Next Article in Journal
Results on Univalent Functions Defined by q-Analogues of Salagean and Ruscheweh Operators
Previous Article in Journal
Well-Posedness for Nonlinear Parabolic Stochastic Differential Equations with Nonlinear Robin Conditions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Self-Weighted Quasi-Maximum Likelihood Estimators for a Class of MA-GARCH Model

1
School of Mathematics and Statistics, Guangxi Normal University, Guilin 541004, China
2
Faculty of Science and Technology, University of Macau, Macau 999078, China
*
Author to whom correspondence should be addressed.
Symmetry 2022, 14(8), 1723; https://doi.org/10.3390/sym14081723
Submission received: 27 July 2022 / Revised: 14 August 2022 / Accepted: 15 August 2022 / Published: 18 August 2022
(This article belongs to the Section Mathematics)

Abstract

:
In financial time series analysis, symmetric and asymmetric GARCH models have become essential models for measuring the characteristics of economic volatility. In this article, we propose the consistency and asymptotic normality properties of the self-weighted quasi-maximum likelihood estimation without assuming the existence of the second moment for the moving average model with a class of GARCH error. Numerical simulation shows that the parameter estimation performs well; empirical analysis shows that the self-weighted quasi-maximum likelihood estimation of the moving average model with a class of GARCH error can improve the data fitting effect and prediction ability.

1. Introduction

With the progress and development of multivariate time series theory and technology, to accurately describe the dynamic time-varying characteristics of the volatility of financial time series data, Engle [1] proposed an ARCH model for the volatility of the UK inflation rate, and Bollerslev [2] extended it to a GARCH model. On this basis, many experts and scholars at home and abroad have done much research on the GARCH model in combination with the asymmetry and long memory of volatility. For example, given that the impact of positive and negative news on stock market volatility is asymmetric, Nelson [3] proposed the exponential GARCH (EGARCH) model; Glosten, Jagannathan, and Runkle [4] proposed the asymmetric GARCH (GJR) model; H. Viet Long and Bin Jebreen et al. [5] employed the generalized autoregressive conditional heteroskedasticity (GARCH) model for risk management as their main contribution. However, in the traditional GARCH model, the conditional heteroscedasticity is a function of the unobservable lag residuals square sequence, which makes the parameter estimation of this model complex. Li and Zhang et al. [6] combined the ideas of a factor model and a symmetric GARCH model to describe the dynamics of a high-dimensional conditional covariance matrix. Therefore, Baillie [7] described the population characteristics of various long memory processes in FIGARCH. The conditional variance of the process implies a slow hyperbolic decay rate for the influence of lagged squared innovations. Ling [8] proposed the DAR(p) model based on the ARCH model. In this way, a better estimate of the parameters can be obtained. Moreover, Ling [9] proposed the consistency and asymptotic properties of self-weighted quasi-maximum likelihood estimation in the presence of fractional moments for the ARMA-GARCH model. Then, Zhu and Ling [10] proposed the strong coincidence and asymptotic properties of global self-weighted quasi-maximal exponential likelihood estimation for the ARMA-GARCH model in the presence of fractional moments. Many models with GARCH also have some properties of similar analysis. For example, Pan and Chen et al. [11] discussed the asymptotic properties of WLAD estimation for the ARFIMA model under stationary and non-stationary conditions. For long-memory and heteroscedastic models, Ramírez-Parietti and Contreras-Reyes et al. [12] computed cross-sample entropy measures for synchrony level, which involved ARMA, ARFIMA, ARMA-GARCH, and FIGARCH processes.
Based on the classical ARCH model, Ling [13] proposed the DAR model to make the parameter estimation of the DAR model become easier to estimate and proved the consistency and asymptotic properties of the QMLE estimation without assuming the existence of the second-order moment of the sequence. Then Zhu and Zhang et al. [14] proposed a class of MA-GARCH models (the DAR(p) model p tends to infinity), proving that Ling has done similar work without assuming the existence of second-order moments. However, when there are extreme data, the parameter weights estimated by QMLE are all 1, which has a specific error. Thus, Ling [9] proved that the self-weighted QMLE estimation of the RAMA-GARCH model is consistent and asymptotic in the presence of fractional moments. The purpose is to use weights to reduce leverage points and make estimates more precise and more efficient.
In this paper, to reduce the error, we perform a self-weighted QMLE estimation based on Zhu and Zhang et al. [14] Giving different weights to different values makes our estimates more robust, proving some properties of the forecast. The rest of the article is organized as follows. Section 2 introduces the consistency and asymptotic properties of estimation; Section 3 presents the numerical simulation of self-weighted QMLE estimation. Section 4 gives a real example; Section 5 concludes the article. The proof is in Appendix A.

