8
Tests Based on the Likelihood

In the previous chapter, we saw that the asymptotic normality of the QMLE of a GARCH model holds true under general conditions, in particular without any moment assumption on the observed process. An important application of this result concerns testing problems. In particular, we are able to test the IGARCH assumption, or more generally a given GARCH model with infinite variance. This problem is the subject of Section 8.1.

The main aim of this chapter is to derive tests for the nullity of coefficients. These tests are complex in the GARCH case because of the constraints that are imposed on the estimates of the coefficients to guarantee that the estimated conditional variance is positive. Without these constraints, it is impossible to compute the Gaussian log‐likelihood of the GARCH model. Moreover, asymptotic normality of the QMLE has been established assuming that the parameter belongs to the interior of the parameter space (assumption A5 in Chapter 7). When some coefficients, α i or β j , are null, Theorem 7.2 does not apply. It is easy to see that, in such a situation, the asymptotic distribution of images cannot be Gaussian. Indeed, the components, images of images , are constrained to be positive or null. If, for instance θ 0i  = 0, then images for all n and the asymptotic distribution of this variable cannot be Gaussian.

Before considering the significance tests, we shall, therefore, establish in Section 8.2 the asymptotic distribution of the QMLE without assumption A5, at the cost of a moment assumption on the observed process. In Section 8.3, we present the main tests (Wald, score and likelihood ratio) used for testing the nullity of some coefficients. The asymptotic distribution obtained for the QMLE will lead to modification of the standard critical regions. Two cases of particular interest will be examined in detail: the test of nullity of only one coefficient and the test of conditional homoscedasticity, which corresponds to the nullity of all the coefficients α i and β j . Section 8.4 is devoted to testing the adequacy of a particular GARCH(p, q) model, using portmanteau tests. The chapter also contains a numerical application in which the pre‐eminence of the GARCH(1, 1) model is questioned.

8.1 Test of the Second‐Order Stationarity Assumption

For the GARCH(p, q) model defined by Eq. (7.1), testing for second‐order stationarity involves testing

equation

Introducing the vector c = (0, 1, …, 1) ∈ ℝ p + q + 1 , the testing problem is

8.1 equation

In view of Theorem 7.2, the QMLE images of θ 0 satisfies

equation

under assumptions which are compatible with H 0 and H 1 . In particular, if c θ 0 = 1 we have

equation

It is thus natural to consider the Wald statistic

equation

where images and images are consistent estimators in probability of κ η and J . The following result follows immediately from the convergence of T n to images when c θ 0 = 1.

Note that for most real series (see, for instance Table 7.4), the sum of the estimated coefficients images and images is strictly less than 1: second‐order stationarity thus cannot be rejected, for any reasonable asymptotic level (when T n  < 0, the p ‐value of the test is greater than 1/2). Of course, the non‐rejection of H 0 does not mean that the stationarity is proved. It is interesting to test the reverse assumption that the data‐generating process is an IGARCH, or more generally that it does not have moments of order 2. We thus consider the problem

8.2 equation

As an application, we take up the data sets of Table 7.4 again, and we give the p ‐values of the previous test for the 11 series of daily returns. For the FTSE (DAX, Nasdaq, and S&P 500), the assumption of infinite variance cannot be rejected at the 5% (3%, 2%, and 1%) level (see Table 8.1). The other series can be considered as second‐order stationary (if one believes in the GARCH(1, 1) model, of course).

Test of the infinite variance assumption for 11 stock market returns.

Index images p ‐Value
CAC 0.983 (0.007) 0.0089
DAX 0.981 (0.011) 0.0385
DJA 0.982 (0.007) 0.0039
DJI 0.986 (0.006) 0.0061
DJT 0.983 (0.009) 0.0023
DJU 0.983 (0.007) 0.0060
FTSE 0.990 (0.006) 0.0525
Nasdaq 0.993 (0.003) 0.0296
Nikkei 0.980 (0.007) 0.0017
SMI 0.962 (0.015) 0.0050
S&P 500 0.989 (0.005) 0.0157

Estimated standard deviations are in parentheses.

8.2 Asymptotic Distribution of the QML When θ 0 is at the Boundary

In view of the form of the parameter space (7.3), the QMLE images is constrained to have a strictly positive first component, while the other components are constrained to be positive or null. A general technique for determining the distribution of a constrained estimator involves expressing it as a function of the unconstrained estimator images (see Gouriéroux and Monfort 1995). For the QMLE of a GARCH, this technique does not work because the objective function

equation

cannot be computed outside Θ (for an ARCH(1), it may happen that images is negative when α 1 < 0). It is thus impossible to define images .

The technique that we will utilise here (see, in particular, Andrews 1999), involves writing images with the aid of the normalised score vector, evaluated at θ 0 :

8.3 equation

with

equation

where the components of l n (θ 0)/∂θ and of J n are right derivatives (see (a) in the proof of Theorem 8.1 below).

