
This paper introduces a comprehensive class of models known as Markov-Switching Threshold Stochastic Volatility (MS-TSV) models, specifically designed to address asymmetry and the leverage effect observed in the volatility of financial time series. Extending the classical threshold stochastic volatility model, our approach expresses the parameters governing log-volatility as a function of a homogeneous Markov chain with a finite state space. The primary goal of our proposed model is to capture the dynamic behavior of volatility driven by a Markov chain, enabling the accommodation of both gradual shifts due to economic forces and sudden changes caused by abnormal events. Following the model's definition, we derive several probabilistic properties of the MS-TSV models, including strict (or second-order) stationarity, causality, ergodicity, and the computation of higher-order moments. Additionally, we provide the expression for the covariance function of the squared (or powered) process. Furthermore, we establish the limit theory for the Quasi-Maximum Likelihood Estimator (QMLE) and demonstrate the strong consistency of this estimator. Finally, a simulation study is presented to assess the performance of the proposed estimation method.
Citation: Ahmed Ghezal, Mohamed balegh, Imane Zemmouri. Markov-switching threshold stochastic volatility models with regime changes[J]. AIMS Mathematics, 2024, 9(2): 3895-3910. doi: 10.3934/math.2024192
[1] | Xin-Jiang He, Sha Lin . Analytical formulae for variance and volatility swaps with stochastic volatility, stochastic equilibrium level and regime switching. AIMS Mathematics, 2024, 9(8): 22225-22238. doi: 10.3934/math.20241081 |
[2] | Omar Alzeley, Ahmed Ghezal . On an asymmetric multivariate stochastic difference volatility: structure and estimation. AIMS Mathematics, 2024, 9(7): 18528-18552. doi: 10.3934/math.2024902 |
[3] | Lin Xu, Linlin Wang, Hao Wang, Liming Zhang . Optimal investment game for two regulated players with regime switching. AIMS Mathematics, 2024, 9(12): 34674-34704. doi: 10.3934/math.20241651 |
[4] | Ahmed Ghezal, Omar Alzeley . Probabilistic properties and estimation methods for periodic threshold autoregressive stochastic volatility. AIMS Mathematics, 2024, 9(5): 11805-11832. doi: 10.3934/math.2024578 |
[5] | Hui Sun, Zhongyang Sun, Ya Huang . Equilibrium investment and risk control for an insurer with non-Markovian regime-switching and no-shorting constraints. AIMS Mathematics, 2020, 5(6): 6996-7013. doi: 10.3934/math.2020449 |
[6] | Luigi Accardi, Amenallah Andolsi, Farrukh Mukhamedov, Mohamed Rhaima, Abdessatar Souissi . Clustering quantum Markov chains on trees associated with open quantum random walks. AIMS Mathematics, 2023, 8(10): 23003-23015. doi: 10.3934/math.20231170 |
[7] | Min-Ku Lee, Jeong-Hoon Kim . Pricing vanilla, barrier, and lookback options under two-scale stochastic volatility driven by two approximate fractional Brownian motions. AIMS Mathematics, 2024, 9(9): 25545-25576. doi: 10.3934/math.20241248 |
[8] | Mohammed Alhagyan, Mansour F. Yassen . Incorporating stochastic volatility and long memory into geometric Brownian motion model to forecast performance of Standard and Poor's 500 index. AIMS Mathematics, 2023, 8(8): 18581-18595. doi: 10.3934/math.2023945 |
[9] | Min Han, Bin Pei . An averaging principle for stochastic evolution equations with jumps and random time delays. AIMS Mathematics, 2021, 6(1): 39-51. doi: 10.3934/math.2021003 |
[10] | Jingen Yang, Zhong Zhao, Xinyu Song . Statistical property analysis for a stochastic chemostat model with degenerate diffusion. AIMS Mathematics, 2023, 8(1): 1757-1769. doi: 10.3934/math.2023090 |
This paper introduces a comprehensive class of models known as Markov-Switching Threshold Stochastic Volatility (MS-TSV) models, specifically designed to address asymmetry and the leverage effect observed in the volatility of financial time series. Extending the classical threshold stochastic volatility model, our approach expresses the parameters governing log-volatility as a function of a homogeneous Markov chain with a finite state space. The primary goal of our proposed model is to capture the dynamic behavior of volatility driven by a Markov chain, enabling the accommodation of both gradual shifts due to economic forces and sudden changes caused by abnormal events. Following the model's definition, we derive several probabilistic properties of the MS-TSV models, including strict (or second-order) stationarity, causality, ergodicity, and the computation of higher-order moments. Additionally, we provide the expression for the covariance function of the squared (or powered) process. Furthermore, we establish the limit theory for the Quasi-Maximum Likelihood Estimator (QMLE) and demonstrate the strong consistency of this estimator. Finally, a simulation study is presented to assess the performance of the proposed estimation method.
Over recent years, Markov-switching models (MSMs) have garnered considerable scholarly attention, emerging as potent tools for modeling and characterizing asymmetric business cycles within the realm of econometrics. The selection of these models is grounded in their notable flexibility to capture stability and/or asymmetric effects in volatility shocks, as well as their efficacy in modeling time series data. Initially highlighted by Hamilton [1,2], these models have been actively employed in statistical applications, addressing various time series phenomena. Several authors have extensively explored aspects such as stationarity, the existence of moments, geometric ergodicity, statistical inference, and asymptotic theory for both linear and nonlinear Markov-switching models, including MS-ARMA models [3,4,5], nonlinear MS-ARMA models [6], MS-GARCH models [7,8], MS-BL models [9,10,11,12,13], MS-BLGARCH models [14,15], doubly MS-AR models [16], MSAR-SV models [17], and MS-AlogGARCH models [18], while also encompassing a distinct case known as the periodic model [19,20]. In our study, we introduce an alternative perspective by presenting a Markov-switching threshold stochastic volatility process. This process incorporates a standard threshold stochastic volatility [21,22] representation within each local regime. Notably, the log-volatility process in this model follows an rth-order Markov-switching threshold autoregression (TAR), with coefficients contingent on a Markov chain. This approach is recognized in the literature as a compelling substitute for MS-ARCH-type models, which rely on exogenous innovations to drive volatility. Our presented model can be viewed as a logical expansion of the MSAR-SV model initially proposed by So et al. [23], thereby incorporating heavy-tailed innovations to describe the observed process. For further nuanced insights, a more qualitative discussion on this approach can be found in the works of Casarin [24]. The primary rationale behind opting for the MS-TSV model is its remarkable enhancement of predictive capabilities compared to the standard TSV model. This model effectively captures pivotal events that impact the oil market, demonstrating superior performance. Additionally, it adeptly accommodates the typical fluctuating behavior of volatility attributable to economic dynamics, while simultaneously addressing abrupt, discrete shifts in volatility resulting from unexpected extraordinary events. The goals of this paper can be summarized as follows: (1) delving into the probabilistic properties of the MS-TSV model. In doing so, we establish the necessary and sufficient assumptions required to ensure the existence of a stationary solution. It's noteworthy that the MS-TSV coefficients linked to the Markov chain can diverge from the conventional stationary assumptions associated with standard TSV models; (2) centering on analyzing the strong consistency of the QMLE for MS-TSV models. Prior to delving into the analysis, we introduce a set of symbols to facilitate the forthcoming discussion.