2. Self-Weighted QMLE

A class of MA-GARCH model can be written as
y t = ϕ ε t 1 + ε t
ε t = e t h t
h t = ω + α y t 1 2 + β h t 1
where | ϕ | < 1 , ω , α > 0 , 0 < β < 1 and when t > s , { y t } and { e t } are independent of each other. Denote δ = ( ω , α , β ) , and θ = ( ϕ , δ ) Θ , and Θ = ( θ : 1 < ϕ ̲ < ϕ < ϕ ¯ < 1 , 0 < ω ̲ < ω < ω ¯ , 0 < α ̲ < α < α ¯ , 0 < β ̲ < β < β ¯ < 1 ) . Denote Φ ( z ) = 1 + ϕ z , β ( z ) = 1 β z , Φ 1 ( z ) = i = 0 a Φ ( i ) z i , s u p Θ a ( i ) = O ( ρ 0 i ) , β 1 ( z ) = i = 0 β i z i , where 0 < ρ 0 < 1 .
Compared to traditional MA-GARCH model, the previous ε t 1 2 in (3) is instead denoted by y t 1 2 . A class of MA-GARCH model is easier to predict than the traditional MA-GARCH model.
Remark 1.
With simple iteration, the model can be transformed into:
y t = i = 1 ( 1 ) i 1 ϕ i y t i + e t ω / ( 1 β ) + α j = 1 β j 1 y t j 2
From this, we can see that, compared with the DAR(p) model proposed by Ling [14], the above formula is equivalent to the order p of the conditional mean, and the conditional variance tends to infinity. Then, we refer to Zhu [15] to obtain the parameter space of the above models (1)–(3).
We introduce the following conditions:
A1:
θ 0 is an interior point in Θ , and { y t } is stationary, ergodic, and identifiable.
A2:
{ e t } is independent and identically distributed with a mean of 0 and a variance of 1.
A3:
{ e t 2 } has a nondegenerate distribution with E e t 4 < .
A4:
ω t = ω ( y t 1 , y t 2 , ) and ω t is a measurable, positive, and bounded function on R Z 0 with E [ ω t ( ξ ρ , t 1 2 + ζ ρ , t 1 2 + ξ ρ , t 1 2 ζ ρ , t 1 2 ) ] < , and Z 0 = { 0 , 1 , 2 , } , ρ ( 0 , 1 ) .
Given the observations { y n , , y 1 } and the initial values { y 0 , y 1 , y 2 , } , we can rewrite the parametric models (1)–(3) as
ε t = y t ϕ ε t 1
ε t = e t h t
h t = ω + α y t 1 2 + β h t 1
According to Yoshida’s [16] research on the asymptotic properties of the likelihood estimators of various nonlinear stochastic processes, for convenience, the weighted log-quasi-likelihood function (ignoring a constant) can be written as follows:
L s n = 1 n t = 1 ω t l t ( θ ) , l t ( θ ) = log h t ( θ ) + ε t 2 ( θ ) h t ( θ )
Remark 2.
This is important because it ensures that the maximum value of the log of the probability occurs at the same point as the original probability function. The quasi-score function and the quasi-information matrix are given in the Appendix A. Equation (A11) in the Appendix A.1 shows that
s u p Θ 2 l t ( θ ) ϕ ϕ O ( 1 ) ξ ρ , t 1 2
It is possible to obtain an estimator such that it is asymptotic normal if we can downweight ξ ρ , t 1 2 .
We look for the minimizer, θ ^ s n = ( ϕ ^ s n , δ ^ s n ) , of L s n on Θ ; that is,
θ ^ s n = a r g m i n Θ L s n ( θ )
We first give two lemmas.
Lemma 1.
Let ξ ρ , t = 1 + Σ i = 0 ρ i | y t i | , where 0 < ρ < 1 . Then there exists a constant ρ and a neighborhood Θ 0 of θ 0 such that:
( i )
s u p Θ | ε t 1 ( θ ) | c ξ ρ , t 1 ;
( i i )
s u p Θ ε t ( θ ) ϕ c ξ ρ , t 1 ;
( i i i )
s u p Θ 2 ε t ( θ ) ϕ ϕ c ξ ρ , t 1 . .
Lemma 2.
Let ζ ρ , t = 1 + i = 0 ρ i y t 1 2 . Then there exists a constant ρ and a neighborhood Θ 0 of θ 0 , and 0 < β < ρ < 1 . Then:
( i )
s u p Θ h t ( θ ) c ζ ρ , t 1
( i i )
s u p Θ 1 h t ( θ ) h t ( θ ) δ c ζ ρ , t 1
( i i i )
s u p Θ 1 h t ( θ ) 2 h t ( θ ) δ δ c ζ ρ , t 1
( i v )
1 h t ( θ ) h t ( θ ) ϕ = 0
We now can state our main results as follows.
Theorem 1.
If Assumptions A1–A4 hold, then
( i )
θ ^ s n p θ 0
( i i )
n ( θ ^ s n θ 0 ) d N ( 0 , 0 1 Ω 0 0 1 ) where
Σ 0 = E 2 ω t h t ( θ 0 ) ε t ( θ 0 ) ϕ ε t ( θ 0 ) ϕ 0 0 ω t h t 2 ( θ ) h t ( θ 0 ) δ h t ( θ 0 ) δ
Ω 0 = E 4 ω t 2 h t ( θ 0 ) ( ε t ( θ 0 ) ϕ ) 2 0 0 ω t 2 ς h t 2 ( θ 0 ) ( h t ( θ 0 ) δ ) 2
ς = E e t 4 1
ε t ( θ ) ϕ = i = 1 ( 1 ) i i ϕ i 1 y t i
h t ( θ ) δ = [ 1 1 β , j = 1 β j 1 y t j 2 , j = 1 β j 1 h t j ]

