-order derivatives of the criterion. Differentiating (7.39), we obtain

(7.50) c07e050_fmt

We begin by studying the integrability of {1− image/image}. This is the most difficult term to deal with. Indeed, the variable image/image is not uniformly integrable on Θ: at θ = (ω, 0′), the ratio image/image is

integrable only if E image exists. We will, however, show the integrability of {1 − image/image} uniformly in θ in the neighborhood of θ0. Let Θ* be a compact set which contains θ0 and which is contained in the interior of Θ (imageθ image Θ*, we have θ ≥ θ* > 0 component by component). Let B0 be the matrix B (defined in (7.26)) evaluated at the point θ = θ0. For all δ > 0, there exists a neighborhood image0) of θ0, included in Θ*, such that for all θ image image0),

c07ue057_fmt

Note that, since image0) image Θ*, we have image 1/αi < ∞. From (7.28), we obtain

c07ue058_fmt

and, again using x/(l + x) ≤ xs for all x ≥ 0 and all s image (0, 1),

(7.51) c07e051_fmt

If s is chosen such that Eimages < ∞ and, for instance, δ = (1 − ρs)/(2ρs), then the expectation of the previous series is finite. It follows that there exists a neighborhood image0) of θ0 such that

c07ue059_fmt

Using (7.51), keeping the same choice of δ but taking s such that Eimages < ∞, the triangle inequality gives

(7.52) c07e052_fmt

Now consider the second term in braces in (7.50). Differentiating (7.46), (7.47) and (7.48), with the arguments used to show (7.43), we obtain

c07ue060_fmt

when the indices i1, i2 and i3 are not all in {q + 1, q + 2, …, q + 1 + p} (that is, when the derivative is taken with respect to at least one parameter different from the βj). Using again the arguments used to show (7.44) and (7.48), and then (7.45), we obtain

c07ue061_fmt

for any s image (0, 1). Since image for some s > 0, it follows that

(7.53) c07e053_fmt

It is easy to see that in this inequality the power 2 can be replaced by any power d:

c07ue062_fmt

Using the Cauchy-Schwarz inequality, (7.52) and (7.53), we obtain

c07ue063_fmt

The other terms in braces in (7.50) are handled similarly. We show in particular that

(7.54) c07e054_fmt

for any integer d. With the aid of Hölder’s inequality, this allows us to establish, in particular, that

c07ue064_fmt

Thus we obtain (c).

(d) Asymptotic decrease of the effect of the initial values. Using (7.29), we obtain the analogs of (7.41) and (7.42) for the derivatives of image:

(7.55) c07e055_fmt

(7.56) c07e056_fmt

(7.57) c07e057_fmt

where image is equal to (0, …, 0)′ when the initial conditions are given by (7.7), and is equal to (1, …, 1)′ when the initial conditions are given by (7.6). The second-order derivatives have similar expressions. The compactness of 0 and the fact that ρ(B) < 1 together allow us to claim that, almost surely,

(7.58) c07e058_fmt

Using (7.30), we obtain

(7.59) c07e059_fmt

Since

c07ue065_fmt

we have, using (7.59) and the first inequality in (7.58),

c07ue066_fmt

It follows that

(7.60) c07e060_fmt

Markov’s inequality, (7.40), and the independence between ηt and image imply that, for all ε > 0,

c07ue067_fmt

which, by (7.60), shows the first part of (d).

Now consider the asymptotic impact of the initial values on the second-order derivatives of the criterion in a neighborhood of θ0. In view of (7.39) and the previous computations, we have

c07ue068_fmt

where

c07ue069_fmt

In view of (7.52), (7.54) and Holder's inequality, it can be seen that, for a certain neighborhood image0), the expectation of imaget is a finite constant. Using Markov’s inequality once again, the second convergence of (d) is then shown.