In the proof of Theorem 7.2, we showed that

8.4 equation

For any value of θ 0 ∈ Θ (even when images ), it will be shown that the vector Z n is well defined and satisfies

8.5 equation

provided J exists. By contrast, when

equation

Result (8.4) is no longer valid. However, we will show that the asymptotic distribution of images is well approximated by that of the vector n 1/2(θ − θ 0) which is located at the minimal distance of Z n , under the constraint θ ∈ Θ. Consider thus a random vector images (which is not an estimator, of course) solving the minimisation problem

8.6 equation

It will be shown that J n converges to the positive definite matrix J . For n large enough, we thus have

equation

where images is a distance between two points x and y of p + q + 1 .

We allow θ 0 to have null components, but we do not consider the (less interesting) case where θ 0 reaches another boundary of Θ. More precisely, we assume that

equation

where images and images . In this case images and images belong to the ‘local parameter space’

8.7 equation

where Λ1 = ℝ, for i = 2, …, p + q + 1, Λ i  = ℝ if θ 0i  ≠ 0 and Λ i  = [0, ∞) if θ 0i  = 0. With the notation

equation

we thus have, with probability 1,

8.8 equation

The vector images is the projection of Z n on Λ, with respect to the norm images (see Figure 8.1). Since Λ is closed and convex, such a projection is unique. We will show that

8.9 equation
Image described by caption and surrounding text.

Figure 8.1 ARCH(1) model with θ 0 = (ω 0, 0) and images : images is the grey area; images ; images and images have the same asymptotic distribution.

Since (Z n , J n ) tends in law to (Z, J) and images is a function of (Z n , J n ) which is continuous everywhere except at the points where J n is singular (that is, almost everywhere with respect to the distribution of (Z, J) because J is invertible), we have images , where λ Λ is the solution of limiting problem

8.10 equation

In addition to B1, we retain most of the assumptions of Theorem 7.2:

  1. B2: θ 0 ∈ Θ and Θ is a compact set.
  2. B3: γ(A 0) < 0 and for all θ ∈ Θ, images .
  3. B4: images has a non‐degenerate distribution with images .
  4. B5: If p > 0, images , and images do not have common roots, images , and α 0q  + β 0p  ≠ 0.
  5. B6: images .

We also need the following moment assumption:

  1. B7: images .

When images , we can show the existence of the information matrix

equation

without moment assumptions similar to B7. The following example shows that, in the ARCH case, this is no longer possible when we allow θ 0 ∈ ∂Θ.

We then have the following result.

8.2.1 Computation of the Asymptotic Distribution

In this section, we will show how to compute the solutions of the optimization problem ( 8.10). Switching the components of θ , if necessary, it can be assumed without loss of generality that the vector images of the first d 1 components of θ 0 has strictly positive elements and that the vector images of the last d 2 = p + q + 1 − d 1 components of θ 0 is null. This can be written as

8.12 equation

More generally, it will be useful to consider all the subsets of these constraints. Let

equation

be the set of the matrices obtained by deleting no, one, or several (but not all) rows of K. Note that the solution of the constrained minimisation problem ( 8.10) is the unconstrained solution λ = Z when the latter satisfies the constraint, that is, when

equation

When Z ∉ Λ, the solution λ Λ coincides with that of an equality constrained problem of the form

equation

An important difference, compared to the initial minimisation program ( 8.10), is that the minimisation is done here on a vectorial space. The solution is given by a projection (non‐orthogonal when J is not the identity matrix). We thus obtain (see Exercise 8.1)

8.13 equation

is the projection matrix (orthogonal for the metric defined by J) on the orthogonal subspace of the space generated by the rows of K i . Note that images does not necessarily belong to Λ because K i λ = 0 does not imply that  ≥ 0. Let images be the class of the admissible solutions. It follows that the solution that we are looking for is

equation

This formula can be used in practice to obtain realisations of λ Λ from realisations of Z . The images can be obtained by writing

8.14 equation

Another expression (of theoretical interest) for λ Λ is

equation

where images and the images form a partition of p + q + 1 . Indeed, according to the zone to which Z belongs, a solution images is obtained. We will make explicit these formulas in a few examples. Let d = p + q + 1, z + = z images (0, + ∞)(z) and z  = z images (− ∞ , 0)(z).

8.3 Significance of the GARCH Coefficients

We make the assumptions of Theorem 8.1 and use the notation of Section 8.2. Assume images , and consider the testing problem

8.19 equation

Recall that under H 0 , we have

equation

where the distribution of λ Λ is defined by

8.20 equation

with images .

8.3.1 Tests and Rejection Regions

For parametric assumptions of the form (8.19), the most popular tests are the Wald, score, and likelihood ratio tests.