3. Simulation

In this section, on the basis of a class of MA-GRACH model, we compared SQMLE with QMLE. In the simulation, we studied the cases when { e t } has N ( 0 , 1 ) , L a p l a c e ( 0 , 1 ) , and t 3 . Then we took the sequences { y t } that are generated by a class of MA-GARCH models, where θ 0 = ( 0.5 , 0.1 , 0.4 , 0.3 ) . We needed to select a weight ω t . It seemed reasonable to use the following weight analog Ling [17] and Huber [18] influence function:
ω t = 1 , a t = 0 ; c 3 a t 3 , a t 0
where a t = | y t 1 | I ( | y t 1 | c ) , c is the 95th quantile of the sequence { y t } . This weight satisfies assumption A4. Then, For the models (1)–(3), we compared the parameter estimation methods of Zhu and Zhang, etc. [14]. We set the sample size at n = 400 , 800 , 1200 and used 1000 replications.
From Table 1, when e t N ( 0 , 1 ) , we can see that all estimators in Table 1 have minimal biases, and the SD and AD of self-weighted QMLE are close. As the sample size n increases, the values of SD and AD both decrease. At the same time, we compare the two estimation methods, the quasi-maximum likelihood estimation(QMLE) and the self-weighted quasi-maximum likelihood estimation (SQMLE), in Table 1. With the increase in n, the bias, AD, and SD shows different trends. The quasi-maximum likelihood estimation and the self-weighted quasi-maximum likelihood estimation values were getting closer and closer. Still, the quasi-maximum likelihood estimation’s bias, AD, and SD are always slightly smaller than the bias, AD, and SD of the self-weighted quasi-maximum likelihood estimate. When this happens, it is reasonable to be pointed out by Ling [9]. Therefore, the simulation results illustrate that self-weighted quasi-maximum likelihood estimates are applicable.
When e t L a p l a c e ( 0 , 1 ) , e t t 3 , respectively, we can see that both Table 2 and Table 3 have similar simulation effects to Table 3. Therefore, the self-weighted quasi-maximum likelihood estimation for a class of MA-GARCH model is valid.
There are many other weights, such as ω t = ( 1 + C | y t | 2 ) 3 / 2 and I ( m a x 1 i p | y t 1 | C ) , that satisfies assumption A4. However, our simulation results that are not reported in this paper show that the self-weighted quasi-maximum likelihood estimated is much more efficient than that based on these weights.

4. A Real Example

In this section, we empirically analyzed the models (1)–(3) by using two real data sets. We used the SQMLE to compare the goodness of fit and prediction effect for a class of MA-GARCH model and traditional MA-GARCH model.
Firstly, we studied the opening price of ST Daji (000564) from 1 January 2018 to 1 January 2022, which has, in total, 837 observations. Let x t be the logarithms of the opening price of ST Daji (000564) in Figure 1, and y t = x t x t 1 , { y t } is plotted in Figure 2. Ling [8] makes a self-weighted quasi-maximum likelihood estimate for ARMA-GARCH.
The ARMA(0,1)-GARCH(0,1) (MA-GARCH) model was adopted to fit { y t } by Ling [8], and the result is
y t = 0.0047 ( 0.0000 ) ε t 1 + ε t
ε t = e t h t = e t 0.0010 ( 0.0059 ) + 0.2703 ( 3.5188 ) ε t 1 2 + 0.0427 ( 0.1955 ) h t 1
where the standard errors are in parentheses.
Similarly, we used models (1)–(3) to fit { y t } , and the result is
y t = 0.0817 ( 0.0487 ) ε t 1 + ε t
ε t = e t h t = e t 0.0010 ( 0.0002 ) + 0.2614 ( 0.1060 ) y t 1 2 + 0.0443 ( 0.1585 ) h t 1
where the standard errors are in parentheses (obtained from Theorem 1).
We compared the root-mean-squared error (RMSE) for the one-step-ahead forecast and log-likelihood value for models (8) and (9) and models (10) and (11), respectively. The results are shown in Table 4. In addition, for the data of the data set from 24 March 2021 to 19 August 2021, we give a 95 percent confidence interval of sequence x t , as shown in Figure 3 below. Figure 4 is a partial enlargement of Figure 3. As can be seen from Table 4, the RMSE value of models (10) and (11) is smaller than that of models (8) and (9), and the log-likelihood value of models (10) and (11) is bigger than that of models (8) and (9). Thus, the goodness of fit of models (10) and (11) is better than models (8) and (9). In addition, as can be seen from Figure 3, at almost all points, models (10) and (11) have a narrower confidence interval than models (8) and (9), and we can see that all points fall within the confidence interval. This shows that under this data, models (10) and (11) are better than models (8) and (9) in terms of goodness of fit and predictive ability. Therefore, in this paper, we consider models (10) and (11) to be better than models (10) and (11) to a certain extent.
Secondly, we studied the 21-day China Interbank Offered Rate from 2 November 2015 to 8 March 2016, which has 178 observations. Let x t be the logarithms of the 21-day China Interbank Offered Rate in Figure 5, and y t = x t x t 1 , { y t } is plotted in Figure 6. Ling [8] makes a self-weighted quasi-maximum likelihood estimate for ARMA-GARCH.
ARMA(0,1)-GARCH(0,1) (MA-GARCH) model was adopted to fit { y t } by Ling [8], and the result is
y t = 0.5143 ( 0.0000 ) ε t 1 + ε t
ε t = e t h t = e t 0.0061 ( 0.0064 ) + 0.0773 ( 5.0485 ) ε t 1 2 + 0.0010 ( 3.8925 ) h t 1
where the standard errors are in parentheses.
Similarly, we used models (1)–(3) to fit { y t } , and the result is
y t = 0.4982 ( 0.0069 ) ε t 1 + ε t
ε t = e t h t = e t 0.0062 ( 0.0092 ) + 0.0496 ( 0.1182 ) y t 1 2 + 0.0010 ( 1.4060 ) h t 1
where the standard errors are in parentheses (From Theorem 1 we get).
From Table 5, we can see that the RMSE and the log-likelihood values of models (12) and (13) and models (14) and (15) are equal. In addition, we give a 95 percent confidence interval of sequence x t , as shown in Figure 7 below. Figure 8 is a partial enlargement of Figure 7. From Figure 7, we can see that models (14) and (15) hasve a narrower confidence interval than models (12) and (13), and we can see that all points fall within the confidence interval. This shows that for this data, models (14) and (15) are better than models (12) and (13) in terms of predictive ability. Therefore, in this paper, we consider models (14) and (15) better than models (12) and (13) to a certain extent.