(e) CLT for martingale increments. The conditional score vector is obviously centered, which can be seen from (7.38), using the fact that image and its derivatives belong to the σ-field generated by {imaget−t, i ≥ 0}, and the fact that image:

c07ue070_fmt

Note also that, by (7.49), image is finite. In view of the invertibility of J and the assumptions on the distribution of ηt (which entail 0 < κn − 1 < ∞), this covariance matrix is nondegenerate. It follows that, for all λ image imagep+q+1, the sequence image is a square integrable ergodic stationary martingale difference. Corollary A.l and the Cramér-Wold theorem (see, for example, Billingsley, 1995, pp. 383, 476 and 360) entail (e).

(f) Use of a second Taylor expansion and of the ergodic theorem Consider the Taylor expansion (7.35) of the criterion at θ0. We have, for all i and j,

(7.61) c07e061_fmt

where imageij is between .θ*ij and θ0. The almost sure convergence of imageij to θ0, the ergodic theorem and (c) Imply that almost surely

c07ue071_fmt

Since image almost surely, the second term on the right-hand side of (7.61) converges to 0 with probability 1. By the ergodic theorem, the first term on the right-hand side of (7.61) converges to J(i, j).

To complete the proof of Theorem 7.2, it suffices to apply Slutsky’s lemma. In view of (d), (e) and (f) we obtain (7.36) and (7.37).            Box_fmt

Proof of the Results of Section 7.1.3

Proof of Lemma 7.1. We have

c07ue072_fmt

Thus

c07ue073_fmt

using (7.18) for the latter inequality. It follows that log ρnhn, and thus ρnhn, tend almost surely to +∞ as n → ∞. Now if ρnhn → +∞ and image, then for any ε > 0, the sequence (image) admits an Infinite number of terms less than ε. Since the sequence (image) is ergodic and stationary, we have (image). Since ε is arbitrary, we have image, which is in contradiction to (7.16).              Box_fmt

Proof of (7.19). Note that

c07ue074_fmt

where

c07ue075_fmt

We have

c07ue076_fmt

For all θ image Θ, we have α ≠ 0. Letting

c07ue077_fmt

and

c07ue078_fmt

we have

c07ue079_fmt

since, by Lemma 7.1, image → ∞almost surely as t → ∞. It is easy to see that this convergence is uniform on the compact set Θ:

(7.62) c07e062_fmt

Let image and image be two constants such that image. It can always be assumed that image. With the notation image, the solution of

c07ue080_fmt

is image. This solution belongs to the interval image when n is large enough. In this case

c07ue081_fmt

is one of the two extremities of the interval image, and thus

c07ue082_fmt

This result and (7.62) show that almost surely

c07ue083_fmt

Since minθ Qn(θ) ≤ Qn(image0) = 0, it follows that

c07ue084_fmt

Since image is an interval which contains α0 and can be arbitrarily small, we obtain the result.       Box_fmt

To prove the asymptotic normality of the QMLE, we need the following intermediate result.

Lemma 7.2 Under the assumptions of Theorem 7.3, we have

(7.63) c07e063_fmt

(7.64) c07e064_fmt

(7.65) c07e065_fmt

(7.66) c07e066_fmt

Proof. Using Lemma 7.1, there exists a real random variable K and a constant σ image (0, 1) independent of θ and of t such that

c07ue085_fmt

Since it has a finite expectation, the series image is almost surely finite. This shows (7.63), and (7.64) follows similarly. We have

c07ue086_fmt

where

c07ue087_fmt

and

c07ue088_fmt

as t → ∞. Thus (7.65) is shown. To show (7.66), it suffices to note that

c07ue089_fmt

Proof of (7.20). We remark that we do not know, a priori, if the derivative of the criterion is equal to zero at image, because we only have the convergence of image to α0. Thus the minimum of the criterion could lie at the boundary of Θ, even asymptotically. By contrast, the partial derivative with respect to the second coordinate must asymptotically vanish at the optimum, since imagen → α0 and θ0 image image. A Taylor expansion of the derivative of the criterion thus gives

(7.67) c07e067_fmt

where Jn is a 2 × 2 matrix whose elements are of the form

c07ue090_fmt

with image. between imagen and θ0. By Lemma 7.1, which shows that image →∞almost surely, and by the central limit theorem of Lindeberg for martingale increment (see Corollary A.l),

(7.68) c07e068_fmt

Relation (7.64) of Lemma 7.2 and the compactness of 0 show that

(7.69) c07e069_fmt

By a Taylor expansion of the function

c07ue091_fmt

we obtain

c07ue092_fmt

where α* is between image and α0. Using (7.65), (7.66) and (7.19), we obtain

(7.70) c07e070_fmt

We conclude using the second row of (7.67), and also using (7.68), (7.69) and (7.70).                 Box_fmt

Proof of Theorem 7.4

The proof follows the steps of the proof of Theorem 7.1. We will show the following points:

(a) image, a.s.