Wald Statistic

The Wald test looks at whether images is close to 0. The usual Wald statistic is defined by

equation

where images is a consistent estimator of Σ = (κ η  − 1)J −1.

Score (or Lagrange Multiplier or Rao) Statistic

Let

equation

denote the QMLE of θ constrained by θ (2) = 0. The score test aims to determine whether images is not too far from 0, using a statistic of the form

equation

where images and images denote consistent estimators of κ η and J .

Likelihood Ratio Statistic

The likelihood ratio test is based on the fact that under H 0 : θ (2) = 0, the constrained (quasi) log‐likelihood images should not be much smaller than the unconstrained log‐likelihood images . The test employs the statistic

equation

Usual Rejection Regions

From the practical viewpoint, the score statistic presents the advantage of only requiring constrained estimation, which is sometimes much simpler than the unconstrained estimation required by the two other tests. The likelihood ratio statistic does not require estimation of the information matrix J , nor the kurtosis coefficient κ η . For each test, it is clear that the null hypothesis must be rejected for large values of the statistic. For standard statistical problems, the three statistics asymptotically follow the same images distribution under the null. At the asymptotic level α , the standard rejection regions are thus

equation

where images is the (1 − α)‐quantile of the χ 2 distribution with d 2 degrees of freedom. In the case d 2 = 1, for testing the significance of only one coefficient, the most widely used test is Student's t test, defined by the rejection region

8.21 equation

where images . This test is equivalent to the standard Wald test because images ( t n being here always positive or null, because of the positivity constraints of the QML estimates) and

equation

Our testing problem is not standard because, by Theorem 8.1, the asymptotic distribution of images is not normal. We will see that, among the previous rejection regions, only that of the score test asymptotically has the level α .

8.3.2 Modification of the Standard Tests

The following proposition shows that for the Wald and likelihood ratio tests, the asymptotic distribution is not the usual images under the null hypothesis. The proposition also shows that the asymptotic distribution of the score test remains the images distribution. The asymptotic distribution of R n is not affected by the fact that, under the null hypothesis, the parameter is at the boundary of the parameter space. These results are not very surprising. Take the example of an ARCH(1) with the hypothesis H 0 : α 0 = 0 of absence of ARCH effect. As illustrated by Figure 8.2, there is a non‐zero probability that images be at the boundary, that is, that images . Consequently images admits a mass at 0 and does not follow, even asymptotically, the images law. The same conclusion can be drawn for the likelihood ratio test. On the contrary, the score n 1/2 l n (θ 0)/∂θ can take as well positive or negative values, and does not seem to have a specific behaviour when θ 0 is at the boundary.

Image described by caption and surrounding text.

Figure 8.2 Concentrated log‐likelihood (solid line) images for an ARCH(1) model. Assume there is no ARCH effect: the true value of the ARCH parameter is α 0 = 0. In the configuration on the right, the likelihood maximum does not lie at the boundary and the three statistics W n , R n , and L n take strictly positive values. In the configuration on the left, we have images and images , whereas images continues to take a strictly positive value.

8.3.3 Test for the Nullity of One Coefficient

Consider the case d 2 = 1, which is perhaps the most interesting case and corresponds to testing the nullity of only one coefficient. In view of relations (8.15), the last component of λ Λ is equal to images . We thus have

equation

where images . Using the symmetry of the Gaussian distribution, and the independence between Z *2 and images when Z * follows the real normal law, we obtain

equation

Testing

equation

can thus be achieved by using the critical region images at the asymptotic level α ≤ 1/2. In view of Remark 8.3, we can define a modified likelihood ratio test of critical region images . Note that the standard Wald test images has the asymptotic level α/2, and that the asymptotic level of the standard likelihood ratio test images is much larger than α when the kurtosis coefficient κ η is large. A modified version of the Student t test is defined by the rejection region

8.25 equation

We observe that commercial software – such as GAUSS, R, RATS, SAS, and SPSS – do not use the modified version (8.25), but the standard version (8.21). This standard test is not of asymptotic level α but only α/2. To obtain a t test of asymptotic level α , it then suffices to use a test of nominal level 2α .

8.3.4 Conditional Homoscedasticity Tests with ARCH Models

Another interesting case is that obtained with d 1 = 1, θ (1) = ω , p = 0, and d 2 = q . This case corresponds to the test of the conditional homoscedasticity null hypothesis

8.26 equation

in an ARCH(q) model

8.27 equation

We will see that for testing problem (8.26), there exist very simple forms of the Wald and score statistics.