5. Conclusions

With the motivation to study the DAP(p) model with p going to , in this article, we propose a class of MA-GARCH model based on Zhu and Zhang et al. [14] The consistency and asymptotic normality properties of the self-weighted quasi-maximum likelihood estimation of a class of MA-GARCH models are established. Through simulation, we found that the the self-weighted quasi-maximum likelihood estimation of a class of MA-GARCH models is feasible. Empirical results show that a class of a MA-GARCH model is better than the MA-GARCH model in terms of goodness of fit and predictive ability. This implies that the model we considered has some applicability to the field of MA-GARCH models.

Author Contributions

Methology and Simulation, D.X.; Guide and Supervision, X.L.; Solfware, R.L. All authors have read and agreed to the published version of manuscript.

Funding

The work is partially supported by National Natural Science Foundation of China under grant no.12161009, The Science and technology project of Guangxi under grant no.2021AC06001, and Scientific Research Foundation of Development Institute of Zhujiang-Xijiang Economic Zone under grant no. ZX2020013.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The empirical data comes from ST Daji (000546) shared on the Shenzhen Stock Exchange; the empirical data comes from the choice financial date terminal.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
MDPIMultidisciplinary Digital Publishing Institute
DOAJDirectory of open access journals
TLAThree letter acronym
LDLinear dichroism

Appendix A

Appendix A.1

Proof of Lemma 1.
(i):
s u p Θ | ε t 1 ( θ ) | = s u p Θ | Φ 1 ( B ) y t 1 | = s u p Θ | i = 0 a Φ ( i ) B i y t 1 | i = 0 O ( ρ i ) | y t 1 i | c ξ ρ , t 1
(ii):
s u p Θ ε t ( θ ) ϕ = s u p Θ Φ 2 ( B ) y t 1 = s u p Θ Φ 1 ( B ) [ Φ 1 ( B ) y t 1 ] Φ 1 ( B ) i = 0 O ( ρ 0 i ) | y t 1 i | j = 0 O ( ρ 0 j ) i = 0 O ( ρ 0 i ) | y t 1 i | = c k = 0 ( k + 1 ) ρ 0 k | y t 1 k | = i = 0 ρ i y t 1 i c ξ ρ , t 1
for some ρ and 1 < k + 1 k ρ 0 ρ < 1 .
(iii):
s u p Θ 2 ε t ( θ ) ϕ ϕ = s u p Θ 2 Φ 3 ( B ) y t 1 = s u p Θ 2 Φ 2 ( B ) ε t 1 c 2 Φ 1 ( B ) i = 0 ρ 0 i ε t 2 i c j = 0 ρ 0 j i = 0 ρ 0 i ε t 2 i j = c k = 0 ( k + 1 ) ρ 0 k ε t 2 k = c k = 0 ( k + 1 ) ρ 0 k [ 1 ϕ y t 1 k + 1 ϕ ε t 1 k ] c k = 0 ( k + 1 ) ρ 0 k | y t 1 k | + c k = 0 ( k + 1 ) ρ 0 k | ε t 1 k | c k = 0 ρ 1 k | y t 1 k | + c k = 0 ρ 1 k ξ ρ , t 1 c k = 0 ρ 1 k | y t 1 k | + c k = 0 ρ 1 k [ 1 + s = 0 ρ s | y t 1 s | c ξ ρ , t 1
for some ρ 0 , ρ 1 and ρ 2 , and 0 < k + 1 k ρ 0 ρ 1 ρ 2 ρ < 1 . □
Proof of Lemma 2.
(i): From the models (1)–(3), we can deduce that h t ( θ ) = β 1 ( B ) ( ω + α y t 1 2 ) .
h t ( θ ) = β 1 ( B ) ( ω + α y t 1 2 ) = i = 0 β i ω + α i = 0 β i y t 1 i 2 i = 0 ρ i ω + α ¯ i = 0 ρ i y t 1 i 2 c ζ ρ , t 1
(ii): From models (1)–(3), one iteratively obtains
h t ( θ ) = ω + α y t 1 2 + β h t 1 = ω + α y t 1 2 + β ( ω + α y t 2 2 + β h t 2 ) = = ω 1 β + α j = 0 β j y t j 1 2
Then
1 h t ( θ ) h t ( θ ) ω = 1 ω 1 β + α j = 0 β j y t j 1 2 1 1 β 1 ω 1 β 1 1 β ω ̲
1 h t ( θ ) h t ( θ ) α = 1 h t ( θ ) β 1 ( B ) y t 1 2 = 1 ω 1 β + α j = 0 β j y t j 1 2 j = 0 β j y t j 1 2 1 ω 1 β j = 0 β j y t j 1 2 c ζ ρ , t 1
because h t ( θ ) β = ω ( 1 β ) 2 + α j = 0 j β j 1 y t j i 2 , so
1 h t ( θ ) h t ( θ ) β = 1 ω 1 β + α j = 0 β j y t j 1 2 [ ω ( 1 β ) 2 + α Σ j = 0 j β j 1 y t j 1 2 ] 1 ω 1 β [ ω ( 1 β ) 2 + α j = 0 j β j 1 y t j 1 2 ] = 1 1 β + 1 β ω α β j = 0 j β j y t j 1 2 1 1 β + 1 β ω α β j = 0 ρ j y t j 1 2 c ζ ρ , t 1
where 0 < β < j j β < ρ < 1 .
(iii): The proof of (iii) is similar to that of (ii).
(vi): Because h t ( θ ) ϕ = 0 , so, 1 h t ( θ ) h t ( θ ) ϕ = 0 . □