(b) image.

(c) If image.

(d) For any imageimage0 there exists a neighborhood V(image) such that

c07ue093_fmt

(a) Nullity of the asymptotic impact of the initial values. Equations (7.10)–(7.28) remain valid under the convention that imaget = imaget (image). Equation (7.29) must be replaced by

(7.71) c07e071_fmt

where image, the “tilde” variables being initialized as indicated before. Assumptions A7 and A8 imply that,

(7.72) c07e072_fmt

It follows that almost surely

c07ue094_fmt

and thus, by (7.28), (7.71) and (7.27),

(7.73) c07e073_fmt

Similarly, we have that almost surely image. The difference between the theoretical log-likelihoods with and without initial values can thus be bounded as follows:

c07ue095_fmt

This inequality is analogous to (7.31), image + 1 being replaced by image. Following the lines of the proof of (a) in Theorem 7.1 (see Exercise 7.2), It suffices to show that for all real r > 0, Etξt)r is the general term of a finite series. Note that7

c07ue096_fmt

since, by Corollary 2.3, images. Statement (a) follows.

(b) Identifiability of the parameter. If imaget(image) = imaget(image0) almost surely, assumptions A8 and A9 imply that there exists a constant linear combination of the variables Xtj, j ≥ 0. The linear Innovation of (Xt), equal to XtE(Xt|Xu, u < t) = ηtσt(image0), is zero almost surely only if ηt = 0 a.s. (since image). This is precluded, since E image. It follows that image = image0 and thus that θ = θ0 by the argument used in the proof of Theorem 7.1.

(c) The limit criterion is minimized at the true value. By the arguments used in the proof of (c) in Theorem 7.1, it can ne shown that, for all image, Eimage0In(image) = Eimage02113_fmtt(image) is defined in image image {+∞}, and in image at image= image0. We have

c07ue097_fmt

because the last expectation Is equal to 0 (noting that imaget(image) − imaget(image0) belongs to the past, as well as σt (image0) and σt(image)), the other expectations being positive or null by arguments already used. This inequality is strict only if imaget(image) = imaget(image0) and if image a.s. which, by (b), implies image = image0 and completes the proof of (c).

(d) Use of the compactness of Ф and of the ergodicity of (2113_fmtt(image)). The end of the proof is the same as that of Theorem 7.1.               Box_fmt

Proof of Theorem 7.5

The proof follows the steps of that of Theorem 7.2. The block-diagonal form of the matrices image and image when the distribution of ηt is symmetric Is shown in Exercise 7.7. It suffices to establish the following properties.

(a) image

(b) image and image are invertible.

(c) image and image tend in probability to 0 as n → ∞.

(d) image.

(e) image s.d., for all image* between image and image0.

Formulas (7.38) and (7.39) giving the derivatives with respect to the GARCH parameters (that is, the vector θ) remain valid in the presence of an ARMA part (writing image=image (image)). The same is true for all the results established in (a) and (b) of the proof of Theorem 7.2, with obvious changes of notation. The derivatives of image with respect to the parameter image, and the cross derivatives with respect to θ and image, are given by

(7.74) c07e074_fmt

(7.75) c07e075_fmt

(7.76) c07e076_fmt

The derivatives of imaget are of the form

c07ue098_fmt

where

(7.77) c07e077_fmt

and

(7.78) c07e078_fmt

where Hk, 2113_fmt(t) is the k × 2113_fmt (Hankel) matrix of general term imagetij, and 0k, 2113_fmt denotes the null matrix of size k × 2113_fmt. Moreover, by (7.28),

(7.79) c07e079_fmt

where imagej denotes the jth component of image, and

(7.80) c07e080_fmt

(a) Integrability of the derivatives of the criterion at φ0 The existence of the expectations in (7.40) remains true. By (7.74)–(7.76), the independence between (imagett)(image0) = ηt and image its derivatives, and the derivatives of imaget(image0), using Eimage < ∞and image(image 0) > ω0 >, it suffices to show that

(7.81) c07e081_fmt

(7.82) c07e082_fmt

to establish point (a), together with the existence of the matrices image and image. By the expressions for the derivatives of imaget, (7.77)–(7.78), and using E image(image 0) < ∞, we obtain (7.81).