Simplified Form of the Wald Statistic

Using Exercise 8.6, we have

equation

Since KΣK  = I q , we obtain a very simple form for the Wald statistic:

8.28 equation

Asymptotic Distribution W and L

A trivial extension of Example 8.3 yields

8.29 equation

The asymptotic distribution of images is thus that of

equation

wherethe Z i 's are independent and images -distributed. Thus, in the case where an ARCH(q) is fitted to a white noise we have

8.30 equation

This asymptotic distribution is tabulated and the critical values are given in Table 8.2. In view of Remark 8.3, Table  8.2 also yields the asymptotic critical values of the modified likelihood ratio statistic images . Table 8.3 shows that the use of the standard images ‐based critical values of the Wald test would lead to large discrepancies between the asymptotic levels and the nominal level α .

Asymptotic critical value c q, α , at level α , of the Wald test of rejection region images for the conditional homoscedasticity hypothesis H 0 : α 1 = ⋯ = α q  = 0 in an ARCH(q) model.

q α (%)
0.1 1 2.5 5 10 15
1  9.5493  5.4119 3.8414 2.7055 1.6424 1.0742
2 11.7625  7.2895 5.5369 4.2306 2.9524 2.2260
3 13.4740  8.7464 6.8610 5.4345 4.0102 3.1802
4 14.9619 10.0186 8.0230 6.4979 4.9553 4.0428
5 16.3168 11.1828 9.0906 7.4797 5.8351 4.8519

Exact asymptotic level (%) of erroneous Wald tests, of rejection region images , under the conditional homoscedasticity assumption H 0 : α 1 = ⋯ = α q  = 0 in an ARCH(q) model.

q α (%)
0.1 1 2.5 5 10 15
1 0.05 0.5 1.25 2.5 5 7.5
2 0.04 0.4 0.96 1.97 4.09 6.32
3 0.02 0.28 0.75 1.57 3.36 5.29
4 0.02 0.22 0.59 1.28 2.79 4.47
5 0.01 0.17 0.48 1.05 2.34 3.81

Score Test

For the hypothesis (8.26) that all the α coefficients of an ARCH(q) model are equal to zero, the score statistic R n can be simplified. To work within the linear regression framework, write

equation

where Y is the vector of length n of the ‘dependent’ variable images , where X is the n × (q + 1) matrix of the constant images (in the first column) and of the ‘explanatory’ variables images (in column i + 1, with the convention ε t  = 0 for t ≤ 0), and images . Estimating J(θ 0) by n −1 X X , and κ η  − 1 by n −1 Y Y , we obtain

equation

and one recognizes n times the coefficient of determination in the linear regression of Y on the columns of X . Since this coefficient is not changed by linear transformation of the variables (see Exercise 5.11), we simply have R n  = nR 2 , where R 2 is the coefficient of determination in the regression of images on a constant and q lagged values images . Under the null hypothesis of conditional homoscedasticity, R n asymptotically follows the images law.

The previous simple forms of the Wald and score tests are obtained with estimators of J which exploit the particular form of the matrix under the null. Note that there exist other versions of these tests, obtained with other consistent estimators of J . The different versions are equivalent under the null, but can have different asymptotic behaviours under the alternative.

8.3.5 Asymptotic Comparison of the Tests

The Wald and score tests that we have just defined are in general consistent, that is, their powers converge to 1 when they are applied to a wide class of conditionally heteroscedastic processes. An asymptotic study will be conducted via two different approaches: Bahadur's approach compares the rates of convergence to zero of the p ‐values under fixed alternatives, whereas Pitman's approach compares the asymptotic powers under a sequence of local alternatives, that is, a sequence of alternatives tending to the null as the sample size increases.

Bahadur's Approach

Let S W (t) = ℙ(W > t) and S R (t) = ℙ(R > t) be the asymptotic survival functions of the two test statistics, under the null hypothesis H 0 defined by ( 8.26). Consider, for instance, the Wald test. Under the alternative of an ARCH(q) which does not satisfy H 0 , the p ‐value of the Wald test S W (W n ) converges almost surely to zero as n → ∞ because

equation

The p‐value of a test is typically equivalent to exp{−nc/2}, where c is a positive constant called the Bahadur slope. Using the fact that

8.31 equation

and that images , the (approximate 1 ) Bahadur slope of the Wald test is thus

equation

To compute the Bahadur slope of the score test, note that we have the linear regression model images is the linear innovation of images . We then have

equation

The previous limit is thus equal to the Bahadur slope of the score test. The comparison of the two slopes favours the score test over the Wald test.

Bahadur's approach is sometimes criticised for not taking account of the critical value of test, and thus for not really comparing the powers. This approach only takes into account the (asymptotic) distribution of the statistic under the null and the rate of divergence of the statistic under the alternative. It is unable to distinguish a two‐sided test from its one‐sided counterpart (see Exercise 8.8). In this sense, the result of Proposition 8.4 must be put into perspective.