Appendix A.2

Proof of Theorem 1.
(i): First, the space Θ is compact, and θ 0 is an interior point in Θ . Second, L s n ( θ ) is continuous in θ Θ and is a measurable function of { y i , i = t , t 1 , } for all θ Θ . Third, by Lemmas 1 and 2, it follows that
| e t | = | ε t ( θ ) h t ( θ ) | = | i = 0 ( 1 ) i ϕ i y t 1 ω 1 β α j = 0 β j y t j 1 2 | c ξ ρ , t 1
s u p Θ | ε t ( θ ) | = s u p Θ | ε t ( θ 0 ) + ( θ θ 0 ) ε t ( θ 0 ) θ | | ε t ( θ 0 ) | + s u p Θ θ θ 0 s u p Θ ε t ( θ ) θ | e t | h t ( θ 0 ) + O ( 1 ) ξ ρ , t 1 O ( 1 ) [ ξ ρ , t 1 ζ ρ , t 1 1 / 2 + ξ ρ , t 1 ]
s u p Θ h t ( θ ) c ζ ρ , t 1
By (A1)–(A3), we can show that
E s u p Θ [ ω t ε t 2 ( θ ) h t ( θ ) ] E s u p Θ [ ω t O ( 1 ) ( ξ ρ , t 1 ζ ρ , t 1 1 / 2 + ξ ρ , t 1 ) 2 ω ̲ ] c E s u p Θ [ ω t ξ ρ , t 1 2 ζ ρ , t 1 + ω t ξ ρ , t 1 2 + 2 ζ ρ , t 1 1 / 2 ξ ρ , t 1 2 ] <
E s u p Θ | ω t l o g h t ( θ ) | = O ( 1 ) E s u p Θ l o g h t ( θ ) O ( 1 ) l o g E s u p Θ h t ( θ ) <
Therefore, by (A4) and (A5), we have
E s u p Θ ω t l t ( θ ) E s u p Θ | ω t l o g h t ( θ ) | + E s u p Θ | ω t ε t 2 ( θ ) h t ( θ ) | <
By ergodic theorem, we have L s n E [ ω t l t ( θ ) ] a . s . θ Θ .
By Theorem 3.1 in Ling and McAleer (2003a) [19], it follows that
s u p Θ | L s n ( θ ) E [ ω t l t ( θ ) ] | p 0
E [ ω t l t ( θ ) ] = E ω t l o g h t ( θ ) + E ω t [ ε t ( θ 0 ) + ( θ θ 0 ) ε t ( θ * ) θ ] 2 h t ( θ ) = E ω t l o g h t ( θ ) + E ω t ε t 2 ( θ 0 ) h t ( θ ) + 2 E ω t ε t ( θ 0 ) ( θ θ 0 ) ε t ( θ * ) θ h t ( θ ) ( θ θ 0 ) E [ ω t h t ( θ ) ε t ( θ * ) θ ε t ( θ * ) θ ] ( θ θ 0 )
Because E e t 2 = 1 , E e t = 0 , thus,
E ω t ε t 2 ( θ 0 ) h t ( θ ) = E ω t e t 2 h t ( θ 0 ) h t ( θ ) = E ω t h t ( θ 0 ) h t ( θ ) [ E e t 2 | F t 1 ] = E ω t h t ( θ 0 ) h t ( θ )
2 E ω t ε t ( θ 0 ) ( θ θ 0 ) ε t ( θ * ) θ h t ( θ ) = 0
.
Therefore, E [ ω t l t ( θ ) ] = A + B , where
A = E ω t l o g h t ( θ ) + E ω t h t ( θ 0 ) h t ( θ )
B = ( θ θ 0 ) E [ ω t h t ( θ ) ε t ( θ * ) θ ε t ( θ * ) θ ] ( θ θ 0 )
.
Considering the function f ( x ) = l o g x + a x when a 0 , it reaches a minimum at x = a . Thus, for A, A reaches the minimum if and only if θ = θ 0 . For B, B 0 , B reaches the minimum if and only θ = θ 0 . Thus, we can claim that E [ ω t l t ( θ ) ] is uniformly minimized at θ = θ 0 .
Next, we use the method in Theorem 4.1.1 in Amemiya [20]. Let V be any open neighborhood of θ 0 Θ . Then V c Θ is the compact set, and E [ ω t l t ( θ ) ] exists as a minimum in θ V c Θ .
We order
ε = min Θ V c Θ E [ ω t l t ( θ ) ] E [ ω t l t ( θ 0 ) ]
and let | L s n ( θ ) E ω t l t ( θ ) | < ε 2 , θ Θ be the event A.
Thus, A E ω t l t ( θ ) > L s n ( θ ) ε 2 , A L s n ( θ ) > E ω t l t ( θ ) ε 2 .
Because θ , θ ^ ( s n ) Θ , thus,
A E ω t l t ( θ 0 ) > L s n ( θ 0 ) ε 2
A L s n ( θ ^ s n ) > E ω t l t ( θ ^ s n ) ε 2
and because L s n ( θ ^ 0 ) L s n ( θ 0 ) , then, by (A2), we have
A E ω t l t ( θ 0 ) > L s n ( θ ^ s n ) ε 2
By (A8) and (A9), we have
A E ω t l t ( θ 0 ) > E ω t l t ( θ ^ s n ) ε
By (A6) and (A10), we have