Throughout the paper, the following symbols are employed:
● The symbol I(.) represents a square matrix in which each main diagonal entry is 1, while all other entries are set to 0. Additionally, O(n,m) signifies a n×m matrix in which all entries are zeros. Meanwhile, F_′:=(I(1),O_(1,r−1),I(1),O_(1,r−1)). The function I{.} refers to an indicator function.
● The notation ρ(Γ) denotes the spectral radius of a square matrix Γ.
● The symbol ‖.‖ represents any norm applicable to m×n matrices (or m×1 vectors). Meanwhile, the symbol ⊗ signifies the Kronecker product operation.
● The sequence (Δt,t∈Z) represents a stationary Markov chain that is both irreducible and aperiodic.
● The matrix Q(n)=(q(n)ij,(i,j)∈E×E) represents the n−step transition probability matrix, where q(n)ij=P(Δt=j|Δt−n=i) with one-step transition probability matrix Q:=(qij, (i,j)∈E×E) where qij:=q(1)ij=P(Δt=j|Δt−1=i) for i,j∈E={1,...,e}.
● The vector Π_′=(π(1),...,π(d)) represents the initial stationary distribution, where π(i)=P(Δ0=i), i=1,...,e, such that Π_′=Π_′Q.
● When considering a collection of deterministic matrices denoted as Γ:={Γ(i),i∈E}, it is important to observe that:
Q(n)(Γ)=(q(n)11Γ(1)…q(n)e1Γ(1)⋮…⋮q(n)1eΓ(e)…q(n)eeΓ(e)), Π_(Γ)=(π(1)Γ(1)⋮π(e)Γ(e)) |
with Q(1)(Γ)=Q(Γ).
The remaining content of the paper is structured in the following manner. In Section 2, we introduce the MS-TSV model, shedding light on its distinctive probabilistic characteristics. Emphasis is placed on the existence of a strictly or second (or higher)-order stationary solution for the MS-TSV model. Additionally, we establish autocovariance functions corresponding to the squared and powered processes. Section 3 unveils our proposition: a meticulously tailored QMLE for the MS-TSV model. This section not only elucidates the essence of QMLE but also establishes its strong consistency within the MS-TSV framework. Dedicated to presenting the outcomes of our simulations, Section 4 provides a comprehensive analysis of the performance of the proposed QMLE within the MS-TSV model framework. Section 5 serves as the conclusion of this paper.
The univariate Markov-switching threshold stochastic volatility model, denoted as MS−TSV(r), is defined by the following equation:
{Xt=σ1/2tεtlogσt=α0(Δt)+ r∑i=1(αi(Δt)I{Xt−i>0}+βi(Δt)I{Xt−i<0})logσt−i+β0(Δt)et. | (2.1) |
In Eq (2.1), the two processes {εt,t∈Z} and {et,t∈Z} represent two independent and identically distributed (i.i.d.) sequences of random variables with zero mean and unit variance. The functions αi(.) and βi(.),i=0,...,r are related to the unobserved Markov chain (Δt,t∈Z). Additionally, we assume that (εt,et) and {(Xu−1,Δt),u≤t} are independent. The objective of this section is to demonstrate some important probabilistic properties of the MS-TSV model. To facilitate the analysis, it is often useful to express Eq (2.1) in an equivalent state-space representation. In this context, we can rewrite Eq (2.1) in the form of a multivariate autoregressions model with Markov-switching dynamics:
Λ_t=Ψ(Δt)Λ_t−1+Υ_t(Δt) | (2.2) |
and
Xt=εtexp(12F_′Λ_t) |
where
Λ_′t:=(I{et>0}logσt,…,I{et−r+1>0}logσt−r+1,I{et<0}logσt,…,I{et−r+1<0}logσt−r+1),Υ_t(Δt):=(α0(Δt)+β0(Δt)et)(I{et>0},O(1,r−1),I{et<0},O(1,r−1))′ |
and
Ψ(Δt):=(A_′(Δt)I{et>0}αr(Δt)I{et>0}B_′(Δt)I{et>0}βr(Δt)I{et>0}I(r−1)O(r−1,1)O(r−1,r−1)O(r−1,1)A_′(Δt)I{et<0}αr(Δt)I{et<0}B_′(Δt)I{et<0}βr(Δt)I{et<0}O(r−1,r−1)O(r−1,1)I(r−1)O(r−1,1))A_′(Δt)=(α1(Δt),…,αr−1(Δt)), B_′(Δt)=(β1(Δt),…,βr−1(Δt)). |
The process ((Λ_′t,Δt)′,t∈Z) represents a Markov chain on R2r×E. However, when investigating the probabilistic properties of the model described in Eq (2.1), it is more convenient and advantageous to utilize the model presented in Eq (2.2). Equation (2.2) is identical to the definition used for the recently studied D-MSAR model by Ghezal [16]. Firstly, we establish the following significant result, implying strict stationarity.
Theorem 2.1. The multivariate model with Markov-switching (2.2) is under consideration. Here, we present the following:
i. Sufficient condition: If
γ(Ψ):=limn→∞E{1nlog‖n−1∏j=0Ψ(Δn−j)‖}a.s=limn→∞{1nlog‖n−1∏j=0Ψ(Δn−j)‖}<0 |
then Eq (2.2) admits a unique, strictly stationary, causal and ergodic solution given by the following series
Xt=εt∞∏k=0exp(12F_′{k−1∏j=0Ψ(Δt−j)}Υ_t−k(Δt−k)) | (2.3) |
which converges absolutely almost surely for all t∈Z.
ii. Necessary condition: If {Υ_t(Δt),Ψ(Δt)} is controllable [17] and the multivariate stochastic volatility model with Markov-switching (2.2) has a strictly stationary solution, then it follows that γ(Ψ)<0.