Pitman's Approach

In the ARCH(1) case, consider a sequence of local alternatives images . We can show that under this sequence of alternatives,

equation

Consequently, the local asymptotic power of the Wald test is

8.32 equation

The score test has the local asymptotic power

8.33 equation

Note that the probability in (8.32) is the power of the test of the assumption H 0 : θ = 0 against the assumption H 1 : θ = τ > 0, based on the rejection region of {X > c 1} with only one observation images . The power (8.33) is that of the two‐sided test {∣X ∣  > c 2}. The tests {X > c 1} and {∣X ∣  > c 2} have the same level, but the first test is uniformly more powerful than the second (by the Neyman–Pearson lemma, {X > c 1} is even uniformly more powerful than any test of level less than or equal to α , for any one‐sided alternative of the form H 1 ). The local asymptotic power of the Wald test is thus uniformly strictly greater than that of Rao's test for testing for conditional homoscedasticity in an ARCH(1) model.

Consider the ARCH(2) case, and a sequence of local alternatives images . Under this sequence of alternatives

equation

with images Let c 1 be the critical value of the Wald test of level α . The local asymptotic power of the Wald test is

equation

Let c 2 be the critical value of the Rao test of level α . The local asymptotic power of the Rao test is

equation

where (U 1 + τ 1)2 + (U 2 + τ 2)2 follows a non‐central χ 2 distribution, with two degrees of freedom and non‐centrality parameter images . Figure 8.5 compares the powers of the two tests when τ 1 = τ 2 .

Image described by caption and surrounding text.

Figure 8.5 Local asymptotic power of the Wald test (solid line) and of the Rao test (dotted line) for testing for conditional homoscedasticity in an ARCH(2) model.

Thus, the comparison of the local asymptotic powers clearly favours the Wald test over the score test, counter‐balancing the result of Proposition 8.4.

8.4 Diagnostic Checking with Portmanteau Tests

To check the adequacy of a given time series model, for instance an ARMA(p, q) model, it is common practice to test the significance of the residual autocorrelations. In the GARCH framework this approach is not relevant because the process images is always a white noise (more precisely a martingale difference) even when the volatility is misspecified, that is, when images with images . To check the adequacy of a volatility model, for instance a GARCH( p, q) of the form (7.1), it is much more fruitful to look at the squared residual autocovariances

equation

where h ∣  < n , images , images is defined by Eq. (7.4) and images is the QMLE given by Eq. (7.9).

For any fixed integer m , 1 ≤ m < n , consider the statistic images . Let images and images be weakly consistent estimators of κ η and J . For instance, one can take

equation

Define also the m × (p + q + 1) matrix images whose (h, k)th element, for 1 ≤ h ≤ m and 1 ≤ k ≤ p + q + 1, is given by

equation

8.5 Application: Is the GARCH(1,1) Model Overrepresented?

The GARCH(1,1) model is by far the most widely used by practitioners who wish to estimate the volatility of daily returns. In general, this model is chosen a priori, without implementing any statistical identification procedure. This practice is motivated by the common belief that the GARCH(1,1) (or its simplest asymmetric extensions) is sufficient to capture the properties of financial series and that higher‐order models may be unnecessarily complicated.

We will show that, for a large number of series, this practice is not always statistically justified. We consider daily and weekly series of 11 returns (CAC, DAX, DJA, DJI, DJT, DJU, FTSE, Nasdaq, Nikkei, SMI and S&P 500) and five exchange rates. The observations cover the period from 2 January 1990 to 22 January 2009 for the daily returns and exchange rates, and from 2 January 1990 to 20 January 2009 for the weekly returns (except for the indices for which the first observations are after 1990). We begin with the portmanteau tests defined in Section 8.4 . Table 8.4 shows that the ARCH models (even with large order q) are generally rejected, whereas the GARCH(1,1) is only occasionally rejected. This table only concerns the daily returns, but similar conclusions hold for the weekly returns and exchange rates. The portmanteau tests are known to be omnibus tests, powerful for a broad spectrum of alternatives. As we will now see, for the specific alternatives for which they are built, the tests defined in Section 8.3 (Wald, score, and likelihood ratio) may be much more powerful.

Portmanteau test p‐values for adequacy of the ARCH(5) and GARCH(1,1) models for daily returns of stock market indices, based on m squared residual autocovariances.