A E ω t l t ( θ 0 ) > E ω t l t ( θ ^ s n ) min Θ V c Θ E [ ω t l t ( θ ) ] + E [ ω t l t ( θ 0 ) ]
So A min Θ V c Θ E [ ω t l t ( θ ) ] > E ω t l t ( θ ^ s n )
Thus, A θ ^ s n V , and then P ( A ) P ( θ ^ s n V ) .
And because lim n P ( A ) = 1 , we have θ ^ s n p θ 0 .
(ii): We further give the quasi-score function and the quasi-information matrix as follows.
Take the derivative of l t ( θ ) :
l t ( θ ) θ = 2 ε t ( θ ) h t ( θ ) ε t ( θ ) ϕ + 1 h t ( θ ) ( 1 ε t 2 ( θ ) h t ( θ ) ) h t ( θ ) θ 2 l t ( θ ) ϕ ϕ = 2 h t ( θ ) ε t ( θ ) ϕ ε t ( θ ) ϕ + 2 ε t ( θ ) h t ( θ ) 2 ε t ( θ ) ϕ ϕ 2 l t ( θ ) δ δ = 1 h t 2 ( θ ) h t ( θ ) δ h t ( θ ) δ + 1 h t ( θ ) 2 h t ( θ ) δ δ + 2 ε t 2 ( θ ) h t 3 ( θ ) h t ( θ ) δ h t ( θ ) δ ε t 2 ( θ ) 2 h t 2 ( θ ) 2 h t ( θ ) δ δ
Then
s u p Θ 2 l t ( θ ) ϕ ϕ O ( 1 ) s u p Θ { ε t ( θ ) ϕ 2 + | e t ( θ ) | 2 ε t ( θ ) ϕ ϕ } O ( 1 ) ξ ρ , t 1 2
s u p Θ 2 l t ( θ ) δ δ O ( 1 ) { ζ ρ , t 1 2 + ζ ρ , t 1 + ζ ρ , t 1 2 | e t | 2 + | e t | 2 ζ ρ , t 1 } O ( 1 ) { ζ ρ , t 1 2 ( 1 + ξ ρ , t 1 2 ) }
Therefore, E s u p Θ ω t 2 l t ( θ ) θ θ <
By ergodic and Theorem 3.1 in Ling and McAleer(2003) [19], it follows that
2 L s n ( θ ) θ θ p E [ ω t 2 l t ( θ ) θ θ ]
if and only if n .
By the double expectation property, it follows that
E [ ω t 2 l t ( θ ) θ θ ] = E { E ( ω t 2 l t ( θ ) θ θ | F t 1 ) } = E [ 2 ω t h t ( θ ) ε t ( θ ) ϕ ε t ( θ ) ϕ + ω t h t 2 ( θ ) h t ( θ ) δ h t ( θ ) δ ] = Σ 0
Then, 2 L s n ( θ 0 ) θ θ = 1 n t = 1 ω t 2 l n ( θ 0 ) θ θ p Σ 0
Similarly, we can show that E s u p Θ l t ( θ ) θ < holds.
Because n L s n ( θ ) θ = 1 n t = 1 n ω t l t ( θ 0 ) θ = t = 1 n 1 n ω t l t ( θ 0 ) θ
Therefore, E ( 1 n ω t l t ( θ 0 ) θ | F t 1 ) = 0
Thus, E ( n L s n ( θ 0 ) θ | F t 1 ) = 0 , V a r ( 1 n ω t l t ( θ 0 ) θ | F t 1 ) = 1 n E ( ω t 2 l t ( θ 0 ) θ l t ( θ 0 ) θ | F t 1 )
Because h t ( θ ) θ ε t ( θ ) θ = 0 , then
l t ( θ 0 ) θ l t ( θ 0 ) θ = 1 h t 2 ( θ ) ( 1 ε t 2 ( θ 0 ) h t ( θ 0 ) 2 h t ( θ 0 ) θ h t ( θ 0 ) θ + 4 ε t 2 ( θ 0 ) h t 2 ( θ 0 ) ε t ( θ 0 ) θ ε t ( θ 0 ) θ + 4 ε t ( θ 0 ) h t 2 ( θ 0 ) ( 1 ε t 2 ( θ 0 ) h t ( θ 0 ) ) h t ( θ 0 ) θ ε t ( θ 0 ) θ = 4 1 h t 2 ( θ ) ( 1 e t 2 ) 2 h t ( θ 0 ) θ h t ( θ 0 ) θ + e t 2 h t ( θ 0 ) ε t ( θ 0 ) θ ε t ( θ 0 ) θ
V a r ( n L s n ( θ 0 ) θ | F t 1 ) = t = 1 V a r ( 1 n ω t l s n ( θ 0 ) θ | F t 1 ) = 1 n t = 1 n E ( ω t 2 l s n ( θ 0 ) θ l s n ( θ 0 ) θ | F t 1 ) = 1 n t = 1 n [ 4 ω t 2 h t ( θ 0 ) ( ε t ( θ 0 ) θ ) 2 + ω t 2 E e t 4 1 h t 2 ( θ 0 ) ( h t ( θ 0 ) θ ) ] p Ω 0
where Ω 0 = E [ 4 ω t 2 h t ( θ 0 ) ( ε t ( θ 0 ) θ ) 2 + ω t 2 ς h t 2 ( θ 0 ) ( h t ( θ 0 ) θ ) ] , ς = E e t 4 1
Then, Ω 0 = E [ 4 ω t 2 h t ( θ 0 ) ( ε t ( θ 0 ) θ ) 2 + ω t 2 s h t 2 ( θ 0 ) ( h t ( θ 0 ) θ ) ] c E [ ω t 2 ξ ρ , t 1 2 + s ω t 2 ζ ρ , t 1 2 ]
Because ω t is bounded, by Assumptions A3 and A4, it follows that Ω 0 <
By the central limit theorem for a martingale different sequence, we have
L s n ( θ 0 ) θ d N ( 0 , Ω 0 ) .
By Theorem 4.1.3 in Amemiya (1985) [20], it follows that
n ( θ ^ s n θ 0 ) d N ( 0 , Σ 0 1 Ω 0 Σ 0 1 )