Proof. i. Sufficient Condition: A sufficient condition is provided by the subadditive ergodic theorem. Almost surely, we have
limsupk‖k−1∏j=0Ψ(Δt−j)‖1/k≤exp{γ(Ψ)}<1. |
Conversely, utilizing the Borel-Cantelli lemma, it follows that
P(limsupk→+∞|et−k|1/k>λ)=0 for all λ>1. |
Consequently,
limsupk→+∞‖{k−1∏j=0Ψ(Δt−j)}Υ_t−k(Δt−k)‖1/k≤exp{γ(Ψ)}<1 |
and by Cauchy's root test, the series (2.3) converges absolutely almost surely.
ii. Necessary Condition: As for the second assertion, we establish a necessary condition. If there exists a strictly stationary solution for Eq (2.2), thus
‖{k−1∏j=0Ψ(Δt−j)}β0(Δt)‖⟶k→∞0 in probability. |
By controllability, we consequently derive ‖{k−1∏j=0Ψ(Δt−j)}‖⟶k→∞0 in probability. Through a straightforward modification of Lemma 3.4 in Picard [25], we deduce that γ(Ψ)<0.
Remark 2.1. If any of the following conditions is satisfied, then it implies that γ(Ψ)<0:
a. E{log‖n−1∏j=0Ψ(Δn−j)‖}<0,
b. E{‖n−1∏j=0Ψ(Δn−j)‖}<1,
c. ρ(|Ψ|)<1, where |Ψ|=E{|Ψ(Δn)|}.
Example 2.1. The MS-TSV(1) model satisfies the following sufficient condition: e∏k=1|α1(k)|κπ(k)|β1(k)|(1−κ)π(k)<1, where κ=P(ε0>0)>0. Consequently, in this state, there exists a unique, strictly stationary, causal, and ergodic solution for the model. Hence, the requirement for local strict stationarity is not essential. In other words, the presence of burst regimes (i.e., |α1(k0)|κπ(k0)|β1(k0)|(1−κ)π(k0)>1) does not preclude the possibility of global strict stationarity. For the specific case of MS-TSV(1) with two-regimes, Xt=σ1/2tεt and
logσt={1+ (aI{Xt−1>0}+bI{Xt−1<0})logσt−1+etifΔt=1((a+1)I{Xt−1>0}+(b−1)I{Xt−1<0})logσt−1+etifΔt=2 |
π(1)=7/9 with et∼N(0,1), the zone of strict stationarity is illustrated in Figure 1 below.
The graphical representation in Figure 1 offers a comprehensive insight into the strict stationarity region of the MS-TSV(1) process under the assumption of et∼N(0,1). The illustration delineates two clearly defined zones:
● The inner zone signifies strict stationarity.
● The outer zone denotes nonstationarity.
This visualization not only facilitates a qualitative assessment of the model's validity but also provides valuable insights into the sensitivity of the model to various inputs.
The distinct delineation of these zones aids in understanding the behavior of the process and contributes to a deeper comprehension of its dynamics.
It's great to hear that other properties of the MS-TSV model, such as second-order stationarity and the existence of moments, are clear and easily obtainable. These properties are essential in understanding the behavior and statistical characteristics of the model. Second-order stationarity ensures that the model's statistical properties remain consistent over time, and the existence of moments indicates that the model's random variables have well-defined statistical properties, such as mean, variance, and higher-order moments.
Theorem 2.2. Consider the MS-TSV(r) model (2.1) with its state-space representation (2.2). If
ρ(Q(Ψ(2)))<1 | (2.4) |
for Ψ(2):={Ψ⊗2(i),i∈E}, hence, Eq (2.2) possesses a unique second-order stationary solution represented by the Series (2.3). This solution demonstrates absolute almost sure convergence and convergence in L2. Moreover, it is both strictly stationary and ergodic.
Proof. The result is derived from the second-order stationarity of the (Λ_t,t∈Z) defined by Eq (2.2). This conclusion is obtained using the findings of Ghezal et al. [10].
To demonstrate this, we provide the explicit expressions of the moments up to the second-order in the following result:
Proposition 2.1. Consider the MS-TSV(r) model (2.1), if Xt∈L2, then
i.E{Xt}=0.
ii.γX(h)=E{XtXt−h}=∑yt,yt−1,...∈E∏k≥0qyt−k−1yt−kE{∞∏k=0exp(F_′{k−1∏j=0Ψ(yt−j)}Υ_t−k(yt−k))}I{h=0}.
Proof. Given the last condition, obtaining the second-order moments becomes straightforward. For brevity, specific details are omitted.
Example 2.2. In the case of the MS-TSV(1) model, the Condition (2.4) simplifies to ρ(Q(ζ_(2)))<1, where ζ_(2):=(ζ(2)(k)=κα21(k)+(1−κ)β21(k),k∈S)′. Specifically, for two regimes with α1(1)=α1(2)−1=a,β1(1)=β1(2)+1=b,q11=q22=1−p,q12=q21=p and et∼N(0,1), the Condition (2.4) can be expressed as the following two equivalent conditions:
{(2p−1)(a2+b2)((a+1)2+(b−1)2)+2(1−p)(2(a+12)2+2(b−12)2+1)<4(1−p)(2(a+12)2+2(b−12)2+1)≤4. |
The zone of second-order stationarity is illustrated in Figure 2.
Figure 2 provides a nuanced understanding of the second-order stationarity region within the MS-TSV(1) process, assuming et∼N(0,1). The graphical representation delineates three crucial zones:
● The inner zone signifies second-order stationarity.
● The boundary curve represents the integrated MS-TSV(1), when ρ(Q(ζ_(2)))=1.
● The outer zone indicates non-second-order stationarity.
This figure prominently highlights that the second-order stationary zone of MS-TSV(1) is more significant when considering a smaller value for p. Additionally, the visual representation of second-order stationary zones serves as a valuable tool to observe the model's behavior under different conditions, enhancing our understanding of its dynamics.
Certainly, for the MS-TSV(r) model with a multivariate representation (2.2), certain assumptions are required to ensure the existence of higher-order moments. These assumptions play a crucial role in understanding the statistical properties and stability of the model.
Remark 2.2. When the odd-order moments of (Xt,t∈Z) exist, they are null. On the other hand, the existence of even-order moments of (Xt,t∈Z) is succinctly summarized in the following theorem.