M
1 2 3 4 5 6 7 8 9 10 11 12
Portmanteau tests for adequacy of the ARCH(5)
CAC 0.194 0.010 0.001 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000
DAX 0.506 0.157 0.140 0.049 0.044 0.061 0.080 0.119 0.140 0.196 0.185 0.237
DJA 0.441 0.34 0.139 0.002 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000
DJI 0.451 0.374 0.015 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000
DJT 0.255 0.514 0.356 0.044 0.025 0.013 0.020 0.023 0.000 0.000 0.000 0.000
DJU 0.477 0.341 0.002 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000
FTSE 0.139 0.001 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000
Nasdaq 0.025 0.031 0.001 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000
Nikkei 0.004 0.000 0.001 0.001 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000
SMI 0.502 0.692 0.407 0.370 0.211 0.264 0.351 0.374 0.463 0.533 0.623 0.700
S&P 500 0.647 0.540 0.012 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000
Portmanteau tests for adequacy of the GARCH(1,1)
CAC 0.312 0.379 0.523 0.229 0.301 0.396 0.495 0.578 0.672 0.660 0.704 0.743
DAX 0.302 0.583 0.574 0.704 0.823 0.901 0.938 0.968 0.983 0.989 0.994 0.995
DJA 0.376 0.424 0.634 0.740 0.837 0.908 0.838 0.886 0.909 0.916 0.938 0.959
DJI 0.202 0.208 0.363 0.505 0.632 0.742 0.770 0.812 0.871 0.729 0.748 0.811
DJT 0.750 0.100 0.203 0.276 0.398 0.518 0.635 0.721 0.804 0.834 0.885 0.925
DJU 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000
FTSE 0.733 0.940 0.934 0.980 0.919 0.964 0.328 0.424 0.465 0.448 0.083 0.108
Nasdaq 0.523 0.024 0.061 0.019 0.001 0.001 0.002 0.001 0.002 0.001 0.001 0.002
Nikkei 0.049 0.146 0.246 0.386 0.356 0.475 0.567 0.624 0.703 0.775 0.718 0.764
SMI 0.586 0.758 0.908 0.959 0.986 0.995 0.996 0.999 0.999 0.999 0.999 0.999
S&P 500 0.598 0.364 0.528 0.643 0.673 0.394 0.512 0.535 0.639 0.432 0.496 0.594

p‐values less than 5% are in bold, those less than 1% are underlined.

The GARCH(1,1) model is chosen as the benchmark model and is successively tested against the GARCH(1,2), GARCH(1,3), GARCH(1,4), and GARCH(2,1) models. In each case, the three tests (Wald, score, and likelihood ratio) are applied. The empirical p‐values are displayed in Table 8.5. This table shows that (i) the results of the tests strongly depend on the alternative; (ii) the p‐values of the three tests can be quite different; (iii) for most of the series, the GARCH(1,1) model is clearly rejected. Point (ii) is not surprising because the asymptotic equivalence between the three tests is only shown under the null hypothesis or under local alternatives. Moreover, because of the positivity constraints, it is possible (see, for instance, the DJU) that the estimated GARCH(1,2) model satisfies images with images . In this case, when the estimators lie at the boundary of the parameter space and the score is strongly positive, the Wald and LR tests do not reject the GARCH(1,1) model, whereas the score does reject it. In other situations, the Wald or LR test rejects the GARCH(1,1), whereas the score does not (see, for instance the DAX for the GARCH(1,4) alternative). This study shows that it is often relevant to employ several tests and several alternatives. The conservative approach of Bonferroni (rejecting if the minimal p‐value multiplied by the number of tests is less than a given level α ), leads to rejection of the GARCH(1,1) model for 16 out of the 24 series in Table  8.5. Other procedures, less conservative than that of Bonferroni, could also be applied (see Wright 1992) without changing the general conclusion.

p‐values for tests of the null of a GARCH(1,1) model against the GARCH(1,2), GARCH(1,3), GARCH(1,4), and GARCH(2,1) alternatives, for returns of stock market indices and exchange rates.