The Cauchy-Schwarz inequality implies that

c07ue099_fmt

Thus, in view of (7.79) and the positivity of ω0,

c07ue100_fmt

Using the triangle inequality and the elementary inequalities image and x/(1 + x2) ≤ 1, it follows that

(7.83) c07e083_fmt

The first inequality of (7.82) follows. The existence of the second expectation in (7.82) is a consequence of (7.80), the Cauchy-Schwarz inequality, and the square integrability of imaget and its derivatives. To handle the second-order partial derivatives of image, first note that image by (7.41). Moreover, using (7.79),

(7.84) c07e084_fmt

By the arguments used to show (7.44), we obtain

(7.85) c07e085_fmt

which entails the existence of the third expectation in (7.82).

(b) Invertibility of image and image. Assume that image is noninvertible. There exists a nonzero vector λ in imagep+Q+p+q+2 such that λ′∂2113_fmtt (image0)/∂image′ = 0 a.s. By (7.38) and (7.74), this implies that

(7.86) c07e086_fmt

Taking the variance of the left-hand side, conditionally on the σ-field generated by {χu, u < t}, we obtain a.s., at image = image0,

c07ue101_fmt

where image. It follows that image and image} a.s. By stationarity, we have either image} a.s. for all t, or image} a.s. for all t. Consider for instance the latter case, the first one being treated similarly. Relation (7.86) implies image a.s. The term in brackets cannot vanish almost surely, otherwise ηt would take at least two different values, which would be in contradiction to assumption A12. It follows that at = 0 a.s. and thus bt = 0 a.s. We have shown that almost surely

(7.87) c07e087_fmt

where λ1 is the vector of the first P + Q + 1 components of λ. By stationarity of (∂imaget/∂image)t, the first equality implies that

c07ue102_fmt

We now use assumption A9, that the ARMA representation is minimal, to conclude that λ1 = 0. The third equality in (7.87) is then written, with obvious notation, as image. We have already shown in the proof of Theorem 7.2 that this entails λ2 = 0. We are led to a contradiction, which proves that image is invertible. Using (7.39) and (7.75)–(7.76), we obtain

c07ue103_fmt

We have just shown that the first expectation is a positive definite matrix. The second expectation being a positive semi-definite matrix, image is positive definite and thus invertible, which completes the proof of (b).

(c) Asymptotic unimportance of the initial values. The initial values being fixed, the derivatives of image, obtained from (7.71), are given by

c07ue104_fmt

with the notation introduced in (7.41)–(7.42) and (7.55)–(7.56). As for (7.79), we obtain

c07ue105_fmt

and, by an obvious extension of (7.72),

(7.88) c07e088_fmt

Thus

c07ue106_fmt

The latter sum converges almost surely because its expectation is finite. We have thus shown that

c07ue107_fmt

The other derivatives of image are handled similarly, and we obtain

c07ue108_fmt

We have, in view of (7.73),

c07ue109_fmt

where image. It is also easy to check that for image = image0

c07ue110_fmt

It follows that, using (7.88),

c07ue111_fmt

Using the independence between ηt and St–1, (7.40), (7.83), the Cauchy-Schwarz inequality and E image < ∞, we obtain

c07ue112_fmt

which shows the first part of (c). The second is established by the same arguments.

(d) Use of a CLT for martingale increments. The proof of this point is exactly the same as that of the pure GARCH case (see the proof of Theorem 7.2).