Theorem 2.3. Consider the MS-TSV(r) model (2.1) with its state-space representation (2.2). For all integer l≥1, assume that E{(max(εt,et))l}<+∞ and
ρ(Q(Ψ(l)))<1 | (2.5) |
where Ψ(l):={Ψ⊗l(j),j∈E}. As a result, the MS-TSV model defined by the state-space representation (2.2) possesses a unique, causal, ergodic, and strictly stationary solution given by (2.3). This solution encompasses moments up to the l−order. Moreover, the closed form expression for the l−th moment of Xt is as follows:
E{Xlt}=E{εlt}∑yt,yt−1,...∈E∏k≥0qyt−k−1yt−kE{exp{l2F_′{k−1∏j=0Ψ(yt−j)}Υ_t−k(yt−k)}}. |
Proof. The proof presented in the previous theorem remains applicable, and the results obtained can be extended accordingly. Therefore, we have decided to omit the details.
The autocovariance function of the (X2t,t∈Z) process is concisely presented in the following theorem
Theorem 2.4. Given the assumptions stated in the previous theorem, we can deduce the following result:
(1) If (Xt,t∈Z) follows the MS-TSV model (2.1) and Xt∈L4, then
γX2(0)=E{ε4t}∑yt,yt−1,...∈E∏k≥0qyt−k−1yt−kE{exp{F_′{k−1∏j=0Ψ(yt−j)}Υ_t−k(yt−k)}}−γ2X(0) |
and γX2(h)=0 otherwise.
(2) If (Xt,t∈Z) follows the MS-TSV model (2.1) and Xt∈L2l, then
γXl(0)=E{ε2lt}∑yt,yt−1,...∈E∏k≥0qyt−k−1yt−kE{exp{lF_′{k−1∏j=0Ψ(yt−j)}Υ_t−k(yt−k)}}−(E{Xlt})2 |
and γXl(h)=0 otherwise.
Proof. Indeed, it suffices to note that both processes (X2t) and (X2lt) are white noise processes.
Estimating Markov-switching models is a complex task, and the literature has considered specific models to address this challenge [15,16,17,18,26]. Various established Markov Chain Monte Carlo procedures exist for estimating certain states of Eq (2.1), as discussed in [23,27], and other works. In our study, we focus on a given realization (X1,X2,...,Xn) generated from a unique, causal, and strictly stationary MS-TSV model. We assume that r and e are known, and (εt) follows a standard Gaussian distribution. The unknown parameters αi(.) and βi(.),i=0,...,r and (qi,j,i,j=1,...,e,i≠j) are combined in a vector θ_ belonging to the parameter space Θ, with θ_0 representing the true values. Xie [28] has advocated using QMLE and proved its strong consistency for MS-GARCH models. Additionally, Ghezal et al. [16,17]) introduced certain assumptions ensuring the strong consistency of QMLE for the doubly MS-AR model and symmetric MSAR-SV. The Gaussian likelihood function can be expressed as
Ln(θ_)=∑Δ1,...,Δn∈Sπ(Δ1){n∏i=2qΔi−1,Δi}{n∏i=1hΔi(X1,...,Xi)} | (3.1) |
where
hΔi(X1,...,Xi)=1(2πσΔi(X1,...,Xi−1))1/2exp{−X2i2σΔi(X1,...,Xi−1)} |
with the log−transformed conditional stochastic variance process, denoted by logσΔi(X1,...,Xi−1),
logσΔi(X1,...,Xi−1)=α0(Δt)+ r∑i=1(αi(Δt)I{Xt−i>0}+βi(Δt)I{Xt−i<0})logσt−i+β0(Δt)et |
is defined by the second equation in Eq (2.1). Furthermore, the expression for this likelihood function can be articulated in the subsequent manner:
Ln(θ_)=1_′(e){n∏i=1Pθ_(h(X1,...Xi))}Π_(h(X1)). | (3.2) |
A quasi-maximum likelihood estimation of θ_0 is determined as any discernible solution, ˆθ_n, in the context of:
ˆθ_n=argmaxθ∈ΘLn(θ_). | (3.3) |
In this section, consider hΔt(Xt|X←t−1) (resp. hΔt(Xt|X_1)) to be the density function characterizing Xt given the all past observations (resp. past observations up to ε1). Similarly, let kθ_(Xt|X←t−1) (resp. kθ_(Xt|X_1)) denote the corresponding logarithmic conditional density of Xt given {Xt−1,Xt−2,...} (resp. {Xt−1,Xt−2,..X1}). Now, we proceed to establish the likelihood function ˜Ln(θ_) based on all past observations. This function, referred to as Ln(θ_) in Eq (3.1), is fashioned by substituting the density hΔt(X1,...Xt) with hΔt(Xt|X←t−1). Elaborating further, ˜Ln(θ_) can be represented as:
˜Ln(θ_)=1_′(e){n∏t=2Pθ_(h(Xt|X←t−1))}Π_(h(X1|X←0)). | (3.4) |
Here, the matrix Pθ_(h(Xi|X←i−1)) (resp. the vector Π_(h(X1|X←0))) takes the place of hΔi(X1,...Xi) by hΔi(Xi|X←i−1) in Pθ_(h(X1,...Xi)) (resp. Π_(h(X1))), for i=1,..,n.
To establish the robust convergence of the QMLE, we rely on the following assumptions:
A1. Θ constitutes a compact subset of R2e(r+1), encompassing the true value θ_0 within its bounds.
A2. For any θ_∈Θ, the sequence Ψ0 (derived by modifying the parameters θ_0) satisfies γL(Ψ0)<0.
A3. Given any θ_ and θ_∗ within Θ, if kθ_(Xt|X←t−1) equals kθ_∗(Xt|X←t−1) almost surely, then it logically follows that θ_ equals θ_∗.
While the first assumption, A1, is a familiar cornerstone adopted extensively in various real analysis results, the second assumption, A2, secures the principle of strict stationarity for the process (Xt,t∈Z). Moreover, A3, our third assumption, guarantees the distinguishability of the parameter θ_. To forge ahead, we lay down the foundation of our discourse through the presentation of pivotal lemmas.