Alternative
GARCH(1,2) GARCH(1,3) GARCH(1,4) GARCH(2,1)
W n R n L n W n R n L n W n R n L n W n R n L n
Daily returns of indices
CAC 0.007 0.033 0.013 0.005 0.000 0.001 0.024 0.188 0.040 0.500 0.280 0.500
DAX 0.002 0.001 0.003 0.001 0.000 0.000 0.001 0.162 0.014 0.350 0.031 0.143
DJA 0.158 0.337 0.166 0.259 0.285 0.269 0.081 0.134 0.064 0.500 0.189 0.500
DJI 0.044 0.100 0.049 0.088 0.071 0.094 0.107 0.143 0.114 0.500 0.012 0.500
DJT 0.469 0.942 0.470 0.648 0.009 0.648 0.519 0.116 0.517 0.369 0.261 0.262
DJU 0.500 0.000 0.500 0.643 0.000 0.643 0.725 0.001 0.725 0.017 0.000 0.005
FTSE 0.080 0.122 0.071 0.093 0.223 0.083 0.213 0.423 0.205 0.458 0.843 0.442
Nasdaq 0.469 0.922 0.468 0.579 0.983 0.578 0.683 0.995 0.702 0.500 0.928 0.500
Nikkei 0.004 0.002 0.004 0.042 0.332 0.081 0.052 0.526 0.108 0.238 0.000 0.027
SMI 0.224 0.530 0.245 0.058 0.202 0.063 0.086 0.431 0.108 0.500 0.932 0.500
SP 500 0.053 0.079 0.047 0.089 0.035 0.078 0.055 0.052 0.043 0.500 0.045 0.500
Weekly returns of indices
CAC 0.017 0.143 0.049 0.028 0.272 0.068 0.061 0.478 0.142 0.500 0.575 0.500
DAX 0.154 0.000 0.004 0.674 0.798 0.674 0.667 0.892 0.661 0.043 0.000 0.000
DJA 0.194 0.001 0.052 0.692 0.607 0.692 0.679 0.899 0.597 0.003 0.000 0.000
DJI 0.173 0.000 0.030 0.682 0.482 0.682 0.788 0.358 0.788 0.000 0.000 0.000
DJT 0.428 0.623 0.385 0.628 0.456 0.628 0.693 0.552 0.693 0.002 0.000 0.004
DJU 0.500 0.747 0.500 0.646 0.011 0.646 0.747 0.038 0.747 0.071 0.003 0.017
FTSE 0.188 0.484 0.222 0.183 0.534 0.214 0.242 0.472 0.272 0.500 0.532 0.500
Nasdaq 0.441 0.905 0.448 0.387 0.868 0.412 0.199 0.927 0.266 0.069 0.961 0.344
Nikkei 0.500 0.140 0.500 0.310 0.154 0.260 0.330 0.316 0.462 0.030 0.138 0.053
SMI 0.500 0.720 0.500 0.217 0.144 0.150 0.796 0.754 0.796 0.314 0.769 0.360
SP 500 0.117 0.000 0.001 0.659 0.114 0.659 0.724 0.051 0.724 0.000 0.000 0.000
Daily exchange rates
$/€ 0.452 0.904 0.452 0.194 0.423 0.181 0.066 0.000 0.015 0.500 0.002 0.500
¥/€ 0.037 0.000 0.002 0.616 0.090 0.618 0.304 0.000 0.227 0.136 0.000 0.000
£/€ 0.439 0.879 0.440 0.471 0.905 0.464 0.677 0.981 0.677 0.258 0.493 0.248
CHF/€ 0.141 0.000 0.012 0.641 0.152 0.641 0.520 0.154 0.562 0.012 0.000 0.000
C$/€ 0.500 0.268 0.500 0.631 0.714 0.631 0.032 0.000 0.002 0.045 0.045 0.029

p‐values less than 5% are in bold, those less than 1% are underlined.

In conclusion, this study shows that the GARCH(1,1) model is certainly overrepresented in empirical studies. The tests presented in this chapter are easily implemented and lead to selection of GARCH models that are more elaborate than the GARCH(1,1).

8.6 Proofs of the Main Results*

8.7 Bibliographical Notes

It is well known that when the parameter is at the boundary of the parameter space, the maximum likelihood estimator does not necessarily satisfy the first‐order conditions and, in general, does not admit a limiting normal distribution. The technique, employed in particular by Chernoff (1954) and Andrews (1997) in a general framework, involves approximating the quasi‐likelihood by a quadratic function, and defining the asymptotic distribution of the QML as that of the projection of a Gaussian vector on a convex cone. Particular GARCH models are considered by Andrews (1997, 1999) and Jordan (2003). The general GARCH(p, q) case is considered by Francq and Zakoïan (2007). A proof of Theorem 8.1, when the moment assumption B7 is replaced by assumption B7 of Remark 8.2, can be found in the latter reference. When the nullity of GARCH coefficients is tested, the parameter is at the boundary of the parameter space under the null, and the alternative is one‐sided. Numerous works deal with testing problems where, under the null hypothesis, the parameter is at the boundary of the parameter space. Such problems have been considered by Chernoff (1954), Bartholomew (1959), Perlman (1969), and Gouriéroux, Holly, and Monfort (1982), among many others. General one‐sided tests have been studied by, for instance Rogers (1986), Wolak (1989), Silvapulle and Silvapulle (1995), and King and Wu (1997). Papers dealing more specifically with ARCH and GARCH models are Lee and King (1993), Hong (1997), Demos and Sentana (1998), Andrews (2001), Hong and Lee (2001), Dufour et al. (2004), Francq and Zakoïan (2009b) and Pedersen and Rahbek (2018).

The portmanteau tests based on the squared residual autocovariances were proposed by McLeod and Li (1983), Li and Mak (1994), and Ling and Li (1997). The results presented here closely follow Berkes, Horváth, and Kokoszka (2003a). Problems of interest that are not studied in this book are the tests on the distribution of the iid process (see Horváth, Kokoszka, and Teyssiére 2004; Horváth and Zitikis 2006).

Concerning the overrepresentation of the GARCH(1, 1) model in financial studies, we mention Stărică (2006). This paper highlights, on a very long S&P 500 series, the poor performance of the GARCH(1, 1) in terms of prediction and modelling, and suggests a non‐stationary dynamics of the returns.