(e) Convergence to the matrix image. This part of the proof differs drastically from that of Theorem 7.2. For pure GARCH, we used a Taylor expansion of the second-order derivatives of the criterion, and showed that the third-order derivatives were uniformly integrable in a neighborhood of θ0. Without additional assumptions, this argument fails in the ARMA-GARCH case because variables of the form image do not necessarily have moments of all orders, even at the true value of the parameter. First note that, since image exists, the ergodic theorem implies that

c07ue113_fmt

The consistency of image having already been established, it suffices to show that for all ε > 0, there exists a neighborhood image(image0) of image0 such that almost surely

(7.89) c07e089_fmt

(see Exercise 7.9). We first show that there exists image(image0) such that

(7.90) c07e090_fmt

By Hölder’s inequality, (7.39), (7.75) and (7.76), it suffices to show that for any neighborhood image(image0) image Ф whose elements have their components αi, and βj bounded above by a positive constant, the quantities

(7.91) c07e091_fmt

(7.92) c07e092_fmt

(7.93) c07e093_fmt

are finite. Using the expansion of the series

c07ue114_fmt

similar expansions for the derivatives, and imageimaget (image0)image4 < ∞, it can be seen that the norms in (7.91) are finite. In (7.92) the first norm is finite, as an obvious consequence of image, this latter term being strictly positive by compactness of Ф. An extension of inequality (7.83) leads to

c07ue115_fmt

Moreover, since (7.41)–(7.44) remain valid when εt is replaced by imaget(image), it can be shown that

c07ue116_fmt

for any d > 0 and any neighborhood image(image0) whose elements have their components αi and βj bounded from below by a positive constant. The norms in (7.92) are thus finite. The existence of the first norm of (7.93) follows from (7.80) and (7.91). To handle the second one, we use (7.84), (7.85), (7.91), and the fact that image. Finally, it can be shown that the third norm is finite by (7.47), (7.48) and by arguments already used. The property (7.90) is thus established. The ergodic theorem shows that the limit in (7.89) is equal almost surely to

c07ue117_fmt

By the dominated convergence theorem, using (7.90), this expectation tends to 0 when the neighborhood image(image0) tends to the singleton {image0}. Thus (7.89) hold true, which proves (e). The proof of Theorem 7.5 is now complete.                    Box_fmt

7.5 Bibliographical Notes

The asymptotic properties of the QMLE of the ARCH models have been established by Weiss (1986) under the condition that the moment of order 4 exists. In the GARCH(1, 1) case, the asymptotic properties have been established by Lumsdaine (1996) (see also Lee and Hansen, 1994) for the local QMLE under the strict stationarity assumption. In Lumsdaine (1996) the conditions on the coefficients α1 and β1 allow to handle the IGARCH(1, 1) model. They are, however, very restrictive with regard to the iid process: it is assumed that Et|32 < ∞ and that the density of ηt has a unique mode and is bounded in a neighborhood of 0. In Lee and Hansen (1994) the consistency of the global estimator is obtained under the assumption of second-order stationarity.

Berkes, Horváth and Kokoszka (2003b) was the first paper to give a rigorous proof of the asymptotic properties of the QMLE in the GARCH (p, q) case under very weak assumptions; see also Berkes and Horváth (2003b, 2004), together with Boussama (1998, 2000). The assumptions given in Berkes, Horváth and Kokoszka (2003b) were weakened slightly in Francq and Zakoäan (2004). The proofs presented here come from that paper. An extension to non-iid errors was recently proposed by Escanciano (2009).

Jensen and Rahbek (2004a, 2004b) have shown that the parameter α0 of an ARCH(l) model, or the parameters α0 and β0 of a GARCH(1, 1) model, can be consistently estimated, with a standard Gaussian asymptotic distribution and a standard rate of convergence, even if the parameters are outside the strict stationarity region. They considered a constrained version of the QMLE, in which the intercept ω is fixed (see Exercises 7.13 and 7.14). These results were misunderstood by a number of researchers and practitioners, who wrongly claimed that the QMLE of the GARCH parameters is consistent and asymptotically normal without any stationarity constraint. We have seen in Section 7.1.3 that the QMLE of ω0 is inconsistent in the nonstationary case.