Lemma 3.1. Given the robust underpinnings of Assumptions A2 and A3, almost surely, we have
limn⟶∞logL1/nn(θ_)=limn⟶∞log˜L1/nn(θ_)=Eθ_0{kθ_(Xt|X←t−1)}. |
Proof. Harnessing the potency of the logarithmic function, we attain:
log˜Ln(θ_)=n∑t=1kθ_(Xt|X←t−1) and logLn(θ_)=n∑t=1kθ_(Xt|X_1). |
Hence,
1nn∑t=1kθ_(Xt|X_1)=1nn∑t=1kθ_(Xt|X←t−1)+1nn∑t=1(kθ_(Xt|X_1)−kθ_(Xt|X←t−1)). |
Presently, for all κ∈R, the process (Ut(s),t∈Z) is defined as Ut(s)=supκ≥s|kθ_(Xt|X_t−κ)−kθ_(Xt|X←t−1)|. For a fixed value of s, the sequence (Ut(s),t∈Z) represents a strictly stationary and ergodic process, with Eθ_0{Ut(s)}<+∞. We have
limsupn|1nn∑t=1(kθ_(Xt|X_1)−kθ_(Xt|X←t−1))|≤limsupn1nn∑t=1|kθ_(Xt|X_1)−kθ_(Xt|X←t−1)|≤limsupn1nn∑t=s+1Ut(s)=Eθ_0{U1(s)} |
the result is established.
The following lemma provides a comparison between the ratios Ln(θ_)Ln(θ_0) and ˜Ln(θ_)˜Ln(θ_0). Define Tn(θ_)=log(L1/nn(θ_)/L1/nn(θ_0)). With this definition, we can observe that:
Lemma 3.2. Given the robust underpinnings of Assumptions A1–A3, we have
limn(˜L1/nn(θ_)/˜L1/nn(θ_0))=limnTn(θ_)≤0 |
with limnTn(θ_)=0 iff θ_=θ_0 for all θ_∈Θ.
Proof. Under assumptions A1–A3, the function Tn(θ_) is well-defined. Additionally, leveraging Lemma 3.1 and Jensen's inequality, we obtain:
limnTn(θ_)=Eθ_0{log(kθ_(Xt|X←t−1)/kθ_0(Xt|X←t−1))}≤logEθ_0{kθ_(Xt|X←t−1)/kθ_0(Xt|X←t−1)}=0. |
Given Assumption A3, it follows that Tn(θ_) converges to the Kullback-Leinbler information, which attains the value of zero only when θ_=θ_0.
Lemma 3.3. Under assumptions A1–A3, for all ˜θ_≠θ_0, there exists a neighborhood V(˜θ_) of ˜θ_such that
limsupnsupθ_∈V(˜θ_)Tn(θ_)<0almost surely. |
Proof. In Eq (3.4), we derive
minjπ(j)hj(X1|X←0)‖{n∏t=2Pθ_(h(Xt|X←t−1))}‖≤˜Ln(θ_)≤maxjπ(j)hj(X1|X←0)‖{n∏t=2Pθ_(f(Xt|X←t−1))}‖. |
Hence, we obtain,
limnlog˜L1/nn(θ_)=limnlog‖{n∏t=2Pθ_(h(Xt|X←t−1))}‖1/n=Eθ_0{kθ_(Xt|X←t−1)}. |
Consider the set Vs(˜θ_)={θ_:‖θ_−˜θ_‖≤s−1} and define Ω2:n(s)=supθ_∈Vs(˜θ_)‖n∏t=2Pθ_(h_(Xt|X←t−1))‖. Due to the multiplicativity of the norm, we derive the following result on Vs(˜θ_),
supθ_‖n+k∏t=2Pθ_(h(Xt|X←t−1))‖≤supθ_‖n∏t=2Pθ_(h(Xt|X←t−1))‖.supθ_‖n+k∏t=n+1Pθ_(h(Xt|X←t−1))‖ |
implying:
logΩ2:n+k(s)≤logΩ2:n(s)+logΩn+1:n+k (s), for all n,k . |
Now, the process (logΩ2:n(s)) is both strictly stationary and ergodic, with Eθ_0{logΩ2:n(s)} being finite. Consequently, we obtain:
ϰs(˜θ_)=limnlogΩ1/n2:n(s)=infn>1Eθ_0{logΩ1/n2:n(s)} almost surely |
where γθ_(H) represents the Lyapunov exponent of the sequence H=(Pθ_0(h(Xt|X←t−1)),t∈Z), that is:
γθ_(H)=infn>1Eθ_0{log‖n∏t=2Pθ_0(h(Xt|X←t−1))‖1/n}a.s.=limnlog‖n∏t=2Pθ_0(h(Xt|X←t−1))‖1/n. |
Therefore, by utilizing Lemma 3.2, we can establish the existence of δ>0 and nδ∈N such that
1nδEθ_0{log‖nδ∏t=2P˜θ_(h(Xt|X←t−1))‖}<γθ_0(H)−δ. |
Applying the dominated convergence theorem, we deduce that for sufficiently large s:
γ˜θ_,s(H)≤Eθ_0{log‖nδ∏t=2P˜θ_(h(Xt|X←t−1))‖1/nδ}+δ2<γθ_0(H)−δ2. |
The final result follows from Lemma 3.1.
Additionally, we present the following main theorem.
Theorem 3.1. Under Assumptions A1–A3, the sequence of QML estimators (ˆθ_n)n satisfying (3.3) exhibits strong consistency, meaning that:
ˆθ_n→θ_0 almost surely when n→+∞. |
Proof. Let's assume that ˆθ_n does not converge to θ_0 almost surely, i.e.,
∀n, ∃δ>0, N>n, such that ‖ˆθ_N−θ_0‖≥δ. |
Using the Lemma 3.3, we establish that Ln(ˆθ_n)<Ln(θ_0). However, according to the QMLE given in (3.3), we have:
Ln(ˆθ_n)=supθ_∈˜ΘLn(θ_)≥Ln(θ_0) |
for any compact subset ˜Θ of Θ containing θ_0. This inconsistency contradicts the result we aim to prove.
In the following remark, we delve into the consideration of an open problem
Remark 3.1. Multifractal processes have emerged as a novel formalism for modeling the time series of returns in finance. The notable appeal of these processes lies in their capacity to generate varying degrees of long memory across different powers of returns, a characteristic prevalent in virtually all financial data. In contrast to MS-TSV-type models, multifractal models, as recently developed, are distinguished by a multiplicative structure inherent in the volatility process. Within the multifractal framework, instantaneous volatility is conceptualized as a product of m volatility components or multipliers and a positive scale factor σ2,
Xt=σ2(σ(1)tσ(2)t⋯σ(m)t)1/2εt. |
The random multipliers or volatility components σ(l)t are non-negative. For simplicity, we assume that the multipliers σ(1)t,σ(2)t,…,σ(m)t at a given time t are statistically independent. This model structure, as outlined by Calvet et al. in [29] and [30], as well as Lux in [31], introduces a new perspective in the representation of financial volatility. To address initial challenges stemming from non-stationarity and the combinatorial nature of the original model, Calvet et al. [29] proposed an iterative multifractal model. This iteration not only overcomes the challenges but also facilitates the estimation of model parameters through methods such as maximum likelihood, providing a robust framework for Bayesian forecasting of volatility in financial time series data.