8.8 Exercises

8.1 (Minimisation of a distance under a linear constraint)

Let J be an n × n invertible matrix, let x 0 be a vector of n , and let K be a full‐rank p × n matrix, p ≤ n . Solve the problem of the minimisation of Q(x) = (x − x 0) J(x − x 0) under the constraint Kx = 0.

  1. 8.2 (Minimisation of a distance when some components are equal to zero) Let J be an n × n invertible matrix, x 0 a vector of n , and p < n . Minimise Q(x) = (x − x 0) J(x − x 0) under the constraints images ( x i denoting the i th component of x , and assuming that 1 ≤ i 1 < ⋯ < i p  ≤ n ).
  2. 8.3 (Lagrangian or method of substitution for optimisation with constraints) Compare the solutions of the optimisation problem of Exercise 8.2, with
    equation

    and the constraints

    1. x 3 = 0,
    2. x 2 = x 3 = 0.
  3. 8.4 (Minimisation of a distance under inequality constraints) Find the minimum of the function
    equation

    under the constraints λ 2 ≥ 0 and λ 3 ≥ 0, when

    1. Z = (−2, 1, 2) ,
    2. Z = (−2, − 1, 2) ,
    3. Z = (−2, 1, − 2) ,
    4. Z = (−2, − 1, − 2) .
  4. 8.5 (Influence of the positivity constraints on the moments of the QMLE) Compute the mean and variance of the vector λ Λ defined by equality (8.16). Compare these moments with the corresponding moments of Z = (Z 1, Z 2, Z 3) .
  5. 8.6 (Asymptotic distribution of the QMLE of an ARCH in the conditionally homoscedastic case) For an ARCH( q ) model, compute the matrix Σ involved in the asymptotic distribution of the QMLE in the case where all the α 0i are equal to zero.
  6. 8.7 (Asymptotic distribution of the QMLE when an ARCH(1) is fitted to a strong white noise) Let images be the QMLE in the ARCH(1) model images when the true parameter is equal to (ω 0, α 0) = (ω 0, 0) and when images . Give an expression for the asymptotic distribution of images with the aid of
    equation

    Compute the mean vector and the variance matrix of this asymptotic distribution. Determine the density of the asymptotic distribution of images . Give an expression for the kurtosis coefficient of this distribution as function of κ η .

  7. 8.8 (One‐sided and two‐sided tests have the same Bahadur slopes) Let X 1, …, X n be a sample from the (θ, 1) distribution. Consider the null hypothesis H 0 :  θ = 0. Denote by Φ the (0, 1) cumulative distribution function. By the Neyman–Pearson lemma, we know that, for alternatives of the form H 1 :  θ > 0, the one‐sided test of rejection region
    equation

    is uniformly more powerful than the two‐sided test of rejection region

    equation

    (moreover, C is uniformly more powerful than any other test of level α or less). Although the previous argument shows that the test C is superior to the test C * in finite samples, we will conduct an asymptotic comparison of the two tests, using the Bahadur and Pitman approaches.

    • The asymptotic Bahadur slope c(θ) is defined as the almost sure limit of −2/n times the logarithm of the p ‐value under P θ , when the limit exists. Compare the Bahadur slopes of the two tests.
    • In the Pitman approach, we define a local power around θ = 0 as being the power at images . Compare the local powers of C and C * . Compare also the local asymptotic powers of the two tests for non‐Gaussian samples.
  8. 8.9 (The local asymptotic approach cannot distinguish the Wald, score and likelihood ratio tests) Let X 1,…, X n be a sample of the images distribution, where θ and σ 2 are unknown. Consider the null hypothesis H 0 :  θ = 0 against the alternative H 1 :  θ > 0. Consider the following three tests:
    • images , where
      equation
      is the Wald statistic;
    • images , where
      equation

      is the Rao score statistic;

    • images , where
      equation

      is the likelihood ratio statistic.

      Give a justification for these three tests. Compare their local asymptotic powers and their Bahadur slopes.

  9. 8.10 (The Wald and likelihood ratio statistics have the same asymptotic distribution) Consider the case d 2 = 1, that is, the framework of Section 8.3.3 where only one coefficient is equal to zero. Without using Remark 8.3, show that the asymptotic laws W and L defined by (8.22) and (8.24) are such that
    equation
  10. 8.11 (For testing conditional homoscedasticity, the Wald and likelihood ratio statistics have the same asymptotic distribution) Repeat Exercise 8.10 for the conditional homoscedasticity test ( 8.26) in the ARCH( q ) case.
  11. 8.12 (The product of two independent random variables is null if and only if one of the two variables is null) Let X and Y be two independent random variables such that XY = 0 almost surely. Show that either X = 0 almost surely or Y = 0 almost surely.

Note

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

You can't read the all page of ebook, please click here login for view all page.
Reset
3.136.233.153