For ARMA-GARCH models, asymptotic results have been established by Ling and Li (1997, 1998), Ling and McAleer (2003a, 2003b) and Francq and Zakoïan (2004). A comparison of the assumptions used in these papers can be found in the last reference. We refer the reader to Straumann (2005) for a detailed monograph on the estimation of GARCH models, to Francq and Zakoïan (2009a) for a recent review of the literature, and to Straumann and Mikosch (2006) and Bardet and Wintenberger (2009) for extensions to other conditionally heteroscedastic models. Li, Ling and McAleer (2002) reviewed the literature on the estimation of ARMA-GARCH models, including in particular the case of nonstationary models.

The proof of the asymptotic normality of the QMLE of ARMA models under the second-order moment assumption can be found, for instance, in Brockwell and Davis (1991). For ARMA models with infinite variance noise, see Davis, Knight and Liu (1992), Mikosch, Gadrich, Klüppelberg and Adler (1995) and Kokoszka and Taqqu (1996).

7.6 Exercises

71 (The distribution of ηt is symmetric for GARCH models) The aim of this exercise is to show property (7.24).

1. Show the result for j < 0.

2. For j ≥ 0, explain why image can be written as image for some function h.

3. Complete the proof of (7.24).

7.2 (Almost sure convergence to zero at an exponential rate)

Let (imaget) be a strictly stationary process admitting a moment order s > 0. Show that if ρ image (0, 1), then image a.s.

7.3 (Ergodic theorem for nonintegrable processes)

Prove the following ergodic theorem. If (Xt) is an ergodic and strictly stationary process and if EX1 exists in image image {+∞}, then

c07ue118_fmt

The result is shown in Billingsley (1995, p. 284) for iid variables.

Hint: Consider the truncated variables image where κ > 0 with κ tending to +∞.

7.4 (Uniform ergodic theorem)

Let {Xt(θ)} be a process of the form

(7.94) c07e094_fmt

where (ηt) is strictly stationary and ergodic and f is continuous in θ image Ф,Ф being a compact subset of imaged.

1. Show that the process {infθimageФ Xt(θ)} is strictly stationary and ergodic.

2. Does the property still hold true if Xt(θ) is not of the form (7.94) but it is assumed that {Xt(θ)} is strictly stationary and ergodic and that Xt(θ) is a continuous function of θ

7.5 (OLS estimator of a GARCH)

In the framework of the GARCH(p, q) model (7.1), an OLS estimator of θ is defined as any measurable solution imagen of

c07ue119_fmt

where

c07ue120_fmt

and image is defined by (7.4) with, for instance, initial values given by (7.6) or (7.7). Note that the estimator is unconstrained and that the variable image can take negative values. Similarly, a constrained OLS estimator is defined by

c07ue121_fmt

The aim of this exercise is to show that under the assumptions of Theorem 7.1, and if image, the constrained and unconstrained OLS estimators are strongly consistent. We consider the theoretical criterion

c07ue122_fmt

1. Show that image almost surely as n → ∞.

2. Show that the asymptotic criterion is minimized at θ0,

c07ue123_fmt

and that θ0 is the unique minimum.

3. Prove that imagen → θ0 almost surely as n → ∞.

4. Show that image almost surely as n → ∞.

7.6 (The mean of the squares of the normalized residuals is equal to 1)

For a GARCH model, estimated by QML with initial values set to zero, the normalized residuals are defined by image. Show that almost surely

c07ue124_fmt

Hint: Note that for all c > 0, there exists image such that image for all t ≥ 0, and consider the function image.

7.7 (image and image block-diagonal)

Show that image and image have the block-diagonal form given in Theorem 7.5 when the distribution of ηt is symmetric.

7.8 (Forms of land Jin the AR(1)-ARCH(1) case)

We consider the QML estimation of the AR(1)-ARCH(1) model

c07ue125_fmt

assuming that ω0 = 1 is known and without specifying the distribution of ηt.

1. Give the explicit form of the matrices image and imagein Theorem 7.5 (with an obvious adaptation of the notation because the parameter here Is (a0, α0)).

2. Give the block-diagonal form of these matrices when the distribution of ηt is symmetric, and verify that the asymptotic variance of the estimator of the ARCH parameter

(i) doe not depend on the AR parameter, and

(ii) is the same as for the estimator of a pure ARCH (without the AR part).

3. Compute Σ when α0 = 0. Is the asymptotic variance of the estimator of a0 the same as that obtained when estimating an AR(1)? Verify the results obtained by simulation in the corresponding column of Table 7.3.

7.9 (A useful result in showing asymptotic normality)

Let (Jt(θ)) be a sequence of random matrices, which are function of a vector of parameters θ. We consider an estimator imagen which strongly converges to the vector θ0. Assume that

c07ue126_fmt

where J is a matrix. Show that if for all ε > 0 there exists a neighborhood V0) of θ0 such that

(7.95) c07e095_fmt

where image · image denotes a matrix norm, then

c07ue127_fmt

Give an example showing that condition (7.95) is not necessary for the latter convergence to hold in probability.

7.10 (A lower bound for the asymptotic variance of the QMLE of an ARCH)

Show that, for the ARCH(q) model, under the assumptions of Theorem 7.2,

c07ue128_fmt

in the sense that the difference is a positive semi-definite matrix.

Hint: Compute image and show that JJθ0θ′0J is a variance matrix.

7.11 (A striking property of J)

For a GARCH(p, q) model we have, under the assumptions of Theorem 7.2,

c07ue129_fmt

The objective of the exercise is to show that

(7.96) c07e096_fmt

1. Show the property in the ARCH case.

Hint: Compute imageand image.

2. In the GARCH case, let image. Show that

c07ue130_fmt

3. Complete the proof of (7.96).

7.12 (A condition required for the generalized Bartlett formula)

Using (7.24), show that if the distribution of ηt is symmetric and if E image < ∞, then formula (B.13) holds true, that is,

c07ue131_fmt

7.13 (Constrained QMLE of the parameter α0 of a nonstationary ARCH(1) process)

Jensen and Rahbek (2004a) consider the ARCH(l) model (7.15), in which the parameter ω0 > 0 is assumed to be known0 = 1 for instance) and where only α0 is unknown. They work with the constrained QMLE of α0 defined by

(7.97) c07e097_fmt

where image. Assume therefore that ω0 = 1 and suppose that the nonstationarity condition (7.16) id satisfied.

1. Verify that

c07ue132_fmt

and that

c07ue133_fmt

2. Prove that

c07ue134_fmt

3. Determine the almost sure limit of

c07ue135_fmt

4. Show that for all image, almost surely

c07ue136_fmt

5. Prove that if image almost surely (see Exercise 7.14) then

c07ue137_fmt

6. Does the result change when image and ω0 ≠ 1?

7. Discuss the practical usefulness of this result for estimating ARCH models.

7.14 (Strong consistency of Jensen and Rahbek’s estimator)

We consider the framework of Exercise 7.13, and follow the lines of the proof of (7.19) on page 169.

1. Show that image (1)converges almost surely to α0 when ω0= 1.

2. Does the result change if image(1) is replaced by image and if ω and ω0 are arbitrary positive numbers? Does it entail the convergence result (7.19)?

1 For the Nasdaq an outlier has been eliminated because the base price was reset on the trading day following December 31, 1993.

2 We use here the fact that (f + g)g for f ≥ 0, and that if fg then fg.

3 To show (7.33) it can be assumed that image and that image (in order to use the linearity property of the expectation), otherwise image and the relation is trivially satisfied.

4 We use the inequality (a + b)sas + bs for all a, b ≥ 0 and any s image (0. 1]. Indeed, xs > x for all x image [0, 1], and if image.

5 If x ≥ 1 then x4 ≥ 1 ≥ x/(l + x). If 0 ≤ x ≤ 1 then xsxx/(l + x).

6 We have

c07ue056_fmt if and only if image a.s., that is, if and only if image a.s.

7 We use the fact that if X and Y are positive random variables, E(X + Y)rE(X)r + E(Y)r for all r image (0, 1], this inequality being trivially obtained from the inequality already used: (a + b)rar + br for all positive real numbers a and b.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
3.17.162.76