We conducted a simulation study to assess the performance of the QML method for parameter estimation. The study was based on the Gaussian MS-TSV(r) model with e=2. We generated 500 data samples with varying lengths. The sample sizes considered in this simulation study were n∈{750,1500,3000}. The chosen parameter values were designed to satisfy the stationarity condition γL(Ψ)<0. For each data trajectory, we estimated the vector θ_ of the parameters of interest using the QMLE, denoted as ˆθ_. The QMLE algorithm was executed using the "fminsearch.m" minimizer function in MATLAB8. In the tables presented below, the root mean square errors (RMSE) of ˆθ are displayed in parentheses. Additionally, the true values (TV) of the parameters for each of the considered data-generating processes are reported.
The primary emphasis of this study centers on analyzing the root mean square errors. The results obtained provide initial insights into the finite sample properties of the QMLE within the framework of the MS-TSV model. It is evident from the analysis that the QMLE method delivers effective parameter estimates. Upon examining Table 1, a noteworthy observation is the strong consistency of the QMLE for the MS-TSV model. The corresponding root mean square errors demonstrate a significant reduction as the sample size increases. This suggests that the estimation method becomes more robust with larger datasets. The outcomes presented in Table 2 further reinforce the strong consistency of the QMLE for the MS-TSV model. Notably, even with a relatively small sample size, the estimation procedure produces favorable and reliable results.
Tv/n | 750 | 1500 | 3000 | |
p11 | 0.25 | 0.2471(0.0053) | 0.2529(0.0032) | 0.2479(0.0014) |
p22 | 0.85 | 0.8534(0.0072) | 0.8521(0.0056) | 0.8509(0.0023) |
α0(1) | 1.00 | 0.9925(0.0355) | 1.0051(0.0169) | 0.9973(0.0087) |
α0(2) | −1.50 | −1.5060(0.0175) | −1.5071(0.0083) | −1.5031(0.0038) |
α1(1) | 0.45 | 0.4457(0.0017) | 0.4493(0.0007) | 0.4487(0.0004) |
α1(2) | 0.25 | 0.2504(0.0012) | 0.2504(0.0006) | 0.2499(0.0003) |
β1(1) | −0.55 | −0.5524(0.0018) | −0.5523(0.0009) | −0.5511(0.0004) |
β1(2) | −0.25 | −0.2480(0.0013) | −0.2496(0.0005) | −0.2506(0.0003) |
β0(1) | 0.85 | 0.8555(0.0076) | 0.8570(0.0034) | 0.8513(0.0015) |
β0(2) | −0.50 | −0.5000(0.0023) | −0.4969(0.0011) | −0.4995(0.0006) |
Tv/n | 750 | 1500 | 3000 | |
p11 | 0.95 | 0.9483(0.0114) | 0.9494(0.0082) | 0.9506(0.0028) |
p22 | 0.15 | 0.1476(0.0089) | 0.1489(0.0043) | 0.1496(0.0017) |
α0(1) | 1.00 | 0.9835(0.0290) | 0.9950(0.0158) | 1.0011(0.0073) |
α0(2) | 1.50 | 1.4887(0.0360) | 1.4910(0.0167) | 1.4978(0.0085) |
α1(1) | −0.45 | −0.4532(0.0029) | −0.4493(0.0014) | −0.4510(0.0007) |
α1(2) | 0.25 | 0.2520(0.0028) | 0.2507(0.0017) | 0.2503(0.0007) |
β1(1) | 0.15 | 0.1499(0.0025) | 0.1487(0.0013) | 0.1497(0.0006) |
β1(2) | 0.55 | 0.5479(0.0033) | 0.5492(0.0015) | 0.5480(0.0007) |
β0(1) | 0.00 | 0.0038(0.0065) | −0.0004(0.0030) | −0.0023(0.0015) |
β0(2) | 0.50 | 0.5053(0.0168) | 0.5032(0.0078) | 0.4998(0.0038) |
In conclusion, this paper has introduced and thoroughly explored the MS-TSV model, a versatile class specifically designed to address asymmetry and the leverage effect in financial time series volatility. Building upon the classical threshold stochastic volatility model, the MS-TSV model incorporates a homogeneous Markov chain to parameterize log-volatility dynamics. The paper derived essential probabilistic properties of MS-TSV models, including strict stationarity, causality, ergodicity, and higher-order moments, along with providing the covariance function of the squared process. The QMLE for the MS-TSV model was introduced and its strong consistency was demonstrated through a simulation study. The MS-TSV model stands out as a robust alternative to traditional models, particularly in capturing nuanced volatility dynamics influenced by economic factors and unexpected events. This research significantly contributes to the broader understanding of modeling time series data and holds practical applications in financial analysis. However, it's crucial to recognize that ongoing research in this field, especially the exploration of multifractal processes, opens avenues for further investigation. The integration of MS-TSV models with multifractal processes represents a promising direction, offering a more nuanced perspective on volatility modeling in finance. Future work could delve into this intersection, advancing our understanding and refining tools for modeling complex financial time series data.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
The authors extend their appreciation to the Deanship of Scientific Research at King Khalid University for funding this work through the Small Groups Project under grant number R.G.P.1/138/44.
All authors declare no conflicts of interest in this paper.