References

  1. Engle, R.F. Autoregressive conditional heteroscdeasticity with estimates of the variance of united kingdom inflation. Econometrica 1982, 50, 897–1007. [Google Scholar] [CrossRef]
  2. Bollerslve, T. Generalized autoregressive condition heteroskedasticity. J. Econom. 1986, 31, 307–327. [Google Scholar] [CrossRef]
  3. Nelson, D.B. Conditional heteroskedasticity in asset returns: A new approach. Econometrica 1991, 59, 347–870. [Google Scholar] [CrossRef]
  4. Glosten, L.R.; Jagannathan, R.; Runkle, D.E. On the relationship between the expected value and the valatility of the nominal excess return on stocks. J. Financ. 1993, 48, 1779–1801. [Google Scholar] [CrossRef]
  5. Long, H.V.; Jebreen, H.B.; Dassios, I.; Baleanu, D. On the Statistical GARCH Model for Managing the Risk by Employing a Fat-Tailed Distribution in Finance. Symmetry 2020, 12, 1698. [Google Scholar] [CrossRef]
  6. Li, X.; Zhang, X.; Li, Y. High-Dimensional Conditional Covariance Matrices Estimation Using a Factor-GARCH Model. Symmetry 2022, 14, 158. [Google Scholar] [CrossRef]
  7. Baillie, R.T. Long memory processes and fractional integration in econometrics. J. Econom. 1996, 73, 5–59. [Google Scholar] [CrossRef]
  8. Ling, S. Estimation and testing stationarity for double autoregressive models. J. R. Stat. Soc. 2004, 66, 63–78. [Google Scholar] [CrossRef]
  9. Ling, S. Self-weighted and local quasi-maximum likelihood estimators for ARMA-GARCH/IGARCH model. J. Econom. 2007, 140, 849–873. [Google Scholar] [CrossRef]
  10. Zhu, K.; Ling, S. Global self-weighted and local quasi-maximum exponential likelihood estimators forARMA-GARCH/IGARCH models. Ann. Stat. 2011, 39, 2131–2163. [Google Scholar] [CrossRef]
  11. Pan, B.; Chen, M.; Wang, Y.; Xia, W. Weighted least absolute deviations estimation for ARFIMA time with finite or infinite variance. J. Korean Stat. Soc. 2015, 44, 1–11. [Google Scholar] [CrossRef]
  12. Ramírez-Parietti, I.; Contreras-Reyes, J.E.; Idrovo-Aguirre, B.J. Cross-sample entropy estimation for time series analysis: A nonparametric approach. Nonlinear Dyn. 2021, 105, 2485–2508. [Google Scholar] [CrossRef]
  13. Ling, S. A double AR(p) model: Structure and estimation. Stat. Sin. 2007, 17, 161–175. [Google Scholar]
  14. Zhu, H.; Zhang, X.; Liang, X.; Li, Y. Moving Average Model with an Alternative GARCH-type Error. J. Syst. Sci. Inf. 2018, 6, 165–177. [Google Scholar] [CrossRef]
  15. Zhu, K.; Ling, S. Quasi-maximum exponential likelihood for a double AR(p) model. Stat. Sin. 2013, 23, 251–270. [Google Scholar]
  16. Yoshida, N. Quasi-likelihood analysis and its applications. Stat. Inference Stoch. Process. 2022, 25, 43–60. [Google Scholar] [CrossRef]
  17. Ling, S. Self-weighted least absolute deviation estimation for infinite variance autoragressive models. R. Stat. Soc. 2005, 67, 381–389. [Google Scholar] [CrossRef]
  18. Huber, P.J. Robust Statistical Procedures; Society for Industrial and Applied Mathematics: Philadelphia, PA, USA, 1997. [Google Scholar]
  19. Ling, S.; McAleer, M. Asymptotic theory for a new vector ARMA-GARCH model. Econom. Theory 2003, 19, 280–310. [Google Scholar] [CrossRef]
  20. Amemiya, T. Advanced Econometrics; Harvard University Press: Cambridge, MA, USA, 1985. [Google Scholar]
Figure 1. x t be the logarithms of the opening price of ST Daji (000564).
Figure 1. x t be the logarithms of the opening price of ST Daji (000564).
Symmetry 14 01723 g001
Figure 2. Time series diagram of y t = x t - x t 1 .
Figure 2. Time series diagram of y t = x t - x t 1 .
Symmetry 14 01723 g002
Figure 3. Time series diagram of the data x t (green line) and 95% forecasting confidence intervals based on models (8) and (9) (red line) and models (10) and (11) (black line), respectively.
Figure 3. Time series diagram of the data x t (green line) and 95% forecasting confidence intervals based on models (8) and (9) (red line) and models (10) and (11) (black line), respectively.
Symmetry 14 01723 g003
Figure 4. Part of Figure 3 is enlarged.
Figure 4. Part of Figure 3 is enlarged.
Symmetry 14 01723 g004
Figure 5. x t , the logarithms of the 21-day China Interbank Offered Rate.
Figure 5. x t , the logarithms of the 21-day China Interbank Offered Rate.
Symmetry 14 01723 g005
Figure 6. Time series diagram of y t = x t - x t 1 .