[1] |
J. D. Hamilton, A new approach to the economic analysis of nonstationary time series and the business cycle, Econometrica, 57 (1989), 357–384. https://doi.org/10.2307/1912559 doi: 10.2307/1912559
![]() |
[2] |
J. D. Hamilton, Analysis of time series subject to changes in regime, J. Economet., 45 (1990), 39–70. https://doi.org/10.1016/0304-4076(90)90093-9 doi: 10.1016/0304-4076(90)90093-9
![]() |
[3] |
M. Cavicchioli, Spectral density of Markov-switching VARMA models, Economic Lett., 121 (2013), 218–220. https://doi.org/10.1016/j.econlet.2013.07.022 doi: 10.1016/j.econlet.2013.07.022
![]() |
[4] |
M. Cavicchioli, Asymptotic Fisher information matrix of Markov switching VARMA models, J. Multivariate Anal., 157 (2017b), 124–135. https://doi.org/10.1016/j.jmva.2017.03.004 doi: 10.1016/j.jmva.2017.03.004
![]() |
[5] |
M. Cavicchioli, Higher order moments of Markov switching VARMA models, Economet. Theory, 33 (2017a), 1502–1515. https://doi.org/10.1017/S0266466616000438 doi: 10.1017/S0266466616000438
![]() |
[6] |
R. Stelzer, On Markov-switching ARMA processes: Stationarity, existence of moments and geometric ergodicity, Economet. Theory, 25 (2009), 43–62. https://doi.org/10.1017/S0266466608090026 doi: 10.1017/S0266466608090026
![]() |
[7] |
M. Cavicchioli, Markov switching GARCH models: higher order moments, kurtosis measures and volatility evaluation in recessions and pandemic, J. Business Economic Statist., 40 (2022), 1772–1783. https://doi.org/10.1080/07350015.2021.1974459 doi: 10.1080/07350015.2021.1974459
![]() |
[8] |
M. Haas, S. Mittnik, M. S. Paolella, A new approach to Markov-switching GARCH models, J. Financial Economet., 2 (2004), 493–530. https://doi.org/10.1093/jjfinec/nbh020 doi: 10.1093/jjfinec/nbh020
![]() |
[9] |
A. Bibi, A. Ghezal, Minimum distance estimation of Markov-switching bilinear processes, Statistics, 50 (2016), 1290–1309. https://doi.org/10.1080/02331888.2016.1229783 doi: 10.1080/02331888.2016.1229783
![]() |
[10] |
A. Bibi, A. Ghezal, On the Markov-switching bilinear processes: stationarity, higher-order moments and β−mixing, Stochast. Int. J. Prob. Stochast. Proc., 87 (2015), 919–945. https://doi.org/10.1080/17442508.2015.1019881 doi: 10.1080/17442508.2015.1019881
![]() |
[11] |
A. Bibi, A. Ghezal, Consistency of quasi-maximum likelihood estimator for Markov-switching bilinear time series models, Stat. Prob. Lett., 100 (2015), 192–202. https://doi.org/10.1016/j.spl.2015.02.010 doi: 10.1016/j.spl.2015.02.010
![]() |
[12] | A. Ghezal, Spectral representation of Markov-switching bilinear processes, São Paulo J. Math. Sci., 2023. https://doi.org/10.1007/s40863-023-00380-w |
[13] | A. Ghezal, I. Zemmouri, The bispectral representation of Markov switching bilinear models, Commun. Faculty Sci. Uni. Ankara Series A1 Math. Statist., http://dx.doi.org/10.31801/cfsuasmas.1232916 |
[14] |
A. Bibi, A. Ghezal, Markov-switching BILINEAR-GARCH models: Structure and estimation, Commun. Stat. Theory Meth., 47 (2018), 307–323. https://doi.org/10.1080/03610926.2017.1303732 doi: 10.1080/03610926.2017.1303732
![]() |
[15] |
A. Ghezal, I. Zemmouri, Estimating MS-BLGARCH models using recursive method, Pan-Amer. J. Math., 2 (2023), 1–7. https://doi.org/10.28919/cpr-pajm/2-6 doi: 10.28919/cpr-pajm/2-6
![]() |
[16] |
A. Ghezal, A doubly Markov switching AR model: Some probabilistic properties and strong consistency, J. Math. Sci., 71 (2023), 66–75. https://doi.org/10.1007/s10958-023-06262-y doi: 10.1007/s10958-023-06262-y
![]() |
[17] | A. Ghezal, I. Zemmouri, On the Markov-switching autoregressive stochastic volatility processes, Sema J., 2023. https://doi.org/10.1007/s40324-023-00329-1 |
[18] | A. Ghezal, I. Zemmouri, On Markov-switching asymmetric logGARCH models: Stationarity and estimation, Filomat, 37 (2023), 1–19. |
[19] | A. Ghezal, I. Zemmouri, M-estimation in periodic Threshold GARCH models: Consistency and asymptotic normality, Miskolc Math. Notes, unpublished. |
[20] |
A. Ghezal, QMLE for periodic time-varying asymmetric logGARCH models, Commun. Math. Statist., 9 (2021), 273–297. https://doi.org/10.1007/s40304-019-00193-4 doi: 10.1007/s40304-019-00193-4
![]() |
[21] |
A. Diop, D. Guegan, Tail behavior of a threshold autoregressive stochastic volatility model, Extremes, 7 (2004), 367–375. https://doi.org/10.1007/s10687-004-3482-y doi: 10.1007/s10687-004-3482-y
![]() |
[22] |
X. Maoa, E. Ruiz, H. Veiga, Threshold stochastic volatility: Properties and forecasting, Int. J. Forecast., 33 (2017), 1105–1123. https://doi.org/10.1016/j.ijforecast.2017.07.001 doi: 10.1016/j.ijforecast.2017.07.001
![]() |
[23] |
M. EC. P. So, K. Lam, W. K. Li, A stochastic volatility model with Markov switching, J. Business Economic Statist., 16 (1998), 244–253. https://doi.org/10.1080/07350015.1998.10524758 doi: 10.1080/07350015.1998.10524758
![]() |
[24] | R. Casarin, Bayesian inference for generalised Markov switching stochastic volatility models, In: Conference Materials at the 4th International Workshop on Objective Bayesian Methodology, CNRS, Aussois, 2003. |
[25] |
P. Bougerol, N. Picard, Strict stationarity of generalized autoregressive processes, Ann. Prob., 20 (1992), 1714–1730. https://doi.org/10.1214/aop/1176989526 doi: 10.1214/aop/1176989526
![]() |
[26] |
C. Francq, J. M. Zakoïan, Deriving the autocovariances of powers of Markov-switching GARCH models, with applications to statistical inference, Comput. Statist. Data Anal., 52 (2008), 3027–3046. https://doi.org/10.1016/j.csda.2007.08.003 doi: 10.1016/j.csda.2007.08.003
![]() |
[27] |
M. T. Vo, Regime-switching stochastic volatility: Evidence from the crude oil market, Energy Economics, 31 (2009), 779–788. https://doi.org/10.1016/j.eneco.2009.05.001 doi: 10.1016/j.eneco.2009.05.001
![]() |
[28] |
Y. Xie, Consistency of maximum likelihood estimators for the regimes witching GARCH model, Statist. J. Theoret. Appl. Statist., 43 (2009), 153–165. https://doi.org/10.1080/02331880701442619 doi: 10.1080/02331880701442619
![]() |
[29] |
L. Calvet, A. Fisher, Forecasting multifractal volatility, J. Economet., 105 (2001), 27–58. https://doi.org/10.1016/S0304-4076(01)00069-0 doi: 10.1016/S0304-4076(01)00069-0
![]() |
[30] |
L. Calvet, A. Fisher, Regime switching and the estimation of multifractal processes, J. Financial Economet., 2 (2004), 49–83. https://doi.org/10.1093/jjfinec/nbh003 doi: 10.1093/jjfinec/nbh003
![]() |
[31] | T. Lux, The Markov-switching multifractal model of asset returns: GMM estimation and linear forecasting of volatility, J. Business Economic Statist., 26 (2008), 194–210. https://www.jstor.org/stable/27638974 |
1. | Omar Alzeley, Ahmed Ghezal, On an asymmetric multivariate stochastic difference volatility: structure and estimation, 2024, 9, 2473-6988, 18528, 10.3934/math.2024902 | |
2. | Najmeddine Attia, Ahmed Ghezal, Global stability and co-balancing numbers in a system of rational difference equations, 2024, 32, 2688-1594, 2137, 10.3934/era.2024097 | |
3. | Ahmed Ghezal, Omar Alzeley, Probabilistic properties and estimation methods for periodic threshold autoregressive stochastic volatility, 2024, 9, 2473-6988, 11805, 10.3934/math.2024578 | |
4. | Hashem Althagafi, Ahmed Ghezal, Global Stability of a System of Fuzzy Difference Equations of Higher-Order, 2024, 1598-5865, 10.1007/s12190-024-02302-1 | |
5. | Hashem Althagafi, Ahmed Ghezal, Solving a system of nonlinear difference equations with bilinear dynamics, 2024, 9, 2473-6988, 34067, 10.3934/math.20241624 |
Tv/n | 750 | 1500 | 3000 | |
p11 | 0.25 | 0.2471(0.0053) | 0.2529(0.0032) | 0.2479(0.0014) |
p22 | 0.85 | 0.8534(0.0072) | 0.8521(0.0056) | 0.8509(0.0023) |
α0(1) | 1.00 | 0.9925(0.0355) | 1.0051(0.0169) | 0.9973(0.0087) |
α0(2) | −1.50 | −1.5060(0.0175) | −1.5071(0.0083) | −1.5031(0.0038) |
α1(1) | 0.45 | 0.4457(0.0017) | 0.4493(0.0007) | 0.4487(0.0004) |
α1(2) | 0.25 | 0.2504(0.0012) | 0.2504(0.0006) | 0.2499(0.0003) |
β1(1) | −0.55 | −0.5524(0.0018) | −0.5523(0.0009) | −0.5511(0.0004) |
β1(2) | −0.25 | −0.2480(0.0013) | −0.2496(0.0005) | −0.2506(0.0003) |
β0(1) | 0.85 | 0.8555(0.0076) | 0.8570(0.0034) | 0.8513(0.0015) |
β0(2) | −0.50 | −0.5000(0.0023) | −0.4969(0.0011) | −0.4995(0.0006) |
Tv/n | 750 | 1500 | 3000 | |
p11 | 0.95 | 0.9483(0.0114) | 0.9494(0.0082) | 0.9506(0.0028) |
p22 | 0.15 | 0.1476(0.0089) | 0.1489(0.0043) | 0.1496(0.0017) |
α0(1) | 1.00 | 0.9835(0.0290) | 0.9950(0.0158) | 1.0011(0.0073) |
α0(2) | 1.50 | 1.4887(0.0360) | 1.4910(0.0167) | 1.4978(0.0085) |
α1(1) | −0.45 | −0.4532(0.0029) | −0.4493(0.0014) | −0.4510(0.0007) |
α1(2) | 0.25 | 0.2520(0.0028) | 0.2507(0.0017) | 0.2503(0.0007) |
β1(1) | 0.15 | 0.1499(0.0025) | 0.1487(0.0013) | 0.1497(0.0006) |
β1(2) | 0.55 | 0.5479(0.0033) | 0.5492(0.0015) | 0.5480(0.0007) |
β0(1) | 0.00 | 0.0038(0.0065) | −0.0004(0.0030) | −0.0023(0.0015) |
β0(2) | 0.50 | 0.5053(0.0168) | 0.5032(0.0078) | 0.4998(0.0038) |
Tv/n | 750 | 1500 | 3000 | |
p11 | 0.25 | 0.2471(0.0053) | 0.2529(0.0032) | 0.2479(0.0014) |
p22 | 0.85 | 0.8534(0.0072) | 0.8521(0.0056) | 0.8509(0.0023) |
α0(1) | 1.00 | 0.9925(0.0355) | 1.0051(0.0169) | 0.9973(0.0087) |
α0(2) | −1.50 | −1.5060(0.0175) | −1.5071(0.0083) | −1.5031(0.0038) |
α1(1) | 0.45 | 0.4457(0.0017) | 0.4493(0.0007) | 0.4487(0.0004) |
α1(2) | 0.25 | 0.2504(0.0012) | 0.2504(0.0006) | 0.2499(0.0003) |
β1(1) | −0.55 | −0.5524(0.0018) | −0.5523(0.0009) | −0.5511(0.0004) |
β1(2) | −0.25 | −0.2480(0.0013) | −0.2496(0.0005) | −0.2506(0.0003) |
β0(1) | 0.85 | 0.8555(0.0076) | 0.8570(0.0034) | 0.8513(0.0015) |
β0(2) | −0.50 | −0.5000(0.0023) | −0.4969(0.0011) | −0.4995(0.0006) |
Tv/n | 750 | 1500 | 3000 | |
p11 | 0.95 | 0.9483(0.0114) | 0.9494(0.0082) | 0.9506(0.0028) |
p22 | 0.15 | 0.1476(0.0089) | 0.1489(0.0043) | 0.1496(0.0017) |
α0(1) | 1.00 | 0.9835(0.0290) | 0.9950(0.0158) | 1.0011(0.0073) |
α0(2) | 1.50 | 1.4887(0.0360) | 1.4910(0.0167) | 1.4978(0.0085) |
α1(1) | −0.45 | −0.4532(0.0029) | −0.4493(0.0014) | −0.4510(0.0007) |
α1(2) | 0.25 | 0.2520(0.0028) | 0.2507(0.0017) | 0.2503(0.0007) |
β1(1) | 0.15 | 0.1499(0.0025) | 0.1487(0.0013) | 0.1497(0.0006) |
β1(2) | 0.55 | 0.5479(0.0033) | 0.5492(0.0015) | 0.5480(0.0007) |
β0(1) | 0.00 | 0.0038(0.0065) | −0.0004(0.0030) | −0.0023(0.0015) |
β0(2) | 0.50 | 0.5053(0.0168) | 0.5032(0.0078) | 0.4998(0.0038) |