Figure 6. Time series diagram of y t = x t - x t 1 .
Symmetry 14 01723 g006
Figure 7. Time series diagram of the data x t (green line) and 95% forecasting confidence intervals based on models (12) and (13) (red line) and models (14) and (15) (black line), respectively.
Figure 7. Time series diagram of the data x t (green line) and 95% forecasting confidence intervals based on models (12) and (13) (red line) and models (14) and (15) (black line), respectively.
Symmetry 14 01723 g007
Figure 8. Part of Figure 7 is enlarged.
Figure 8. Part of Figure 7 is enlarged.
Symmetry 14 01723 g008
Table 1. Estimators for the models (1)–(3) when e t N ( 0 , 1 ) .
Table 1. Estimators for the models (1)–(3) when e t N ( 0 , 1 ) .
θ 0 = ( 0.5 , 0.1 , 0.4 , 0.3 )
ϕ ^ ω ^ α ^ β ^
n = 400QMLE(Bias)−0.00080.0053−0.0024−0.0167
SQMLE(Bias)0.00010.00510.0048−0.0196
QMLE(SD)0.04700.02630.08600.1030
SQMLE(SD)0.04560.02600.09290.1037
QMLE(AD)0.04160.02240.08480.0694
SQMLE(AD)0.04230.02220.09380.0691
n = 800QMLE(Bias)−0.00140.00280.0008−0.0099
SQMLE(Bias)−0.00120.00280.0033−0.0111
QMLE(SD)0.03300.01770.05290.0707
SQMLE(SD)0.03340.01770.06350.0718
QMLE(AD)0.03240.02190.06030.0722
SQMLE(AD)0.03290.02270.06530.0757
n = 1200QMLE(Bias)−0.00110.00130.0000−0.0054
SQMLE(Bias)−0.00060.00120.0010−0.0052
QMLE(SD)0.02720.01390.04870.0564
SQMLE(SD)0.02760.01400.05390.0578
QMLE(AD)0.02410.01400.04410.0592
SQMLE(AD)0.02440.01450.04920.0612
Table 2. Estimators for the models (1)–(3) when e t L a p l a c e ( 0 , 1 ) .
Table 2. Estimators for the models (1)–(3) when e t L a p l a c e ( 0 , 1 ) .
θ 0 = ( 0.5 , 0.1 , 0.4 , 0.3 )
ϕ ^ ω ^ α ^ β ^
n = 400QMLE(Bias)−0.00030.0080−0.0002−0.0304
SQMLE(Bias)0.00240.00600.0254−0.0334
QMLE(SD)0.05060.04030.17090.1809
SQMLE(SD)0.05290.04050.20180.1876
QMLE(AD)0.05760.06140.13450.2946
SQMLE(AD)0.07020.05610.29020.2805
n = 800QMLE(Bias)−0.00050.00570.0026−0.0219
SQMLE(Bias)−0.00080.00410.0184−0.0221
QMLE(SD)0.03680.03230.12990.1447
SQMLE(SD)0.03830.03230.16170.1519
QMLE(AD)0.03330.01560.09080.0812
SQMLE(AD)0.03620.01700.10830.0952
n = 1200QMLE(Bias)−0.00010.00320.0003-0.0119
SQMLE(Bias)−0.00060.00210.0065−0.0100
QMLE(SD)0.02970.02490.10420.1174
SQMLE(SD)0.03070.02570.13050.1260
QMLE(AD)0.02890.02040.09660.0880
SQMLE(AD)0.03080.02000.12180.0895
Table 3. Estimators for the models (1)–(3) when e t t 3 .
Table 3. Estimators for the models (1)–(3) when e t t 3 .
θ 0 = ( 0.5 , 0.1 , 0.4 , 0.3 )
ϕ ^ ω ^ α ^ β ^
n = 400QMLE(Bias)−0.00140.0072−0.0160−0.0485
SQMLE(Bias)0.00030.00600.0293−0.0490
QMLE(SD)0.05360.04260.22230.1944
SQMLE(SD)0.05550.04300.24630.2016
QMLE(AD)0.05810.01660.14370.1026
SQMLE(AD)0.06380.01590.19930.2042
n = 800QMLE(Bias)−0.0010−0.0046−0.0034-0.0265
SQMLE(Bias)−0.00100.00420.0191−0.0303
QMLE(SD)0.04080.03570.19210.1769
SQMLE(SD)0.04360.03610.21910.1846
QMLE(AD)0.03710.02440.11280.1032
SQMLE(AD)0.03930.02280.13040.0967
n = 1200QMLE(Bias)−0.00090.00390.0059−0.0243
SQMLE(Bias)−0.00100.00370.0188−0.0279
QMLE(SD)0.03220.03280.17670.1600
SQMLE(SD)0.03340.03240.20170.1603
QMLE(AD)0.03040.02170.12430.0885
SQMLE(AD)0.03160.02290.14060.0941
Table 4. Comparison of the fitting effect of models (8) and (9) and models (10) and (11).
Table 4. Comparison of the fitting effect of models (8) and (9) and models (10) and (11).
ModelsRMSELog-Likelihood Value
(8) and (9)0.03612315.2
(10) and (11)0.03602316.4
Table 5. Comparison of the fitting effect of models (12) and (13) and models (14) and (15).
Table 5. Comparison of the fitting effect of models (12) and (13) and models (14) and (15).
ModelsRMSELog-Likelihood Value
(12) and (13)0.0818340.91
(14) and (15)0.0818340.91
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Xie, D.; Liang, X.; Liang, R. Self-Weighted Quasi-Maximum Likelihood Estimators for a Class of MA-GARCH Model. Symmetry 2022, 14, 1723. https://doi.org/10.3390/sym14081723

AMA Style

Xie D, Liang X, Liang R. Self-Weighted Quasi-Maximum Likelihood Estimators for a Class of MA-GARCH Model. Symmetry. 2022; 14(8):1723. https://doi.org/10.3390/sym14081723

Chicago/Turabian Style

Xie, Danni, Xin Liang, and Ruilin Liang. 2022. "Self-Weighted Quasi-Maximum Likelihood Estimators for a Class of MA-GARCH Model" Symmetry 14, no. 8: 1723. https://doi.org/10.3390/sym14081723

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop